.. . alt: . . .3. . \. am . . _ 5:1! ill.- . t. .a: 0n- ”Emu? zildfifi... ink—.13.. . 3 . .1. {up . 2:13.. 311.3 U‘?! *A--. u. - 1..Iz.n.?1... 1.. 312833.... . 5 r}... .3. :13) 1’14 alianv 1. . . v3fnrll l. . ; .11! . 95.9. (.32.... . u .1 . I , .. f . {seat .. ya,"3{s , f (‘04; ..{ 1. .0V. 4. .53 5...: “gal :24... .. 1.3:? L). ... . :3?! .1. 13.2.1 :1; :2 53:... .33: 2&0 UBRARY *7 Michigan State I University I This is to certify that the dissertation entitled CELL DEATH AND INSULIN SIGNALING IN LIVER CELLS: MOLECULAR AND SYSTEMS BIOLOGY STUDY presented by Xuerui Yang has been accepted towards fulfillment of the requirements for the PhD. degree in Chemical Engineering and Biochemistry & Molecular Biology C. y/‘l/uaC/ék, (iffy... Major Professor’s Signature j/?/09 Date MSU is an Affinnative Action/Equal Opportunity Employer 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 —AU- griffin” I ~ C 5/08 K:IProj/Aoc&Pres/ClRC/DateDue.indd CELL DEATH AND INSULIN SIGNALING IN LIVER CELLS: MOLECULAR AND SYSTEMS BIOLOGY STUDY By Xuerui Yang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemical Engineering and Biochemistry & Molecular Biology 2009 u I I ‘ g I E f ‘. a i“ ‘1 s g 1' A E I I“ .15 _‘3 i- R! 79- “I v. '1 a £3 5“ $3 ’5'; ‘3” ‘n :4 .1, ‘1 i? .h 8‘ 3: x,‘ _’;. Ei‘ 9-7 A. .; ABSTRACT CELL DEATH AND INSULIN SIGNALING IN LIVER CELLS: MOLECULAR AND SYSTEMS BIOLOGY STUDY By Xuerui Yang Liver disorders, such as hepatic insulin resistance, non-alcoholic fatty liver disease (NAFLD), and non-alcoholic steatohepatitis (NASH) are closely related with each other, and their initiation and development are attributed to the dysregulations of the hepatic cellular activities including lipotoxicity and insulin signaling. This thesis focused on the regulatory mechanisms that control these hepatocellular activities. Saturated free fatty acids (FFAs), e.g., palmitate, are known to induce lipotoxicity, which is partially due to apoptosis. Using the molecular and cellular biology techniques, novel mechanisms involved in palmitate-induced apoptosis were identified in this thesis. In brief, double-stranded RNA-dependent protein kinase (PKR) was found, for the first time, to exploit an anti-apoptotic role in human hepatocellular carcinoma cells by regulating the protein level and phosphorylation of Bcl-2. Palmitate suppresses this PKR—mediated anti-apoptotic machinery in HepGZ cells, thereby resulting in apoptosis. It is well recognized that most cellular functions are regulated by networks of genes, proteins, and other small molecules, rather than single, isolated factor(s). Given the complexity of biological systems, a system-level understanding could provide a broader view on the cellular activities and thereby complement the molecular biology studies. In collaboration with statisticians and computational experts, we developed and applied two methodologies for system- Ievel analyses of the palmitate-induced cytotoxicity. These strategies, based on the principles of systems biology, recovered gene networks that are specifically responsible for the phenotype of palmitate-induced cytotoxicity, thereby shedding light into the potential mechanisms of the phenotype. Another part of this thesis focused on the regulation of insulin signaling. This thesis identified a novel player of insulin signaling, PKR, which differentially regulates the two major insulin receptor substrates (IRS1 and IRSZ). For lRS1, PKR induces the inhibitory serine phosphorylation at 312, resulting in down- regulation of the tyrosine phosphorylation of IRS1. On the other hand, PKR regulates "282 at the transcriptional rather than the post-translational level. In summary, the hepatic insulin signaling and saturated FFA—induced cytotoxicity were investigated in this thesis. The protein PKR was found, for the first time, to be highly involved in regulating these hepatic cellular events, and detailed mechanisms were identified. In addition to the molecular and cellular biology studies, novel systems biology methodologies were developed and implemented to obtain a boarder systems level view of palmitate-induced cytotoxicity. These systems biology approaches suggested potential research targets and novel mechanisms that were supported by the literature as well as our experiments. mam: wetter Iv Y'Nr‘O‘ 'I J,” - i". I u!» .:.. Y a k D .3. ~ _ I ‘l l.- 5'9 . I1: - r‘ h.- a E a v .‘l c O szv’ i6. -3’ , ‘. I- I D I: .‘v -‘s .g. I; P 33" L“ -A .v“ m .2 I "tip-O ' INN" ',- v V J! n- O .0 he. 3| 9' 'p ' n ’1‘! a; .4 r 3 "rm! « 41- 51.“ I _ ‘ u 1 I .~ tie- DEDICATION To my parents Weihua Yang, June Zhang and my wife Xian Cao -WI¢“ i- l'u-v In...” In "I‘_.( H . h u I a r 1. u < u n . q...» n; ... ,4...“ "I I I‘m". 4.1. n .l 2‘ .. h ‘ L *4 .. a... ,_ "It“ " (.I I?" " LIL. II Pu J P'- 7 In v 2‘.“ ., ,. iv ' 7 03:13 ACKNOWLEDGEMENTS Five and half years ago, I joined Dr. Christina Chan’s group at Michigan State University, with the passion of becoming a research scientist but no idea of what I was going to experience during my PhD study. Now, I am sitting here finishing my doctoral dissertation, thinking back of the joy and the pain in the last five and half years. I have always been grateful to myself for what I chose to do five and half years ago. I feel more grateful to Dr. Chan, who is my research mentor and supported me with all the patience and kindness during the successful and frustrating times. Thanks to Dr. Chan, the dream does come true. I want to say thank you to Dr. Chan, for being such a great advisor. Thanks for the generous support and priceless academic independence you gave me, the innovative and brilliant ideas that sparkle my research, the helpful advice whenever I encountered problems during my research, and of course the pushing when I sometimes relaxed too much. You will always be my greatest advisor and the most important person in my professional career. In addition to Dr. Chan, I would also like to thank Dr. Kathleen Gallo, Dr. Dennis Miller, Dr. James Pestka, Dr. Patrick Walton, and Dr. Honggao Yan, who served on my dissertation committee. Thank you all for your valuable time and efforts in helping with my dissertation project. The discussions with all of you have always been very productive and helpful. Special thanks to Dr. Walton, who keeps a close relationship with our group and offered many suggestions and creative ideas for my research projects. I. .. ..- . . Iv»-‘W‘ lu, I4 aw- _. .. . i it - 1 «a rfifflrwll “are. » «I, Q . . I I -‘ .{Tfi‘wvm‘ I {Ott'f‘J _“V‘9:‘ -r‘ "O- _, ,I .. _‘? _‘I b I Many thanks to the members in Dr. Chan and Dr. Walton’s groups. Dr. Zheng Li and his wife, Lufang Sheng, Dr. Shireesh Srivastava and his wife, Dr. Sheenu Mittal kindly taught me many basic techniques in statistical analysis, cell culturing, Western blotting, RNA isolation, various biochemistry assays, and etc. Discussions with different group members with different specifies has been an important way of getting new knowledge and generating creative ideas. I also want to thank Michael Opperrnan, who helped doing some of the experiments. A part of my PhD work was done by collaborating with other people inside and outside our group. Special thanks to Dr. Zheng Li, Dr. Shireesh Srivastava, Dr. Xuewei Wang, Ming Wu, Dr. Rong Jin, and Yang Zhou. I enjoyed the happy and productive collaborations with all of you. Without your great contributions, a lot of excellent research works would have been very difficult to do. I would also like to thank the Department of Chemical Engineering and Materials Science, and the Department of Biochemistry & Molecular Biology. Special thanks to the Quantitative Biology & Modeling Initiative (QBMI) at MSU, for founding the Interdisciplinary PhD Program between Chemical Engineering and Biochemistry & Molecular Biology. I feel very lucky being the first PhD student enrolled in this brilliant interdisciplinary program. Last but not the least, I would like to thank my wife Xian Cao, and my parents Weihua Yang and June Zhang. Thanks to my wife, for your support, help, and love. You are the greatest thing that ever happened to me. Thanks to my parents. You have been my guide, my role model, and you will always be. Thank you all again! TABLE OF CONTENTS LIST OF TABLES ................................................................................................. ix LIST OF FIGURES ................................................................................................ x KEY TO SYMBOLS AND ABBREVIATIONS ...................................................... xii CHAPTER 1: Introduction ..................................................................................... 1 Background ................................................................................................ 1 Hepatic Lipotoxicity .................................................................................... 2 Insulin Signaling and Insulin Resistance .................................................... 9 Summary .................................................................................................. 1 1 References ............................................................................................... 1 3 CHAPTER 2: Repression of PKR Mediates Palmitate-Induced Apoptosis in HepG2 Cells through Bcl-2 ................................................................................. 23 Abstract .................................................................................................... 23 Introduction .............................................................................................. 24 Materials and Methods ............................................................................. 27 Results ..................................................................................................... 33 Discussion ................................................................................................ 54 References ............................................................................................... 60 CHAPTER 3: Construction of Phenotype-Specific Gene Network for Palmitate- lnduced Cytotoxicity by Synergy Analysis ........................................................... 68 Abstract .................................................................................................... 68 Introduction .............................................................................................. 69 Design of the Methodology ....................................................................... 73 Materials and Methods ............................................................................. 74 Results ..................................................................................................... 81 Discussion ................................................................................................ 90 References ............................................................................................... 92 CHAPTER 4: Reconstruct Modular Phenotype-Specific Gene Networks by Knowledge-Driven Matrix Factorization ............................................................. 100 Abstract .................................................................................................. 100 Introduction ............................................................................................ 101 KMF Algorithm ........................................................................................ 104 Experimental Methods ............................................................................ 1 1 1 Results and Discussion .......................................................................... 112 Conclusions ............................................................................................ 130 References ............................................................................................. 132 CHAPTER 5: PKR Differentially Regulates IRS1 and IRSZ in HepG2 Cells ..... 140 vii Abstract .................................................................................................. 140 Introduction ............................................................................................ 141 Materials and Methods ........................................................................... 144 Results ................................................................................................... 148 Discussion .............................................................................................. 165 References ............................................................................................. 1 70 CHAPTER 6: Conclusion and Discussion ......................................................... 179 References ............................................................................................. 191 viii LIST OF TABLES Table 3.1. The hub genes in the synergy network. ............................................. 83 Table 4.1. Predefined gene modules. ............................................................... 110 Table 4.2. Gene clusters identified by KMF. ..................................................... 115 Table 4.3. C Matrix of the clusters. ................................................................... 115 Table 4.4. C Matrix in toxic conditions ............................................................... 118 Table 4.5. C Matrix in nontoxic conditions ......................................................... 118 Table 4.6. Difference C Matrix. ......................................................................... 118 Table 4.7. Top 10 out of 33 genes in module 4121 Table 4.8. All 10 genes in module 6. ................................................................. 122 ix LIST OF FIGURES Figure 2.1. Effects of palmitate on the cytotoxicity and apoptosis of Hesz cells. ............................................................................................................................ 33 Figure 2.2. Effects of palmitate and oleate on the activity of PKR ....................... 34 Figure 2.3. Role of PKR in the cytotoxicity and apoptosis of HepGZ cells. ......... 36 Figure 2.4. Effects of palmitate and oleate on the protein level of Bcl-2 .............. 38 Figure 2.5. Involvement of PKR in regulating the protein level of Bcl-2 ............... 40 Figure 2.6. Role of PKR in regulating the Nuclear NF-KB p65 level. ................... 42 Figure 2.7. Effect of Palmitate on the phosphOrylation of Bcl-2 ........................... 44 Figure 2.8. Involvement of PKR in regulating the phosphorylation of Bcl-2 at Ser70. ................................................................................................................. 46 Figure 2.9. Involvement of JNK in regulating the phosphorylation of Bcl-2. ........ 50 Figure 2.10. Effect of inhibiting JNK activity on the cytotoxicity and apoptosis of HepG2 cells. ....................................................................................................... 52 Figure 2.11. Proposed signaling pathways from PKR to Bel-2 induced by palmitate. ............................................................................................................ 53 Figure 3.1. Flowchart of the Proposed Methodology ........................................... 74 Figure 3.2. Four Representative Trends of the Metabolites. ............................... 79 Figure 3.3. The Synergy Network. ...................................................................... 82 Figure 3.4. The Distribution of Shortest Path Lengths in the Synergy Network...82 Figure 3.5. The Degree Distribution of the Synergy Network. ............................. 83 Figure 3.6. P4HA1 catalyzes the formation of 4-hydroxyproline. ........................ 85 Figure 3.7. INSIG2 plays important role in lipid synthesis. .................................. 87 Figure 3.8. The Protein Domains of SH3RF2 ...................................................... 89 Figure 4.1. Sorted difference regression coefficients (DRC) for all the genes...105 Figure 4.2. Gene module interaction network. .................................................. 116 Figure 4.3. Effects of the fatty acids on the expression level of RABGGTA and the role of RABGGTA in cytotoxicity. ................................................................ 125 Figure 4.4. Effects of the fatty acids on the expression level of HSP105B. ....... 126 Figure 4.5. Effects of the fatty acids on the expression level of ATP6IP1 and the role of ATP6IP1 in cytotoxicity. ......................................................................... 128 Figure 4.6. Effect of the fatty acids on the expression level of GMPS ............... 130 Figure 5.1. Effects of ceramide and PKR inhibitors on the phosphorylation of IRS1. ................................................................................................................. 149 Figure 5.2. Involvement of PKR in regulating the phosphorylation of IRS1. ...... 150 Figure 5.3. Involvement of JNK and IKK in regulating the phosphorylation of IRS1. .......................................................................................................................... 153 Figure 5.4. Involvement of PKR in regulating IRSZ. .......................................... 154 Figure 5.5. Involvement of FoxO1 in mediating the effect of PKR on IRS2 ....... 159 Figure 5.6. Effect of insulin on the activity of PKR. ........................................... 162 Figure 5.7. Involvement of IRS1 and IRSZ in mediating the effect of insulin on the phosphorylation of PKR. ................................................................................... 163 Figure 5.8. Involvement of PP1 in mediating the effect of insulin on the phosphorylation of PKR. ................................................................................... 164 Figure 5.9. Proposed signaling pathways through which PKR is involved in insulin signaling network in HepG2 cells. .......................................................... 165 xi ASKl ATPGIPI Bel-2 Bcl-X(L) Birn BSA CRAB dsRNA ECM eIF-Za ER ERK ETC FFA FoxO GATA4 GMPS GTP HCV KEY TO SYMBOLS AND ABBREVIATIONS apoptosis signal-regulating kinase 1 ATPase, H+ transporting, Iysosomal interacting protein 1 B-cell leukemia/lymphoma 2 Bcl-2-like l Bcl-2-interacting mediator of cell death bovine serum albumin heart and neural crest derivatives expressed 2 CAMP response element-binding protein double-stranded RNA extracellular matrix eukaryotic initiation factor 2-a1pha endoplasmic reticulum extracellular-signal-regulated kinase electron transport chain free fatty acid Forkhead box-containing protein, class 0 GATA binding protein 4 guanine monophosphate synthetase gene ontology guanine triphosphate hepatitis C virus xii HepGZ HSPIOSB Hsp70 IGF-l IKK IR IRS LDH MAPK Mda7 NAFLD NASH PCR PKR PP 1 PP 1 c PP2A RABGGTA RAX/PACT RNAi hepatocellular carcinoma heat shock lOSkD 70-kDa heat shock protein insulin-like growth factor 1 I-kappaB kinase insulin receptor insulin receptor substrate c-Jun N-terminal kinase knowledge-driven matrix factorization lactate dehydrogenase mitogen-activated protein kinase melanoma differentiation-associated gene-7 non alcoholic fatty liver disease non-alcoholic steatohepatitis nuclear factor kappa B polymerase chain reaction double-stranded RNA-dependent protein kinase protein phosphatase 1 catalytic domain of PP1 protein phosphatase 2A Rab geranylgeranyltransferase PKR activator X RNA interference xiii ROS SCAP SH3 siRNA SREBPS TCA TG TNF-a UCP UPR reactive oxygen species SREBP Cleavage-Activating Protein Src homology 3 small interfering RNA sterol regulatory element binding proteins tricarboxylic acid tn'acyl glycerol Tumor necrosis factor -alpha uncoupling proteins unfolded protein response xiv tor. . fl - . ‘F . Q . I- 'l ,.. C ‘O ,.. .. .- .. I-x. . r.- r... I. - .'. .. ,..- .. . . . . D'. .- O >-.' i I} 40 4"— ’3. W ~. ". . “F l. rIn'nJ "' "Y'TTJT 7"! Tim g‘tiign‘ :17“.- Dig {A} ~ .= miter-mime. ,‘ A.“ all?” _. r. 0" no I l CHAPTER 1: Introduction Background As one of the vital organs in vertebrates, the liver plays a major role in metabolism (i.e. carbohydrates, lipids, amino acids, and proteins), thereby functioning in glycogen storage, plasma protein synthesis, hormone production, and detoxification (1). Primary hepatocytes make up 70-80% of the cytoplasmic mass of the liver, and perform most of the functions of the liver (1). Dysregulation of the hepatocyte function can lead to liver diseases, e. g., non-alcoholic fatty liver disease (NAFLD), which consists of different stages: steatosis and non—alcoholic steatohepatitis (NASH), and, at its most severe stages, may progress to cirrhosis and liver failure (2). NAFLD is the most common liver disease in the United States. It is believed that over next 20 years, NAFLD and NASH will surpass hepatitis C as the leading cause for liver transplantation in the United States (3). The exact cause of NAFLD is still unknown. However, it has been recognized that excess circulating free fatty acids (F FAs), in particular long-chain saturated fatty acids, e. g. palmitate, play significant roles in the development and progression of NAFLD by inducing hepatic lipotoxicity (4, 5). Hepatic lipid metabolism is highly regulated by insulin, through the insulin signaling pathway (6). Disruption of insulin signaling, i.e., insulin resistance, is common in patients with NAFLD (7), and is believed to play an important role in the pathogenesis of NAFLD by impairing mitochondrial F FA B-oxidation and promoting fat accumulation in the liver (8). Insulin signaling is a central signaling pathway that regulates many cellular activities, such as glucose and lipid metabolism, protein synthesis and degradation, cell growth and differentiation (6). Insulin resistance leads to dysregulation of glucose homeostasis and lipid metabolism in hepatocytes (9), thereby playing a major role in the development of type 2 diabetes as well as NAFLD. Type 2 diabetes is characterized by high blood glucose in response to insulin resistance as well as insulin deficiency (10). According to the NIH, in 2007 up to 23.6 million people (7.8% of the population) have diabetes, 90-95% of which are type 2 diabetes (11). In this thesis, my research was focused on the mechanisms of hepatic lipotoxicity and the regulation of insulin signaling in HepGZ cells, a study model of the hepatocytes. Hepatic Lipotoxicity Chapters 2 to 4 discussed the investigations on hepatic lipotoxicity. Chronic treatments of elevated FFAs induce cytotoxicity in a variety of cells, including skeletal muscle myotubes (12), cardiomyocytes (13), pancreatic cells (14, 15), and hepatocytes (I6, 17). Previous studies have identified that saturated FFAs, e. g. palmitate, are much more toxic to hepatocytes and hepatoma cells than unsaturated FFAs, e. g. oleate (l 7, 18). Over the past decades, researchers have put much effort in uncovering the mechanisms of saturated FFAs induce cytotoxicity, which is believed to largely relate to apoptosis (18- 20). Apoptosis, programmed cell death, is one of the major cellular activities involved in metazoan development, inflammatory responses, tumor growth control, and many other processes (21, 22). Saturated FFAs, e. g. palmitate, induce apoptosis in many cell types, such as cardiac cells, pancreatic beta cells, breast cancer cells, and hepatocytes. The \IF; ' a ‘l 'f.‘ 02-. H. 1),. 3: v... 7o. 3‘! -;.4‘ '7} 8.- 1- .1‘ “I r I win“! fulll'l‘ 9 II “I 1.9".‘1‘I "r t“. 'I,‘ are . , We .0 A «a 'ttr-x r: Wklhflfkixtli IO.’ ' . 'l'l'v WW . 0' mechanism by which palmitate induces apoptosis in different types of cells has been under intense investigation. Studies of palmitate-induced apoptosis in liver cells have focused predominantly on Iysosomal permeabilization (23), ROS production (24), and intracellular metabolic pathways such as beta oxidation (25), TC accumulation (26, 27), and ceramide production (1 8, 28). More recent studies revealed the involvement of certain Bel-2 family proteins in mediating saturated FFA-induced apoptosis of liver cells (19, 20, 23). The Bel-2 family proteins play critical roles in the intrinsic and extrinsic apoptosis pathways by regulating the release of apoptogenic molecules from the mitochondria to the cytosol or the activities of caspases (29, 30). It has been proposed that palmitate-induced apoptosis in liver cells is related to some of the Bcl-2 family proteins, such as Bax and the Bax antagonist Bcl-X(L) (19, 23), and Bcl-2-interacting mediator of cell death (Bim), a pro- apoptotic Bel-2 family protein (19, 20). In addition to these pro-apoptotic Bel-2 family proteins, an anti-apoptotic member, Bel-2, has been identified as an important factor in regulating the apoptosis of HepGZ cells (31-33). As one of the most studied and important anti-apoptotic members of the Bcl-2 family, Bel-2 was found to protect cells against intrinsic (mitochondria) apoptosis by maintaining the integrity of the mitochondrial membrane (29). In Chapter 2 of this thesis, the potential effects of FFAs, palmitate and oleate, on the Bcl-2 protein in HepG2 cells were investigated. The findings suggest a negative effect of palmitate on the protein level of Bcl-2 in HepGZ cells, and further show that palmitate also down-regulates the phosphorylation of Bel-2 at Ser70, which also plays a role in determining the anti-apoptotic function of Bel-2 (34). Furthermore, 2 signal transduction pathways were identified, through which palmitate exploits its regulatory effect on Bel-2. The double-stranded RNA-dependent protein kinase (PKR) was found to be central in mediating these 2 pathways. Initially identified as an anti-viral protein, PKR is best known for triggering cell defense responses and initiating innate immune responses by arresting general protein synthesis and inducing apoptosis during virus infection (35). Foreign double-stranded RNA (dsRNA), a by—product of viral RNA polymerases during virus replication, binds PKR, facilitating the homo-dimerization and auto-phosphorylation at Thr451 and Thr446 and thereby the activation of PKR (36, 37). Independent of dsRNA, PKR can also be activated by other cellular proteins such as RAX/PACT (38, 39) and Mda7 (40), thereby mediating stresses signals induced by serum starvation, H202, or ceramide (38, 39, 41). The activated PKR, known as a eukaryotic initiation factor 2-alpha (eIF-2a) kinase, induces the phosphorylation of eIF-2a at SerSl (3 7), which inhibits the initiation of RNA translation by the tRNA 4OS ribosomal subunit. This function of PKR results in the inhibition of general protein synthesis and the induction of cellular apoptosis in many types of eukaryotic cells (37, 42). However, evidence is emerging, albeit controversial‘at this point, suggesting that PKR also has an anti-apoptotic role in mouse embryo fibroblasts and certain tumor cells (43-47). In this thesis, an anti-apoptotic role of PKR in HepG2 cells was identified, mediated through Bcl-2, an anti-apoptotic protein. The findings further showed that palmitate suppresses the phosphorylation of PKR, which serves as a mechanism of palmitate-induced apoptosis. Thus, in Chapter 2 using research tools from molecular and cellular biology, I successfully uncovered a mechanism by which palmitate induces apoptosis and cytotoxicity in liver cells. A number of mechanisms have already been proposed, in which many important genes were suggested to play roles in saturated F FA-induced cell death. They include Iysosomal permeabilization (23), ROS production (24), and altered intracellular metabolic pathways (18, 25-28). Considering these multi-faceted effects, it is unlikely that hepatic lipotoxicity can be attributed to a single or a few genes functioning on an isolated cellular activity. For example, cytotoxicity, as well as apoptosis, were shown to be related to ROS production, but they were not completely prevented upon treatment with mitochondrial complex inhibitors or fi‘ee radical scavengers, suggesting that mechanisms other than ROS production contributed to the toxicity of palnritate (24). It is now well-recognized that most cellular fimctiorrs are regulated by complex networks of genes, proteins, and small molecules, including metabolites and signaling molecules (48). Focusing on a limited number of proteins/ genes, molecular biology has made great strides in elucidating detailed mechanistic information of cellular activities. Given the complexity of biological systems, a system-level understanding could complement and help unravel the complex interactions in biological systems and thereby provide a broader view on the cellular activities (48). Recent development in biotechnology has made it possible to systematically measure profiles of genes, proteins and metabolites of cells, enabling system-level analyses of cellular activities facilitated by statistical or mathematical tools and grounded in molecular-level understanding (49). With the technique of cDNA microarray, our group has obtained comprehensive gene-expression profiles of HepGZ cells in response to different FFA exposures (50). To systematically study the overall effect of FFAs on liver cells, in Chapters 3 and 4, I, together with my colleague-mathematicians and computational experts, evaluated methodologies that could extract mechanistic information from gene-expression profiles. Assuming that saturated FFA-induced cytotoxicity is coordinately regulated by a set of genes that interact in a complex network, we developed methods to 1) select a pool of the genes that are potentially relevant to the phenotype of interest, i.e. saturated F FA-induced lipotoxicity, and 2) reconstruct gene networks, with the selected genes, that give rise to the observed phenotype. These gene networks should only recover the interactions that are related to the phenotype of interest, herein denoted as phenotype-specific gene networks. By reconstructing phenotype-specific gene networks we hypothesize that a broader view of cellular activities could provide novel insights into the mechanisms involved, and ultimately, in the mechanisms at play in diseases (51, 52). Selection of the Genes——Selection of the genes relevant to a phenotype is usually viewed as a feature selection problem (53, 54) and have been based predominantly on statistical tests (55) or correlation measures (56), which were used to assess the statistical differences among the gene profiles and the correlations between the genes and the phenotype, respectively. These methods, purely based on microarray data, are susceptible to the noise level in the data and usually computationally expensive. In addition, they are likely to miss some important genes that are not differentially expressed. One strategy to ameliorate these issues is to incorporate prior knowledge, such as domain knowledge and functional (e. g. gene ontology (60)) information of the genes, in the gene selection process (57). In Chapter 4 of this thesis, a Bayesian mixture regression model was developed to quantitatively incorporates the prior knowledge of the gene functions, upfront, in the process of gene-selection (58). This method is more efficient in incorporating the prior knowledge, as compared to the typical knowledge-based methods, in which the prior information is incorporated in the post-processing of the genes that are selected by the data-driven approaches. In addition to the Bayesian mixture regression model, an alternative strategy of gene-selection was also evaluated in Chapter 3. This strategy integrated multiple levels of information, i.e., gene expression and metabolite profiles, in the process of gene selection. It is known that some alterations in metabolism are involved in the induction of cytotoxicity by saturated FFAs (26, 59, 60). Therefore, this approach better reflected the “multi-level” characteristic of cellular activities and took advantage of the additional accessible biological information in the selection of genes that are involved in the phenotype. Reconstruction of the Gene Network—Various methods, such as correlation (61), mutual information (62, 63), and Bayesian network analysis (64), have been used to assess gene interactions and reconstruct gene networks. These methods do not directly incorporate the phenotype information in identifying the gene interactions. Instead, gene networks were built for each of the conditions, and compared across the conditions to identify the gene interactions that are specific to a condition or a phenotype. Consequently, these methods are computationally expensive and sensitive to the quality of data To address this short-coming, Chapter 3 introduced an alternative method, which identifies the synergistic gene pairs that cooperatively function on a phenotype, thereby avoiding comparison of the networks under different conditions. This method is based on the concept of synergy. Synergy is defined as the “additional” contribution provided by the ‘fivhole” as compared to the sum of the contributions of the individual “parts” (65, 66). Biological activities are regulated by multiple factors, many of which function cooperatively, i.e., synergistically. The basic idea is that the whole (i.e., the regulatory system) is greater than the sum of the individual parts (i.e., regulators) of a system (67). We assess the synergistic effects between gene pairs on the phenotype, namely palmitate- induced cytotoxicity, and with the identified synergistic gene pairs, built a synergy network that was phenotype-specific. The topological analyses revealed the structural characteristics of the network while the hub genes provided insights into potential mechanism(s) involved in the induction of the phenotype. Given that biological networks, e.g., protein-protein interaction network (68), metabolic network (69) and transcriptional regulation network (70), are modular in structure, identifying the gene modules and their interplay in a modular network should provide insights into the differential mechanisms involved in different cellular states, i.e., normal vs. disorder. Several clustering methods, such as self-organizing mapping (71, 72), hierarchical clustering (61) and K-means (73), have been commonly used to identify gene modules. However, most of these approaches cannot uncover the interactions among the modules or clusters. To address this limitation, several studies integrated clustering methods with structure learning algorithms, such as Graphical Gaussian Modeling and Bayesian network learning (74-76). These approaches are predominantly data-driven and thus susceptible to noise in the expression data, and suffer from the sparse data problem associated with limited number of experimental conditions (77, 78). Prior knowledge has been shown to be helpful in reconstructing modular networks with sparse and noisy expression data (70, 79-84). To exploit the prior knowledge and reconstruct modular gene networks, we developed a framework based on knowledge-driven matrix factorization, termed KMF, in Chapter 4. This KMF fiamework efficiently incorporates the prior knowledge of co-regulation relationships into the network reconstruction using a regularization scheme and successfully derives both gene modules and their interaction simultaneously, while the other matrix factorization methods can only identify gene clusters. Supported by the literature, these modules and their interactions potentially may be important in saturated-FFA induced cytotoxicity. In addition, evaluating the contribution of the genes within the highly relevant modules revealed potential genes that may be involved in palmitate-induced cytotoxicity. Further experiments in Chapter 4 confirmed the involvement of these genes in conferring a cytotoxic phenotype and suggested novel research targets for addressing the cytotoxicity. In summary, in this thesis saturated FFA-induced hepatic cytotoxicity was investigated using 2 different types of approaches, experimental and computational. In Chapter 2, with the help of molecular and cellular biology techniques, 2 pathways through which PKR mediates the regulation of the protein level and the phosphorylation status of Bcl-2 were identified, providing a novel mechanism by which palmitate induced apoptosis in HepG2 cells. On the other hand, fundamental ideas in systems biology were applied in Chapters 3 and 4, and 2 strategies were developed, to reconstruct phenotype- specific gene networks, one based on synergy analysis between gene pairs (Chapter 3) and the other one based on gene clustering (Chapter 4). Both of these systematic strategies were able to provide a systematic view and capture certain features of the phenotype, further shedding light into potential mechanisms of lipotoxicity. Insulin Signaling and Insulin Resistance Chapter 5 discussed my investigations on insulin signaling. As a central signaling pathway that regulates many cellular activities including glucose and lipid metabolism, . .. . ..... . r .. ‘ I - If." v r‘ ,n s I.-' ' . ,.. - ,.. .r "fl-Itch“ ‘.'£’l‘i . I' (I I we" IV... .1 rig 13‘1;15;6;‘ if»; ’v' s . :- kiln-s;- . I .‘ '1" ”I’J'iy' . mlilllz‘u protein synthesis and degradation, cell growth and differentiation (6), insulin signaling is initiated upon binding of insulin to the insulin receptor (IR), a receptor tyrosine kinase (85), and transmitted intracellularly by the insulin receptor substrates (IRS) (6, 85). At least 4 of the 1R substrates belong to the IRS group, with R81 and IRSZ being predominant and expressed in most tissues, including the liver (86, 87). Upon phosphorylation of the tyrosine residues catalyzed by IR, the IRS proteins initiate, through different binding mechanisms (87), various downstream signal transduction cascades, including mitogen-activated protein kinase (MAPK) pathways (c-Jun N- terminal kinase (JNK), extracellular signal-related kinase (ERK), and p38) (88, 89) and phosphoinositide 3-kinase (PI3K) (90), which in turn activates Akt/protein kinase B (Akt/PKB) (91), and atypical protein kinase C (aPKC) (92). Insulin signaling is sophisticatedly tuned by a large number of regulators. Pathologically, dysregulation of insulin signaling has been linked to multiple diseases and disorders. In particular, disruption of insulin signaling leads to insulin resistance, a pathological state in which the target cells fail to respond to physiological levels of insulin (93). In combination with the failure of B cells to compensate, insulin resistance directly contributes to the development of type 2 diabetes (94, 95). Insulin resistance is also related to multiple other metabolic, endocrine and cardiovascular disorders (96, 97). At the molecular level, dysregulation of insulin signaling or insulin resistance could occur at several possible stages, e.g., degradation or mutation of IR (98, 99), inhibitory phosphorylation or degradation of IRS (100, 101), or suppression of down-stream signaling molecules, such as PI-3 kinase or Alct/PKB (reviewed in ref. (102)). However, most of the insulin signaling disruption and insulin resistance have been attributed to 10 dysregulation of the phosphorylation of IR and, in particular, IRS (100, 101). IRS proteins mediate intracellular insulin signaling, through the tyrosine residues, which facilitate recruitment of IRS substrates and promote insulin signaling (87), and the serine residues, which generally suppress the activities of IRS by blocking the interaction between IRS and IR (103), inhibiting the tyrosine phosphorylation of IRS (104), or inducing the degradation of IRS (105). A number of serine residues have been identified to negatively regulate the activity of IRS], in particular, Ser307 (equivalent to Ser312 in human IRS 1 ). Ser307 has been extensively investigated and characterized as a key indicator of inhibitory phosphorylation of R81 and insulin resistance, and confirmed in insulin-resistant rodent models (106). In Chapter 5, the potential involvement of PKR in regulating the phosphorylation of IRSl was investigated. PKR was found to up-regulate the inhibitory phosphorylation of IRSl at Ser312, which is believed to suppress insulin signaling and induce insulin resistance. In addition, further investigation shoWed that, IRSZ, another predominant IRS family protein in the liver, is also affected by PKR, through a different mechanism. In addition, we as well as others showed that PKR activity is down-regulated by insulin through the IRS proteins (107). Therefore this feedback loop between PKR and IRS proteins may have potential implications in the regulation of insulin signaling and the development of insulin resistance. Summary Liver disorders such as NAFLD and hepatic insulin resistance are correlated with each other and involve major cellular activities such as lipotoxicity and insulin signaling 11 in liver cells. In this thesis, my research focused on the regulatory mechanisms that control these hepatic cellular activities. Using molecular and cellular biology techniques, mechanisms involved in palmitate-induced apoptosis and the regulation of insulin signaling through IRS proteins were identified. In addition, in collaboration with other colleagues inside and outside our research group, we applied novel systems biology methodologies to provide potential insights into lipotoxicity. These methods could be easily adapted to other complex biological systems, thereby serving as powerful research tools in extracting mechanistic information and identifying potential research and drug targets. 12 10. 11. References Davey, B., Halliday, T., and Hirst, M. (2001) Human biology and health : an evolutionary approach, Open University, Buckingham [UK] ; Phildelphia Clark, J. M., and Diehl, A. M. (2003) Nonalcoholic fatty liver disease: an underrecognized cause of cryptogenic cirrhosis. JAMA 289, 3000-3004 Contos, M. J ., and Sanyal, A. J. (2002) The clirricopathologic spectrum and management of nonalcoholic fatty liver disease. Adv Anat Pathol 9, 37-51 Feldstein, A. B., Wemeburg, N. W., Canbay, A., Guicciardi, M. B., Bronk, S. F., Rydzewski, R., Burgart, L. J., and Gores, G. J. (2004) Free fatty acids promote hepatic lipotoxicity by stimulating TNF-alpha expression via a lysosomal pathway. Hepatology 40, 185-194 Wang, D., Wei, Y., and Pagliassotti, M. J. (2006) Saturated fatty acids promote endoplasmic reticulum stress and liver injury in rats with hepatic steatosis. Endocrinology 147, 943-951 Saltiel, A. R., and Kahn, C. R. (2001) Insulin signalling and the regulation of glucose and lipid metabolism. Nature 414, 799-806 Hamaguchi, M., Kojima, T., Takeda, N., Nakagawa, T., Taniguchi, H., Fujii, K., Omatsu, T., Nakajima, T., Sarui, H., Shimazaki, M., Kato, T., Okuda, J ., and Ida, K. (2005) The metabolic syndrome as a predictor of nonalcoholic fatty liver disease. Ann Intern Med 143, 722-728 Salt, W. B., 2nd (2004) Nonalcoholic fatty liver disease (NAFLD): a comprehensive review. J Insur Med 36, 27-41 Klover, P. J ., and Mooney, R. A. (2004) Hepatocytes: critical for glucose homeostasis. Int J Biochem Cell Biol 36, 753-758 Kumar, V., Abbas, A. K., Fausto, N., Robbins, S. L., and Cotran, R. S. (2005) Robbins and Cotran pathologic basis of disease, Elsevier Saunders, Philadelphia Rena, 6., Woods, Y. L., Prescott, A. R., Peggie, M., Unterman, T. G., Williams, M. R., and Cohen, P. (2002) Two novel phosphorylation sites on FKHR that are critical for its nuclear exclusion. EMBO J 21, 2263-2271 13 12. l3. 14. 15. 16. 17. l8. 19. 20. 21. 22. 23. Turpin, S. M., Lancaster, G. I., Darby, I., Febbraio, M. A., and Watt, M. J. (2006) Ap0ptosis in skeletal muscle myotubes is induced by ceramides and is positively related to insulin resistance. Am J Physiol Endocrinol Metab 291, E1341-1350 Hickson-Bick, D. L., Sparagna, G. C., Buja, L. M., and McMillin, J. B. (2002) Palmitate-induced apoptosis in neonatal cardiomyocytes is not dependent on the generation of ROS. Am J Physiol Heart Circ Physiol 282, H656-664 Landau, Z., Forti, B., Alcaly, M., and Birk, R. Z. (2006) Palmitate induced lipoapoptosis of exocrine pancreas AR42J cells. Apoptosis 11, 717-724 Shimabukuro, M., Zhou, Y. T., Levi, M., and Unger, R. H. (1998) Fatty acid- induced beta cell apoptosis: a link between obesity and diabetes. Proc Natl Acad Sci U S A 95, 2498-2502 Ji, J., Zhang, L., Wang, P., Mu, Y. M., Zhu, X. Y., Wu, Y. Y., Yu, H., Zhang, B., Chen, S. M., and Sun, X. Z. (2005) Saturated free fatty acid, palmitic acid, induces apoptosis in fetal hepatocytes in culture. Exp T oxicol Pathol 56, 369-376 Srivastava, S., and Chan, C. (2007) Hydrogen peroxide and hydroxyl radicals mediate palmitate-induced cytotoxicity to hepatoma cells: relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 Wei, Y., Wang, D., Topczewski, F., and Pagliassotti, M. J. (2006) Saturated fatty acids induce endoplasmic reticulum stress and apoptosis independently of ceramide in liver cells. Am J Physiol Endocrinol Metab 291, E275-281 Barreyro, F. J., Kobayashi, S., Bronk, S. F., Wemeburg, N. W., Malhi, H., and Gores, G. J. (2007) Transcriptional Regulation of Birn by FoxO3A Mediates Hepatocyte Lipoapoptosis. J Biol Chem 282, 27141-27154 Malhi, H., Bronk, S. F., Wemeburg, N. W., and Gores, G. J. (2006) Free fatty acids induce JNK-dependent hepatocyte lipoapoptosis. J Biol Chem 281, 12093-12101 Lawen, A. (2003) Apoptosis-an introduction. Bioessays 25, 888-896 Elmore, S. (2007) Apoptosis: a review of programmed cell death. T oxicol Pathol 35, 495-5 1 6 Feldstein, A. B., Wemeburg, N. W., Li, Z., Bronk, S. F., and Gores, G. J. (2006) Bax inhibition protects against free fatty acid-induced Iysosomal permeabilization. Am J Physiol Gastrointest Liver Physiol 290, G1 339-1 346 14 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Srivastava, S., and Chan, C. (2006) Hydrogen peroxide and hydroxyl radicals mediate palrrritate-induced cytotoxicity to hepatoma cells: Relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 Giordano, A., Calvani, M., Petillo, O., Grippo, P., Tuccillo, F., Melone, M. A., Bonelli, P., Calarco, A., and Peluso, G. (2005) tBid induces alterations of mitochondrial fatty acid oxidation flux by malonyl-CoA-independent inhibition of carnitine palmitoyltransferase-l. Cell Death Dijfir 12, 603-613 Listenberger, L. L., Han, X., Lewis, S. B., Cases, S., Farese, R. V., Jr., Ory, D. S., and Schaffer, J. E. (2003) Triglyceride accumulation protects against fatty acid- induced lipotoxicity. Proc Natl Acad Sci U S A 100, 3077-3082 Li, Z., Srivastava, S., Yang, X., Mittal, 8., Norton, P., Resau, J ., Haab, B., and Chan, C. (2007) A hierarchical approach employing metabolic and gene expression profiles to identify the pathways that confer cytotoxicity in HepG2 cells. BMC Syst Biol 1, 21 Assimacopoulos-Jeannet, F. (2004) Fat storage in pancreas and in insulin-sensitive tissues in pathogenesis of type 2 diabetes. Int J Obes Relat Metab Disord 28 Suppl 4, $53-57 Breckenridge, D. G., and Xue, D. (2004) Regulation of mitochondrial membrane permeabilization by BCL-2 family proteins and caspases. Curr Opin Cell Biol 16, 647-652 Li, H., Zhu, H., Xu, C. J., and Yuan, J. (1998) Cleavage of BID by caspase 8 mediates the mitochondrial damage in the Fas pathway of apoptosis. Cell 94, 491- 501 Feng, L. F., Zhong, M., Lei, X. Y., Zhu, B. Y., Tang, S. S., and Liao, D. F. (2006) Bcl-2 siRN A induced apoptosis and increased sensitivity to 5-fluorouracil and HCPT in HepG2 cells. J Drug Target 14, 21-26 Handayani, T., Sakinah, S., Nallappan, M., and Pihie, A. H. (2007) Regulation of p53-, Bcl-2- and caspase-dependent signaling pathway in xanthorrhizol-induced apoptosis of HepG2 hepatoma cells. Anticancer Res 27, 965-971 Huang, D. S., Shen, K. 2., Wei, J. F., Liang, T. B., Zheng, S. S., and Xie, H. Y. (2005) Specific COX-2 inhibitor NS398 induces apoptosis in human liver cancer cell line HepG2 through BCL-2. World J Gastroenterol 11, 204-207 15 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. Blagosklonny, M. V. (2001) Unwinding the loop of Bel-2 phosphorylation. Leukemia 15, 869-874 Barber, G. N. (2005) The dsRNA-dependent protein kinase, PKR and cell death. Cell Death Difler 12, 563-570 Zhang, R, Romano, P. R., Nagamura-Inoue, T., Tian, B., Dever, T. B., Mathews, M. B., Ozato, K., and Hinnebusch, A. G. (2001) Binding of double-stranded RNA to protein kinase PKR is required for dimerization and promotes critical autophosphorylation events in the activation loop. J Biol Chem 276, 24946-24958 Taylor, S. S., Haste, N. M., and Ghosh, G. (2005) PKR and eIF2alpha: integration of kinase dimerization, activation, and substrate docking. Cell 122, 823-825 Patel, C. V., Handy, 1., Goldsmith, T., and Patel, R. C. (2000) PACT, a stress- modulated cellular activator of interferon-induced double-stranded RNA-activated protein kinase, PKR. J Biol Chem 275, 37993-37998 Ito, T., Yang, M., and May, W. S. (1999) RAX, a cellular activator for double- stranded RNA-dependent protein kinase during stress signaling. J Biol Chem 274, 15427-15432 Pataer, A., Vorburger, S. A., Chada, S., Balachandran, 8., Barber, G. N., Roth, J. A., Hunt, K. K., and Swisher, S. G. (2005) Melanoma difi‘erentiation-associated gene-7 protein physically associates with the double-stranded RNA-activated protein kinase PKR. MoI Ther 11, 717-723 Ruvolo, P. P., Gao, F., Blalock, W. L., Deng, X., and May, W. S'. (2001) Cerarnide regulates protein synthesis by a novel mechanism involving the cellular PKR activator RAX. J Biol Chem 276, 11754-11758 Stark, G. R., Kerr, 1. M., Williams, B. R., Silverman, R. H., and Schreiber, R. D. (1998) How cells respond to interferons. Annu Rev Biochem 67, 227-264 Donze, O., Deng, J ., Curran, J ., Sladek, R., Picard, D., and Sonenberg, N. (2004) The protein kinase PKR: a molecular clock that sequentially activates survival and death programs. EMBO J 23, 564-571 Kim, S. H., Forrnan, A. P., Mathews, M. B., and Gunnery, S. (2000) Human breast cancer cells contain elevated levels and activity of the protein kinase, PKR. Oncogene 19, 3086-3094 16 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. Kim, S. H., Gunnery, S., Choc, J. K., and Mathews, M. B. (2002) Neoplastic progression in melanoma and colon cancer is associated with increased expression and activity of the interferon-inducible protein kinase, PKR. Oncogene 21, 8741- 8748 Hiasa, Y., Kamegaya, Y., Nuriya, H., Onji, M., Kohara, M., Schmidt, E. V., and Chung, R. T. (2003) Protein kinase R is increased and is functional in hepatitis C virus-related hepatocellular carcinoma. Am J Gastroenterol 98, 2528-2534 Delgado Andre, N., and De Lucca, F. L. (2007) Knockdown of PKR expression by RNAi reduces pulmonary metastatic potential of B16-F10 melanoma cells in mice: Possible role of NF-kappaB. Cancer Lett 258, 118-125 Kitano, H. (2002) Computational systems biology. Nature 420, 206-210 Kitano, H. (2002) Systems biology: a brief overview. Science 295, 1662-1664 Li, Z., Srivastava, S., Mittal, S., Yang, X., Sheng, L., and Chan, C. (2007) A Three Stage Integrative Pathway Search (TIPS) framework to identify toxicity relevant genes and pathways. BMC Bioinformatics 8, 202 Said, M. R., Begley, T. J ., Oppenheim, A. V., Lauffenburger, D. A., and Samson, L. D. (2004) Global network analysis of phenotypic effects: protein networks and toxicity modulation in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 101, 18006-1801 1 Srivastava, S., Li, Z., Yang, X., Yedwabnick, M., Shaw, 8., and Chan, C. (2007) Identification of genes that regulate multiple cellular processes/responses in the context of lipotoxicity to hepatoma cells. BMC Genomics 8, 364 Ressom, H. W., Varghese, R. S., Zhang, Z., Xuan, J., and Clarke, R. (2008) Classification algorithms for phenotype prediction in genomics and proteorrrics. Front Biosci 13, 691-708 Saeys, Y., Inza, 1., and Larranaga, P. (2007) A review of feature selection techniques in bioinforrnatics. Bioinformatics 23, 2507-2517 Tusher, V. G., Tibshirani, R., and Chu, G. (2001) Significance analysis of microarrays applied to the ionizing radiation response. Proc Natl Acad Sci U S A 98, 51 16-5121 Li, Z., and Chan, C. (2004) Integrating gene expression and metabolic profiles. J Biol Chem 279, 27124-27137 17 57. 58. 59. 60. 61. 62. 63. 65. 66. 67. 68. Le Phillip, P., Bah], A., and Ungar, L. H. (2004) Using prior knowledge to improve genetic network reconstruction from rrricroarray data. In Silico Biol 4, 335-353 Srivastava, S., Zhang, L., Jin, R., and Chan, C. (2008) A novel method incorporating gene ontology information for unsupervised clustering and feature selection. PLoS ONE 3, e3 860 Lam, T. K., Carpentier, A., Lewis, G. F., van de Werve, G., Fantus, I. G., and Giacca, A. (2003) Mechanisms of the free fatty acid-induced increase in hepatic glucose production. Am J Physiol Endocrinol Metab 284, E863-873 Srivastava, S., and Chan, C. (2008) Application of metabolic flux analysis to identify the mechanisms of free fatty acid toxicity to human hepatoma cell line. Biotechnol Bioeng 99, 399-410 Eisen, M. B., Spellman, P. T., Brown, P. O., and Botsteirr, D. (1998) Cluster analysis and display of genome-wide expression patterns. Proc Natl Acad Sci U S A 95, 14863-14868 Liang, K. C., and Wang, X. (2008) Gene regulatory network reconstruction using conditional mutual information. E URASIP journal on bioinformatics & systems biology, 253894 Basso, K., Margolin, A. A., Stolovitzky, G., Klein, U., Dalla-Favera, R., and Califano, A. (2005) Reverse engineering of regulatory networks in human B cells. Nature genetics 37, 382-390 Pe'er, D., Regev, A., Elidan, G., and Friedman, N. (2001) Inferring subnetworks from perturbed expression profiles. Bioinformatics 17 Suppl 1, 8215-224 Varadan, V., and Anastassiou, D. (2006) Inference of disease-related molecular logic from systerns-based microarray analysis. PLoS Comput Biol 2, e68 Anastassiou, D. (2007) Computational analysis of the synergy among multiple interacting genes. Mol Syst Biol 3, 83 Varadan, V., Miller, D. M., 3rd, and Anastassiou, D. (2006) Computational inference of the molecular logic for synaptic connectivity in C. elegans. Bioinformatics 22, 6497-506 Han, J. D. (2008) Understanding biological functions through molecular networks. Cell Res 18, 224-237 18 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. Ravasz, B., Somera, A. L., Mongru, D. A., Oltvai, Z. N., and Barabasi, A. L. (2002) Hierarchical organization of modularity in metabolic networks. Science 297, 1551- 1555 Ihmels, J ., Friedlander, G., Bergrnann, S., Sarig, O., Ziv, Y., and Barkai, N. (2002) Revealing modular organization in the yeast transcriptional network. Nat Genet 31, 370-377 Yin, L., Huang, C. H., and Ni, J. (2006) Clustering of gene expression data: performance and similarity analysis. BMC Bioinformatics 7 Suppl 4, S19 Toronen, P., Kolehmainen, M., Wong, G., and Castren, E. (1999) Analysis of gene expression data using self-orgarrizing maps. FEBS Lett 451, 142-146 Ma, S. F., Grigoryev, D. N., Taylor, A. D., Nonas, S., Sammani, S., Ye, S. Q., and Garcia, J. G. (2005) Bioinfonnatic identification of novel early stress response genes in rodent models of lung injury. Am J Physiol Lung Cell Mol Physiol 289, L468-477 Toh, H., and Horimoto, K. (2002) Inference of a genetic network by a combined approach of cluster analysis and graphical Gaussian modeling. Bioinformatics 18, 287-297 Segal, B., Shapira, M., Regev, A., Pe'er, D., Botstein, D., Koller, D., and Friedman, N. (2003) Module networks: identifying regulatory modules and their condition- specific regulators fiom gene expression data. Nat Genet 34, 166-176 Li, Z., and Chan, 'C. (2004) Inferring pathways and networks with a Bayesian framework. FASEB J 18, 746-748 Husmeier, D. (2003) Sensitivity and specificity of inferring genetic regulatory interactions fiom microarray experiments with dynamic Bayesian networks. Bioinformatics 19, 2271-2282 Yu, J., Smith, V. A., Wang, P. P., Hartemink, A. J., and Jarvis, E. D. (2004) Advances to Bayesian network inference for generating causal networks from observational biological data. Bioinformatics 20, 3594-3603 Bar-Joseph, Z., Gerber, G. K., Lee, T. 1., Rinaldi, N. J., Yoo, J. Y., Robert, R, Gordon, D. B., Fraenkel, B., Jaakkola, T. 8., Young, R. A., and Gifford, D. K. (2003) Computational discovery of gene modules and regulatory networks. Nat Biotechnol 21, 1337-1342 19 __'L 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. Berman, B. P., Nibu, Y., Pfeiffer, B. D., T omancak, P., Celniker, S. E., Levine, M., Rubin, G. M., and Eisen, M. B. (2002) Exploiting transcription factor binding site clustering to identify cis-regulatory modules involved in pattern formation in the Drosophila genome. Proc Natl Acad Sci U S A 99, 757-762 Hartemink, A. J., Gifford, D. K., Jaakkola, T. S., and Young, R. A. (2002) Combining location and expression data for principled discovery of genetic regulatory network models. Pac Symp Biocomput, 43 7-449 Ideker, T., Thorsson, V., Ranish, J. A., Christmas, R., Buhler, J., Eng, J. K., Bumgarner, R., Goodlett, D. R., Aebersold, R., and Hood, L. (2001) Integrated genomic and proteomic analyses of a systematically perturbed metabolic network. Science 292, 929-934 Pilpel, Y., Sudarsanam, P., and Church, G. M. (2001) Identifying regulatory networks by combinatorial analysis of promoter elements. Nat Genet 29, 153-159 Li, F ., and Yang, Y. (2004) Recovering genetic regulatory networks fiom micro- array data and location analysis data. Genome Inform 15, 131-140 Patti, M. B., and Kahn, C. R. (1998) The insulin receptor--a critical link in glucose homeostasis and insulin action. J Basic Clin Physiol Pharmacol 9, 89-109 Thirone, A. C., Huang, C., and Klip, A. (2006) Tissue-specific roles of IRS proteins in insulin signaling and glucose transport. Trends Endocrinol Metab 17, 72-78 White, M. F. (1998) The IRS-signalling system: a network of docking proteins that mediate insulin action. Mol Cell Biochem 182, 3-11 Lowenstein, E. J., Daly, R. J., Batzer, A. G., Li, W., Margolis, B., Lammers, R., Ullrich, A., Skolnik, E. Y., Bar-Sagi, D., and Schlessinger, J. (1992) The SH2 and SH3 domain-containing protein GRB2 links receptor tyrosine kinases to ras signaling. Cell 70, 431-442 Skolnik, E. Y., Lee, C. H., Batzer, A., Vicentini, L. M., Zhou, M., Daly, R., Myers, M. J., Jr., Backer, J. M., Ullrich, A., White, M. F., and et a1. (1993) The SH2/SH3 domain-containing protein GRB2 interacts with tyrosine-phosphorylated IRS1 and She: implications for insulin control of ras signalling. EMBO J 12, 1929-1936 Backer, J. M., Schroeder, G. G., Kahn, C. R., Myers, M. G., Jr., Wilden, P. A., Cahill, D. A., and White, M. F. (1992) Insulin stimulation of phosphatidylinositol 3-kinase activity maps to insulin receptor regions required for endogenous substrate phosphorylation. J Biol Chem 267, 1367-1374 20 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. Alessi, D. R., James, S. R., Downes, C. P., Holmes, A. B., Gaffney, P. R., Reese, C. B., and Cohen, P. (1997) Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and activates protein kinase Balpha. Curr Biol 7, 261-269 Standaert, M. L., Galloway, L., Karnarn, P., Bandyopadhyay, G., Moscat, J., and Farese, R. V. (1997) Protein kinase C-zeta as a downstream effector of phosphatidylinositol 3-kinase during insulin stimulation in rat adipocytes. Potential role in glucose transport. J Biol Chem 272, 30075-30082 Le Roith, D., and Zick, Y. (2001) Recent advances in our understanding of insulin action and insulin resistance. Diabetes Care 24, 588-597 Kahn, B. B. (1998) Type 2 diabetes: when insulin secretion fails to compensate for insulin resistance. Cell 92, 593-596 Kahn, S. E., Hull, R. L., and Utzschneider, K. M. (2006) Mechanisms linking obesity to insulin resistance and type 2 diabetes. Nature 444, 840-846 Reaven, G. M. (1988) Banting lecture 1988. Role of insulin resistance in human disease. Diabetes 37, 1595-1607 Sone, H., Suzuki, H., Takahashi, A., and Yarnada, N. (2001) Disease model: hyperinsulinernia and insulin resistance. Part A-targeted disruption of insulin signaling or glucose transport. Trends Mol Med 7, 320-322 Irnamura, T., Takata, Y., Sasaoka, T., Takada, Y., Morioka, H., Haruta, T., Sawa, T., Iwanishi, M., Hu, Y. G., Suzuki, Y., and et a1. (1994) Two naturally occurring mutations in the kinase domain of insulin receptor accelerate degradation of the insulin receptor and impair the kinase activity. J Biol Chem 269, 31019-31027 McElduff, A., Hedo, J. A., Taylor, S. I., Roth, J., and Gorden, P. (1984) Insulin receptor degradation is accelerated in cultured lymphocytes fi'om patients with genetic syndromes of extreme insulin resistance. J Clin Invest 74, 1366-1374 Zick, Y. (2005) Ser/Thr phosphorylation of IRS proteins: a molecular basis for insulin resistance. Sci SYKE 2005, pe4 Herschkovitz, A., Liu, Y. F., Ilan, E., Ronen, D., Boura-Halfon, S., and Zick, Y. (2007) Common inhibitory serine sites phosphorylated by IRS-1 kinases, triggered by insulin and inducers of insulin resistance. J Biol Chem 282, 18018-18027 21 102. 103. 104. 105. 106. 107. Taylor, S. I., and Arioglu, E. (1998) Syndromes associated with insulin resistance and acanthosis nigricans. J Basic Clin Physiol Pharmacol 9, 419-439 Paz, K., Henri, R., LeRoith, D., Karasik, A., Elhanany, E., Kanety, H., and Zick, Y. (1997) A molecular basis for insulin resistance. Elevated serine/threonine phosphorylation of IRS-1 and IRS-2 inhibits their binding to the juxtamembrane region of the insulin receptor and impairs their ability to undergo insulin-induced tyrosine phosphorylation. J Biol Chem 272, 29911-29918 Hotarrrisligil, G. S., Peraldi, P., Budavari, A., Ellis, R., White, M. F., and Spiegehnan, B. M. (1996) IRS-l-mediated inhibition of insulin receptor tyrosine kinase activity in TNF-alpha- and obesity-induced insulin resistance. Science 271, 665-668 Pederson, T. M., Kramer, D. L., and Rondinone, C. M. (2001) Serine/threonine phosphorylation of IRS-1 triggers its degradation: possible regulation by tyrosine phosphorylation. Diabetes 50, 24-31 Hirosumi, J., Tuncman, G., Chang, L., Gorgun, C. 2., Uysal, K. T., Maeda, K., Karin, M., and Hotamisligil, G. S. (2002) A central role for JNK in obesity and insulin resistance. Nature 420, 333-336 Eley, H. L., Russell, S. T., and Tisdale, M. J. (2008) Role of the dsRNA-dependent protein kinase (PKR) in the attenuation of protein loss from muscle by insulin and insulin-like growth factor-1(IGF-I). Mol Cell Biochem 22 CHAPTER 2: Repression of PKR Mediates Palmitate-Induced Apoptosis in HepGZ Cells through BcI-2 The work in this chapter has been published in Cell Research: Yang, X. and Chan, C. (2009) Repression of PKR mediates palmitate- induced apoptosis in HepGZ cells through regulation of Bel-2. Cell Research 19, 469-486 Abstract The study in this chapter shows that the double-stranded RNA-dependent protein kinase (PKR) regulates the protein expression level and phosphorylation of Bcl-2 and exploits an anti-apoptotic role in human hepatocellular carcinoma cells (HepG2). In various types of cells, saturated fiee fatty acids (FF As), such as palrrritate, have been shown to induce cellular apoptosis by several mechanisms. Palmitate down-regulates the activity of PKR and thereby decreases the level of Bcl-2 protein, mediated in part by the NF-KB transcription factor. In addition to the level of Bel-2 protein, the phosphorylation of Bcl-2 at different amino acid residues, such as Ser70 and Ser87, is also important in regulating cellular apoptosis. The decrease in the phosphorylation of Bcl-2 at Ser70 upon exposure to palmitate is mediated by PKR and possibly JN K, whereas the phosphorylation of Bel-2 at Ser87 is unaffected by palmitate or PKR. In summary, PKR mediates the regulation of the protein level and the phosphorylation status of Bel-2, providing a novel mechanism of palmitate-induced apoptosis in HepG2 cells. 23 Introduction The double-stranded RNA-dependent protein kinase (PKR), activated by foreign double-stranded RNA (dsRNA) (l , 2), is best known for triggering cell defense responses by inhibiting general protein synthesis and inducing cellular apoptosis in many types of eukaryotic cells during viral infection (2, 3). However, evidence is emerging, albeit controversial at this point, suggesting that PKR also has an anti-apoptotic role in mouse embryo fibroblasts and certain tumor cells (4-8). As a Ser/Thr protein kinase, PKR is also known for its role in mediating signaling pathways (9) by interacting with proteins such as NF-KB, MAPKs, and PP2A (10-13). PKR phosphorylates and thereby releases I-KB from NF-ch, thus activating NF-ch and promoting the translocation of NF-KB into the nucleus (10, 11). Upon activation of NF-KB by PKR, the transcription of a number of apoptosis-regulating genes, such as FasL (14), p53 (15), and cIAPs (4), has been shown to be up-regulated. PKR plays a crucial role in the phosphorylation of the 3 mitogen- activated protein kinases (MAPK) (JNK, ERK, and p3 8) upon ribotoxic stress (12). However, the mechanism by which PKR interacts with the major MAPKs, as well as whether its role is apoptotic or anti-apoptotic, is unclear. PKR can also phosphorylate the PP2A regulatory subunit B56a and thus activate the catalytic subunit of PP2A, potentially leading to the dephosphorylation of eIF—4E and the arrest of translation (13). In addition to dsRNA, PKR has been shown to be responsive to many other factors (e.g., endoplasmic reticulum (ER) stress (16, 17), cytokines such as tumor necrosis factor (TNF)-or (18) and interleukin (IL)-1 (19), deoxynivalenol (DON, or vomitoxin) (12) and lipopolysaccharide (19)). The present study shows for the first time that palmitate down-regulates the activity of PKR in HepG2 cells. 24 Saturated FFAs (e.g., palmitate) induce apoptosis in many cell types, such as cardiac cells, pancreatic beta cells, breast cancer cells, and hepatocytes. They are associated with the development of a variety of diseases, including diabetes, heart disease, and non alcoholic fatty liver disease (NAFLD) (20, 21). The mechanism by which palmitate induces apoptosis depends on the cell type. For example, palmitate induces apoptosis by generating intracellular reactive oxygen species (ROS) in microvascular endothelial cells (EC) and retinal pericytes (22), but not in neonatal rat cardiomyocytes (23). Studies of palmitate-induced apoptosis in liver cells have focused predominantly on Iysosomal permeabilization (24), intracellular metabolic pathways such as beta oxidation (25), TG accumulation (26, 27), and ceramide production (28, 29). In addition, increased hydrogen peroxide (H202) and hydroxyl (*OH) radicals mediated the palmitate-induced lipotoxicity in the human hepatocellular carcinoma (HepG2/C3A) cell line (30). However, cytotoxicity, as well as apoptosis, was not completely prevented upon treatment with mitochondrial complex inhibitors or free radical scavengers, suggesting that mechanisms other than ROS production in mitochondria contribute to the toxicity of palmitate. More recent investigations implicate certain Bel-2 family proteins in mediating the saturated FFA-induced apoptosis of liver cells. For example, palmitate-induced apoptosis in liver cells is related to the activation (24, 31) of Bax and a decrease in the Bax antagonist Bcl-X(L) (24). In addition, elevated Bcl-2-interacting mediator of cell death (Bim), a pro-apoptotic Bel-2 family protein, plays a role in stearic and palmitic acid-induced apoptosis of several liver cell lines including HepG2 (31, 32). This process depends on the transcription factor F oxO3a (31). In addition to these pro-apoptotic Bel-2 family proteins, an anti-apoptotic member, Bel-2, has been identified as an important 25 factor in regulating the apoptosis of HepG2 cells (33-35). In pancreatic cells, the induction of apoptosis by palmitate is associated with reduced anti-apoptotic Bcl-2 levels (36). We previously showed that palmitate also decreased the protein level of Bel-2 in HepG2 cells (3 7). The present study confirms a similar effect of palmitate on the levels of Bcl-2 and provides a potential mechanism; furthermore, it shows that palnritate also down-regulates the phosphorylation of Bel-2 at Ser70, but not at Ser87, in HepG2 cells. As one of the most important anti-apoptotic members in the Bel-2 family, Bcl-2 protects cells against intrinsic apoptosis by maintaining the integrity of the mitochondrial membrane (3 8). The expression of the Bcl-2 gene is regulated by different transcription factors depending on the cell type (39-41). In liver tumor cell lines such as U937 and HepG2 cells, NF-ch has been identified as the central regulator of the transcription of the Bcl-2 gene (42, 43). Bel-2 is expressed in the progenitor cells of several self-renewmg tissues and some tumor cells, including human hepatocellular carcinoma (HepG2) cells (44-47). Irnmunohistochemical studies indicate that Bel-2 is not expressed in primary hepatocytes (44), although a more recent study with “hi gh-power’ ’ staining shows Bel-2 expression in primary hepatocytes (48). In the present study, I confirm by both RT-PCR and western blotting that HepG2 cells express Bel-2, and I propose a pathway through which palmitate regulates the Bcl-2 protein in HepG2 cells. Post-translational modification (e.g., phosphorylation) of Bcl-2 also plays a role in determining the anti-apoptotic role of Bcl-2 (49). Phosphorylation of Bel-2 at the anti- apoptotic site, Ser70, sustains the anti-apoptotic role of Bcl-2 (50). On the other hand, phosphorylation of Bcl-2 at SerS7 is believed to reduce the anti-apoptotic fiinction of Bel-2, possibly by inhibiting the phosphorylation of Bel-2 at Ser70 or destabilizing the 26 Bcl-2 protein (51, 52). MAPKs have been proposed to mediate the phosphorylation of Bel-2 (53-57) because the sequences surrounding both Ser70 and Ser87 residues of the Bel-2 protein represent the consensus motif, X-X-S-P, recognized by MAPKs (57). PP2A, a Ser/Thr-specific protein phosphatase, has been shown to dephosphorylate Bcl-2 at both Ser70 (58) and Ser87 (52) residues. The data in the present study support the action of PKR as an anti-apoptotic factor and demonstrate that PKR is involved in regulating the protein level and phosphorylation of Bcl-2 in HepG2 cells. There has been no evidence in the literature to date indicating whether palrrritate has an effect on the activity of PKR. I show that palmitate down- regulates the activity of PKR, which, however, does not alter the phosphorylation level of eIF-Za, and I propose that the repression of PKR mediates apoptosis through the regulation of Bel-2. Materials and Methods Cell Culture and Reagents—Human hepatocellular carcinoma cells (HepG2) were cultured in Dulbecco's Modified Eagle Medium (DMEM) (Invitrogen, Carlsbad, CA) with 10% fetal bovine serum (FBS) (Biomeda Corp, Foster City, CA) and penicillin- streptomycin (penicillin: 10,000 U/ml, streptomycin: 10,000 rig/ml) (Invitrogen, Carlsbad, CA). Freshly trypsinized HepG2 cells were suspended at 5 X 105 cells/ml in standard HepGZ culture medium and seeded at a density of 106 cells per well in standard six-well tissue culture plates. After being seeded, the cells were incubated at 37°C in a 90% air/10% CO2 atmosphere, and two milliliters of fresh medium was supplied every other day to the cultures after removal of the supernatant. The HepG2 cells were cultured in 27 standard medium for 5-6 days to achieve 90% confluence before treatment with FF As or other additives. HepGZ cell number was assessed by trypan blue dye exclusion using a hematocytometer. Phosphate-buffered saline (PBS) was purchased from Invitrogen, poly- (I:C) and trypan blue from Sigrna-Aldrich (St. Louis, MO), and NF-KB SN50 and its inactive control, NF-KB SNSOM, PKR inhibitor, JNK inhibitor (SP600125) and their analogues, used as negative controls, from EMD Biosciences (San Diego, CA). Fatty Acid Salt Treatment—Sodium salts of palmitate (P9767) and oleate (07501) were purchased fi'om Sigma-Aldrich. Palrrritate and oleate were complexed to 0.7 mM bovine serum albumin (BSA, fatty acid free) dissolved in the medium, which mirrrics the physiological concentration of albumin in human blood (3.5-5%, (59)). Fatty acid free BSA was purchased from MP Biomedicals (Chillicothe, OH). Dose responses of palnritate at 0, 0.2, 0.4, and 0.7 mM were performed in most of the experiments. If only one concentration was reported, the concentration of palmitate or oleate was 0.7 mM. In all experiments, FFAs were given for 24 hours, and the vehicle (0.7 mM BSA) was used as the control. RNA Interference for PKR and Reverse T ransfection—Silencer® Validated siRNA targeting human PKR mRN A was purchased fi'om Ambion (Austin, TX). The synthesized oligonucleotides for siRNA are 5'-GGUGAAGGUAGAUCAAAGATT-3‘ and 5'-UCUUUGAUCUACCUUCACCTI‘-3'. Reverse transfection of siRNA was performed. In general, the scrambled non-targeting siRNA as a negative control or the siRNA targeting PKR was diluted in serum and antibiotic free Opti-MEM (Invitrogen) and then mixed with the transfection reagent, Lipofectamine RNAiMAX (Invitrogen). The mixture of siRNA and Lipofectamine RNAiMAX in Opti-MEM was then added into 28 6-well plates and incubated at room temperature for 20 minutes. HepG2 cells suspended in antibiotic fi'ee medium were counted and plated into the 6-well plates at the same cell- number per well. The cells were then incubated at 37°C for 24 hours. After transfection, the cells were incubated in regular medium for another 24 hours and then collected. The mRNA and protein levels of PKR were measured by RT-PCR and western blot analysis, respectively. Titration of the siRNA and the transfection reagent was performed (not shown), and the lowest working amounts of the siRN A and the transfection reagent were applied in the loss-of-function (LOF) experiments in the present study. Over-expression of PKR and Forward T ransfection—The PKR plasmid, pCMV6- XLS-hPKR, and the empty vector, pCMV6-XL5, were purchased from Origene (Rockville, MD). Transient transfection was performed according to the Lipofectamine 2000 (Invitrogen) method. In general, regular HepG2 cells (Fig. 2.5C and 2.8B), or the cells in which the PKR geneiwas silenced by the siRNA of PKR (Fig. 2.36), were washed twice with phosphate buffered saline, and the medium was replaced with 2 ml of Opti-MEM with 1% FBS. Two micrograms per well of pCMV6-XL5-hPKR or the empty vector pCMV6-XL5 was then mixed with 10 ul/well of Lipofectamine 2000 in Opti- MEM and 20 minutes later the mixture was added to the wells. After 6 hours of transfection, the cells were then cultured in regular medium for 48 hours and subsequently harvested (Fig. 2.3C) or treated with palmitate (Fig. 2.5C and 2.8B). Cytotoxicity Measurement—HepG2 cells were cultured in different media for 24 hours and the supematants were collected. Cells were washed with PBS and kept in 1% triton-X-lOO in PBS for 24 hours at 37°C. The cell lysate was then collected, vortexed for 15 seconds, and centrifuged at 7000 rpm for 5 minutes. A cytotoxicity detection kit 29 (Roche Applied Science, Indianapolis, IN) was used to measure the LDH levels in the supematants and in the cell lysates. The fraction of LDH released into the medium was normalized to the total LDH (LDH released into the medium + LDH remaining in the cell lysates) (30). DNA Fragmentation—Treated HepG2 cells were lysed, and the DNA was extracted using the DNA purification kit from Promega (Madison, WI). Two rrricrograms of DNA samples was analyzed by electrophoresis on 1.5% agarose gels and visualized by SYBR gold staining for 4 hours. Caspase Analysis—For the caspase-3 substrate cleavage assay, the cells were washed with PBS, lysed, and assayed in a 96-well plate using the Caspase-3 cellular assay kit (Biomol, Plymouth Meeting, PA). Fluorescence was measured at emission and excitation settings of 360 and 460 nm, respectively, with a Microplate Spectrofluorometer from Molecular Device (Sunnyvale, CA). The caspase-3 activities were normalized by relative beta-actin levels. Nuclear Extraction and Detection of Nuclear NF -ch levels—Nuclear extracts from HepG2 cells were prepared using the Nuclear/Cytosol Fraction Kit from BioVision (Mountain View, CA). The extracted nuclear and cytoplasmic protein fractions were subjected to western blot analysis with anti-NF-kB p65 and anti-TBP, as a loading control for the nuclear extracts, and anti-beta actin for the cytoplasmic fractions. Western Blot Analysis and Immunoprecipitation—The HepG2 cells were washed twice with cold PBS and lysed in 300 uUwell of CelLytic M cell lysis buffer (Sigrna- Aldrich) supplemented with protease inhibitor cocktail (Roche Applied Science, Indianapolis, IN) and Ser/Thr phosphatase inhibitor cocktail (Sigrna-Aldrich). The cell 30 lysate was clarified by centrifugation at 10,000 rpm for 15 minutes, and the supernatant was collected. Total protein levels were quantified by BCA assay kit from Pierce Inc (Rockford, IL). A total of 20-40 ug of total protein was resolved by SDS-PAGE gels fi'om Bio-Rad, transferred to nitrocellulose membranes, and probed with primary and secondary antibodies. Biotinylated protein ladders (Cell Signaling, Beverly, MA) were loaded to one well of each SDS-PAGE gel, and anti-biotin antibody was used to detect the protein ladders on the western blots. Antibody detection was performed using the enhanced chemiluminescence kit from Pierce Biotechnology and imaged on the Molecular Imager ChemiDoc XRS System from Bio-Rad. For irnmunoprecipitation, the cell lysates were incubated with appropriate primary antibodies at 4°C for 1-2 hours, and the immunocomplexes were precipitated in a mixture with protein A affinity gel (Sigrna- Aldrich) by overnight incubation. The irnmunoprecipitates were washed three times with the cell lysis buffer and boiled in SDS-PAGE sample buffer, and the immune complexes were analyzed by western blot analysis. Phospho site-specific anti-eIF2a (SerSl) and JNK (T183/Y 185), anti-eIFZa, anti-PKR, anti-NF-kB p65, and anti-JNK rabbit polyclonal antibodies were purchased from Cell Signaling (Beverly, MA), phospho site- specific anti-PPZA/C (T yr307) from Abcarn (Cambridge, MA), phospho site-specific anti-Bcl-Z (Ser70) from Upstate (Charlottesville, VA), phospho site-specific anti-phospho PKR (Thr451) and Bcl-2 (Ser87) polyclonal antibodies from EMD Biosciences, and anti- Bcl-2, anti-Birn, anti-TBP, and anti-beta actin antibodies from Sigma-Aldrich. Anti- PP2A-B56a and anti- goat secondary antibodies were purchased fi'om Santa Cruz Biotechnology (Santa Cruz, CA). Secondary anti-rabbit and anti-mouse antibodies were purchased from Pierce Biotechnology Inc. 31 Real-time Quantitative RT -PCR Analysis—Total RNA was extracted from cells with the RNeasy mini kit (Qiagen, Valencia, CA) and depleted of contaminating DNA with RNase-fiee DNase (Qiagen). Equal amounts of total RNA (1 pg) were reverse- transcribed using an iScript cDNA synthesis kit (Bio-RAD). The first-strand cDNA was used as a template. The primers used for quantitative RT-PCR analyses of human PKR (5'-CCT GTCCT CT GG’I'I‘CTTIT GCT-3' and 5'-GATGATI‘CAGAAGCGAGTGTGC-3') (60), human Bcl-2 (5'-ACATCGCCCTGTGGATGACT-3' and 5'- TCAC'I'I‘GTGGCCCAGATAGG—3'), and human GAPDH (5'- AACT'I‘TGGTATCGTGGAAGGA-3' and 5'-CAGTAGAGGCAGGGATGATGT-3') were synthesized by Operon Biotechnologies, Inc. (Huntsville, AL). RT-PCR was performed in 25-ul reactions using 1/ 10 of the cDNA produced by reverse transcription, 0.2 uM of each primer, 1 X SYBR green superrnix from Bio-RAD, and an annealing temperature of 60 °C for 40 cycles. Each sample was assayed in three independent RT reactions and triplicate reactions were performed and normalized to the GAPDH expression levels. Negative controls included the absence of enzyme in the RT reaction and the absence of template during PCR. The cycle threshold (CT) values, corresponding to the PCR cycle number at which the fluorescence emission in real time reaches a threshold above the baseline emission, were determined using MyIQ1M Real-Time PCR Detection System (Bio-RAD). Statistical Analysis—All experiments were performed at least three times, and representative results are shown. All data, unless specified, are shown as the mean :t SD. for the indicated number of experiments. One-way AN OVA with Tukey’s post hoc 32 method was used to evaluate statistical significances between different treatment groups. Statistical significance was set at p<0.01. A B. 30% * 10 g 25% a} 3 8 8 20% '3 w a “ 5 6 '3 15% E '5 Z a I 2 4 3 10A: * a 2‘3 -' 5% .3 2 0% 0 0 0.2 0.4 0.7 Oleate 0 0.2 0.4 0.7 Oleate Palmitate Concentration (mM) Palmitate Concentration (mM) Figure 2.1. Effects of palmitate on the cytotoxicity and apoptosis of HepG2 cells. HepG2 cells were exposed to different levels of palmitate or 0.7 mM oleate for 24 hours. The vehicle for the FFAs (0.7 mM BSA) was used as the control (i.e., regular medium with BSA) (A, B). LDH release (A) and caspase-3 activity (B) were measured after treatment with palmitate. Data are expressed as the averages of nine samples i SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used to analyze the differences between the treatment groups. *, significantly higher than control (i.e., regular medium with BSA), p<0.01. Results Palmitate Induces Cytotoxicity and Apoptosis of HepGZ Cells—Previous work in our lab showed that palmitate induced cytotoxicity in HepG2 cells (30, 37), whereas unsaturated FFAs (e.g., oleate and linoleate) were not cytotoxic (30, 37). In a separate study, we found that upon exposure to palmitate, the HepG2 cells stained for Annexin-V (data not shown), which indicates phosphatidylserine extemalizatiorr, a sign of early stage apoptosis. In the present study, I further found that palmitate increased LDH release and caspase-3 activity of HepG2 cells in a dose-dependent manner, whereas oleate did not have a significant effect on LDH release and caspase-3 activity (Fig. 2.1 A and B), supporting the idea that palmitate induces cytotoxicity and apoptosis in HepG2 cells. It 33 has also been confirmed in the literature that palmitate induces apoptosis of HepG2 cells (31,32). A. we: E 1. , n. 1 p-PKRThr451—V-z p. an... L” 32"» . .. . = g '3 o PKR hunch—”w 220.5 E Q beta Actin w—va— n.- a. Palmitate (mM) 0 0.2 0.4 0.7 0 0,2 0.4 0-7 Palmitate Concentration (mM) B. we: g 1-2 ”"1 I ‘L 1 '5 A p"°KRT'"4~"1- § § 0.8 T. a ,5 ,3; 0.4 betaActin y.» n: 2; E 0.2 Control Palm. Ole. a 0 Control Palm. Ole. Figure 2.2. Effects of palmitate and oleate on the activity of PKR. HepG2 cells were exposed to different levels of palmitate (A) or 0.7 mM palmitate or oleate (B) for 24 hours. The vehicle for the FFAs (0.7 mM BSA) was used as the control (i.e., regular medium with BSA), in which the concentration of FFAs was 0 (A, B). Afier treatment, the cells were harvested and western blot analysis was performed to detect the phosphorylated level of PKR. The level of p-PKR Thr451 was quantified by normalizing to the levels of total PKR and is expressed as the average of three samples 3: SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used to analyze the differences between the treatment groups. *, significantly lower than control (i.e., regular medium with BSA), p<0.01. Palmitate Decreases the Activity of PKR in HepG2 Cells—Palmitate decreased the phosphorylation of PKR at Thr451, which indicates the activity of PKR, in a dose- dependent manner (Fig. 2.2A), whereas the unsaturated FFA oleate did not significantly affect the phosphorylation of PKR (Fig. 2.28) in HepG2 cells. Considering the apoptotic effects of palrrritate on HepG2 cells (Fig. 2.1B), I hypothesize that in HepG2 cells PKR is 34 involved in mediating the apoptosis induced by palmitate. To uncover the role of PKR in regulating apoptosis, gene silencing and gene over-expression studies were performed. PKR is Anti-apoptotic in HepGZ Cells—The siRNA targeting PKR that was employed in the present study markedly inhibited the gene and protein expression of PKR and thereby reduced the level of phosphorylated PKR (Fig. 2.3A). Silencing PKR with this siRNA increased the activity of caspase-3 significantly (Fig. 2.33) but did not increase the release of LDH (Fig. 2.3B), and it also induced the fragmentation of chromatin DNA (Fig. 2.33), suggesting that PKR has an anti-apoptotic role in HepG2 cells. To confirm the role of PKR in apoptosis, I over-expressed and rescued the PKR expression level in PKR-silenced cells and found that the caspase-3 activity was reduced to levels close to that of the control (Fig. 2.3 C). Taken together, these results suggest that PKR plays an anti-apoptotic role in HepG2 cells. To further confirm a catalytic role of PKR in regulating apoptosis, I inhibited the activity of PKR with a pharmaceutical inhibitor of PKR (61-63) and found that, similar to the siRNA of PKR, the PKR inhibitor also induced apoptosis in HepG2 cells, as evidenced by caspase-3 activity and DNA fragmentation (not shown). Considering the negative effect of palmitate on the activity of PKR, I therefore proposed that palmitate induces apoptosis, in part by repressing PKR. 35 Figure 2.3. Role of PKR in the cytotoxicity and apoptosis of HepG2 cells. Reverse transfection of suspended HepG2 cells was performed with scrambled siRNA (Control) or siRNA of PKR for 24 hours, and the transfected cells were cultured in regular medium for another 24 hours (A, B). Cells were then harvested, and RT-PCR and western blot analysis were performed to detect the gene, protein and phosphorylation levels of PKR to confirm that the PKR gene was silenced and the activity of PKR was suppressed (A). LDH release, Caspase—3 activity, and DNA fragmentation were assayed (B). In (C), reverse transfection of scramble siRNA (I, control) or siRNA of PKR (E, siPKR) was performed followed by forward transfection of empty vector pCMV6-XL5 (pCMV) or the plasmid containing PKR cDNA sequence (hPKR). Cells were then harvested and caspase-3 activity was assayed (C). Data are expressed as the average of three (A) or nine (B, C) samples i SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used to analyze the differences between the treatment groups. * , significantly higher (B, C) or lower (A, C), than control, p<0. 01. #, significantly lower than siPKR-CMV in Fig. C, p<0. 01. 36 2 1 ._ a. a. ._._ 0 o o o .358 28. :otafinxm 0:00 ”in. PKR beta Actin p-PKR Thr451 —> * — A 2 0 siPKR Ctrl siPKR Ctrl siPKR Clrl . _ 1 5 5- 1. o. $955 22. 33:3. @8380 . 0 15% 10% ~ % Se 883. In... siPKR Ctrl siPKR Chi IControl 1 5. 0 3955 En: 22.2 983.8 1.5 pCMVG hPKR hPKR pCMV6 37 1.2 0.8 BCI-2 0.6 * 0.4 Palmitate(mM) o 0.2 0.4 0.7 0, 801-2 Protein Level (fold change) 0 0.2 0.4 0.7 Palmitate Concentration (mM) 0.8 0.6 0.4 0.2 WB: beta Actin \_ ~ Ctrl Palm. Ole. Bel-2 Protein Level (fold change) Control Palm. Ole. Figure 2.4. Effects of palmitate and oleate on the protein level of Bcl-2. HepG2 cells were exposed to different levels of palmitate (A) or 0.7 mM palmitate or 0.7 mM oleate (B) for 24 hours. The vehicle for the FFAs (0.7 mM BSA) was used as the control (i.e., regular medium with BSA), in which the concentration of FFAs was 0 (A, B). After treatment, the cells were harvested and western blot analysis was performed to detect the protein level of Bel-2. Bel-2 protein levels were quantified by normalizing to the beta actin levels and are expressed as the average of three samples i SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used to analyze the differences between the treatment groups. *, significantly lower than control (i.e., regular medium with BSA), p<0.01. Palmitate Down-regulates the Protein Expression Level of Bcl-2—Palmitate decreased the protein level of Bcl-2 (Fig. 2.4A). Oleate did not have a significant effect on the protein level of Bel-2 (Fig. 2.43). It was unclear from the literature how palmitate would regulate the Bel-2 protein. However, as illustrated in Figs. 2.2 and 2.4, palmitate concomitantly decreased the phosphorylation of PKR and the protein level of Bel-2, suggesting a potential association between PKR and Bel-2. To confirm this association 38 and test the involvement of PKR in mediating the effect of palmitate on Bel-2, gene silencing and over-expression of PKR were performed. PKR is Involved in Mediating the Effects of Palmitate on the Protein Level of Bel-2 in HepGZ Cells—Silencing the expression of the PKR gene with siRNA of PKR (Fig. 2.3A) down-regulated the mRNA and protein levels of Bel-2 (Fig. 2.5 A and B), suggesting that PKR positively regulates the mRNA expression (Fig. 2.5A) and, in turn, the protein level (Fig. 2.53) of Bel-2. Similarly, inhibiting the activity of PKR with the PKR inhibitor also suppressed the mRN A and protein levels of Bel-2 (not shown). Considering the repressive effect of palmitate on PKR activity (Fig. 2.2), I hypothesize that the suppression of PKR activity mediates the negative effects of palmitate on Bel-2 levels. To confirm my hypothesis that PKR is involved in the decrease of Bel-2 by palnritate, I over-expressed the PKR gene in HepG2 cells (Fig. 2.56) and exposed the cells to palmitate. I indeed found that the Bel-2 level was rescued by over-expression of PKR (Fig. 2.5C). This suggests that the suppression of PKR activity indeed mediates the decrease of Bcl-2 induced by palmitate. 39 Figure 2.5. Involvement of PKR in regulating the protein level of Bcl-2. Reverse transfection of suspended HepG2 cells was performed with scrambled siRNA (Control) or siRNA of PKR for 24 hours and the transfected cells were cultured in regular medium for another 24 hours (A, B). Cells were harvested, and RT-PCR (A) and western blot analysis (B) were performed to detect the gene (A) and protein (B) expression levels of PKR (I, also shown in Fig. 2.3C) and Bcl-2 (a). In HepG2 cells, the forward transfection of empty vector pCMV6-XL5 (pCMV6) or the plasmid containing PKR cDNA sequence (pCMV6-hPKR) was performed, and the cells were then treated with different concentrations of palmitate for 24 hours (C). The vehicle for palmitate (0.7 mM BSA) was used as the control (i.e., regular medium with BSA), in which the concentration of palmitate was 0 (C). Cells were then harvested and western blot analysis was performed to detect the phosphorylation and protein levels of PKR and the protein levels of Bel-2 (C). Gene expression data are expressed as the average of nine samples :1: SD from three independent experiments. The protein levels of Bel-2 were quantified by normalizing to beta actin and are expressed as the average of three samples :l: SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used for analyzing the differences between the treatment groups. * and #, PKR and Bel-2 levels, respectively, significantly lower than control, p<0.01. 40 > ,6 1.2 3 12 .. g 1 3 IPKR IPKR .: 1 UBcl-2 1 o . ,, , .: DBcl—Z. E 0-8 3 0.8 ‘c’ 0 . # beta Actin WM 3% , g -6 - a, 0.6 # s 3 0.4 Control siPKR g 04 a. E _| " ' r: ‘k l: 0.2 . * g 0.2 5 ; e c o n. 0 cm siPKR cm siPKR .0 WB: mpam_~aaq betaActin [wwmeW ,.. I p-PKR Thr451 PKR Palmitate (mM) 0 0.2 0.4 0.7 0 0.2 0.4 0.7 pCMV6 pCMV6-hPKR 3.5 3 N r001 (fold change) 2; Bel-2 protein level —A 0.5 0 Palmitate(mM) 0 0.2 0.4 0.7 0 0.2 0.4 0.7 pCMV6 pCMV6-hPKR 41 A. B. Nuclear extract NF-KB p65 w w ' ‘ 4 Bcl-2 «OI—v” “M beta Actin ~ w m TBP h u _ '“' “- l Ctrl SN50M SN50 Cytoplasmic fraction Bel-2 “WW ”4 Nuclear extract NF-KB p65 ~ H "I TBP P" hi- beta ACti" W Ctrl Ole. Palm. pCMV6 hPKR pCMV6 hPKR Control siPKR NF-ch p65 W w v u.- Figure 2.6. Role of PKR in regulating the Nuclear NF-KB p65 level. Reverse transfection of scramble siRNA (control, the first two lanes) or siRNA of PKR (siPKR, the third and fourth lanes) was performed, followed by the forward transfection of empty vector pCMV6-XL5 (pCMV) or the plasmid containing PKR cDNA sequence (hPKR) (A). Cells were then harvested, the nuclear extract was separated from the cytoplasmic fiaction, and western blot analysis was performed to detect the level of Bel-2 in the cytoplasmic fiaction and the levels of NF-kB p65 in both the nuclear extract and the cytoplasmic fiaction. TBP and beta actin were also measured as loading controls for the nuclear extract and the cytoplasmic fi'action, respectively (A). HepG2 cells were exposed to a cell-permeable inhibitor of NF-kB, SN50 (18 uM), or its negative control, SNSOM (18 uM), in regular medium for 24 hours (B). After treatment, the cells were harvested, and western blot analysis was performed to detect the protein levels of Bcl-2 in the whole cell lysates (B). HepG2 cells were exposed to 0.7 mM palmitate or oleate for 24 hours (C). The vehicle for the FFAs (0.7 mM BSA) was used as the control (i.e., regular medium with BSA). After treatment, the cells were harvested, and western blot analysis was performed to detect the protein level of NF-kB p65 in the nuclear extracts (C). Thus far, I showed that palmitate down-regulated the activity of PKR, which played an anti-apoptotic role in HepG2 cells. By suppressing PKR activity, palnritate down-regulated the expression levels of Bel-2. Although the mechanism is not fully understood, the positive effect of PKR on the anti-apoptotic protein Bel-2 could serve as a potential pathway by which PKR protects HepG2 cells from apoptosis. PKR 42 phosphorylates I-KB, which then releases and activates NF-KB (10, 11), the key transcription factor that up-regulates the transcription of Bel-2 in HepG2 cells (42, 43). Indeed, silencing the gene expression of PKR decreased the level of NF-KB in the nucleus (Fig. 2.6A, comparing lanes 1 and 3), whereas over-expressing PKR increased it (Fig. 2.6A, comparing lanes 1 and 2). Moreover, the decreased level of NF-KB in the nucleus was restored by rescuing the PKR expression in PKR-silenced cells (Fig. 2.6A, comparing lanes 3 and 4). Furthermore, the PKR inhibitor also decreased the level of NF- KB p65 in the nuclear extract of HepG2 cells (not shown). These results suggest that PKR regulates the activity of NF-ch in HepG2 cells. NF-KB plays a key role in facilitating the transcription of the Bel-2 gene in liver tumor cell lines, such as U937 and HepG2 cells, and the inhibition of NF-KB results in the down-regulation of Bcl-2 gene expression (42, 43). Therefore, it is expected that the protein level of Bel-2 changes in correspondence with the nuclear level of NF-KB, as shown in Fig. 2.6A, in which the gene expression level of PKR is modulated. I further confirmed the role of NF-KB in regulating the expression level of Bcl-2 with an inhibitor of NF-KB, NF-KB SN50 (Fig. 2.6B). Thus, from my results and the literature data, I propose that the transcription factor, NF-KB, mediates the PKR regulation of Bel-2 expression in HepG2 cells. Indeed, I also observed that palmitate decreased the level of NF-KB in the nucleus (Fig. 2.66). In summary, palmitate induces apoptosis of HepG2 cells, in part by reducing the Bel-2 level, which is mediated by the repression of PKR and NF-KB activities (Fig. 2.6). However, in addition to the protein level of Bcl-2, the anti-apoptotic role of Bcl-2 is also regulated by the post-translational modification of Bel-2. The phosphorylation of Bcl-2 at Ser70 sustains the anti-apoptotic role of Bel-2, whereas phosphorylation of Bcl-2 at 43 Ser87 attenuates the anti-apoptotic role of Bel-2 (50-52). Therefore, I further investigated the involvement of palmitate and PKR in regulating the phosphorylation of Bel-2. A. 1.2 IP: Bcl-2. IB: p-BcI-2 Ser70 I--- - -~-l lP: Bcl—2, IB: Bel-2 p Bel-2 Ser70 (fold change) lmuw Palmitate (mM) 0 0.2 0.4 0.7 0 0.2 0.4 0.7 Palmitate Concentration B. IP: Bel-2, IB: p-Bcl-2 Ser87 IP: Bcl-Z. IB: Bcl-2 Control Palmitate Figure 2.7. Effect of Palmitate on the phosphorylation of Bel-2. HepG2 cells were exposed to regular medium with different levels of palmitate (A, B) for 24 hours. The vehicle for palmitate (0.7 mM BSA) was used as the control (i.e., regular medium with BSA), in which the concentration of palmitate was 0 (A, B). After treatment, the cells were harvested and immunoprecipitated with anti-Bcl—2 and detected for the phosphorylation of Bel-2 at Ser 70 (A) and Ser87 (B). The phosphorylation levels of Bcl- 2 were quantified by normalizing to total Bel-2 levels and are expressed as the average of three samples i SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used for analyzing the differences between the treatment groups. *, significantly lower than control (i.e., regular medium with BSA), p<0.01. Palmitate Decreases the Phosphorylation of Bel-2 at Ser70, whereas the Phosphorylation of Bcl-2 at Ser87 is Not Affected—Palmitate decreased the phosphorylation of Bel-2 at Ser 70 in HepG2 cells (Fig. 2.7A). 0n the other hand, the phosphorylation of Bel-2 at the pro-apoptotic residue, Ser87, was not affected by palmitate (Fig. 2.73). Considering the concomitant repression by palmitate of the phosphorylation of Bel-2 and the PKR activity (Fig. 2.2), I propose that PKR may be involved in mediating the effects of palmitate on the phosphorylation of Bel-2. PKR is Involved in Mediating the Effecm of Palmitate on the Phosphorylation of Bel-2 at Ser70—Silencing the PKR gene using siRNA of PKR decreased the phosphorylation of Bel-2 at Ser70 (Fig. 2.8A), indicating that PKR has a positive effect on (i.e., enhances) the phosphorylation of Bel-2 at Ser70, whereas the phosphorylation of Bcl-2 at Ser87 was not affected by silencing PKR (Fig. 2.8A). The effects of silencing or inhibiting PKR on the phosphorylation of Bel-2 at Ser70 and Ser87 are consistent with those of palmitate, and considering the negative effect of palmitate on PKR activity (Fig. 2.2), I propose that the suppression of the PKR activity mediates the negative effects of palmitate on the phosphorylation of Bel-2 at Ser70, as well as the protein level of Bcl-2. To confirm this hypothesis, I over-expressed the PKR gene in HepG2 cells and exposed the cells to palmitate (Fig. 2.5C). The phosphorylation of Bcl-2 at Ser70 was restored by over-expressing PKR in the palmitate treatment (Fig. 2.83), supporting the involvement of PKR in mediating the effect of palmitate on the phosphorylation of Bel-2 at Ser70. However, a co-irnmunoprecipitation study showed that PKR did not directly interact with the Bcl-2 protein (Fig. 2.80; therefore, other intermediate signaling molecules are involved in mediating the effect of PKR on the phosphorylation of Bel-2. PKR has been reported to positively signal MAPKs (12), and sequence analysis of both Ser70 and Ser87 residues of Bcl-2 suggests that JNK could regulate the phosphorylation of Bel-2 (57). Taken together, the evidence suggests that JNK may act as one of the intermediates between PKR and Bel-2. 45 Figure 2.8. Involvement of PKR in regulating the phosphorylation of Bel-2 at Ser70. Reverse transfection of suspended HepG2 cells was performed with scrambled siRNA (Control) or siRNA of PKR for 24 hours and the transfected cells were cultured in regular medium for another 24 hours (A). Cells were harvested and immunoprecipitated with anti-Bcl-2 and detected for the phosphorylation of Bel-2 at Ser 70 and Ser87 (A). In HepG2 cells, the forward transfection of empty vector pCMV6-XL5 (pCMV 6) or the plasmid containing PKR cDNA sequence (pCMV6-hPKR) was performed and the cells were then treated with different concentrations of palmitate for 24 hours (B). The vehicle for palmitate (0.7 mM BSA) was used as the control (i.e., regular medium with BSA), in which the concentration of palnritate was 0 (B). After the transfections and palmitate treatment, the cells were harvested, and cell lysates were immunoprecipitated with anti- Bcl-2 and detected for phosphorylation of Bel-2 at Ser 70 (B) (Please see Fig. 2.5C for the over-expressed PKR levels). Confluent HepG2 cells were harvested and immunoprecipitated with anti-Bcl-2 or anti-PKR, and western blot analysis was performed to detect the coirnmunoprecipitation of Bel-2 and PKR (C). The phosphorylation levels of Bel-2 were quantified by normalizing to total Bcl-2 levels and are expressed as the average of three samples :t SD from three independent experiments. One-way AN OVA with Tukey’s post hoc method was used for analyzing the differences between the treatment groups. *, significantly different fiom control, p<0.01. 46 IP: Bel-2, IB: p-Bcl-2 Ser87 3 1 8 - 3 IP: Bel-2, IB: p-BCl—Z Ser70 g g 0'8 * 2 .9 0.4 IP: Bcl-2, lB: Bel-2 73 - . l=E' 0.2 H' ”with 0- Ctrl siPKR ° . Ctrl SIPKR B. IP: Bel-2, lB: p-Bcl-2 Ser70 Lw ¢ ~— - “nu-«o ~ _ n —l IP: Bel-2, IB: Bcl-2 It-N- what—“w m w—‘MH‘ Palmitate(mM) 0 0.2 0.4 0.7 O 0.2 0.4 0.7 pCMV6 pCMV6-hPKR 1.6 1.4 1.2 1 0.8 0.6 0.4 p-Bcl2 Ser70/total Bcl2 (fold change) 0.2 O Palmitate(mM) o 0.2 0.4 0.7 o 0.2 0.4 0.7 pCMV6 pCMV6-hPKR IB: Bcl- 2 IB: PKR IP: Ctrl Bcl-2 PKR 47 PKR Positively Regulates JNK, and JNK Regulates the Phosphorylation of Bcl- 2 at Ser 70, but not at Ser8 7—A co-immunoprecipitation study indicated that JNK directly interacted with both PKR and Bel-2 (Fig. 2.9A), suggesting that JNK may be a potential intermediate protein kinase that mediates the positive effect of PKR on the phosphorylation of Bel-2 at Ser70. Indeed, silencing the PKR gene repressed the activity of JNK (Fig. 2.93, comparing lanes 1 and 3), whereas over-expressing PKR enhanced it (Fig. 2.9B, comparing lanes 1 and 2). Moreover, the suppressed JNK activity was restored by rescuing the PKR expression in PKR-silenced cells (Fig. 2.93, comparing lanes 3 and 4), confirming the positive connection between PKR and JNK. Considering the negative effect of palmitate on PKR activity (Fig. 2.2), I then propose that palmitate decreases the activity of JNK by repressing PKR. Indeed, I found that treatment of palmitate for 24 hours decreased the phosphorylation of JNK (Fig. 2.9C), and over- expressing PKR in the palmitate treatment restored the phosphorylation of JN K (Fig. 2.9D), supporting the involvement of PKR in mediating the effect of palmitate on the phosphorylation of JNK. In addition, to firrther assess the connection between JNK and Bel-2, I performed a JNK inhibition study and found that the JNK inhibitor, SP600125, suppressed the phosphorylation of Bel-2 at Ser70, but not the phosphorylation at Ser87 (Fig. 2.9 E and F), which is consistent with the distinct effects of palnritate as well as PKR on the phosphorylation of Bel-2 at the two amino acid residues (Figs. 2.7, 2.8). Therefore, the results in Figure 2.9 suggest JNK is an intermediate protein that mediates the effects of pahnitate and PKR on the phosphorylation of Bcl-2 at Ser70. I showed that PKR had an anti-apoptotic role in HepG2 cells (Fig. 2.3). To firrther confirm the positive correlation between PKR and JNK, I investigated the role of JNK in regulating 48 cytotoxicity and apoptosis in HepG2 cells. The inhibition of JNK using SP600125 significantly increased LDH release (Fig. 2.104) and caspase-3 activity (Fig. 2.103), suggesting that, similar to PKR, JNK has an anti-apoptotic role in HepG2 cells. In summary, palmitate inhibits the phosphorylation of Bel-2 at the anti-apoptotic residue, Ser70, without affecting the pro-apoptotic residue, Ser87. My data suggest that this effect of palmitate on the phosphorylation of Bcl-2 is mediated by the signaling of PKR and JNK. I therefore propose another anti-apoptotic pathway, which is suppressed by palmitate, consisting of PKR, JNK, and phosphorylation of Bcl-2. My investigation of the association between PKR and Bel-2 revealed two different but complementary anti-apoptotic pathways that connect PKR and Bel-2. First, PKR up-regulates the transcription of the Bel-2 gene, possibly through the transcription factor, NF-KB. Second, PKR up-regulates the phosphorylation of Bel-2 at the anti- apoptotic residue, Ser70, mediated by JNK (Fig. 2.113). These two pathways were down- regulated in HepG2 cells upon exposure to palmitate, and they may constitute one of the mechanisms by which palmitate induces apoptosis in HepG2 cells. 49 Figure 2.9. Involvement of JN K in regulating the phosphorylation of Bel-2. Confluent HepG2 cells were harvested and immunoprecipitated with anti-Bcl-2 or anti- PKR, and western blot analysis was performed to detect the protein level of JN K (A). Reverse transfection of scramble siRNA (control, the first two lanes) or siRNA of PKR (siPKR, the third and fourth lanes) was performed, followed by the forward transfection of empty vector pCMV6-XL5 (pCMV) or the plasmid containing PKR cDNA sequence (hPKR) (B). HepGZ cells were exposed to 0.7 mM palmitate or oleate for 24 hours (C). The vehicle for the FFAs, 0.7 mM BSA, was used as the control (i.e., regular medium with BSA) (C). In HepG2 cells, the forward transfection of empty vector pCMV6-XL5 (pCMV6) or plasmid containing PKR cDNA sequence (pCMV6-hPKR) was performed and the cells were then treated with 0.7 mM palmitate for 24 hours (D). The vehicle for palmitate, 0.7 mM BSA, was used as the control (i.e., regular medium with BSA) (D). After treatments, cells were then harvested and western blot analysis was performed to detect the phosphorylation and protein levels of JNKl/2 after treatment (B, C, D). HepG2 cells were exposed in regular medium for 24 hours to the pharmaceutical inhibitor of JNK, SP600125 (25 uM) (E, F), or its analogue, SP600125 (25 uM) (E), as a negative control. The control in F is the vehicle of palnritate, BSA. After treatment, the cells were harvested and western blot analysis was performed to detect the protein level of Bel-2 and the phosphorylation of Bcl-2 at Ser70 (E), or the cell lysate was immunoprecipitated with anti-Bcl-2 and western blot analysis was performed to detect the phosphorylation level of Bcl-2 at Ser87 (F). 50 A. B. leJNK p-JNK —+ q... - T1831Y185 _’ . —9 j IP. Ctrl Bel-2 PKR JNK . pCMV6 hPKR pCMV6 hPKR Control siPKR C. D. p-JNK —> f j i " ' ' P-JNK -’ ~.3--- ’ T183IY185—D .. . T183IY185 -* g JNK" . "T r ,..} - _> , . v , . ..7 Control Palm. Ole. Ct rl Palm. Ct ” Palm. pCMV6 pCMV6-hPKR E. F IP: Bel-2, IB: p-BcI-2 Ser87 IP: Bel-2, IB: Bcl-2 Bel-2 W «null-b W W, . Control 3P600125 Control Palm. SP600125 51 .> 5” 1.60% 2.5 'k g 1.20% - :EA 2 v .5 o g 2 g 1.5 a 0.80% 3 5 . §£ 0.40%. o 0.5 0.00% 0 cm SP600125 Ctrl SP600125 Figure 2.10. Effect of inhibiting JNK activity on the cytotoxicity and apoptosis of HepG2 cells. HepG2 cells were exposed for 24 hours to the pharmaceutical inhibitor of JNK, SP600125 (25 pM), or its analogue, SP600125 (ZSuM), as a negative control (A, B). LDH release (A) and caspase-3 activity (B) were measured afier treatment with the JNK inhibitor. Data are expressed as the average of nine samples :t SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used for analyzing the differences between the treatment groups. *, significantly higher than control, p<0.01. 52 Bim-EL—> m w Bim-L —> ~ Ctrl siPKR Palmitate cu ‘......... FoxO3a other factors? V Bim Ser70 genes including Bel-2 Figure 2.11. Proposed signaling pathways from PKR to Bel-2 induced by palmitate. A. Reverse transfection of suspended HepG2 cells was performed with scrambled siRNA (Control) or siRNA of PKR for 24 hours and the transfected cells were cultured in regular medium for another 24 hours. Cells were harvested, and western blot analysis was performed to detect the protein level of Bim. B. A summary of the signaling pathways identified in this study. First, PKR up-regulates the transcription of the Bel-2 gene through a transcription factor, likely NF-KB. Second, PKR up-regulates the phosphorylation of Bel-2 at the anti-apoptotic residue Ser70, mediated by JNK. By suppressing PKR, and in turn the two downstream pathways, palmitate regulates the protein level and phosphorylation of Bcl—2 at Ser70 and therefore induces apoptosis in HepGZ cells. Another Bel-2 family protein, Bim, which also mediates palmitate-induced apoptosis (31, 32) (dashed lines), and the potential effect of PKR on Bim (line with the question mark) are included. Apoptosis regulating 1 53 DiscuSsion In this chapter, PKR is found to be anti-apoptotic in human hepatoma cells. PKR is best known for its pro-apoptotic role by phosphorylating eIF-Za and thereby inhibiting general protein synthesis (2, 3). Thus, it has previously been suggested that PKR acts as a tumor suppressor by inhibiting cell growth and inducing apoptosis (64, 65). In contrast, more recent studies suggest that PKR has an anti-apoptotic role in regulating tumor development and tumor cell apoptosis (5-8). Elevated PKR protein levels and activity were observed in certain tumor cells (e.g., human breast cancer cells (5), melanoma cells (6, 8), and hepatitis C virus (HCV)-related hepatocellular carcinoma (7)). The over- expression and elevated activity of PKR have been attributed to the development of tumors and the proliferation of tumor cells (5-8), but the mechanism is still unclear. Research suggests that PKR may suppress apoptosis by activating the NF-KB pathway before phosphorylating eIF-Za, therefore inducing cell survival initially and, subsequently, cell death in NIH3T3 cells expressing PKR (4). The present study supports an anti-apoptotic role of PKR in human hepatoma cells (HepG2) and suggests potential pathways by which PKR mediates anti-apoptosis through Bel-2, thereby further providing a potential mechanism of palmitate-induced apoptosis in HepG2 cells. Interestingly, although the phosphorylation of PKR was significantly suppressed in HepG2 cells, the phosphorylation level of eIF-Za was not altered by palmitate (not shown), suggesting that the protein synthesis arrest machinery and related apoptosis activity were not affected. Similarly, another substrate of PKR, BS6a-PP2A, was not affected by palmitate (not shown). This suggests that other substrates of PKR, which I 54 propose may be NF-KB and JN K, are mediating the effects of altered PKR activity on the regulation of the Bel-2 protein and its phosphorylation levels in HepGZ cells. PKR activates several transcription factors, such as IRF-l, p53, and NF-ch (66, 67). In the present study, I show that PKR is involved in controlling the transcription of Bel-2 in HepGZ cells, mediated by the transcription factor NF-KB. This result suggests a novel mechanism by which palmitate down-regulates the protein level of Bel-2 (Fig. 2.4) in HepGZ cells. By modulating the transcription of a number of anti-apoptotic genes, such as Bel-2 (68), Bcl-xL (69), cIAPs (70), and p53 (71), NF-KB is commonly considered an anti-apoptotic transcription factor (72, 73), which is consistent with the anti-apoptotic role of PKR in HepG2 cells. NF -KB is the key regulator of transcription of the Bel-2 gene in liver tumor cell lines, including HepG2 (42, 43). However, other transcription factors that may regulate the transcription of the Bel-2 gene, for example, c- Myb (39, 40) and STAT5 (41), have not been evaluated in the present study. It is not known in the literature whether these transcription factors are altered by palmitate or regulated by PKR. Therefore, it remains to be determined whether other transcription factors, in addition to NF-kB, mediate the effect of PKR on the gene expression and protein levels of Bel-2. Furthermore, other studies have revealed that elevated Bim, a pro-apoptotic Bel-2 family protein, also plays an important role in the stearic and palmitic acid-induced apoptosis of several liver cell lines, including HepG2 (31, 32), and this process has been determined to be dependent on the transcription factor FoxO3a (31). In the present study, I propose that the suppression of PKR mediates palmitate-induced apoptosis through Bcl- 2; therefore, to test the potential connection between my results and the previous findings 55 on Bim, I evaluated the effect of PKR on the expression of Bim. Interestingly, silencing PKR significantly increased the protein level of Bim (Fig. 2.11A), suggesting that PKR represses Bim expression. As a BH3-only Bel-2 protein, Bim promotes apoptosis by binding and inactivating anti-apoptotic proteins (74); therefore, the negative effect of PKR on Bim supports the role of PKR as an anti-apoptotic protein and could serve as another potential mechanism by which PKR inhibits apoptosis in HepGZ cells. In addition, I show in the present study that palmitate decreased PKR activity (Fig. 2.2), and thus it is possible that the up-regulation of Bim by palmitate (31) may be due, in part, to the repression of PKR activity. However, the mechanism by which PKR down-regulates the expression of Bim is not known at this point. Future experiments will determine whether PKR has any effect on the key transcription factor of Bim, Fox03a (31). Previously, it was shown that PP2A, which was activated by palmitate, mediated the effect of palmitate on the activity of Fox03a and thereby the expression of Bim (31). However, in my study palmitate did not affect the activity of PP2A (data not shown), and this contradictory result may be due to the different treatment time (6 vs. 24 hours) and/or the different types of palmitate used (palmitic acid dissolved in isopropyl alcohol vs. sodium palmitate complexed with BSA). In the present study, I uncovered a second anti-apoptotic pathway, namely, that PKR up-regulates the phosphorylation of Bel-2 at Ser70 (Fig. 2.8), mediated by JNK (Fig. 2.9). Although sequence analysis suggests that both Ser70 and Ser87 residues of Bel-2 can be recognized by JN K (57), I found that only phosphorylation at Ser70 is affected by JNK (Fig. 2.9). Concomitantly, palmitate (Fig. 2.7) and PKR silencing (Fig. 2.8) both decreased only the phosphorylation of Bel-2 at Ser70, which lends support to the 56 hypothesis that palmitate down-regulates the phosphorylation of Bel-2 at Ser70 through PKR and JNK. In fact, all 3 MAPKs are activated by PKR upon ribotoxic stress, and the investigators proposed that MAPKs respond to PKR in the rank order of JNK>p3 8>ERK (12). However, I did not test the effects of PKR on the other two less responsive MAPK proteins, ERK and p38, which were also proposed to phosphorylate Bel-2 at Ser70 (54- 56). Thus, JNK is one of the intermediates involved in mediating the signaling pathway from PKR to Bel-2 phosphorylation. Further investigations of the 3 MAPKs are required to fully understand the interaction between PKR and Bel-2. Currently, it remains unclear how PKR interacts with MAPKs (e. g., JNK). An association between PKR and apoptosis signal-regulating kinase 1 (ASKI), one of the MAPK signaling proteins, has been previously established (75). ASK], a MAPK kinase kinase (MAPKKK), phosphorylates SEKI/MKK4 and MKK3/MKK6, which, in turn, activate IN K and p38 MAPK, respectively (76). My coirnmunoprecipitation study suggests an interaction between JNK and PKR (Fig. 2.9). Nevertheless, it must be noted that the coirnmunoprecipitation of PKR and IN K does not necessarily indicate a “direct” binding of these two proteins. It is believed that the activation of the MAPK cascade requires a scaffold protein that assembles the MAPKKK, MAPKK, and MAPK proteins together into a certain signaling module (77). For instance, J IPl organizes upstream kinase HPK-l, MAPKKK MLKI, MKK7, and JNKl/2, and it facilitates the activation of JNK (7 8). Therefore, the coirnmunoprecipitation study suggests an interaction, but not necessarily a direct one, between PKR and JNK. Instead, the coirnmunoprecipitation of PKR and JNK may recruit a third protein, possibly a scaffold protein that assembles JNK and facilitates the activation of JNK by PKR. In other words, the pathway that I identified 57 consisting of PKR and JNK does not preclude the possibility of other intermediates between PKR and JN K. To understand how PKR interacts with JNK, the scaffold protein and the assembly of the signaling proteins with their upstream activators must be investigated. JNK is involved in many signaling pathways that control diverse cellular activities such as proliferation, differentiation, and apoptosis. Through different pathways and substrates (79), JNK has been shown to have either pro- or anti-apoptotic functions, depending on the cell type, stimulus, duration of its activation and activity of other signaling pathways, and it is therefore considered a “double-edged sword” (80). In liver tumor cell lines such as HepG2, conflicting views on the role of JNK in regulating apoptosis have been reported in the literature. For example, JNK was shown to mediate anti-apoptotic signals in transforming growth factor-beta l- (81) and tumor necrosis factor-induced apoptosis (82) in HepGZ cells, whereas other researchers have proposed pro-apoptotic roles for JNK in HepG2 cells (32, 83). In these studies on the role of JNK in regulating apoptosis, the JNK inhibitor SP600125 was used, and the conflicting results may be due to the different conditions applied by the investigators, such as the concentrations and treatment times of SP600125 and the accompanying additives. In my study, I inhibited JNK in regular medium for 24 hours and found that the IN K inhibitor (SP600125, at 25 uM) was apoptotic, as evidenced by caspase-3 activity (Fig. 2.103). Zhang et al. (81), treated liver tumor cell lines with 20 uM SP600125 for 50 hours, resulting in a significant increase of apoptosis. On the other hand, Chen et al. (83) treated the HepGZ cells with 1 uM SP600125 for 48 hours and showed that it blocked the norcantharidin-induced apoptosis. Interestingly, in their controls without norcantharidin, 58 SP600125 or ERK inhibitor (U0126 or PD98059) slightly increased the apoptosis of the HepGZ cells, which supports my finding with SP600125 in the control medium. Malhi et al. (32) used SP600125 to inhibit JNK in the presence of palmitic acid for 24 hours, but the concentration of the inhibitor used was not specified. My study suggests that JNK is involved in phosphorylating Bel-2, an anti-apoptotic member of the Bel-2 family, at its anti-apoptotic residue, Ser70 (Fig. 2.9C), suggesting that JNK may act as an anti- apoptotie factor. The anti-apoptotic role exhibited by JNK is consistent with that of PKR and therefore lends support to the hypothesis that JNK may mediate, in part, the signaling between PKR and Bel-2. However, inhibiting JNK with SP600125 also increased the LDH release significantly (Fig. 2.10A), whereas silencing PKR did not have a significant effect on the LDH release (Fig. 2.33), suggesting that JNK may be involved in other cellular activities related to cytotoxicity, in addition to the pathway fiom PKR to Bel-2. On the other hand, it is also possible that the inhibitor SP600125 itself has other non- specific effects that lead to cytotoxicity, despite its widespread use as an inhibitor of JNK. In summary, I identified an anti-apoptotic role of PKR and found that it is involved in regulating the protein level and phosphorylation status of Bel-2 in HepGZ cells. The transcription factor NF-KB and the MAP kinase JNK appear to be involved in mediating the effects of PKR on the protein level and the phosphorylation of Bel-2, respectively. I propose that palmitate suppresses these two pathways (Fig. 2.4 and 2.7) by inhibiting PKR (Fig. 2.2) and thereby attenuates the anti-apoptotic machinery (Fig. 2.113). 59 10. References Zhang, R, Romano, P. R, Nagamura-Inoue, T., Tian, B., Dever, T. B., Mathews, M. B., Ozato, K., and Hinnebusch, A. G. (2001) Binding of double-stranded RNA to protein kinase PKR is required for dimerization and promotes critical autophosphorylation events in the activation loop. J Biol Chem 276, 24946-24958 Taylor, S. 8., Haste, N. M., and Ghosh, G. (2005) PKR and eIF2alpha: integration of kinase dimerization, activation, and substrate docking. Cell 122, 823-825 Stark, G. R., Kerr, 1. M., Williams, B. R., Silverman, R. H., and Schreiber, R. D. (1998) How cells respond to interferons. Annu Rev Biochem 67, 227-264 Donze, 0., Deng, J ., Curran, J., Sladek, R., Picard, D., and Sonenberg, N. (2004) The protein kinase PKR: a molecular clock that sequentially activates survival and death programs. EMBOJ 23, 564-571 Kim, S. H., Forrnan, A. P., Mathews, M. B., and Gunnery, S. (2000) Human breast cancer cells contain elevated levels and activity of the protein kinase, PKR. Oncogene 19, 3086-3094 Kim, S. H., Gunnery, S., Choe, J. K., and Mathews, M. B. (2002) Neoplastic progression in melanoma and colon cancer is associated with increased expression and activity of the interferon-inducible protein kinase, PKR. Oncogene 21, 8741- 8748 Hiasa, Y., Kamegaya, Y., Nuriya, H., Onji, M., Kohara, M., Schmidt, E. V., and Chung, R. T. (2003) Protein kinase R is increased and is functional in hepatitis C virus-related hepatocellular carcinoma. Am J Gastroenterol 98, 2528-2534 Delgado Andre, N., and De Lucca, F. L. (2007) Knockdown of PKR expression by RNAi reduces pulmonary metastatic potential of Bl6-F10 melanoma cells in mice: Possible role of NF-kappaB. Cancer Lett 258, 1 18-125 Williams, B. R. (2001) Signal integration via PKR. Sci STKE 2001, RE2 Kumar, A., Haque, J., Lacoste, J., Hiscott, J ., and Williams, B. R. (1994) Double- stranded RN A-dependent protein kinase activates transcription factor NF-kappa B by phosphorylating I kappa B. Proc Natl Acad Sci U S A 91, 6288-6292 60 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. Gil, J., Alcami, J., and Esteban, M. (2000) Activation of NF-kappa B by the dsRNA-dependent protein kinase, PKR involves the I kappa B kinase complex. Oncogene 19, 1369-1378 Zhou, H. R., Lau, A. S., and Pestka, J. J. (2003) Role of double-stranded RNA- activated protein kinase R (PKR) in deoxynivalenol-induced ribotoxic stress response. Toxicol Sci 74, 335-344 Xu, 2., and Williams, B. R. (2000) The B56alpha regulatory subunit of protein phosphatase 2A is a target for regulation by double-stranded RNA-dependent protein kinase PKR. Mol Cell Biol 20, 5285-5299 Watanabe, M. A., Rodrigues Souza, L., Murad, J. M., and De Lucca, F. L. (2003) Antitumor activity induced by regulatory RNA: possible role of RNA—dependent protein kinase and nuclear factor-kappaB. Eur J Pharmacol 465, 205-210 Yeung, M. C., and Lau, A. S. (1998) Tumor suppressor p53 as a component of the tumor necrosis factor-induced, protein kinase PKR-mediated apoptotic pathway in human promonocytic U937 cells. J Biol Chem 273, 25198-25202 Onuki, R., Bando, Y., Suyama, B., Katayama, T., Kawasaki, H., Baba, T., Tohyama, M., and Taira, K (2004) An RNA-dependent protein kinase is involved in tunicamycin-induced apoptosis and Alzheimer’s disease. EMBO J23, 959-968 He, B. (2006) Viruses, endoplasmic reticulum stress, and interferon responses. Cell Death Differ 13, 393-403 Meusel, T. R., Kehoe, K. E., and Imani, F. (2002) Protein kinase R regulates double-stranded RNA induction of TNF-alpha but not IL-1 beta mRNA in human epithelial cells. J Immunol 168, 6429-6435 Goh, K. C., deVeer, M. J ., and Williams, B. R. (2000) The protein kinase PKR is - required for p38 MAPK activation and the innate immune response to bacterial endotoxin. EMBO J 19, 4292-4297 Feldstein, A. B., Werneburg, N. W., Canbay, A., Guicciardi, M. B., Bronk, S. F., Rydzewski, R., Burgart, L. J ., and Gores, G. J. (2004) Free fatty acids promote hepatic lipotoxicity by stimulating TNF-alpha expression via a lysosomal pathway. Hepatology 40, 185-194 Lu, Z. H., Mu, Y. M., Wang, B. A., Li, X. L., Lu, J. M., Li, J. Y., Pan, C. Y., Yanase, T., and Nawata, H. (2003) Saturated free fatty acids, pahnitic acid and 61 22. 23. 24. 25. 26. 27. 28. 29. 30. stearic acid, induce apoptosis by stimulation of ceramide generation in rat testicular Leydig cell. Biochem Biophys Res Commun 303, 1002-1007 Yamagishi, S., Okamoto, T., Amano, S., Inagaki, Y., Koga, K., Koga, M., Choei, H., Sasaki, N., Kikuchi, S., Takeuchi, M., and Makita, Z. (2002) Palmitate-induced apoptosis of microvascular endothelial cells and pericytes. Mol Med 8, 179-184 Hickson-Bick, D. L., Sparagna, G. C., Buja, L. M., and McMillin, J. B. (2002) Palmitate-induced apoptosis in neonatal cardiomyocytes is not dependent on the generation of ROS. Am J Physiol Heart Circ Physiol 282, H656-664 Feldstein, A. B., Werneburg, N. W., Li, Z., Bronk, S. F., and Gores, G. J. (2006) Bax inhibition protects against free fatty acid-induced Iysosomal permeabilization. Am J Physiol Gastrointest Liver Physiol 290, G1339-1346 Giordano, A., Calvani, M., Petillo, 0., Grippo, P., Tuccillo, F., Melone, M. A., Bonelli, P., Calarco, A., and Peluso, G. (2005) tBid induces alterations of mitochondrial fatty acid oxidation flux by malonyl-CoA-independent inhibition of carnitine palmitoyltransferase-l. Cell Death Difler 12, 603-613 Listenberger, L. L., Han, X., Lewis, S. B., Cases, S., Farese, R. V., Jr., Ory, D. S., and Schaffer, J. E. (2003) Triglyceride accumulation protects against fatty acid- induced lipotoxicity. Proc Natl Acad Sci U S A 100, 3077-3082 Li, Z., Srivastava, S., Yang, X., Mittal, S., Norton, P., Resau, J ., Haab, B., and Chan, C. (2007) A hierarchical approach employing metabolic and gene expression profiles to identify the pathways that confer cytotoxicity in HepG2 cells. BMC Syst Biol 1, 21 Wei,.Y., Wang, D., Topczewski, F., and Pagliassotti, M. J. (2006) Saturated fatty acids induce endoplasmic reticulum stress and apoptosis independently of ceramide in liver cells. Am J Physiol Endocrinol Metab 291, 13275-281 Assimacopoulos-Jeannet, F. (2004) Fat storage in pancreas and in insulin-sensitive tissues in pathogenesis of type 2 diabetes. Int J Obes Relat Metab Disord 28 Suppl 4, 853-57 Srivastava, S., and Chan, C. (2006) Hydrogen peroxide and hydroxyl radicals mediate palmitate-induced cytotoxicity to hepatoma cells: Relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 62 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. Barreyro, F. J., Kobayashi, S., Bronk, S. P., Werneburg, N. W., Malhi, H., and Gores, G. J. (2007) Transcriptional Regulation of Bim by FoxO3A Mediates Hepatocyte Lipoapoptosis. J Biol Chem 282, 27141-27154 Malhi, H., Bronk, S. P., Werneburg, N. W., and Gores, G. J. (2006) Free fatty acids induce JNK-dependent hepatocyte lipoapoptosis. J Biol Chem 281, 12093-12101 Feng, L. F., Zhong, M., Lei, X. Y., Zhu, B. Y., Tang, S. S., and Liao, D. F. (2006) Bcl-2 siRN A induced apoptosis and increased sensitivity to 5-fluorouracil and HCPT in HepG2 cells. J Drug Target 14, 21-26 Handayani, T., Sakinah, S., Nallappan, M., and Pihie, A. H. (2007) Regulation of p53-, Bcl-2- and caspase-dependent signaling pathway in xanthorrhizol-induced apoptosis of HepG2 hepatoma cells. Anticancer Res 27, 965-971 Huang, D. S., Shen, K. Z., Wei, J. F., Liang, T. B., Zheng, S. S., and Xie, H. Y. (2005) Specific COX-2 inhibitor NS398 induces apoptosis in human liver cancer cell line HepG2 through BCL—2. World J Gastroenterol 11, 204-207 Landau, Z., Forti, B., Alcaly, M., and Birk, R. Z. (2006) Palmitate induced lipoapoptosis of exocrine pancreas AR42J cells. Apoptosis 11, 717-724 Srivastava, S., Li, Z., Yang, X., Yedwabnick, M., Shaw, S., and Chan, C. (2007) Identification of genes that regulate multiple cellular processes/responses in the context of lipotoxicity to hepatoma cells. BMC Genomics 8, 364 Breckenridge, D. G., and Xue, D. (2004) Regulation of mitochondrial membrane permeabilization by BCL-2 family proteins and caspases. Curr Opin Cell Biol 16, 647-652 Salomoni, P., Perrotti, D., Martinez, R., Franceschi, C., and Calabretta, B. (1997) Resistance to apoptosis in CTLL—2 cells constitutively expressing c-Myb is associated with induction of BCL-2 expression and Myb-dependent regulation of bcl-2 promoter activity. Proc Natl Acad Sci U S A 94, 3296-3301 Taylor, D., Badiani, P., and Weston, K. (1996) A dominant interfering Myb mutant causes apoptosis in T cells. Genes Dev 10, 2732-2744 Weber-Nordt, R. M., Egen, C., Wehinger, J ., Ludwig, W., Gouilleux-Gruart, V., Mertelsmann, R., and Finke, J. (1996) Constitutive activation of STAT proteins in primary lymphoid and myeloid leukemia cells and in Epstein-Barr virus (EBV)- related lymphoma cell lines. Blood 88, 809-816 63 42. 43. 45. 46. 47. 48. 49. 50. 51. D'Alessio, M., Cerella, C., Amici, C., Pesce, C., Coppola, S., Fanelli, C., De Nicola, M., Cristofanon, S., Clavarino, G., Bergamaschi, A., Magrini, A., Gualandi, G., and Ghibelli, L. (2004) Glutathione depletion up-regulates Bcl-2 in BSO-resistant cells. FASEB J 18, 1609-1611 Nishikawa, T., Nakajima, T., Moriguchi, M., Jo, M., Sekoguchi, S., Ishii, M., Takashima, H., Katagishi, T., Kimura, H., Minami, M., Itoh, Y., Kagawa, K., and Okanoue, T. (2006) A green tea polyphenol, epigalocatechin-3 -gallate, induces apoptosis of human hepatocellular carcinoma, possibly through inhibition of Bcl-2 family proteins. J Hepatol 44, 1074-1082 Charlotte, F., L'Hermine, A., Martin, N., Geleyn, Y., Nollet, M., Gaulard, P., and Zafrani, E. S. (1994) Immunohistochemical detection of bcl-2 protein in normal and pathological human liver. Am J Pathol 144, 460-465 Terada, T., and Nakanuma, Y. (1996) Expression of apoptosis, proliferating cell nuclear antigen, and apoptosis-related antigens (bcl-2, c-myc, Fas, Lewis(y) and p53) in human cholangiocarcinomas and hepatocellular carcinomas. Pathol Int 46, 764-770 Zeppa, P., Benincasa, G., Fulciniti, F., Cochand-Priollet, B., Troncone, G., Vetrani, A., Celentano, B., Nardone, G., and Palombini, L. (1996) Apoptosis and cytologic differentiation in hepatocellular carcinoma on fine needle aspiration samples. Acta Cytol 40, 861-866 Zhao, M., Zhang, N. X., Economou, M., Blaha, I., Laissue, J. A., and Zimmermann, A. (1994) Immunohistochemical detection of bcl-2 protein in liver lesions: bcl-2 protein is expressed in hepatocellular carcinomas but not in liver cell dysplasia. Histopathology 25, 237-245 Pianko, S., Patella, S., and Sievert, W. (2000) Alcohol consumption induces hepatocyte apoptosis in patients with chronic hepatitis C infection. J Gastroenterol Hepatol 15, 798-805 Blagosklonny, M. V. (2001) Unwinding the loop of Bcl-2 phosphorylation. Leukemia 15, 869-874 Ito, T., Deng, X., Carr, B., and May, W. S. (1997) Bcl-2 phosphorylation required for anti-apoptosis function. J Biol Chem 272, 11671-11673 De Chiara, G., Marcocci, M. B., Torcia, M., Lucibello, M., Rosini, P., Bonini, P., Higashimoto, Y., Damonte, G., Armirotti, A., Amodei, S., Palamara, A. T., Russo, T., Garaci, B., and Cozzolino, F. (2006) Bc1-2 Phosphorylation by p38 MAPK: 64 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. identification of target sites and biologic consequences. J Biol Chem 281, 21353- 21361 Simizu, S., Tamura, Y., and Osada, H. (2004) Dephosphorylation of Bcl-2 by protein phosphatase 2A results in apoptosis resistance. Cancer Sci 95, 266-270 Srivastava, R. K., Mi, Q. S., Hardwick, J. M., and Longo, D. L. (1999) Deletion of the loop region of Bcl-2 completely blocks paclitaxel-induced apoptosis. Proc Natl Acad Sci USA 96, 3775-3780 Deng, X., Xiao, L., Lang, W., Gao, F., Ruvolo, P., and May, W. S., Jr. (2001) Novel role for JNK as a stress-activated Bc12 kinase. J Biol Chem 276, 23681- 23688 Deng, X., Ruvolo, P., Carr, B., and May, W. 8., Jr. (2000) Survival function of ERK1/2 as IL-3 -activated, staurosporine-resistant Bc12 kinases. Proc Natl Acad Sci USA 97, 1578-1583 Mai, H., May, W. S., Gao, F., Jin, Z., and Deng, X. (2003) A fimctional role for nicotine in Bc12 phosphorylation and suppression of apoptosis. J Biol Chem 278, 1886-1891 Basu, A., You, S. A., and Haldar, S. (2000) Regulation of Bc12 phosphorylation by stress response kinase pathway. Int J Oncol 16, 497-500 .Ruvolo, P. P., Deng, X., Ito, T., Carr, B. K., and May, W. S. (1999) Ceramide induces Bc12 dephosphorylation via a mechanism involving mitochondrial PP2A. J Biol Chem 274, 20296-20300 Peters, T. (1996) All about albumin : biochemistry, genetics, and medical applications, Academic Press, San Diego Pataer, A., Vorburger, S. A., Chada, S., Balachandran, 8., Barber, G. N., Roth, J. A., Hunt, K. K, and Swisher, S. G. (2005) Melanoma differentiation-associated gene-7 protein physically associates with the double-stranded RNA-activated protein kinase PKR. Mol Thar 11, 717-723 Jammi, N. V., Whitby, L. R., and Beal, P. A. (2003) Small molecule inhibitors of the RNA-dependent protein kinase. Biochem Biophys Res Commun 308, 50-57 Eley, H. L., and Tisdale, M. J. (2007) Skeletal muscle atrophy, a link between depression of protein synthesis and increase in degradation. J Biol Chem 282, 7087- 7097 65 63. 65. 66. 67. 68. 69. 70. 71. 72. Gilfoy, F. D., and Mason, P. W. (2007) West Nile virus-induced interferon production is mediated by the double-stranded RN A-dependent protein kinase PKR. J Virol 81, 11148-11158 Koromilas, A. R., Roy, S., Barber, G. N., Katze, M. G., and Sonenberg, N. (1992) Malignant transformation by a mutant of the IFN-inducible dsRNA-dependent protein kinase. Science (New York, NY 257, 1685-1689 Meurs, E. F., Galabru, J., Barber, G. N., Katze, M. G., and Hovanessian, A. G. (1993) Tumor suppressor function of the interferon-induced double-stranded RNA- activated protein kinase. Proceedings of the National Academy of Sciences of the United States of America 90, 232-236 Kumar, A., Yang, Y. L., Flati, V., Der, S., Kadereit, S., Deb, A., Haque, J ., Reis, L., Weissmann, C., and Williams, B. R. (1997) Deficient cytokine signaling in mouse embryo fibroblasts with a targeted deletion in the PKR gene: role of [RF-l and NF- kappaB. Embo J 16, 406-416 Cuddihy, A. R., Li, S., Tam, N. W., Wong, A. H., Taya, Y., Abraham, N., Bell, J. C., and Koromilas, A. E. (1999) Double-stranded—RNA-activated protein kinase PKR enhances transcriptional activation by tumor suppressor p53. Mol Cell Biol 19, 2475-2484 Tamatani, M., Che, Y. H., Matsuzaki, H., Ogawa, S., Okado, H., Miyake, S., Mizuno, T., and Tohyama, M. (1999) Tumor necrosis factor induces Bcl-2 and Bcl- x expression through NFkappaB activation in primary hippocampal neurons. J Biol Chem 274, 8531-8538 Chen, C., Edelstein, L. C., and Gelinas, C. (2000) The Rel/NF-kappaB family directly activates expression of the apoptosis inhibitor Bcl-x(L). Mol Cell Biol 20, 2687-2695 Wang, C. Y., Mayo, M. W., Komeluk, R. G., Goeddel, D. V., and Baldwin, A. S., Jr. (1998) NF-kappaB antiapoptosis: induction of TRAFl and TRAF2 and c-IAPl and c-IAP2 to suppress caspase-8 activation. Science 281, 1680-1683 Wu, H., and Lozano, G. (1994) NF-kappa B activation of p53. A potential mechanism for suppressing cell growth in response to stress. J Biol Chem 269, 20067-20074 Kucharczak, J ., Simmons, M. J ., Fan, Y., and Gelinas, C. (2003) To be, or not to be: NF-kappaB is the answer--role of Rel/NF-kappaB in the regulation of apoptosis. Oncogene 22, 8961-8982 66 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. Li, M., Shillinglaw, W., Henzel, W. J., and Beg, A. A. (2001) The Rela(p65) subunit of NF -kappaB is essential for inhibiting double-stranded RNA-induced cytotoxicity. J Biol Chem 276, 1185-1194 Willis, S. N., and Adams, J. M. (2005) Life in the balance: how BH3-only proteins induce apoptosis. Curr Opin Cell Biol 17, 617-625 Takizawa, T., Tatematsu, C., and Nakanishi, Y. (2002) Double-stranded RNA- activated protein kinase interacts with apoptosis signal-regulating kinase 1. Implications for apoptosis signaling pathways. Eur J Biochem 269, 6126-6132 Ichijo, H., Nishida, B., Irie, K., ten Dijke, P., Saitoh, M., Moriguchi, T., Takagi, M., Matsurnoto, K., Miyazono, K., and Gotoh, Y. (1997) Induction of apoptosis by ASKl, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275, 90-94 Chang, L., and Karin, M. (2001) Mammalian MAP kinase signalling cascades. Nature 410, 37-40 Whitmarsh, A. J ., Cavanagh, J ., Toumier, C., Yasuda, J ., and Davis, R. J. (1998) A mammalian scaffold complex that selectively mediates MAP kinase activation. Science 281, 1671-1674 Davis, R. J. (2000) Signal transduction by the JNK group of MAP kinases. Cell 103, 239-252 Liu, J ., and Lin, A. (2005) Role of JNK activation in apoptosis: a double-edged sword. Cell Res 15, 36-42 Zhang, H., Ozaki, I., Mizuta, T., Yoshimura, T., Matsuhashi, S., Eguchi, Y., Yasutake, T., Hisatomi, A., Sakai, T ., and Yamamoto, K. (2004) Transforming growth factor-beta l-induced apoptosis is blocked by beta l-integrin-mediated mitogen-activated protein kinase activation in human hepatoma cells. Cancer Sci 95, 87 8-886 Liedtke, C., Plumpe, J ., Kubicka, S., Bradham, C. A., Manns, M. P., Brenner, D. A., and Trautwein, C. (2002) Jun kinase modulates tumor necrosis factor-dependent apoptosis in liver cells. Hepatology 36, 315-325 Chen, Y. N., Cheng, C. C., Chen, J. C., Tsauer, W., and Hsu, S. L. (2003) Norcantharidin-induced apoptosis is via the extracellular signal-regulated kinase and c-Jun-NH2-terminal kinase signaling pathways in human hepatoma HepG2 cells. Br J Pharmacol 140, 461-470 67 CHAPTER 3: Construction of Phenotype-Specific Gene Network for Palmitate-Induced Cytotoxicity by Synergy Analysis The work in this chapter has been accepted to be published in the book Systems Analysis of Biological Networks: Yang, X., Wang, X., Dalkic, E., Wu, M., and Chan, 0., Construction of Phenotype-Specific Gene Network by Synergy Analysis. Systems Analysis of Biological Networks, Chapter 5, 2009, in press Abstract Complex cellular activities, such as the saturated free fatty acid (FFA)-induced cytotoxicity, are believed to be coordinately regulated by genes that function in a network. Reconstructing these gene networks can provide insights into the molecular mechanisms of cell physiology and thus represents a fundamental challenge in systems biology. Elucidating the phenotype— gene interaction through the reconstruction of context-specific gene networks remains elusive. In this chapter, I present a methodology that integrates multi-level biological data to infer a cooperative gene network with respect to a specific phenotype, palmitate-induced cytotoxicity. Our method introduces the concept of synergy and builds a network that consists of gene pairs with significant synergistic relations, which implies cooperation in producing cytotoxicity in response to palmitate. Scale-free characteristics and multiple-hub genes are found in the network, revealing many important cooperative candidate genes in regulating the FF A-induced cytotoxicity. These candidates are supported by the literature. 68 Introduction The development of most diseases can be traced to abnormal activities of the cells in specific tissues or parts of the body. Hepatic disorders, such as non alcoholic fatty liver disease (NAFLD) and non-alcoholic steatohepatitis (NASH), are associated with saturated free fatty acid (F FA)-induced cytotoxicity of liver cells (1-3). Cellular activities are tuned by regulatory machineries involving genes, proteins, and metabolites. For instance, saturated FFA-induced cytotoxicity is coordinately regulated by a set of genes that interact in a complex network (4). Therefore, reconstructing the gene networks that give rise to the different phenotypes may provide insights into the cellular mechanisms involved, ultimately, in the deve10pment of diseases and disorders (5, 6). This chapter describes a methodology to reconstruct phenotypic and context- specific gene networks based on the assumption that only a subset of the genes is relevant to the target phenotype. The phenotype addressed in this chapter is saturated FFA- induced cytotoxicity. Methods of gene selection relevant to a phenotype have been based predominantly on fold changes in the genes across different conditions or correlations between the genes and the phenotype, using statistical tests (7) or correlation measures (8), respectively. Statistical tests typically yield too many genes for analysis, while correlation measures only select genes that are statistically correlated with the phenotype, thereby missing potentially relevant genes that are not highly correlated to the phenotype. Incorporating prior information, such as gene set enrichment analysis (4) or trend profiles, into the gene selection methods can help to alleviate these problems, however, the source and quality of the prior knowledge will affect the results. 69 FF As modulate intracellular metabolic pathways involved in glucose (9), triglyceride (TG) (10), and amino acid (1 l), metabolism. Tuned by the gene network (4, 8), some of these alterations are involved in the induction of cytotoxicity by saturated FFAs (9-11). Therefore, integrating multiple levels of information, i.e., gene expression and metabolite profiles, would better reflect the “multi-level” characteristic of cellular activities, such as saturated FFA-induced cytotoxicity, and thereby aid in the selection of genes involved in the observed phenotype, and in the reconstruction of a phenotype- specific gene network. Various methods, such as correlation (12), mutual information (l3, l4), and Bayesian network analysis (15), have been used to construct gene networks. These methods do not directly incorporate the phenotype in identifying the gene interactions. Instead, they typically build a gene network for each of the conditions, and compare the networks to identify the gene interactions that are specific to a condition or phenotype. Consequently, these methods are computationally expensive and sensitive to the quality of data. Since the size (i.e., the number of genes included in the network reconstruction) and the noise level of the samples can affect the networks reconstructed for each of the conditions, it is difficult to determine whether the differences in the networks across the conditions are real changes in the mechanisms or simply an artifact due to the size or noise levels. Alternatively, methods have been developed to select sets of gene pairs relevant to a phenotype based on classification models, such as support vector machine (16, 17), decision tree (18), and probabilistic model (19). Intuitively, if a phenotype- prediction based on a pair of genes performs better than that based on either one of the genes, then the pair of genes is suggested to have cooperative effects on the phenotype. 7O However, these classification methods fail to differentiate the cooperative effects of the genes pairs from the independent contributions of the individual genes (20). To address this short-coming, in collaboration with Xuewei Wang, in this chapter I present a method that distinguishes the difference in the cooperative vs. individual effects of the genes on the phenotype. Biological activities are regulated by multiple factors, many of which function cooperatively, i.e., synergistically. The basic idea is that the whole (i.e., the regulatory system) is greater than the sum of the individual parts (i.e., regulators) of a system. Synergy is defined as the “additional” contribution provided by the “whole” as compared to the sum of the contributions of the individual “parts”. An example of synergy can been seen with transcription factors, such as GAT A4 and dHAND, which together cooperatively and dramatically up-regulate cardiac (target) gene expression levels more significantly than the sum of the effects fi'om either of the transcription factors alone (21). The concept of synergy will be used in this chapter to assess the cooperative effect of two genes on a phenotype. In a multivariate system, the synergistic effect of two factors on a phenotype is the gain in the “mutual information” over the sum of the information provided by each factor on a phenotype. A positive synergy denotes that two factors regulate a phenotype, either cooperatively (e.g., co-activating) or antagonistically (e.g., competitive inhibiting). Thus, one can predict the phenotype with a certain confidence from either of the two factors; however, knowing both factors brings additional information, which enhances the confidence of the prediction. Negative synergy denotes redundancy, thus knowing both factors brings redundant information to the prediction of the phenotype. Zero synergy denotes that at least one of the two factors 71 has no effect on the phenotype, and therefore brings neither additional nor redundant information to the prediction of the phenotype. Systematic assessment of synergy was first applied in neuroscience, where the goal was to understand the neuron code by evaluating the strength of correlations between the neurons upon activation by a stimulus (22, 23). More recently the concept of synergy has been applied to the field of systems biology (24-26). Investigators developed an information theoretic measure of synergy fi'om discretized gene expression data, and applied this measure to identify cooperative gene interactions associated with neural interconnectivity (24) and prostate cancer development (25). More recently, the concept of synergy and the information theoretic measure of synergy have been applied directly to continuous gene expression data (20). In this chapter we introduce an integrative methodology to reconstruct phenotype- specific gene networks based on synergy analysis. First, we select the phenotype-specific genes by integrating the gene expression and metabolite profiles in the context of saturated FFA-induced cytotoxicity. Next, we assess the synergistic effects between the gene pairs. Unlike other computational methods used to identify gene interactions, the fimdamental concept of synergy is to identify the cooperative gene interactions responsible for the phenotype, and these cooperative gene interactions may or may not be direct interactions. Finally, with the identified synergistic gene pairs, we build a synergy network. Topological analyses reveal the structural characteristics of the network while the hub genes provide insights into potential mechanism(s) involved in the induction of the phenotype, i.e., saturated F FA-induced cytotoxicity. 72 Design of the Methodology The Framework of the proposed experiments and methodology, described in a flowchart (Fig. 3.1), for reconstructing the phenotype-specific gene network with synergy analysis are as follows. (1) Identify the phenotype of interest: cytotoxicity levels induced by saturated FFA; (2) Obtain metabolite data; (3) Obtain gene-expression data with cDNA microarray; (4) Select the metabolites that are associated with the phenotype, i.e. cytotoxicity, (5) Select the genes by matching the trend of the metabolite and gene profiles; (6) Obtain synergistic gene pairs by calculating their synergy scores; (7) Construct a synergy network with the gene pairs that are sigrificantly synergistic. In the proposed methodology, the phenotype (cytotoxicity), metabolite and gene expression profiles were collected as “inputs” and integrated as described above. The anticipated results include the representative trends of the metabolites relevant to the phenotype, genes that match the representative trends of the metabolites, and gene pairs with significant synergy scores. Gene pairs that have statistically sigiificant synergy scores indicate possible membership of the gene pairs in a shared pathway or potential cross-talk between different pathways. Graphical representation of these synergistic gene pairs yields a network of gene-gene interactions that are associated with the phenotype. Topology analysis of the synergy network reveals the characteristics of the network, such as degree 73 distribution, modularity and centrality. The hub genes, which have the highest number of edges, suggest that they may be central regulators in the induction of the phenotype. FFA treatment 11 HepGZ cells (‘1/ (2); \(f) Cytotoxicity Metabolites Gene expression (4) | Phenotype-related metabolites | (5) l Phenotype-related genes 1(6) Synergistic gene pairs in) Synergy Network Figure 3.1. Flowchart of the Proposed Methodology. Materials and Methods Cell Culture and Reagents, Fatty Acid Salt Treatment, and Cytotoxicity Measurement—HepG2 cells were cultured and treated by sodium salts of palmitate and oleate as previously described in Chapter 2. Cytotoxicity of the FFAs were measured using the cytotoxicity detection kit (Roche Applied Science, Indianapolis, IN) as previously described in Chapter 2 (27). RNA Isolation for Gene Expression Profiling—Cells were cultured in 10 cm tissue culture plates until confluence and then exposed to different treatments. RNA was 74 isolated with Trizol reagent. In detail, after the treatments, cells were washed twice with cold phosphate buffered saline (PBS). 10 ml trizol reagent was added and kept for 5 - 10 minutes. The cell lysate was then transferred to 15 ml conical tubes and vortexed. Meanwhile, phase-lock gel (heavy) in 15 ml tubes was centrifuged at 1000 g for 10 min. Trizol cell lysate was added to tubes containing phase-lock gel. 3 ml of chloroform was added to the tubes and mixed. The tubes were then centrifuged at 4000 g for 30 minutes to separate the aqueous and organic phases, with the phase-lock gel forming a central layer. The aqueous phase was poured off into fi'esh 15 ml conical tubes. 5 ml of isopropyl alcohol was then added to the tubes containing the aqueous phase and tubes were mixed gently by inversion. Tubes were then centrifuged at 4000 g for 20 minutes. RNA forms a pellet at the bottom. The supernatant liquid is decanted and the pellets were washed thrice with ice-cold 75% ethanol. Ethanol was then carefully removed and the pellet suspended in 1 ml of water and the solution was transferred to 1.5 ml microcenuifuge tubes. For LiCl precipitation, 0.5 ml of 7.5 M LiCl was added to the tubes and vortexed. The tubes were then kept overnight at -20C. The next morning, tubes were centrifuged at 13000 g for 45 min. The pellet was washed thrice with ice-cold 75% ethanol. Finally, all the ethanol was removed and the pellet suspended in 100 ul RNAase-free water. Micro-array Analysis—The gene expression profiles were obtained with cDNA microarray. Analyses were done at the Van Andel Institute, Grand Rapids, MI. The procedure of the microarray analysis was described previously (6). The detailed procedure is available on line (28). In brief, labeled cDNAs were generated with Reverse I Transcriptase (RT) reaction using Cy3 or Cy5 -labeled dCTP and low-dCTP dNTP mix. After generating labeled cDNA, the template was degraded using Rnase and the cDNA 75 further purified using QIAquick columns (Qiagen, Valencia, CA). Hybridization reactions were performed in a 50C water bath for 16 h, following which the microarrays were washed and read. There were two biological replicates for each condition and each replicate was measured with the Cy3 and Cy5 dyes, i.e. there were two technical replicates/color swaps for each biological replicate. Color swaps indicate the arrays in which the cDNA from the treated sample was labeled with Cy3 dye and the cDNA from the control sample was labeled with the Cy5 dye. Metabolites Measurements—The concentrations of the various metabolites were measured according to (l 1, 29-31). The metabolites include glucose, lactate, glycerol, beta-hydroxybutyrate, triglycerides, acetoacetate, Cystine, and all the amino acids. All the measured fluxes were normalized to total protein in the cell extract, measured with the bicinchoninic acid (BCA) method (Pierce Chemicals, Rockford, IL). Gene Selection Based on Trends of Metabolites—The statistical significance of the changes in the metabolite levels across the conditions, i.e. BSA (control), Palmitate and Oleate, were assessed using two-sample t-test for each metabolite. The metabolites altered by saturated FFAs were previously identified (1 1). Eleven metabolites differed sigrificantly across the three conditions, and four representative trends were extracted from these metabolites (Fig. 3.2). 7394 genes remained after removing the EST/hypothetical proteins and ORF of unknown functions fiom the list of ~20,000 genes. Genes with expression patterns that matched the four representative metabolite trends were selected. Two-sample t-test was applied to each gene to assess the sigiificance in 76 their fold change across the different conditions. Finally, 610 genes were selected from the hill list of 7394 genes. The p-value cutoff was set at 0.05. Calculation of the Synergy Scores of Gene Pairs—An information theory based score was calculated to quantify the synergy between the genes (26). Given two genes, G1 and G2, and a phenotype P, the synergy score between G1 and G2 with respect to the phenotype P is defined as: Syn(Gl,G2;P) = I (G1,G2;P) — [I (G1;P) + I (GZ;P)] where I(G1;P) is the mutual information between G1 and P, I(G2;P) is the mutual information between G2 and P, and I(G1,G2;P) is the mutual information between (G1,G2) and P. This equation reflects the definition of synergy, the additional contribution provided by the “whole” as compared to the sum of the contributions of the individual “parts”. Mutual information (I) was calculated using a clustering-based method from continuous data (20). The synergy scores range from [-1 1]. A positive synergy score indicated that two genes jointly provided additional information on the phenotype, a negative synergy score indicated that the two genes provided redundant information about the phenotype, and a zero score indicated that the two genes provided no additional information about the phenotype. The 610 genes that were selected based on the metabolite trends generated 185745 gene pairs, of which 436 pairs had sigrificant synergy scores. Permutation Test to Evaluate the Significance of the Synergy—A permutation test was performed to assess the statistical sigrificance of the synergy of the gene pairs. The phenotypes, i.e., toxic and nontoxic, were randomly permutated to be un-correlated with the gene expression profiles. The synergy scores of the genes were then re-calculated 77 based on the permutated phenotype. This process was repeated 100 times to calculate the p-values of the synergy score for each gene pair. Finally, Benjamin-Hochberg false discovery rate procedure (32) was performed to adjust the p-values for all the gene pairs and thereby control the expected false discoveries. The p-value cutoff was set at 0.05. Characterization of the Network Topology—A synergy network was built with gene pairs that have statistically significant synergy scores. The network composed of nodes that represented the genes, and edges that represented the synergy of the gene pairs. Graph theoretical (topological) analysis of the reconstructed gene networks was used to assess the generated network and how it compared with other biological networks (33). We characterized the topology of the synergy network by its degree distribution and shortest path length. Degree distribution provides a distribution of the number of edges associated with the nodes. Shortest path length is the lowest number of edges that connect two nodes, and is measured using a bread-first search algorithm (34). 78 Figure 3.2. Four Representative Trends of the Metabolites. The metabolites altered by saturated FFAs were previously identified (11). Eleven metabolites differed significantly across the three conditions, and four representative trends were extracted fi'om these metabolites. Trend 1: BSA < Palm and Palm > Ole; Trend II: BSA > Palm and Palm < Ole; Trend 111: BSA < Palm < Ole; Trend IV: BSA > Palm > Ole. 79 ow >_ .82» 20 8.5.". .= 2:5 Sin. 0.0 __ 32p _ .82» 20 Ein— Results Based upon the concept of synergy, we reconstructed a synergy network specifically for the phenotype of saturated FFA-induced cytotoxicity. We first selected the phenotype-relevant genes by integrating the metabolites altered by saturated FFAs (11) with the global gene expression profile and extracted the genes that followed the trends of the metabolites. From the selected genes, the synergy analysis revealed synergistic gene pairs, which were used to build a synergy network. The reconstructed network suggested potential gene targets that may play important roles in the induction of the phenotype. Topological characteristics of the synergy network—The synergy network, shown in Figure 3.3, is composed of 292 genes with 436 connection edges. The synergy network is characterized by relatively short path lengths, ranging fi'om 2 to 10 (Fig. 3.4), while the characteristic path length, or average diameter, of the network is 4.872. The network demonstrates small world characteristics of real networks (35), suggesting that the propagation of communication between the genes is relatively fast. The degree distribution, P(k), provides the probability that a randomly selected vertex has k links to its neighbors. A power law distribution suggests that P(k) ~k'7, where k is the degree and y is the degree exponent, and in most biological, scale-free networks y ranges around 2 and 3 (35). The degree distribution of our synergy network is shown in Figure 3.5, and 7 ~ 2, suggesting the synergy network, similar to other biological networks, is scale-free. Therefore, most of the genes are sparsely connected, while few of the genes (hubs) are connected to many genes and play important roles in sustaining the integity of the network, which suggest their importance in the biological function or the 81 phenotype. In summary, the topology of the synergy network differs from the bell-like Poisson distribution characteristic of a random and statistically homogeneous network and suggests the existence of hub genes. tau-00». Figure 3.3. The Synergy Network. The synergy network is composed of 292 genes with 436 connection edges. The size of the nodes indicates its degree. 20000" 16000- 12000 - 8000 - Frequency 4000‘ 7 8 9 10 11 12 13 1 2 3 4 5 6 Shortest path length Figure 3.4. The Distribution of Shortest Path Lengths in the Synergy Network. 82 N or 2 if g 1.5 - Log(Ndegree) = -1.9Log(degree) + 2.3005 8 e E ‘l l . O 0.5 - o I I I r 0 0.2 0.4 0.6 0.8 1 1 .2 1 .4 1 .6 Log(degree) Figure 3.5. The Degree Distribution of the Synergy Network. For a degee value k, Ndegree is the number of genes with degree k. in the network. Hub genes in the network—The genes in the synergy network are listed and ranked by their degree. Table 3.1 lists the genes with the highest degree, which are therefore “hub genes” in the synergy network. These genes include P4HA1, AHDC1, MACF1, INSIG2, and SH3RF2. Table 3.1. The hub genes in the synergy network. Gene Symbol Degree Full Name P4HA1 22 procollagen-proline, 2-oxoglutarate 4-dioxygenase (proline 4-hydroxylase), alpha polypeptide AHDC1 20 AT hook, DNA binding motif, containing 1 MACF1 19 microtubuIe-actin crosslinking factor 1 lNSlGZ 18 insulin induced gene 2 SH3RF2 13 SH3 domain containing ring finger 2 P4HA1, proline 4-hydroxylase, alpha polypeptide, is the alpha subunit of the protein proline 4-hydroxylase (P4H). P4H is a key enzyme involved in the biosynthesis 83 of collagens (36, 37). Collagens, the most abundant protein in the extracellular matrix (ECM) and the main protein of connective tissues, play important roles in regulating cellular activities and are linked to multiple diseases, including cardiovascular diseases and cancers (3 8-40). In liver, alteration in the synthesis of collagen is related to liver cell apoptosis, hepatic fibrosis and cirrhosis (41 -43). Physiologically, the synthesis, processing, secretion, and deg'adation of collagens are tightly modulated by regulatory factors, including; P4HA1. As a critical functional subunit of P4H, P4HA1 is involved in the post-translational modification of procollagen (36, 44). As shown in Figure 3.6, P4HA1 is located within the endoplasmic reticulum and catalyzes the post-translational formation of 4-hydroxyproline in the -Xaa-Pro-Gly- sequences in procollagens, which is essential for proper folding of the procollagen polypeptide (44). Inhibiting P4H generates unstable intracellular collagens which cannot be secreted (45). On the other hand, over- expressing P4HA1 causes excess synthesis of collagen (46). The deregulation of P4H or P4HA1 in collagen synthesis has been attributed to cytotoxicity and apoptosis in various types of cells (47-50). P4HA1 regulates the ECM components by controlling the synthesis and secretion of procollagens, which modulates fibrosis, cell proliferation, and apoptosis (47-50). The role of P4HA1 in palmitate-induced lipotoxicity is unclear. However in our data, the level of P4HA1 is sigrificantly down-regulated in palmitate as compared to the control (p= 0.0014) and oleate (p=0.018) samples. Therefore, it is plausible that palmitate-induced cytotoxicity is mediated, in part, by P4HA1 through altered synthesis or improper folding of collagen peptides, although experimental validations are needed to confirm this hypothesis. 84 o o 2-ketoglutarate procollagen-L-proline proline 4-hydroxy1ase 0 '1" Fit N H0 0 C 02 + NW + O HO succinate procollagen trans 4-hydroxy-L-proline P4HA1: prolyl 4-hydroxylase is a key enzyme in collagen synthesis Figure 3.6. P4HA1 catalyzes the formation of 4-hydroxyproline. Graph adapted from MetaCyc database (http://www.metacyc.org/) (51). In addition, a number of proteins have been identified or predicted to interact with P4HA1, fi'om publicly available protein-protein interaction databases, such as KEGG (52) and STRING (51, 52). An interaction network (not shown) obtained fi'om STRING shows some of the proteins that potentially interact with P4HA1. P4HA1 is centrally positioned and integates different gene clusters in this network, supporting our result that P4HA1 is a hub gene (with the highest degree, 22) in the synergy network. AHDC1, AT hook, DNA binding motif, containing 1, contains 2 AT hook DNA binding domains, and can be phosphorylated upon DNA damage, probably by ATM or ATR (55). Although the in vivo function or phenotype of this gene has not been identified, it is known that AHDC1 encodes 7 different isoforrns, some of which contain HMG-l and 85 HMG-Y, DNA-binding domains (56). HMG proteins are involved in nucleosome phasing, 3' end processing of mRNA transcripts, and transcription of genes close to AT rich regions, and are thereby related to the pathogenesis of inflammatory and autoimmune diseases (57-59). Although the fimction of AHDC1 is currently unknown, it may be involved in DNA damage and inflammatory responses. The significant alteration of this gene by palmitate (p=0.015 for palmitate vs. BSA and 0.014 for palmitate vs. oleate) and the identification of this gene in the synergy network hints at the possibility that palmitate may affect DNA damage and inflammatory responses through AHDC1. MACF1, microtubule-actin crosslinking factor 1, also called ACF7 (actin cross- linking factor 7), is a member of the spectrale family of cytoskeletal cross—linking proteins that possess actin- and microtubule-binding domains (60, 61). It may be involved in microtubule dynamics to facilitate actin-microtubule interactions at the cell periphery and in coupling the microtubule network to cellular junctions (56). Cell-cell contact and cell-surface interactions through the cytoskeleton and ECM are involved in the control and regulation of cell motility, tissue remodeling, gene expression, differentiation, and proliferation (62). In the literature, a large number of cytoskeletal and ECM genes were found to be down-regulated by palmitate treatment (63, 64). Therefore, it is possible that our method has identified two central genes, i.e. P4HA1 and MACF1, in mediating the effect of palmitate on the ECM and cytoskeletal structure. INSIG2, insulin-induced gene 2, encodes a protein that blocks the processing of sterol regulatory element binding proteins (SREBPs), which regulate human lipogenic and adipocyte metabolism (65). As shown in Figure 3.8, in endoplasmic reticulum (ER), SREBP Cleavage-Activating Protein (SCAP) can bind to the regulation domain of 86 SREBP and transfer SREBP into Golgi Apparatus, where SCAP activates protease SlP to cleave the regulation domain and activate the transcription activation/DNA binding domain of SREBP. Activated SREBP then can be transported to the nucleus to bind to the cis-element SRE to promote the expression of a series of enzymes that are involved in lipid synthesis (66). INSIG2 can bind to SCAP and inhibit its function, thereby blocking lipid synthesis (67). Indeed, reduced INSIG2 levels in adipocytes resulted in SREBP activation, which increased the expression of genes involved in adipogenesis (68). In our system of HepG2 cells, microarray analysis found that the gene expression levels of IN SIG2 were down-regulated by both palrrritate (p=0.011) and oleate (p=0.14). Therefore, the suppression of INSIG2 expression likely contributes to the increased TG synthesis observed in the FFA cultures (10). Figure 3.7. INSIG2 plays important role in lipid synthesis. (66, 67) Figure adapted from Ref (66). By binding to SCAP, INSIG2 blocks the processing of SREBPs and therefore suppresses lipid synthesis. ' 87 SH3RF2, SH3 domain containing ring finger 2, is a putative protein whose function is unknown, but from sequence analysis SH3RF2 contains 3 Src homology 3 (SH3) domains and a RING-type zinc finger domain (Fig. 3.9, from the InterPro database). RING-type zinc finger domain is found in many E3 ubiquitin—protein ligases (69). E3 ubiquitin-protein li gases determine the substrate specificity for ubiquitination and are therefore involved in targeting proteins for degradation by the Ubiquitin- Proteasome System (70). During this process, RING fingers, by interacting with E2 ubiquitin-conjugating enzymes, promotes ubiquitination (71 , 72). Although the exact function of SH3RF2 is not known, the RING-type zinc finger domain suggests SH3RF2 as a putative E3 ubiquitin-protein ligase, which may be involved in the protein degradation pathway through the Ubiquitin-Proteasome System. In addition, SH3RF2 contains another type of domain, SH3 domains, which are present in many proteins involved in intracellular sigral transduction pathways (73, 74). SH3 domains recognize and bind to the proline-rich motifs (-X-P-P-X-P-) on the associated proteins. Therefore, SH3 domains are recruited by the sigraling proteins to direct protein-protein interactions and thereby specify distinct regulatory pathways mediated by different protein binding domains (75, 76). Taken together, the SH3 domains of SH3RF2 could potentially serve as a targeting domain that determines the substrate specificity of SH3RF2 as a putative E3 ubiquitin—protein ligase, therefore facilitating SH3RF2 to subject certain types of proteins to degradation via the ubiquitin-proteasome system. Interestingly, this gene is significantly up-regulated in the palmitate culture (p=0.019 for palmitate vs. BSA and 0.025 for palmitate vs. oleate). This result is consistent with the literature suggesting that palmitate induces protein degradation. 88 ——an RING SH3 m: — Figure 3.8. The Protein Domains of SH3RF2. Figure modified fiom the InterPro database. SH3RF2 contains 3 Src homology 3 (SH3) domains and a RING-type zinc finger domain. Chronic treatment of palmitate increases the levels of unfolded or misfolded proteins, which induces Endoplasmic reticulum (ER) stress (77, 78). This lends support to our finding that P4HA1 was down-regulated in palmitate, suggesting the possibility that procollagen may be misfolded in the palmitate culture as compared to the oleate and control cultures. The accumulation of unfolded or misfolded proteins activates the ubiquitin-proteasome system. Indeed, recent studies found palmitate strongly enhances ubiquitination by activating E3 ubiquitin ligases, resulting in enhanced protein degradation. Therefore, SH3 RF2, putative ubiquitin-protein ligase, may be potentially involved in palmitate-induced cytotoxicity by triggering ubiquitination. In addition, SH3RF2, composed of 3 SH3 and a zinc finger domain which are all protein-protein interaction domains, could recruit a diverse set of proteins for degradation, supporting its central position in our synergy network. Therefore, the synergy network identified a novel protein, SH3RF 2, which potentially plays a central role in palmitate-induced cytotoxicity. The domain knowledge of this protein suggests SH3RF2 may be involved in palmitate-induced cytotoxicity by recruiting unfolded proteins and triggering ubiquitination. In summary, we built a synergy network specifically for palmitate-induced cytotoxicity. This network is scale-free and has multiple hub genes. The hub genes are related to cellular activities, such as cellular contact, cytotoxicity, metabolic pathways, protein degradation, which may play important roles in palmitate-induced cytotoxicity. 89 Therefore, these hub genes suggest potential mechanisms involved in palnnitate-induced cytotoxicity. Discussion Phenotype-specific gene network reconstruction is a useful approach to extract gene interaction information from nnicro-array data and to help provide insight into disease mechanisms. Multiple methods, e. g., meta-analysis (79) and pair-wise relevance (80), have been developed to reconstruct gene networks associated with diseases. The information theoretic measure of synergy provides a convenient method to identify cooperative gene pairs with respect to a phenotype. Therefore in this chapter, we present an alternative strategy to reconstruct phenotype-specific networks based on the concept of synergy. A major concern that often arises in network reconstruction using microarray data is the high computational cost. To alleviate this limitation, we pre-select a subset of genes by matching the trend of their expression profiles, across the different conditions, to the profiles of the phenotype-relevant metabolites. This step reduces the nmnber of genes to be analyzed. Concomitantly, the trend-based analysis allows the incorporation of prior knowledge of the gene expression patterns. In other words, it permits the inclusion of genes of particular interest that are known to be related to the phenotype, whether or not these gene profiles are statistically correlated with the phenotype. Gene pairs with statistically significant synergy scores suggest potential combinatorial effects of those genes on the phenotype. However, the scores cannot distinguish between the types of combinatorial effects, such as additive or antagonistic. 9O Nevertheless, this limitation can be addressed by integating physical interaction data (i.e. protein-protein and protein-DNA interaction) into the analysis. In addition, the proposed analysis pipeline consists of several steps (see Fig. 3.1), including selection of genes relevant to the phenotype, calculation of the synergy scores for each gene pair, evaluation of the statistical significance of the synergy score, and analysis of the network t0pology to identify biologically relevant genes. The methods for these steps are not limited to those proposed in this chapter; alternative methods for each of the steps in the framework can be used. For example, pattern recognition methods can be used to select the genes, discretization-based entropy estimation can be used to calculate the synergy score, Bayesian FDR control procedure can be used to evaluate the statistical significance of the synergy score, etc. A comparison of the different methods for each step could be performed to determine the optimal procedure for each step. In summary, in this chapter we have achieved the following goals: 1). Integrated gene and the metabolite profiles to identify a select group of genes that may be involved in palmitate-induced cytotoxicity; 2). Reconstructed a phenotype-specific synergy network. Topology analysis of the synergy network revealed scale-flee characteristics and multiple hub genes, which are typical characteristics shared by many biological networks. These hub genes suggest potential mecharnisms and may be targets for modulating palmitate-induced cytotoxicity. 91 References Feldstein, A. B., Werneburg, N. W., Canbay, A., Guicciardi, M. B., Bronk, S. F., Rydzewski, R., Burgart, L. J ., and Gores, G. J. (2004) Free fatty acids promote hepatic lipotoxicity by stimulating TN F-alpha expression via a lysosomal pathway. Hepatology 40, 185-194 Lu, Z. H., Mu, Y. M., Wang, B. A., Li, X. L., Lu, J. M., Li, J. Y., Pan, C. Y., Yanase, T., and Nawata, H. (2003) Saturated free fatty acids, palmitic acid and stearic acid, induce apoptosis by stimulation of ceramide generation in rat testicular Leydig cell. Biochem Biophys Res Commun 303, 1002-1007 Farrell, G. C., and Latter, C. Z. (2006) Nonalcoholic fatty liver disease: from steatosis to cirrhosis. Hepatology 43, 899-5112 Li, Z., Srivastava, S., Yang, X., Mittal, S., Norton, P., Resau, J., Haab, B., and Charn, C. (2007) A hierarchical approach employing metabolic and gene expression profiles to identify the pathways that confer cytotoxicity in HepG2 cells. BMC Syst Biol 1, 21 Said, M. R., Begley, T. J ., Oppenheim, A. V., Lauffenburger, D. A., and Samson, L. D. (2004) Global network analysis of phenotypic effects: protein networks and toxicity modulation in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 101, 18006-18011 Srivastava, S., Li, Z., Yang, X., Yedwabnick, M., Shaw, S., and Chan, C. (2007) Identification of genes that regulate multiple cellular processes/responses in the context of lipotoxicity to hepatoma cells. BMC Genomics 8, 364 Tusher, V. G., Tibshirani, R., and Chu, G. (2001) Significance analysis of microarrays applied to the ionizing radiation response. Proc Natl Acad Sci U S A 98, 5116-5121 Li, Z., and Chan, C. (2004) Integrating gene expression and metabolic profiles. J Biol Chem 279, 27124-27137 Lam, T. K., Carpentier, A., Lewis, G. F., van de Werve, G., Fantus, I. G., and Giacca, A. (2003) Mecharnisms of the free fatty acid-induced increase in hepatic glucose production. Am J Physiol Endocrinol Metab 284, E863-873 92 10. ll. 12. l3. 14. 15. l6. 17. 18. 19. 20. 21. Listenberger, L. L., Han, X., Lewis, S. E., Cases, 8., Farese, R. V., Jr., Ory, D. S., and Schaffer, J. E. (2003) Triglyceride accumulation protects against fatty acid- induced lipotoxicity. Proc Natl Acad Sci U S A 100, 3077-3082 Srivastava, S., and Chan, C. (2008) Application of metabolic flux analysis to identify the mechanisms of fi'ee fatty acid toxicity to human hepatoma cell line. Biotechnol Bioeng 99, 399-410 Eisen, M. B., Spellman, P. T., Brown, P. O., and Botstein, D. (1998) Cluster analysis and display of genome-wide expression patterns. Proc Natl Acad Sci U S A 95, 14863-14868 Liang, K. C., and Wang, X. (2008) Gene regulatory network reconstruction using conditional mutual information. E URASIP journal on bioinformatics & systems biology, 253 894 Basso, K., Margolirn, A. A., Stolovitzky, G., Klein, U., Dalla-Favera, R., and Califano, A. (2005) Reverse engineering of regulatory networks in human B cells. Nature genetics 37, 382-390 Pe'er, D., Regev, A., Elidan, G., and Friedman, N. (2001) Inferring subnetworks from perturbed expression profiles. Bioinformatics 17 Suppl 1, 8215-224 Furey, T. S., Cristianini, N., Duffy, N., Bednarski, D. W., Schurnmer, M., and Haussler, D. (2000) Support vector machine classification and validation of cancer tissue samples using microarray expression data. Bioinformatics 16, 906-914 Tang, E. K., Sugantlnan, P. N., and Yao, X. (2006) Gene selection algorithms for microarray data based on least squares support vector machine. BM C Bioinformatics 7, 95 Diaz-Uriarte, R., and Alvarez de Andres, S. (2006) Gene selection and classification of microarray data using random forest. BMC Bioinformatics 7, 3 Paul, T. K., and Iba, H. (2005) Gene selection for classification of cancers using probabilistic model building genetic algorithm. Bio Systems 82, 208-225 Watkinson, J., Wang, X., Zheng, T., and Anastassiou, D. (2008) Identification of gene interactions associated with disease from gene expression data using synergy networks. BMC Syst Biol 2, 10 Dai, Y. S., Cserjesi, P., Markham, B. E., and Molkentin, J. D. (2002) The transcription factors GATA4 and dHAND physically interact to synergistically 93 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. activate cardiac gene expression through a p3 OO-dependent mechanism. J Biol Chem 277, 24390-24398 Schneidrnan, B., Bialek, W., and Berry, M. J., 2nd (2003) Synergy, redundancy, and independence in population codes. J Neurosci 23, 11539-11553 Brenner, N., Strong, S. P., Koberle, R., Bialek, W., and de Ruyter van Steveninck, R. R. (2000) Synergy in a neural code. Neural Comput 12, 1531-1552 Varadan, V., Miller, D. M., 3rd, and Anastassiou, D. (2006) Computational inference of the molecular logic for synaptic connectivity in C. elegans. Bioinformatics 22, e497-506 Varadan, V., and Anastassiou, D. (2006) Inference of disease-related molecular logic fi'om systems-based microarray analysis. PLoS Comput Biol 2, e68 Anastassiou, D. (2007) Computational analysis of the synergy among multiple interacting genes. Mol Syst Biol 3, 83 Srivastava, S., and Chan, C. (2006) Hydrogen peroxide and hydroxyl radicals mediate palmitate-induced cytotoxicity to hepatoma cells: Relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 Institute, V. A. R. Standard Operating Procedures Chan, C., Berthiaume, F., Lee, K., and Yarmush, M. L. (2003) Metabolic flux analysis of hepatocyte function in horrnone- and amino acid-supplemented plasma. Metab Eng 5, 1-15 Olsen, C. (1971) An enzymatic fluorimetric nnicromethod for the determination of acetoacetate, -hydroxybutyrate, pyruvate and lactate. Clin Chim Acta 33, 293-300 Fiskerstrand, T., Refsum, H., Kvalheinn, G., and Ueland, P. M. (1993) Homocysteine and other tlniols in plasma and urine: automated determination and sample stability. Clin Chem 39, 263-271 ’ Benjarnini, Y., and Hochberg, Y. (1995) Controlling the False Discovery Rate - a Practical and Powerful Approach to Multiple Testing. J Roy Stat Soc B Met 57, 289-300 Christensen, C., Gupta, A., Maranas, C. D., and Albert, R. (2007) Large-scale inference and gaph-theoretical analysis of gene-regulatory networks in B-Subtilis. Physica A 373, 796-810 94 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. Cormern, T. H., Leiserson, C. E., and Rivest, R. L. (2001) Introduction to Algorithms, MIT press Barabasi, A. L., and Oltvai, Z. N. (2004) Network biology: Understanding the cell's functional organization. Nat Rev Genet 5, 101-U1 15 Chen, L., Shen, Y. H., Wang, X., Wang, J., Gan, Y., Chen, N., LeMaire, S. A., Coselli, J. S., and Wang, X. L. (2006) Human prolyl-4-hydroxylase alpha(l) transcription is mediated by upstream stimulatory factors. J Biol Chem 281, 10849- 10855 Annunen, P., Autio-Harmainen, H., and Kivirikko, K. I. (1998) The novel type II prolyl 4-hydroxylase is the main enzyme fornn in chondrocytes and capillary endotlnelial cells, whereas the type I enzyme predominates in most cells. J Biol Chem 273, 5989-5992 Bedossa, P., and Paradis, V. (2003) Liver extracellular matrix in health and disease. J Pathol 200, 504-515 Rodriguez-Feo, J. A., Sluijter, J. P., de Kleijn, D. P., and Pasterkarnp, G. (2005) Modulation of collagen turnover in cardiovascular disease. Curr Pharm Des 11, 2501-2514 Stone, P. J. (2000) Potential use of collagen and elastin degradation markers for monitoring liver fibrosis in schistosonniasis. Acta Trop 77, 97-99 Bickel, M., Baringhaus, K. H., Gerl, M., Gunzler, V., Kanta, J ., Schmidts, L., Stapf, M., Tschank, G., Weidmann, K., and Werner, U. (1998) Selective inhibition of hepatic collagen accumulation in experimental liver fibrosis in rats by a new prolyl 4-hydroxylase inhibitor. Hepatology 28, 404-411 Clement, B., Chesne, C., Satie, A. P., and Guillouzo, A. (1991) Effects of the prolyl 4-hydroxylase proinhibitor HOE 077 on human and rat hepatocytes in primary culture. J Hepatol 13 Suppl 3, 841-47 Faouzi, 8., Le Bail, B., Neaud, V., Boussarie, L., Saric, J., Bioulac-Sage, P., Balabaud, C., and Rosenbaum, J. (1999) Myofibroblasts are responsible for collagen synthesis in the stroma of human hepatocellular carcinoma: an in vivo and in vitro study. J Hepatol 30, 275-284 Myllyharju, J. (2003) Prolyl 4-hydroxylases, the key enzymes of collagen biosyntlnesis. Matrix Biol 22, 15-24 95 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. Rocnik, E. F., Chan, B. M., and Pickering, J. G. (1998) Evidence for a role of collagen synthesis in arterial smooth muscle cell migration. J Clin Invest 101, 1889- 1898 John, D. C., Watson, R., Kind, A. J., Scott, A. R., Kadler, K. E., and Bulleid, N. J. (1999) Expression of an engineered form of recombinant procollagen in mouse milk. Nat Biotechnol 17, 385-389 Xia, S. H., Wang, J., and Kang, J. X. (2005) Decreased n-6/n-3 fatty acid ratio reduces the invasive potential of human lung cancer cells by downregulation of cell adhesion/invasion-related genes. Carcinogenesis 26, 779-784 Huet, C., Monget, P., Pisselet, C., and Monniaux, D. (1997) Changes in extracellular matrix components and steroidogenic enzymes during growth and atresia of antral ovarian follicles in the sheep. Biol Reprod 56, 1025-1034 Dong, M. S., Jung, S. H., Kim, H. J., Kim, J. R., Zhao, L. X., Lee, E. S., Lee, E. J., Yi, J. B., Lee, N., Cho, Y. B., Kwak, W. J., and Park, Y. I. (2004) Structure-related cytotoxicity and anti-hepatofibric effect of asiatic acid derivatives in rat hepatic stellate cell-line, HSC-T6. Arch Pharm Res 27, 512-517 Ju, H., Hao, J ., Zhao, S., and Dixon, 1. M. (1998) Antiproliferative and antifibrotic effects of mimosine on adult cardiac fibroblasts. Biochim Biophys Acta 1448, 51-60 Caspi, R., Foerster, H., Fulcher, C. A., Kaipa, P., Krurnmenacker, M., Latendresse, M., Paley, S., Rhee, S. Y., Shearer, A. G., Tissier, C., Walk, T. C., Zhang, P., and Karp, P. D. (2008) The MetaCyc Database of metabolic pathways and enzymes and the BioCyc collection of Pathway/Genome Databases. Nucleic Acids Res 36, D623- 631 Aoki, K. F., and Kanehisa, M. (2005) Using the KEGG database resource. Curr Protoc Bioinformatics Chapter 1, Unit 1 12 von Mering, C., Jensen, L. J., Snel, B., Hooper, S. D., Krupp, M., Foglierini, M., Jouffre, N., Huynen, M. A., and Bork, P. (2005) STRING: known and predicted protein-protein associations, integrated and transferred across organisms. Nucleic Acids Res 33, D43 3-437 von Mering, C., Jensen, L. J ., Kuhn, M., Chaffron, S., Doerks, T., Kruger, B., Snel, B., and Bork, P. (2007) STRING 7--recent developments in the integration and prediction of protein interactions. Nucleic Acids Res 35, D358-362 96 55. 56. 57. 58. 59. 60. 61. '62. 63. 65. Matsuoka, S., Ballif, B. A., Smogorzewska, A., McDonald, E. R., 3rd, Hurov, K. B., Luo, J ., Bakalarski, C. B., Zhao, Z., Solimirni, N., Lerenthal, Y., Shiloh, Y., Gygi, S. P., and Elledge, S. J. (2007) ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316, 1160-1166 Thierry-Mieg, D., and Thierry-Mieg, J. (2006) AceView: a comprehensive cDNA- supported gene and transcripts annotation. Genome Biol 7 Suppl 1, 312 11-14 Voll, R. E., Urbonaviciute, V., Herrmann, M., and Kalden, J. R. (2008) High mobility group box 1 in the pathogenesis of inflammatory and autoimmune diseases. Isr Med Assoc J 10, 26-28 Tesniere, A., Panaretakis, T., Kepp, 0., Apetoh, L., Ghiringhelli, F., Zitvogel, L., and Kroemer, G. (2008) Molecular characteristics of immunogenic cancer cell death. Cell Death Differ 15, 3-12 Jiang, W., and Pisetsky, D. S. (2007) Mechanisms of Disease: the role of high- mobility group protein 1 in the pathogenesis of inflammatory arthritis. Nat Clin Pract Rheumatol 3, 52-58 Kodama, A., Karakesisoglou, I., Wong, B., Vaezi, A., and Fuchs, E. (2003) ACF7: an essential integrator of microtubule dynamics. Cell 115, 343-354 Gong, T. W., Besirli, C. G., and Lomax, M. I. (2001) MACF1 gene structure: a hybrid of plectin and dystrophin. Mamm Genome 12, 852-861 Zarnir, E., and Geiger, B. (2001) Molecular complexity and dynamics of cell-matrix adhesions. J Cell Sci 114, 3583-3590 Draghici, S., Khatri, P., Tarca, A. L., Amin, K., Done, A., Voichita, C., Georgescu, C., and Romero, R. (2007) A systems biology approach for pathway level analysis. Genome Res 17, 1537-1545 Swagell, C. D., Henly, D. C., and Morris, C. P. (2005) Expression analysis of a human hepatic cell line in response to palrrnitate. Biochem Biophys Res Commun 328, 432-441 Krapivner, S., Popov, S., Chemogubova, B., Hellenius, M. L., Fisher, R. M., Hamsten, A., and van't Hooft, F. M. (2008) Insulin-induced gene 2 involvement in human adipocyte metabolism and body weight regulation. J Clin Endocrinol Metab 93, 1995-2001 97 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. Horton, J. B., Goldstein, J. L., and Brown, M. S. (2002) SREBPs: activators of the complete mom of cholesterol and fatty acid synthesis in the liver. J Clin Invest 109,1125-1131 Yabe, D., Brown, M. S., and Goldsteirn, J. L. (2002) Insig-2, a second endoplasmic reticulum protein that binds SCAP and blocks export of sterol regulatory elernent- binding proteins. Proc Natl Acad Sci U S A 99, 12753-12758 Rosen, E. D., Walkey, C. J., Puigserver, P., and Spiegelman, B. M. (2000) Transcriptional regulation of adipogenesis. Genes Dev 14, 1293-1307 Freemont, P. S. (1993) The RING finger. A novel protein sequence motif related to the zinc finger. Ann N Y Acad Sci 684, 174—192 Hershko, A., Heller, H., Eytan, E., and Reiss, Y. (1986) The protein substrate binding site of the ubiquitin-protein ligase system. J Biol Chem 261, 11992-11999 Freemont, P. S. (2000) RING for destruction? Curr Biol 10, R84-87 Barinaga, M. (1999) A new finger on the protein destruction button. Science 286, 223, 225 Cohern, G. B., Ren, R., and Baltimore, D. (1995) Modular binding domains in signal transduction proteins. Cell 80, 237-248 Pawson, T. (1995) Protein modules and signalling networks. Nature 373, 573-580 Ren, R., Mayer, B. J ., Cicchetti, P., and Baltimore, D. (1993) Identification of a ten- arnino acid proline-rich SH3 binding site. Science 259, 1157-1161 Morton, C. J ., and Campbell, I. D. (1994) SH3 domains. Molecular 'Velcro'. Curr Biol 4, 615-617 Guo, W., Wong, S., Xie, W., Lei, T., and Luo, Z. (2007) Palmitate modulates intracellular signaling, induces endoplasmic reticulum stress, and causes apoptosis in mouse 3T3-Ll and rat primary preadipocytes. Am J Physiol Endocrinol Metab 293, E576-586 Karaskov, B., Scott, C., Zhang, L., Teodoro, T., Ravazzola, M., and Volchuk, A. (2006) Chronic palnnitate but not oleate exposure induces endoplasmic reticulum stress, which may contribute to INS-1 pancreatic beta-cell apoptosis. Endocrinology 147, 3398-3407 98 79. 80. Rasche, A., Al-Hasani, H., and Herwig, R. (2008) Meta-analysis approach identifies candidate genes and associated molecular networks for type-2 diabetes mellitus. BMC Genomics 9, 310 Jiang, W., Li, X., Rao, S., Wang, L., Du, L., Li, C., Wu, C., Wang, H., Wang, Y., and Yang, B. (2008) Constructing disease-specific gene networks using pair-wise relevance metric: application to colon cancer identifies interleukin 8, desmin and enolase 1 as the central elements. BMC Syst Biol 2, 72 99 CHAPTER 4: Reconstruct Modular Phenotype-Specific Gene Networks by Knowledge-Driven Matrix Factorization The work in this chapter has been published in Bioinformatics: Yang, X., Zhou, Y., Jin, R. and Chan, C. (2009) Reconstruct Modular Phenotype-Specific Gene Networks by Knowledge-Driven Matrix Factorization, Bioinformatics. Abstract Reconstructing gene networks from rrnicroarray data is believed to be able to provide mechanistic information on cellular processes. A popular structure learning method, Bayesian network inference, has been used to determine network topology despite its shortcomings, i.e., the high computational cost when analyzing a large number of genes and the inefficiency in exploiting prior knowledge, such as the coregulation information of the genes. To address these limitations, in collaboration with computer scientists, we introduced an alternative method, krnowledge-driven matrix factorization (KMF) fiamework, to reconstruct phenotype-specific modular gene networks. Considering the reconstruction of gene network as a matrix factorization problem, we first use the gene expression data to estimate a correlation matrix, and then factorize the correlation matrix to recover the gene modules and the interactions between them. Prior knowledge from Gene Ontology is integrated into the matrix factorization. We applied this KMF algorithm to hepatocellular carcinoma (HepG2) cells treated with free fatty acids (FFAs). By comparing the module networks for the different conditions, we lOO identified the specific modules that are involved in conferring the cytotoxic phenotype induced by palnnitate. Furtlner analysis of the gene modules of the different conditions suggested individual genes that play important roles in palnnitate-induced cytotoxicity. In summary, KMF can efficiently integrate gene expression data with prior knowledge, thereby providing a powerful method of reconstructing phenotype-specific gene networks and valuable insights into the mechanisms that govern the phenotype. Introduction Cellular activities are believed to be coordinately regulated by genes and proteins that function in complex networks. Reconstructing the gene networks that give rise to the different phenotypes may provide insights into the cellular mechanisms involved (1, 2). Biological networks of protein-protein interaction (3), metabolic pathways (4) and transcriptional regulation (5) are modular in structure, enabling mutations to be isolated to specific modules without affecting the overall viability of the system (6-8). Since organized modularity is ubiquitous in biological systems, identifying the gene modules and their interplay in a modular network should provide insights into the differential mechanisms involved in normal vs. disease states. Previously we identified that saturated FFA, e. g. palrrnitate, induced cytotoxicity in liver cells, while unsaturated FFAs, e.g. oleate and linoleate, were not significantly cytotoxic (9-12). Palnnitate-induced cytotoxicity of liver cells has been implicated in the pathogenesis of many obesity-related metabolic disorders, such as non-alcoholic fatty liver disease (NAFLD) and non-alcoholic steatohepatitis (NASH) (13, 14). Tumor necrosis factor (TNF)-o., a proinflarnmatory cytokine often is involved, along with 101 elevated FFA, in these diseases (15), and further potentiated the cytotoxicity induced by palnnitate (2, 9, 10). Multiple mechanisms have been proposed for palnnitate-induced cytotoxicity (16, 17). To study the multi-faceted effects of palmitate and provide insights into potential mecharnism of saturated F FA-induced alterations, we obtained gene expression profiles of hepatocellular carcinoma (HepG2) cells upon exposure to different FFAs and TNF-a, and applied a module-based gene network reconstruction method that integates prior knowledge and phenotypic information. The proposed methodology consists of two phases. The first phase, the “gene selection phase”, selects a subset of genes that are relevant to the phenotype, palnnitate-induced cytotoxicity, using a mixture regession model. The second phase, the “network reconstruction phase”, clusters the selected genes into modules, and reconstructs a module network based upon the interactions between the modules. Selecting the genes that are potentially relevant to the desired metabolic/phenotypic response of the cells can be viewed as a feature selection problem (1 8, 19), which is extensively studied in machine learrning (20, 21). Most feature selection methods, such as the Wilcoxon's rank sum test (22) and Fisher‘s Discriminant Analysis (FDA) (23), are data driven, and thus susceptible to the noise level of the nnicroarray data. One strategy to ameliorate this problem is to incorporate domain knowledge and functional information of the genes (24). Typically these knowledge-based methods qualitatively incorporate the prior krnowledge in post-processing the genes that are selected by the data-driven approaches. In the present work, we address this limitation with a Bayesian mixture regession model that quantitatively incorporates the prior knowledge of the gene firnctions, upfront, in the gene-selection phase. 102 Clustering methods, such as Self-organizing mapping (25, 26) hierarchical clustering (27) and K-means (28), commonly used to identify gene modules, cannot uncover the interactions among the modules or clusters. To address this limitation, several studies integated clustering methods with structure learrning algorithms, such as Graphical Gaussian Modeling and Bayesian network learning (29-31). These approaches are predominantly data-driven and thus susceptible to noise in the expression data, and suffer from the sparse data problem associated with limited number of experimental conditions (32, 33). Previous studies recognized the importance of exploiting prior knowledge in reconstructing networks with sparse and noisy expression data (5, 34-39). Similarly, we developed a fi'amework based on knowledge-driven matrix factorization, termed KMF, to exploit the domain knowledge and reconstruct modular gene networks. This fiamework views the gene network reconstruction as a matrix factorization problem. In brief, the pair-wise correlation coefficients between any two genes are computed from their expression data and used to construct a correlation matrix. This correlation matrix is decomposed into a product of three matrices, from which the gene clusters and module interaction information are extracted. During this process, the Gene Ontology (GO) information is introduced as regularization in matrix factorization, which affects the decomposed matrices and eventually the derived gene modules and interaction among modules. Compared to the existing approaches for gene network reconstruction, the key features of the proposed KMF fiamework are: (1) derives both the gene modules and their interactions from a combination of expression data and GO information, (2) incorporates the prior knowledge of co-regulation relationships into the network 103 reconstruction using a regularization scheme, and (3) presents an efficient learning algorithm based on non-negative matrix factorization and semi-definite pro gamming. The difference between our work and previous works on matrix factorization is that the proposed framework is able to derive both gene modules and their interaction simultaneously while the other matrix factorization methods can only identify gene clusters. KMF Algorithm Gene Selection by Knowledge-driven Bayesian Mixture Regression Model— Regession models are often used for gene selection, in which genes with large assigned weights are deemed to be important for given cellular responses. However, these methods usually suffer from the sparse data problem and are sensitive to noise in the expression data To overcome these limitations, we incorporated the prior knowledge from GO into the regession model with a Bayesian prior. As described previously (40), the main idea is to integate the ontology information of the genes with their expression data (X) to perform unsupervised classification and to identify the genes that regulate the cellular responses (Y). A central assumption behind this method is that the genes within a G0 category should have similar function or effect on a cellular process. Thus, genes belonging to the same GO category were constrained to have similar regession weights. The second problem using the linear regession model for gene selection is that the genes with the largest regession weights may not be specific to the target cellular process since they may also be important to a number of biological processes other than the target process. We addressed this problem by extending the linear regession model into a 104 mixture regression model. The main idea behind the mixture models is to cluster the experimental conditions into two subgoups: the subgroup of conditions that produce the target phenotype, i.e. cytotoxicity, and the subgroup of conditions that do not. A different regession model was built for each subgroup of conditions, and the genes with the largest difference in their regession weights between these two subgoups of conditions were deemed to be specific to the target phenotype. Figure 4.1 shows the absolute difference in their regession coefficients (denoted by DRC) of the genes between the two regession models. The DRC values decrease exponentially for the top ranked genes followed by a slow linear reduction. To identify a subset of genes with large DRC values, we extracted only the genes that are related to the exponential component of the DRC curve. This was done by identifying the point of the DRC curve where the second order derivative starts to approach zero or a negative value. Thus, 250 genes are selected by this process. 0.2 l 0.18 ’ 0.16- 0.14 , 0.12 0.1 0.08 0.06 0.04 0.02 0 abe(DRC) 0 100 200 300 400 500 600 700 Gene Number Figure 4.1. Sorted difference regression coefficients (DRC) for all the genes. Knowledge-driven Matrix Factorization (KMF) Algorithm—KMF is a technique based on matrix factorization. It first computes pair wise correlation between two genes 105 based on their expression levels across different experimental conditions. The matrix of pair wise gene correlation, denoted by W, is approximated by the product of three matrices, M X C X M. A gene modular network, including gene modules and their interaction, is derived from the decomposed matrices M and C. We denote the gene expression data by X = (x1, x2, , xn) where n is the number of genes, and each xi = (xi, 1, xi, 2, , xi,m) E Rm is the expression levels of the ith gene measured under m conditions. We can compute the pairwise correlation between any two genes using statistical correlation metrics such as Pearson correlation, mutual information and chi-square statistics. In our experiment, we use RBF kernel function. This computation results in a symmetric matrix W = [wi,fln X n where wi,j measures the correlation between gene xi and xj. This estimated correlation matrix Wprovides valuable information about the structure of the gene network since a high correlation wi, j between two genes xi and xj could suggest that: 1) genes xi and xj belong to the same module, or 2) gene xi regulates the expression levels of gene xj or vice versa. To derive these two types of interactions simultaneously, we factorize W as W e M x C x MT where M is a matrix of size n X r and C is a matrix of size r X r, and r is the number of modules that can be determined empirically. Matrix M = [miJJn xr represents the memberships of the n genes in r modules where mij Z 0 indicates the confidence of assigning the ith gene to the jth module. Matrix C = [cierxr represents the relationships among r modules where ci,j Z 0 indicates the confidence of the two gene modules to interact (regulate) with each other. Note that in this study, we focus on the undirected network since the gene module regulation matrix C is symmetric. 106 The appropriate factorization of matrix W, the solution of M using the prior knowledge from GO information, and the solution for C by the regularizer lc(C) =||C| |F were performed according to (41). Solving the Optimization Problem—Optimize M by fixing C: The related optimization problem is: arg min Fm(M) = ”W — MGM-r”? + atr(MTL(S)M) MeRan 8.13. M,,j20,i,j=1,2,...,n To find an optimal solution for M, we use the following bound optimization algorithm (42). The key idea is to upper bound the objective firnction using the properties of convex functions, and iteratively update the solution based on the solution of previous iteration. Let M represent the solution of the previous iteration, and our goal is to find a solution of M for the current iteration. Taking the derivative of the upper bound of Fm(M) with respect to Mi, k, we have the optimal solution for M as follows: NIH where am = [MCMTMClak’bak = aMi,sz'7 Cue = a[SM]z‘,k + [WMCk’k 107 Optimize C by fixing M: The related optimization problem is: arg min 77 + fig — 2tr(MTWMC) CERrXr $.13. Ci,i=1,i=1,2,...,7“, QJZQiJZIJVHJ,Ct0 B = MTMC, 77 2 ET: Bi,ij.ia 52 Zr: 029' i,j = 1 i,j = 1 The above problem can be solved using semi-definite progarnming technique (43). The KMF algorithm was applied to toxic conditions and non-toxic conditions, separately. It was also applied to the combination of botln conditions. We denoted by Ct and Cu the interaction matrices of toxic and non-toxic conditions, respectively, and by Call the interaction matrix derived from all the conditions. In order to ensure that matrices Ct and Cu are comparable, we align Ct and Cu with Call. Parameter Tuning—There are three key parameters that can significantly affect the outcome of the proposed algorithm: a, which is used to weight the knowledge from the GO database, [3, which is used to control the sparseness of the interaction matrix C, and the number of modules. Tuning the parameters a and B: In the KMF model, parameter a weights the prior knowledge from GO, and [3 controls the sparseness of the interaction matrix C. To tune these two parameter for optimal result, we apply the supervised learrning technique. First, from the 250 genes, we pick some genes that obviously should belong to the same goup based on their biological functions, and thereby predefined 7 virtual modules composed of 32 genes (see Table 4.1). Second, we gadually change the parameter within a predefined range (i.e., [0.1 . . . 10] in our experiment), and choose the parameters a and 108 B that make the best prediction of the module membership compared to the predefined modules. Determining the number of modules: We determine the optimal number of modules using stability analysis. The basic assumption of stability analysis is that if the estimated number of modules is close to the true number of modules in the data, we would expect the KMF model to yield similar results for matrix C and M with different random initialization. In addition, the stability analysis is applied in a hierarchical fashion. More specifically, in our experiment, the first application of stability analysis results in two major modules. The application of the stability analysis to the two major module yields four and five sub-modules, respectively. 109 Table 4.1. Predefined gene modules. 7 virtual modules composed of 32 genes fiom the 250 genes selected in the gene selection phase. TCA cycle (90) isocitrate dehydrogenase 3 (NAD+) gamma (IDH3G), mRNA. (90) succinate-00A ligase, GDP-forming, beta subunit (gC) succinate-00A ligase, GDP-forming, alpha subunit (SUCLG1), mRNA. (90) citrate synthase (CS), nuclear gene encoding mitochondrial protein, mRNA. Glycolfiis (gC) aldolase C, fructose-bisphosphate (ALDOC), mRNA. (gM) phosphoglycerate kinase 1 (PGK1), mRNA. (90) lactate dehydrogenase C (LDHC), transcript variant 2, mRNA. (gM) glucose phosphate isomerase (GPl), mRNA. (gM) ketohexokinase (fructokinase) (KHK), transcript variant b, mRNA. (90) hexokinase 1 (HK1), transcript variant 5, nuclear gene encoding mitochondrial protein, mRNA. Post-translational: Ubiguitin-Proteasome Pathway (gC) ubiquitin D (UBD), mRNA. (gC) ubiquitin-conjugating enzyme E23 (RAD6 homolog) (UBEZB), mRNA. (9) proteasome (prosome, macropain) subunit, alpha type, 6 (PSMA6), mRNA. (90) proteasome (prosome, macropain) subunit, alpha type, 1 (PSMA1), transcript variant 1, mRNA. (9C) 268 proteasome-associated pad1 homolog Phosphatases (gM) dual specificity phosphatase 3 (vaccinia virus phosphatase VH1-related) (90) protein phosphatase 2, regulatory subunit B (B56), beta isoform (PPP2R5E), mRNA. (90) protein tyrosine phosphatase, non-receptor type substrate 1 (PTPNSt ), mRNA. (90) protein phosphatase 3 (formerly 28), regulatory subunit B, alpha isoform (calcineurin B, type I) (90) protein phosphatase 2 (formerly 2A), catalytic subunit, alpha isoform (PPPZCA), mRNA. (gC) protein phosphatase 2, regulatory subunit B (B56), epsilon isoform (PPP2R5E), mRNA. (gM) protein phosphatase 2, regulatory subunit B (856), gamma isoform (90) protein tyrosine phosphatase, non-receptor type 9 (PTPNQ), mRNA. Translation initiation (gC) eukaryotic translation initiation factor 48 (EIF4B), mRNA. (gC) eukaryotic translation initiation factor 4E binding protein 2 (gN) eukaryotic translation initiation factor 3, subunit 4 delta, 44kDa (EIF3S4), mRNA. (90) eukaryotic translation initiation factor 1A, Y chromosome (ElF1AY), mRNA. Urea cycle (90) argininosuccinate lyase (ASL), mRNA. (90) argininosuccinate synthetase (ASS), transcript variant 2, mRNA. Protein folding, transmrtatlon (gC) heat shock 105kD (HSP1OSB), mRNA. (QC) heat shock 70kDa protein 8 _(gC) heat shock 10kDa protein 1 (chaperonin 10) (HSPE1), mRNA. 110 Experimental Methods Cell Culture and Reagents, Fatty Acid Salt Treatment, and Cytotoxicity Measurement—HepG2 cells were cultured and treated by sodium salts of palrrnitate and oleate as previously described in Chapter 2. Cytotoxicity of the FFAs were measured as previously described in Chapter 2 (44). Gene Expression Profiling—As previously described in Chapter 3 (2), cells were cultured and exposed to different treatments. RNA was isolated with Trizol reagent, and the gene expression profiles were obtained with cDNA microarray at the Van Andel Institute, Grand Rapids, MI. RNA Interference and Reverse T ransfection—Silencer® select predesigned siRNAs targeting human ATP6IP1 and RABGGTA mRN As were purchased from Ambion (Austin, TX). The synthesized oligonucleotides are 5'- AGUCCGAAGAUGUCCCAUATT-3' and 5'-UAUGGGACAUCUUCGGACUTG-3' for the siRNA of ATP6IP1, and 5'-AAAGAAUGCGUGCUUUUAATT-3' and 5'- UUAAAAGCACGCAUUCUUUCT-B' for the siRNA of RABGGTA. Reverse transfection of siRNA was performed as described in Chapter 2 (12), with the transfection reagent, Lipofectamine RNAiMAX (Invitrogen). Western Blot Analysis—HepG2 cell lysate was obtained and Western Blot analysis was performed as previously described in Chapter 2 (12). Anti-RABGGTA, anti- ATP6AP1, and anti-GMPS antibodies were purchased from Abcam Inc. (Cambridge, MA). Anti-beta actin antibody was purchased from Si gna-Aldrich. Secondary anti-rabbit and anti-mouse antibodies were purchased fi'om Pierce Biotechnology Inc. Real-time Quantitative RT -PCR Analysis—Total RNA was extracted from cells as 111 previously described in Chapter 2 (12). Equal amounts of total RNA (1 pg) were reverse- transcribed using an iScript cDNA synthesis kit (Bio-RAD). The primers used for quantitative RT-PCR analyses of human RABGGTA (5'- G'ITTCCTGGAGGTGGATGAG—3' and 5'-GCCAGGAAGAGTAGTTGGAGAA-3'), human HSPIOSB (5'-GGAGTTCCATATCCAGAAGCA-3' and 5'- TCCACCATAGATGCCGTAGA-3'), human GMPS (5'- CAAAGCCTGCACAACAGAAG—3' and 5'—ACTGGAGATTCCACACACGTAA-3'), human ATP6IP1 (5'-CCTGCTCTGCTGCTCATTC-3' and 5'- CGCTGTGTATGGGACATCTT-3 '), and human GAPDH (5'- AACTI'I‘GGTATCGTGGAAGGA-3' and 5'-CAGTAGAGGCAGGGATGATGT-3') were synthesized by Operon Biotechnologies, Inc. (Huntsville, AL). RT-PCR was performed as previously described in Chapter 2 (12). Statistical Analysis—All experiments were perfornned at least tlnree times, and representative results are shown. All data are shown as the mean :t SD. for indicated number of experiments. Student’s t-test was used to evaluate statistical significances between different treatment goups. Statistical significance was set at p<0.01. Results and Discussion We applied the proposed KMF framework to HepG2 cells cultured in different FFAs, with or without TNF-u for 24 hours. 250 genes selected in the gene selection phase (see Methods section) were used to reconstruct the network. Prior knowledge can be incorporated to help reconstruct networks with sparse and noisy expression data (5, 34- 39). Typically, the prior knowledge of the gene interaction is encoded in a Bayesian prior, 112 in which a high probability is given for each gene regulatory relationship derived from prior krnowledge. By incorporating a Bayesian prior, Bayesian Network analysis penalizes any gene relationship (i.e., gives a low score) when it violates the prior krnowledge of the gene relationships, thus improving both the accuracy and efficiency of BN analysis. In this study, the prior knowledge of the genes is taken fiom the GO database. Although GO information does not directly reveal the gene relationships, nevertheless it does provide coregulation relationships and functional information of the genes, both of which are still potentially useful for reconstructing gene networks. Unlike existing methods that apply the GO information to generate pre-defined sets of genes based on supervised feature selection (45), our KMF algorithm applies an alternative unsupervised feature selection, which allows us to identify the feature genes when the classification of the experimental conditions is unknown. In addition, KMF tunes the impact of the GO information on the model selection to obtain optimal results (see Methods section). This is in contrast to the other methods where the GO information takes precedence over the subsequent analysis (40). The KMF algorithm yields two matrices, M and C. M is the module matrix. Each element Mij in matrix M represents the confidence of assigning the ith gene to the jth module. From the module matrix M, we can derive the member genes for each module by only including the genes whose confidences of being assigned to the module exceeds a certain threshold. These member genes will furthermore allow us to infer the overall biological functions of each module. C is the network structure matrix that indicates the connectivity between gene modules. In particular, each element Cij in matrix C represents the strength of the interaction between modules i and j. The interaction information 113 revealed by the C matrix may shed light onto how biological information is processed and passed between different cellular activities. Furthermore, comparing the C matrices for the different conditions suggests structural changes in the module network in response to the toxic conditions, and these changes may confer the cytotoxic phenotype. Application of KMF to identify gene modules and the interactions between the modules—Nine modules (Table 4.2) were identified by KMF. These gene modules are highly enriched with genes involved in specific cellular firnctions or activities. A full list of the genes in each module is available online at http://www.chems.msu.edu/goups/chan/GO_KMFjgenecluster.xls. Next, KMF identified the interactions between the modules, namely the connections between different cellular fimctions, in the form of the C matrix (Table 4.3), and thereby recovered a module network (Fig. 4.2). The bottom row (“sum”) of Table 4.3 sums the correlation coefficients (Cg) between a module with the other 8 modules, tlnereby capturing an overall snapshot of the module connections. A higher “sum” value indicates that the module is more highly correlated with the other modules and thereby takes a more central position in the overall gene module network. A map of the module network is provided in Fig. 4.2, where the strengths of the interactions between the gene modules are indicated by darkness and thickness of the edges. From the C matrix (Table 4.3) and the module interaction network (Fig. 4.2), module 6 (ATP and GTP metabolism) has the highest “sum” value among the 9 modules, and the largest node in the module interaction map. Indeed, as the molecular currency of intracellular energy transfer, ATP (as well as GTP) is either produced or consumed by most of cellular activities, e.g., metabolism (catabolism and anabolism) and signaling 114 pathways. Module 6 has the highest interaction values witln modules 3, 5 and 8 in the C matrix, reflecting that glucose metabolism (module 3) and TCA cycle (module 5) are the major metabolic pathways that produce ATP, the electron transport chain (ETC) (module 5) produces the proton gadient across the mitochondria membrane to provide the driving force for ATP production (46), and amino acid metabolism (module 8) is highly dependent on the ATP levels. Therefore, KMF recovered a high connectivity between ATP (and GTP) synthesis and the major cellular activities that are known to be related to energy production and consumption. Table 4.2. Gene clusters identified by KMF. Module 1 Lipid metabolism and lipid processing 2 Signaling proteins, intracellular and membrane protein-mediated: G protein-coupled receptor signaling, chemokine/TNF-a receptor signaling, ion channel-related signaling Glucose metabolism: glycolysis and pentose phosphate pathway Post-translational modification: ubiquitin-proteasome pathway, protein folding, transportation, phosphorylation/dephosphorylation ROS homeostasis, redox system regulation and the TCA cycle Energy: ATP and GTP metabolism Protein synthesis regulation: translation initiation and transcription Amino acid metabolism and urea cycle Apoptosis: executors and regulators Function uh“ @QNQUI Table 4.3. C Matrix of the clusters. Rows 1-9 represent the interaction between two clusters. Bottom row (sum) is the summation of each column. Clusters 1 2 3 4 5 6 7 8 9 @Q‘IOOI-wa-l 0.152 0.234 0.195 0.191 0.275 0.101 0.236 0.176 1.56 0.152 0.177 0.155 0.152 0.214 0.092 0.183 0.14 1.265 0.234 0.177 0.236 0.215 0.305 0.107 0.284 0.209 1.767 0.195 0.155 0.236 0.204 0.295 0.12 0.249 0.188 1.642 115 0.191 0.152 0.215 0.204 0.302 0.122 0.253 0.186 1.625 0.275 0.214 0.305 0.295 0.302 0.17 0.36 0.267 2.188 0.101 0.092 0.107 0.12 0.122 0.17 0.138 0.108 0.958 0.236 0.183 0.284 0.249 0.253 0.36 0.138 0.227 1 .93 0.176 0.14 0.209 0.188 0.186 0.267 0.108 0.227 1.501 Figure 4.2. Gene module interaction network. Interactions among the nine gene modules are visualized according to the C matrix. The nodes represent modules and the edges indicating the strength of the interaction between modules. A higher Cij value in the C matrix, suggesting stronger interaction, is indicated by a thicker and darker edge line, whereas a higher “sum” value in the C matrix, suggesting more relevant module, is indicated by a larger and darker node. From the C matrix (Table 4.3) and the module interaction network (Fig. 4.2), module 6 (ATP and GTP metabolism) has the highest “sum” value among the 9 modules, and the largest node in the module interaction map. Indeed, as the molecular currency of intracellular energy transfer, ATP (as well as GTP) is either produced or consumed by most of cellular activities, e.g., metabolism (catabolism and anabolism) and signaling pathways. Module 6 has the highest interaction values with modules 3, 5 and 8 in the C matrix, reflecting that glucose metabolism (module 3) and TCA cycle (module 5) are the major metabolic pathways that produce ATP, the electron transport chain (ETC) (module 5) produces the proton gadient across the mitochondria membrane to provide the driving force for ATP production (46), and arrnino acid metabolism (module 8) is highly dependent on the ATP levels. Therefore, KMF recovered a high connectivity between 116 ATP (and GTP) synthesis and the major cellular activities that are krnown to be related to energy production and consumption. Application of KMF to identify the interactions involved in palmitate-induced cytotoxicity—KMF, if applied to the different conditions separately, yields different C matrices specifically for the toxic (saturated FFAs and TNF-a) (Table 4.4) and nontoxic (control, unsaturated FFAs and TNF-a) (Table 4.5) conditions. This is in contrast to the average C matrix obtained using all the conditions discussed above (Table 4.3). Sinnilarly, these condition-specific C matrices indicate module networks composed of interactions between cellular activities for their corresponding condition. The C matrix in the toxic conditions differs significantly fi'om the nontoxic conditions, suggesting that the interactions between the gene modules in the toxic (saturated FFAs and TNF-or) case are altered significantly, and these changes potentially may help explain the phenotype, palmitate-induce cytotoxicity. To quantitatively assess tlnese changes, we subtracted the C matrix for the nontoxic conditions from the C matrix for the toxic conditions, and obtained a matrix we denoted as the “difference C matrix” (Table 4.6). This matrix indicates the differences in the interactions between the gene modules for the toxic vs. the nontoxic conditions. Positive values indicate stronger interactions between the modules under the toxic than the nontoxic conditions, and vice versa. The surmnation of each column in the difference C Matrix indicates the overall difference in the module interactions of a module with the other modules for the toxic vs. the nontoxic conditions. As shown in the difference C matrix (Table 4.6), modules 2, 3, 4 and 5 are more highly connected to the other modules in the toxic than in the nontoxic conditions, while clusters 6 and 9 are less connected in the toxic than in the non-toxic conditions. Since modules 4 117 and 6 have the largest positive and negative “sum” values, 0.144 and -0.294, respectively, we focused on these two modules in the supplemented discussion of their potential involvement in palmitate-induced cytotoxicity. Table 4.4. C Matrix in toxic conditions. Rows 1-9 represent the interaction between two clusters under toxic conditions. Bottom row is the summation of each column. Clusters 1 2 3 4 5 6 7 8 9 1 0.158 0.252 0.218 0.199 0.229 0.088 0.242 0.17 2 0.158 0.212 0.184 0.168 0.197 0.087 0.203 0.137 3 0.252 0.212 0.272 0.252 0.255 0.106 0.291 0.218 4 0.218 0.184 0.272 0.236 0.271 0.128 0.274 0.194 5 0.199 0.168 0.252 0.236 0.281 0.134 0.266 0.181 6 0.229 0.197 0.255 0.271 0.281 0.162 0.3 0.213 7 0.088 0.087 0.106 0.128 0.134 0.162 0.142 0.083 8 0.242 0.203 0.291 0.274 0.266 0.3 0.142 0.216 9 0.17 0.137 0.218 0.194 0.181 0.213 0.083 0.216 sum 1.556 1.346 1.858 1.777 1.717 1.908 0.93 1.934 1.412 Table 4.5. C Matrix in nontoxic conditions. Rows 1-9 represent the interaction between two clusters under nontoxic conditions. Bottom row is the summation of each column. Clusters @ONOUIhNN-h 1 0.153 0.236 0.195 0.191 0.279 0.104 0.236 0.177 1.571 2 0.153 0.174 0.153 0.15 0.214 0.095 0.18 0.139 1.258 3 0.236 0.174 0.234 0.213 0.307 0.108 0.282 0.207 1.761 4 0.195 0.153 0.234 0.201 0.295 0.121 0.246 0.186 1.631 5 0.191 0.15 0.213 0.201 0.302 0.123 0.251 0.185 1.616 6 0.279 0.214 0.307 0.295 0.302 0.172 0.362 0.27 2.201 7 0.104 0.095 0.108 0.121 0.123 0.172 0.139 0.11 0.972 8 0.236 0.18 0.282 0.246 0.251 0.362 0.139 0.225 1 .921 9 0.177 0.139 0.207 0.186 0.185 0.27 0.11 0.225 1 .499 Table 4.6. Difference C Matrix. Subtract the C Matrix of the nontoxic conditions from the toxic conditions. Clusters 1 2 3 4 5 6 7 8 9 CONGOI‘WN-h 0.006 0.016 0.023 0.008 -0.05 -0.016 0.006 -0.007 -0.014 0.006 0.037 0.031 0.017 -0.017 —0.007 0.023 -0.002 0.088 0.016 0.037 0.038 0.039 -0.053 -0.001 0.01 0.01 1 0.097 0.023 0.031 0.038 0.034 -0.024 0.007 0.027 0.008 0.144 0.008 0.01 7 0.039 0.034 -0.021 0.011 0.015 -0.005 0.098 -0.05 -0.017 -0.053 -0.024 -0.021 -0.01 -0.062 -0.057 -0.294 -0.016 -0.007 -0.001 0.007 0.011 -0.01 0.003 -0.027 -0.04 0.006 0.023 0.01 0.027 0.015 -0.062 0.003 -0.01 0.012 -0.007 -0.002 0.01 1 0.008 -0.005 -0.057 -0.027 -0.01 -0.089 118 Module 4: Palnnitate and cytokines, such as TNF-or, have been shown to increase the levels of unfolded or rrnisfolded proteins, which induces endoplasnnic reticulum (ER) stress (47, 48) and triggers the unfolded protein response (UPR) (48, 49) and ubiquitin- proteasome system (50, 51). Indeed, it has been suggested that ER stress together with UPR and ubiquitin-proteasome system serve as self-healing processes for cells upon toxic or stressed conditions (52-54). In addition, palrrnitate-induced ER stress is also correlated with the redistribution of the ER chaperon proteins (48), which are highly involved in post-translational modifications, such as protein transportation and folding, and the chaperon proteins have been suggested to counteract the ER stress induced by palnnitate (55). Therefore, although the mechanism is not fully understood, the ubiquitin- proteasome pathway and the post-translational modifications (folding/unfolding, transportation, degadation) of proteins have been suggested to play important roles in cellular responses to saturated FFAs and cytokines, such as TNF-or. Indeed, module 4 (post-translational modification of proteins: ubiquitin-proteasome pathway, protein folding, transportation, phosphorylation/dephosphorylation) has the highest positive ‘sum’ value in the difference C matrix, reflecting the fact that post-translational modifications of proteins (folding, transportation, and recycling) are highly activated, playing significant roles, and therefore more closely connected with other cellular activities in the toxic conditions. Module 6: Module 6 (ATP and GTP metabolism) has the largest negative ‘sum’ value in the difference C matrix, indicating that ATP metabolism is less correlated with the other cellular processes in the toxic than the non-toxic conditions. Long-term exposure of saturated FFAs, e.g. palrrnitate, leads to activation of the uncoupling proteins 119 (UCP) (56-58). Located in the mitochondrial inner membrane, UCPs act as proton channels to dissipate the proton gadient before it is used by ATP synthases to produce ATP. Thus, UCPs uncouple mitochondrial oxidative phosphorylation and produce heat instead of ATP (58, 59). Therefore in the toxic conditions, the highly activated UCPs may serve as anotlner key regulator that controls the level of ATP, in addition to the other cellular processes that produce proton gradients or consume ATP, such as glucose and amino acid metabolism and TCA cycle. As a result, with this additional regulation through UCPs, the level of ATP is less connected with the cellular activities identified by KMF, but more related to regulators such as UCPs in the toxic than the nontoxic conditions. This could provide an explanation for the values in column 6 (ATP and GTP metabolism) being negative in the difference C matrix (Table 4.6). In summary, the proposed KMF algorithm identified the gene modules and their interactions, as well as how they change in the toxic vs. non-toxic conditions. The results suggested that post-translational modification and uncoupling proteins (U CP) may play important roles in mediating the palmitate/TNF-a induced cellular responses, thereby shedding light on potential mechanisms involved in palnnitate-induced cytotoxicity. Thus far, this methodology has focused on the module network. To further uncover the specific genes that may be responsible for the palrrnitate-induced cytotoxicity, we performed further analysis to assess the contribution of each gene in the two gene modules that were deemed important. Identifying potential genes responsible for palmitate-induced cytotoxicity—As described above, the values in the M matrix (My) indicate the strength or contribution of gene i to module j. The rank of the genes in a module by their My- values provides a 120 relative index of the importance of a gene to the cellular function that corresponds to that module. Under different conditions, the modules remained relatively stable with respect to their size and gene members, however, the rank of certain genes changed significantly in some of the modules. This suggests that the importance or the weights of these genes in their corresponding modules varied across the different conditions; therefore these genes may play important roles in conferring a phenotype. Table 4.7. Top 10 out of 33 genes in module 4. Top 10 out of 33 genes in module 4 ranked according to their contributions to the module under toxic conditions. The ranking difference was calculated by subtracting the ranking number of the specific gene under toxic conditions from non-toxic conditions. Positive ranking difference numbers indicate bigger ranking numbers and less contribution in non-toxic conditions. Rank in Rank In Rank toxic non-toxic difference Gene 1 4 3 leucine carboxyl methyltransferase (LCMT) 2 5 3 mitogen-activated protein kinase kinase kinase 12 (MAP3K12) 3 3 0 protease, serine, 2 (trypsin 2) (PRSSZ) 4 2 -2 suppression of tumorigenicity 13 (colon carcinoma) (ST13) 5 25 20 heat shock 105kD (HSP1058) 6 6 0 apolipoprotein C-l (APOC1) 7 30 23 Rab geranylgeranyltransferase, alpha subunit (RABGGTA) 8 14 6 UV radiation resistance associated gene (UVRAG) 9 7 _2 dolichyl-phosphate mannosyltransferase polypeptide 2, regulatory subunit (DPM2) 10 23 13 mitogen-activated protein kinase-activated protein kinase 3 (MAPKAPKS) Given the importance of modules 4 (Post-translational modification of proteins) and 6 (ATP and GTP metabolism), we ranked the genes in these two modules according to their Mi,- values in the toxic conditions. The top 10 out of 33 genes in module 4 and all the genes in module 6 are listed in Tables 4.7 and 4.8, respectively. The ranking numbers 121 of these genes in the toxic and nontoxic conditions are listed, as is the difference in the rankings of the genes between these conditions (Tables 4.7 and 4.8). Table 4.8. All 10 genes in module 6. The genes in module 6 ranked according to their contributions to the module under toxic conditions. The ranking difference was calculated by subtracting the ranking number of the specific gene under toxic conditions from non- toxic conditions. Positive ranking difference numbers indicate bigger ranking numbers and less contribution in non-toxic conditions. Rank in Rank in Rank toxic non-toxic difference Gene 1 5 4 ATP citrate lyase (ACLY) 1 7 6 guanine monphosphate synthetase (GMPS) 1 8 7 ATPase, H+ transporting, Iysosomal interacting protein 1 (ATP6IP1) 1 1 0 ATP synthase, H+ transporting, mitochondrial F0 complex, subunit c (subunit 9), isoform 2 (ATPSG2) _3 ATP synthase, H+ transporting, mitochondrial F0 complex, 5 2 subunit F6 (ATPSJ) 6 4 -2 adenosine kinase (ADK), transcript variant ADK-long 7 3 -4 adenylosuccinate synthase 8 9 1 adenosine monophosphate deaminase (isoform E) (AMPD3) 9 6 -3 Iysosomal apyrase-like 1 10 10 0 ATPase, H+ transporting, Iysosomal 13kDa, V1 subunit G isoform 1 (ATP6V1G1) In module 4, the positions of two genes, Rab geranylgeranyltransferase (RABGGTA) and heat shock 105kD (HSPlOSB) changed significantly in the toxic conditions, as indicated by high positive differences in rankings, 23 and 20 respectively (Table 4.7), suggesting that tlnese two genes may be more involved in the toxic than non- toxic conditions and tlnereby play a role in conferring the toxic phenotype. RABGGT catalyzes the transfer of a geranyl-geranyl moiety from geranyl-geranyl pyrophosphate to Rab proteins (GTPases) such as RABlA, RAB3A and RABSA (60). As a member of the Ras superfarnily of monomeric G proteins, Rab proteins regulate membrane traffic, which facilitates the trafficking of cell membrane proteins from the Golgi to the plasma 122 membrane and the recycling of the membrane proteins (61 , 62). RABGGT, by facilitating the prenylation of Rab proteins (60), ensures that the Rab proteins are insoluble and correctly anchored in the membrane. The response of RABGGT to saturated FF As and its potential role, if any, in the saturated F F A-induced cytotoxicity has never been studied. The mRNA level of RABGGT is not affected by oleate and increased by palmitate albeit insignificantly (Fig. 4.3A). However, furtlner analysis by silencing the gene expression level of RABGGT revealed a very interesting feature of RABGGT in regulating cytotoxicity (Fig. 4.33). In the non-toxic conditions, i.e. BSA (vehicle of the FFAs) or oleate, the LDH release was increased by the siRN A of RABGGT (Fig. 4.33), suggesting that RABGGT may help to maintain normal healthy cellular activities under physiological and non-toxic conditions. Indeed, membrane traffic patlnways, regulated by RABGGT through Rab GTPases, are important in maintaining normal vesicle formation and movement and membrane protein trafficking and recycling. By contrast, in the toxic condition, i.e. palrrnitate, the LDH release was decreased by the siRNA of RABGGT (Fig. 4.38), suggesting that RABGGT may be involved in mediating the cytotoxic effect of palmitate. The potential mechanism of the distinct roles of RABGGT under the different conditions is unclear at this point. Given that RABGGT catalyzes the prenylation and therefore the activation of Rab GTPases, we hypothesize that the toxic conditions (i.e., palnnitate) induce disordered trafficking and recycling of the membrane proteins and disrupt the membrane integity, through RABGGT and Rab proteins, thereby enhancing the cytotoxicity. Containing a HSP70-like motif, HSPIOSB belongs to the Hsp105/ 110 family, a diverged subgoup of the 70-kDa heat shock protein (Hsp70) family (63). HSP105B has 123 been found to inhibit the aggegation of denatured proteins under severe stress conditions (64). In addition, HSP105B is usually associated with Hsp70 or Hsc70 (a constitutive form of Hsp70) in mammalian cells and suppresses the chaperone activity of Hsp70/Hsc70 (65). The role of HSP105B in cytotoxicity and apoptosis has been suggested to be cell type-dependent (66—69). HSP105 suppressed staurosporine-induced apoptosis in HeLa cells (66) and stress-induced apoptosis in neuronal PC12 cells (67). However, HSP105 was also found to enhance the apoptosis induced by oxidative stress in mouse embryonal F9 cells (68) and promote ER stress-induced caspase-3 activation, an indicator of apoptosis (69). As an important chaperon protein involved in processing denatured proteins under stress conditions (64, 65), HSP105B may also be involved in r the cellular responses induced by the toxic conditions, potentially by regulating the post- translational modifications, such as denaturation, folding/unfolding, transportation and degadation. Indeed, we found that both the mRNA (Fig. 4.4A) and protein (Fig. 4.4B) expression levels of HSP105B are significantly increased by palmitate but not by oleate, suggesting that this gene potentially plays a role in the cytotoxicity induced by saturated FFA. More detailed investigation is needed to clarify the exact role of HSP105 in palnnitate-induced cytotoxicity. 124 P 1.6 < 'g 1.4- m I; 1.2 a . E 1: 1 '- U 3 g 0.8 1 e E E 0.6 — < "' 0.4 - E 0.2 - E 0 BSA Palm Ole B. 16% M DNeg. Ctrl. 312% . IsiRABGGTA a **4: 0 80/ . 2 ° * E .1 4% 0% BSA ' Palm ‘ Ole Figure 4.3. Effects of the fatty acids on the expression level of RABGGTA and the role of RABGGTA in cytotoxicity. HepG2 cells were exposed to 0.7 mM palmitate or oleate for 24 hours (A). After treatment, the cells were harvested, and RT-PCR analysis was performed to detect the mRNA levels of RABGGTA (A). Reverse transfection of suspended HepG2 cells were performed with scrambled siRNA ([1, negative control) or siRNA of RABGGTA (I, siRABGGTA) for 24 hours and the transfected cells were then cultured in 0.7 mM pahnitate or oleate for another 24 hours (B). Cells were then harvested, and the LDH release was assayed (B). In all the experiments with the FFAs, the vehicle (0.7 mM BSA) was used as the control (i.e., regular media with BSA). Data expressed as average of nine samples i SD from three independent experiments. One- way ANOVA with Tukey’s post hoc method was used to analyze the differences between treatment groups. *, **, ***, significantly higher (* and ***) or lower (**) than negative control, i.e., scrambled siRNA, p<0.0l. 125 P 12 a: g 10 - * D. A g 81 8— .. 5 0 s: 6 _ a 0 In; E 4 _ i8 “é " 0 _ BSA Palm Ole B. HSP105B beta Actin BSA Palm. Ole. Figure 4.4. Effects of the fatty acids on the expression level of HSP105B. HepG2 cells were exposed to 0.7 mM palmitate or oleate for 24 hours. The vehicle for the FFAs (0.7 mM BSA) was used as the control (i.e., regular media with BSA). After treatment, the cells were harvested, and RT-PCR (A) and western blot analysis (B) were performed to detect the mRNA (A) and the protein (B) expression levels of HSP105B. Gene expression data expressed as average of nine samples :t SD from three independent experiments. Student’s t-test was used to analyze the differences between treatment goups. *, significantly higher than control, i.e., regular media with BSA, p<0.01. In module 6, the positions of two genes, ATPase, H+ transporting, Iysosomal interacting protein 1 (ATP6IPl) and guarnine monophosphate synthetase (GMPS) are changed significantly by the toxic conditions, indicated by their highest positive difference rankings, 7 and 6 respectively (see Table 4.8), suggesting that these two genes may be involved in conferring the toxic phenotype. As an essential component of most eukaryotic cells, ATP6IP1 is located on the vacuole membrane, responsible for acidifying 126 vacuoles by transporting H+ into the vacuoles at the expense of ATP (70, 71). Vacuoles are involved in removing and recycling unwanted or harmful substances, such as misfolded proteins and foreign invaders such as bacteria (72, 73). Together with lysosomes, vacuoles play major roles in autophagy and maintaining the balance between biogenesis and degadation of many cellular products. It has been shown that palmitate induces Iysosomal pernneabilization, which contributes to the cytotoxicity induced by palmitate (16, 74), however the effect of palmitate on the vacuole membrane has not been reported. In fact, vacuoles and lysosomes share similarities in their structures, internal pH and major functions, and these two organelles sometimes fuse together to exchange their internal substances. Considering the similarity between vacuoles and lysosomes, and more importantly based on our result that suggests this vacuole membrane protein ATP6IP1 plays an important role in the toxic (palmitate) conditions, we hypothesize that I palrrnitate may similarly perturb the vacuole membrane, and thereby interrupt the internal pH, and induce vacuole permeabilization, which may lead to apoptosis and cytotoxicity. Since ATP6IP1 is the major membrane protein that produces the pH difference across the vacuole membrane, we further propose that palmitate perturbs the vacuole membrane by interacting with ATP6IP1. The mRNA (Fig. 4.5A) and protein (Fig. 4.5B) expression levels of ATP6IP1 were not significantly affected by either palnritate or oleate. However, silencing the expression of ATP6IP1 with the siRNA of ATP6IPl decreased the cytotoxicity induced by palmitate, as evidenced by the LDH release (Fig. 4.56). This result suggests that down-regulating vacuole membrane protein reduces palnnitate- induced cytotoxicity, suggesting that palnrnitate may alter the vacuole membrane and induce cytotoxicity through the vacuole membrane protein ATP6IP1. 127 E 1.6 — g A 1.4 — '2 g, 1.2 — - 5 1 - 3 5 08 O . r > 2 33 0.6 — < a; z 0.4 “ “E5 0.2 — 0 _ BSA Palm Ole C. 16% ** EINeg. Ctrl. B. . A ElsnATP6AP1 WBI °\o 2 ' I 3 BSA Palm. Ole. BSA Palm Ole Figure 4.5. Effects of the fatty acids on the expression level of ATP6IP1 and the role of ATP6IP1 in cytotoxicity. HepG2 cells were exposed to 0.7 mM palmitate or oleate for 24 hours (A, B). After treatment, the cells were harvested, and RT-PCR (A) and western blot analysis (B) were performed to detect the mRNA (A) and the protein (B) expression levels of ATP6IPl. Reverse transfection of suspended HepG2 cells were performed with scrambled siRNA (D, negative control) or siRNA of ATP6IP1 (I, siATP6IPl) for 24 hours and the transfected cells were then cultured in 0.7 mM palmitate or oleate for another 24 hours (C). Cells were then harvested, and the LDH release was assayed (C). In all the experiments with the FFAs, the vehicle (0.7 mM BSA) was used as the control (i.e., regular media with BSA). Data expressed as average of nine samples i SD from three independent experiments. One-way ANOVA with Tukey’s post hoc method was used to analyze the differences between treatment goups. *, significantly lower than negative control, i.e., scrambled siRNA, p<0.01. GMPS is the key enzyme involved in the de novo synthesis of guanine nucleotides by catalyzing the amination of xantlnosine monophosphate (XMP) to 128 guanosine monophosphate (GMP) (75). Guanine nucleotides are essential for DNA and RNA synthesis, and GMP is also the substrate for GTP production, which is a source of energy, similar to ATP, especially in protein synthesis. Another major role of GTP is mediating signal transduction pathways, particularly G-protein signaling pathways (76). Therefore, in addition to RNA and DNA replication, GMPS is also potentially involved in other cellular processes by regulating the GTP level. The regulation of guanine nucleotide synthesis by GMPS in liver cell cytotoxicity has never been studied. Interestingly, we found that the mRNA (Fig. 4.6A) and protein (Fig. 4.68) expression levels of GMPS were both significantly enhanced by the saturated FFA, palnnitate, but not by unsaturated FFA, oleate. Currently it is unclear what role this enzyme plays in the toxicity induced by saturated FFA. However, the changes at the gene and protein expressions of GMPS in response to pahnitate suggested that GMPS may play a role in saturated FFA induced cellular activities. 129 P u (D * n. 3“ 5325 For 0 g 2. '6: 531a 3e1— % 0e 0. BSA Palm Ole B. GMPS BSA Palm. Ole. Figure 4.6. Effect of the fatty acids on the expression level of GMPS. HepG2 cells were exposed to 0.7 mM palmitate or oleate for 24 hours. The vehicle for the FFAs (0.7 mM BSA) was used as the control (i.e., regular media with BSA). After treatment, the cells were harvested, and RT-PCR (A) and western blot analysis (B) were performed to detect the mRNA (A) and the protein (B) expression levels of GMPS. Gene expression data expressed as average of nine samples i SD fiom three independent experiments. Student’s t-test was used to analyze the differences between treatment groups. * , significantly higher than control, i.e., regular media with BSA, p<0.01. Conclusions KMF was used to reconstruct a gene module network, composed of fimctional gene modules and their interactions. Comparing the gene module networks for the different conditions revealed changes in module interactions across the conditions. Our results showed that modules 2-6 and 9 played important roles in palnnitate-induced cytotoxicity. These modules covered most of the popular areas of the research on the cytotoxicity induced by saturated F FAs and TNF-or in liver cells, including the regulation 130 of apoptosis pathways (module 9) (12, 16, 77-79) and redox system (module 5) (10, 80). In a separate study, we evaluated how saturated FFAs and TNF-or affected some of the genes in module 9, such as PKR and Bcl-2 family proteins (11, 12), and found that the gene expression level of Bcl-2 was suppressed by palmitate and TNF-a through PKR (l 1, 12), providing a potential mechanism by which palmitate- and T'NF-or-induced cytotoxicity. We also investigated some genes in module 5, such as NADH dehydrogenases, which we found are also highly involved in palmitate-induced cytotoxicity by inducing ROS production (80). In addition, assessing the contribution of genes within some of the highly relevant modules (4 and 6) revealed potential genes that may be involved in palmitate-induced cytotoxicity. Furtlner experiments confirmed the involvement of these genes in conferring the phenotype, palmitate-induced cytotoxicity, and suggested novel research targets for addressing the palmitate-induced cytotoxicity. In summary, by quantitatively integating gene expression profile with prior knowledge extracted from GO database, KMF provides a powerful tool to reconstruct modular and phenotype-specific gene networks that elucidate possible mechanisms involved in producing a phenotype. 131 10. References Said, M. R., Begley, T. J ., Oppenheim, A. V., Lauffenburger, D. A., and Samson, L. D. (2004) Global network analysis of phenotypic effects: protein networks and toxicity modulation in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 101, 18006-1801 1 Srivastava, S., Li, Z., Yang, X., Yedwabnick, M., Shaw, S., and Chan, C. (2007) Identification of genes that regulate multiple cellular processes/responses in the context of lipotoxicity to hepatoma cells. BMC Genomics 8, 364 Han, J. D. (2008) Understanding biological functions through molecular networks. Cell Res 18, 224-237 Ravasz, E., Somera, A. L., Mongu, D. A., Oltvai, Z. N., and Barabasi, A. L. (2002) Hierarchical orgarnization of modularity in metabolic networks. Science 297, 1551- 1555 Ilnmels, J ., Friedlander, G., Bergnann, S., Sarig, 0., Ziv, Y., and Barkai, N. (2002) Revealing modular organization in the yeast transcriptional network. Nat Genet 31, 370-377 Thieffry, D., and Romero, D. (1999) The modularity of biological regulatory networks. Biosystems 50, 49-59 Yook, S. H., Oltvai, Z. N., and Barabasi, A. L. (2004) Functional and topological characterization of protein interaction networks. Proteomics 4, 928-942 Jeong, H., Tombor, B., Albert, R., Oltvai, Z. N., and Barabasi, A. L. (2000) The large-scale orgarnization of metabolic networks. Nature 407, 651-654 Li, Z., Srivastava, S., Mittal, S., Yang, X., Sheng, L., and Chan, C. (2007) A Three Stage Integative Pathway Search (TIPS) framework to identify toxicity relevant genes and pathways. BMC Bioinformatics 8, 202 Srivastava, S., and Chan, C. (2007) Hydrogen peroxide and hydroxyl radicals mediate palnnitate-induced cytotoxicity to hepatoma cells: relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 132 ll. 12. l3. 14. 15. l6. 17. 18. 19. 20. 21. 22. Srivastava, S., Li, Z., Yang, X., Yedwabnick, M., Shaw, S., and Chan, C. (2007) Identification of genes that regulate multiple cellular processes/ responses in the context of lipotoxicity to hepatoma cells. BMC Genomics 8, 364-379 Yang, X., and Charn, C. (2009) Repression of PKR mediates palrrnitate-induced apoptosis in HepG2 cells through regulation of Bcl-2. Cell Res Scheern, A. J ., and Luyckx, F. H. (2002) Obesity and liver disease. Best Pract Res Clin Endocrinol Metab 16, 703-716 Farrell, G. C., and Latter, C. Z. (2006) Nonalcoholic fatty liver disease: from steatosis to cirrhosis. Hepatology 43, S99-S112 Bruce, C. R., and Dyck, D. J. (2004) Cytokine regulation of skeletal muscle fatty acid metabolism: effect of interleukin-6 and tumor necrosis factor-alpha. Am J Physiol Endocrinol Metab 287, E616-621 Feldstein, A. B., Werneburg, N. W., Canbay, A., Guicciardi, M. E., Bronk, S. P., Rydzewski, R., Burgart, L. J ., and Gores, G. J. (2004) Free fatty acids promote hepatic lipotoxicity by stimulating TNF-alpha expression via a lysosomal pathway. Hepatology 40, 185-194 Wang, T., and Weinman, S. A. (2006) Causes and consequences of mitochondrial reactive oxygen species generation in hepatitis C. J Gastroenterol Hepatol 21 Suppl 3, S34-37 Ressom, H. W., Varghese, R. S., Zhang, Z., Xuarn, J., and Clarke, R. (2008) Classification algorithms for phenotype prediction in genomics and proteonrics. Front Biosci 13, 691-708 Saeys, Y., Inza, 1., and Larranaga, P. (2007) A review of feature selection techniques in bioinformatics. Bioinformatics 23, 2507-2517 Bhaskar, H., Hoyle, D. C., and Singh, S. (2006) Machine learning in bioinformatics: a brief survey and recommendations for practitioners. Comput Biol Med 36, 1104- 1 125 Inza, 1., Larranaga, P., Blanco, R., and Cerrolaza, A. J. (2004) Filter versus wrapper gene selection approaches in DNA microarray domains. ArtifIntell Med 31, 91-103 Troyanskaya, O. G., Garber, M. E., Brown, P. O., Botstein, D., and Altman, R. B. (2002) Nonparametric methods for identifying differentially expressed genes in microarray data. Bioinformatics 18, 1454-1461 133 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. Chan, C., Hwang, D., Stephanopoulos, G. N., Yarrnush, M. L., and Stephanopoulos, G. (2003) Application of multivariate analysis to optimize function of cultured hepatocytes. Biotechnol Prog 19, 580-598 Le Phillip, P., Bah], A., and Ungar, L. H. (2004) Using prior knowledge to improve genetic network reconstruction from rrnicroarray data. In Silico Biol 4, 335-353 Yirn, L., Huang, C. H., and Ni, J. (2006) Clustering of gene expression data: performance and similarity analysis. BMC Bioinformatics 7 Suppl 4, S19 Toronern, P., Kolehmainen, M., Wong, G., and Castren, E. (1999) Analysis of gene expression data using self-organizing maps. FEBS Lett 451, 142-146 Eisen, M. B., Spellman, P. T., Brown, P. 0., and Botstein, D. (1998) Cluster analysis and display of genome-wide expression patterns. Proc Natl Acad Sci U S A 95, 14863-14868 Ma, S. F., Grigoryev, D. N., Taylor, A. D., Nonas, S., Sarnmarni, S., Ye, S. Q., and Garcia, J. G. (2005) Bioinforrnatic identification of novel early stress response genes in rodent models of lung injury. Am J Physiol Lung Cell Mol Physiol 289, L468-477 Toh, H., and Horimoto, K. (2002) Inference of a genetic network by a combined approach of cluster analysis and gaphical Gaussian modeling. Bioinformatics 18, 287-297 Segal, E., Shapira, M., Regev, A., Pe'er, D., Botsteirn, D., Koller, D., and Friedmarn, N. (2003) Module networks: identifying regulatory modules and their condition- specific regulators from gene expression data. Nat Genet 34, 166-176 Li, Z., and Chan, C. (2004) Inferring patlnways and networks with a Bayesian framework. FASEB J 18, 746-748 Husmeier, D. (2003) Sensitivity and specificity of inferring genetic regulatory interactions from microarray experiments with dynamic Bayesian networks. Bioinformatics 19, 2271-2282 Yu, J., Snnith, V. A., Wang, P. P., Hartemink, A. J., and Jarvis, E. D. (2004) Advances to Bayesian network inference for generating causal networks from observational biological data. Bioinformatics 20, 3594-3603 Bar-Joseph, Z., Gerber, G. K., Lee, T. 1., Rinaldi, N. J., Yoo, J. Y., Robert, R, Gordon, D. B., Fraenkel, E., Jaakkola, T. S., Young, R. A., and Gifford, D. K. 134 35. 36. 37. 38. 39. 40. 41. 42. 43. (2003) Computational discovery of gene modules and regulatory networks. Nat Biotechnol 21, 1337-1342 Berman, B. P., Nibu, Y., Pfeiffer, B. D., Tomancak, P., Celrniker, S. E., Levine, M., Rubin, G. M., and Eisern, M. B. (2002) Exploiting transcription factor binding site clustering to identify cis-regulatory modules involved in pattern formation in the Drosophila genome. Proc Natl Acad Sci U S A 99, 757-762 Hartemink, A. J., Gifford, D. K., Jaakkola, T. S., and Young, R. A. (2002) Combining location and expression data for principled discovery of genetic regulatory network models. Pac Symp Biocomput, 43 7—449 Ideker, T., Thorsson, V., Ranish, J. A., Christmas, R., Buhler, J., Eng, J. K., Bumgarner, R., Goodlett, D. R., Aebersold, R., and Hood, L. (2001) Integrated genomic and proteomic analyses of a systematically perturbed metabolic network. Science 292, 929-934 Pilpel, Y., Sudarsanarn, P., and Church, G. M. (2001) Identifying regulatory networks by combinatorial analysis of promoter elements. Nat Genet 29, 153-159 Li, R, and Yang, Y. (2004) Recovering genetic regulatory networks fi'om rrnicro- array data and location analysis data. Genome Inform 15, 131-140 Srivastava, S., Zhang, L., Jin, R., and Chan, C. (2008) A novel method incorporating gene ontology information for unsupervised clustering and feature selection. PLoS ONE 3, e3860 Yang, X., Zhou, Y., Jin, R., and Chan, C. (2009) Reconstruct Modular Phenotype- Specific Gene Networks by Knowledge-Driven Matrix Factorization. Bioinformatics Zhou, Y., Li, Z., Yang, X., Zhang, L., Srivastava, S., Jin, R., and Chan, C. (2008) Using knowledge driven matrix factorization to reconstruct modular gene regulatory network. In AAA] Conference on Artificial Intelligence pp. 811-816, The AAAI Press, Menlo Park, CA., Chicago, IL Boyd, S. P., and Vandenberghe, L. (2004) Convex optimization, Cambridge University Press, Cambridge, UK ; New York Srivastava, S., and Chan, C. (2006) Hydrogen peroxide and hydroxyl radicals mediate palrrnitate-induced cytotoxicity to hepatoma cells: Relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 135 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. Subramaniarn, A., Tarnayo, P., Mootha, V. K., Mukherjee, S., Ebert, B. L., Gillette, M. A., Paulovich, A., Pomeroy, S. L., Golub, T. R., Lander, E. S., and Mesirov, J. P. (2005) Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci U S A 102, 15545-15550 Lehninger, A. L., Nelson, D. L., and Cox, M. M. (2005) Lehninger principles of biochemistry, W.H. Freeman, New York Guo, W., Wong, S., Xie, W., Lei, T., and Luo, Z. (2007) Palmitate modulates intracellular signaling, induces endoplasmic reticulum stress, and causes apoptosis in mouse 3T3-L1 and rat primary preadipocytes. Am J Physiol Endocrinol Metab 293, E576-586 Karaskov, E., Scott, C., Zhang, L., Teodoro, T., Ravazzola, M., and Volchuk, A. (2006) Chronic palnnitate but not oleate exposure induces endoplasmic reticulrun stress, which may contribute to INS-1 pancreatic beta-cell apoptosis. Endocrinology 147, 3398-3407 Zhang, Z., Lee, Y. C., Kim, S. J., Choi, M. S., Tsai, P. C., Xu, Y., Xiao, Y. J., Zhang, P., Heifer, A., and Mukherjee, A. B. (2006) Palrrnitoyl-protein tlnioesterase-l deficiency mediates the activation of the unfolded protein response and neuronal apoptosis in INCL. Hum Mol Genet 15, 337-346 Zhou, Q., Du, J., Hu, Z., Walsh, K., and Wang, X. H. (2007) Evidence for adipose- muscle cross talk: opposing regulation of muscle proteolysis by adiponectin and Fatty acids. Endocrinology 148, 5696-5705 Ding, W. X., Ni, H. M., Gao, W., Yoshimori, T., Stolz, D. B., Ron, D., and Yin, X. M. (2007) Linking of autophagy to ubiquitin-proteasome system is important for the regulation of endoplasmic reticulum stress and cell viability. Am J Pathol 171, 513-524 Zhao, L., and Ackennan, S. L. (2006) Endoplasmic reticulum stress in health and disease. Curr Opin Cell Biol 18, 444-452 Shen, X., Zhang, K., and Kaufinan, R. J. (2004) The unfolded protein response--a stress signaling patlnway of the endoplasmic reticulum. J Chem Neuroanat 28, 79- 92 Ron, D. (2002) Translational control in the endoplasmic reticulum stress response. JClin Invest 110, 1383-1388 136 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. Lai, B., Bikopoulos, G., Wheeler, M. B., Rozakis-Adcock, M., and Volchuk, A. (2008) Differential activation of ER stress and apoptosis in response to chronically elevated free fatty acids in pancreatic beta-cells. Am J Physiol Endocrinol Metab 294, E540-550 Gonzalez-Barroso, M. M., Fleury, C., Arechaga, 1., Zaragoza, P., Levi-Meyrueis, C., Raimbault, S., Ricquier, D., Bouillaud, F., and Rial, E. (1996) Activation of the uncoupling protein by fatty acids is modulated by mutations in the C-terrninal region of the protein. Eur J Biochem 239, 445—450 Larneloise, N., Muzzin, P., Prentki, M., and Assimacopoulos-Jeannet, F. (2001) Uncoupling protein 2: a possible link between fatty acid excess and impaired glucose-induced insulin secretion? Diabetes 50, 803-809 Nedergaard, J ., Ricquier, D., and Kozak, L. P. (2005) Uncoupling proteins: current status and therapeutic prospects. EMBO Rep 6, 917-921 Breen, E. P., Gouin, S. G., Murphy, A. F., Haines, L. R., Jackson, A. M., Pearson, T. W., Murphy, P. V., and Porter, R. K. (2006) On the mechanism of mitochondrial uncoupling protein 1 function. J Biol Chem 281, 2114-2119 Leung, K. F., Baron, R., and Seabra, M. C. (2006) Thematic review series: lipid posttranslational modifications. geranylgeranylation of Rab GTPases. J Lipid Res 47, 467-475 Seabra, M. C., Mules, E. H., and Hume, A. N. (2002) Rab GTPases, intracellular traffic and disease. Trends Mol Med 8, 23-30 Stenmark, H., and Olkkonen, V. M. (2001) The Rab GTPase family. Genome Biol 2, REVIEWS3007 Ishihara, K., Yasuda, K., and Hatayama, T. (1999) Molecular cloning, expression and localization of human 105 kDa heat shock protein, hsp105. Biochim Biophys Acta 1444, 138-142 Yamagishi, N., Ishihara, K., Saito, Y., and Hatayama, T. (2003) Hsp105 but not Hsp70 family proteins suppress the aggegation of heat-denatured protein in the presence of ADP. FEBS Lett 555, 390-396 Yamagishi, N., Nishihori, H., Ishihara, K., Ohtsuka, K., and Hatayama, T. (2000) Modulation of the chaperone activities of Hsc70/Hsp40 by Hsp105alpha and Hsp105beta. Biochem Biophys Res Commun 272, 850-855 137 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. Yamagishi, N., Ishihara, K., Saito, Y., and Hatayama, T. (2006) Hsp105 family proteins suppress staurosporine-induced apoptosis by inhibiting the translocation of Bax to mitochondria in HeLa cells. Exp Cell Res 312, 3215-3223 Hatayama, T., Yamagishi, N., Minobe, E., and Sakai, K. (2001) Role of hsp105 in protection against stress-induced apoptosis in neuronal PC12 cells. Biochem Biophys Res Commun 288, 528-534 Yamagishi, N., Saito, Y., Ishihara, K., and Hatayama, T. (2002) Enhancement of oxidative stress-induced apoptosis by Hsp105alpha in mouse embryonal F9 cells. Eur J Biochem 269, 4143-4151 Meares, G. P., Zmijewska, A. A., and Jope, R. S. (2008) HSP105 interacts with GRP78 and GSK3 and promotes ER stress-induced caspase-3 activation. Cell Signal 20, 347-358 Nelson, N. (1987) The vacuolar proton-ATPase of eukaryotic cells. Bioessays 7, 251-254 Nelson, N. (1992) Evolution of organellar proton-ATPases. Biochim Biophys Acta 1100, 109-124 Ganong, W. F. (2003) Review of medical physiology, McGraw-Hill, New York London Alberts, B. (2002) Molecular biology of the cell, Garland Science, New York Acosta, D., and Wenzel, D. G. (1974) Injury produced by fiee fatty acids to lysosomes and nnitochondria in cultured heart muscle and endothelial cells. Atherosclerosis 20, 417-426 Nakarnura, J ., and Lou, L. (1995) Biochemical characterization of human GMP synthetase. J Biol Chem 270, 7347-7353 Berg, J. M., Tymoczko, J. L., and Stryer, L. (2002) Biochemistry, W.H. Freeman, New York Lu, Z. H., Mu, Y. M., Wang, B. A., Li, X. L., Lu, J. M., Li, J. Y., Pan, C. Y., Yanase, T., and Nawata, H. (2003) Saturated free fatty acids, palnnitic acid and stearic acid, induce apoptosis by stimulation of ceramide generation in rat testicular Leydig cell. Biochem Biophys Res Commun 303, 1002-1007 138 78. 79. 80. Landau, Z., Forti, E., Alcaly, M., and Birk, R. Z. (2006) Palnnitate induced lipoapoptosis of exocrine pancreas AR42J cells. Apoptosis 11, 717-724 Barreyro, F. J., Kobayashi, S., Bronk, S. F., Werneburg, N. W., Malhi, H., and Gores, G. J. (2007) Transcriptional Regulation of Bim by FoxO3A Mediates Hepatocyte Lipoapoptosis. J Biol Chem 282, 27141-27154 Li, Z., Srivastava, S., Findlan, R., and Chan, C. (2008) Using Dynamic Gene Module Map Analysis To Identify Targets That Modulate Free Fatty Acid Induced Cytotoxicity. Biotechnol Prog 24, 29-37 139 CHAPTER 5: PKR Differentially Regulates IRS1 and IRSZ in HepG2 Cells The work in this chapter has been submitted for publication: Yang, X., Opperman, M. and Chan, 0., PKR Differentially Regulates IRS1 and lRSZ in HepGZ Cells, submitted to Molecular and Cellular Biology, under review Abstract Irnitially identified to be activated upon virus infection, the double-stranded RNA- dependent protein kinase (PKR) is best krnown for triggering cell defense responses by phosphorylating eIF-Za, thus suppressing RNA translation. Here, I uncovered a novel function of PKR in regulating insulin signaling activity. I found that PKR up-regulates the inhibitory phosphorylation of IRS1 at Ser312, which is believed to suppress the activity of IRS1. This effect of PKR on the phosphorylation of IRS1 is mediated by 2 other protein kinases, JNK and IKK. By contrast, PKR regulates IRSZ, another predominant IRS family protein in the liver, at the transcriptional rather than the post- translational level, and this effect is mediated by a transcription factor, FoxOl, which plays a critical role in hyperglycemia during insulin resistance. In summary, I found, for the first time, that PKR regulates the transmitters of insulin signaling, IRS1 and IRS2, tlnrough different mechanisms. In addition, I as well as other researchers showed that PKR activity is down-regulated by insulin through the IRS proteins. Therefore this 140 feedback loop between PKR and IRS proteins may have potential implications in the regulation of insulin signaling and the development of insulin resistance. Introduction Insulin signaling, a central signaling pathway that regulates many cellular activities, such as glucose and lipid metabolism, protein synthesis and degadation, cell gowth and differentiation (1), has been extensively studied over the past decades. Insulin signaling is initiated upon binding of insulin to the insulin receptor (IR), a receptor tyrosine kinase (2), and transmitted intracellularly by the insulin receptor substrates (IRS) (1, 2). At least 4 of the IR substrates belong to the IRS goup, with IRS1 and IRS2 being predominant and expressed in most tissues, including the liver (3, 4). Upon phosphorylation of the tyrosine residues catalyzed by IR, the IRS proteins initiate, through different binding mechanisms (4), various downstream signal transduction cascades, including mitogen-activated protein kinase (MAPK) pathways (c-Jun N- terrninal kinase (JNK), extracellular signal-related kinase (ERK), and p3 8) (5, 6) and phosphoinositide 3-kinase (PI3 K) (7), which in turn activates Akt/protein kinase B (Akt/PKB) (8), and atypical protein kinase C (aPKC) (9). Insulin signaling is sophisticatedly tuned by a large number of regulators. Pathologically, dysregulation of insulin signaling is closely related to the development of insulin resistance (10), and contributes to multiple diseases and disorders, such as type 2 diabetes as well as other metabolic, endocrine and cardiovascular disorders (IO-12). At the molecular level, dysregulation of insulin signaling could occur at several possible stages, e.g., degadation or mutation of IR (13, 14), inhibitory phosphorylation or 141 degadation of IRS (15, 16), or suppression of down-stream signaling molecules, such as PI-3 kinase or Akt/PKB (reviewed in ref. (17)). However, most of the insulin signaling disruption and insulin resistance have been attributed to dysregulation of the phosphorylation of IR and, in particular, IRS (15, 16). [RS proteins mediate intracellular insulin signaling, through the tyrosine residues, which facilitate recruitment of IRS substrates and promote insulin signaling (4), and the serine residues, which generally suppress the activities of IRS by blocking the interaction between IRS and IR (18), inhibiting the tyrosine phosphorylation of IRS (19), or inducing the degadation of IRS (20). A number of serine residues have been identified to negatively regulate the activity of IRS], in particular, Ser307 (equivalent to Ser312 in human IRSl). Ser307 has been extensively investigated and characterized as a key indicator of inhibitory phosphorylation of IRS1 and insulin resistance, and confirmed in insulin-resistant rodent models (21). Although the mechanisms are not well understood, a number of factors, such as cerarrnide (22, 23), and Hepatitis C viral (HCV) infection (24, 25), have been associated with the down-regulation of insulin signaling and development of insulin resistance in liver cells, potentially through their regulation of IRS1 Ser312 phosphorylation. Interestingly, these factors share a common link; they both can activate the double- stranded RNA-dependent protein kinase (PKR). Cerarrnide induces the phosphorylation of PKR by activating the PKR activator X (RAX) (or its human homolog PACT) (26). HCV infection induces phosphorylation and activation of PKR, mediated by the HCV core protein, which directly binds and activates PKR (27). As an anti-viral protein, PKR is best known for triggering the cell’s defense response and initiating the innate immune 142 response by arresting general protein synthesis and inducing apoptosis during virus infection (28, 29). In addition, as a serine-threonine protein kinase that is ubiquitously expressed, PKR is actively involved in multiple signaling pathways by regulating the activities of transcription factors, protein kinases, and phosphatases (30). Therefore, PKR serves as a signal integator, regulating various cellular activities, such as protein syntlnesis and degadation, stress and immune responses, and apoptosis (30, 31). Thus far, it has not been determined whether PKR plays a role in regulating insulin signaling. Given that ceramide and HCV, botln known as activators of PKR, have been identified as regulators of insulin signaling and inducers of insulin resistance by promoting inhibitory phosphorylations of IRS1, I hypothesize that PKR plays a role in the induction of insulin resistance by phosphorylating IRS1 at Ser312, a key indicator of inhibitory phosphorylation of IRS1 and insulin resistance. A number of signaling molecules, such as Ich kinase (IKK) (32), mammalian target of raparnycin (mTOR) (33), PKCC (34), S6 Kinase 1 (S6Kl) (35) and JNK (36, 37) have been shown to function as IRS serine kinases that induce insulin resistance by promoting the inhibitory phosphorylation of IRS (reviewed in (15, 38)). PKR, as a signal integator of many intracellular signaling events (3 O), has been shown to activate certain IRS kinases, e.g., IKK (39) and JNK (40, 41). Therefore, I investigated whether PKR is involved in the induction of inhibitory Ser312 phosphorylation of IRS1, and if so, which IRS kinase mediates the phosphorylation. Since IRS1 and IRS2 are both expressed in liver cells, I investigated whether PKR also affected IRS2. Our results indicated that PKR regulates the gene transcriptiorn, rather than the post-translational modification, of IRS2. The trarnscription of the IRS2 gene has 143 been shown to be up-regulated by several transcription factors, such as forkhead box 01 (FoxOl) (42, 43) and cAMP response element binding (CREB) (44). I found that PKR enhanced the expression of IRS2 through FoxOl, which controls the transcription of IRS2 but not IRS1. PKR did not affect the activity of CREB. In addition to showing that PKR differentially regulates the two major IRS proteins, IRS1 and IRS2. I confirmed that insulin suppressed the activity of PKR in liver cells, mediated by IRS, thereby identifying PKR as a downstream target of insulin signaling. Indeed, a recent study showed insulin suppressed the activity of PKR in muscle cells (45). Taken together, our results suggest a novel feedback mechanism between PKR and the IRS proteins. Materials and Methods Cell Culture and Reagents—HepG2 cells were cultured as previously described in Chapter 2 (46). Human insulin and 2-aminopurine (2-AP) were purchased from Signa- Aldrich (St. Louis, MO), C2 ceramide, Tautomycetin, SC-514, okadaic acid (0A), PKR inhibitor, JNK inhibitor (SP600125) and their analogues, used as negative controls, from EMD Biosciences (San Diego, CA). Insulin treatment—Human insulin was stocked in HEPES buffer, which was therefore used in controls for all the experiments with the insulin treatment. I treated the cells with insulin at the concentrations lower than 1 nM to mimic the physiological concentrations (47). At 95% confluence, cells were deprived of serum for 16 hours prior to each experiment and subjected to insulin treatments for the indicated doses and times at 37 °C in serum-free medium. 144 Chemical inhibitors—In the present study, I used the commercially available 2- AP and PKR inhibitor as one of the tools to elucidate the role of PKR. However, even though these chemical have been widely used in the studies of PKR for a variety of systems, the specificity of these inhibitor has not been extensively tested in the literature. Keeping in mind the potential non-specific targets of these PKR inhibitors, 1 performed more PKR gene silencing and over-expressing studies to test our hypothesis and draw the conclusions. The JNK inhibitor, SP600125, IKK inhibitor, SC-514, PPlc inhibitor, TMC, and PP2A inhibitor, OA are popularly used and proven to be specific to their targets at the concentrations I used (48-51). RNA interference for PKR and reverse transfection—Silencer® Validated siRN A targeting human PKR mRNA and Silencer® Pre-designed siRNAs targeting human FoxOl, IRS1 and IRS2 mRNAs were purchased fiom Ambion (Austin, TX). The synthesized oligonucleotides are 5'-GGUGAAGGUAGAUCAAAGATT-3' and 5'- UCUUUGAUCUACCUUCACCTT-3' for siRNA of PKR, 5'- CCCAAGAGCAUGCACAAACTT-B' and 5'-GUUUGUGCAUGCUCUUGGGTI‘-3' for siRNA of IRS1, 5'-GCGAGUACAUCAACAUCGATT-3’ and 5'- UCGAUGUUGAUGUACUCGCCG-3' for siRNA of IRS2, and 5'- GCUCAAAUGCUAGUACUAUTT-3' and 5'-AUAGUACUAGCAUUUGAGCTA-3' for siRNA of FoxOl. reverse transfection of siRNA was performed with the transfection reagent, Lipofectamine RNAiMAX (Invitrogen) Reverse transfection of siRNA was performed as described in Chapter 2 (40), with the transfection reagent, Lipofectannine RNAiMAX (Invitrogen). 145 Over-expression of PKR and forward transfection—The PKR plasmid, pCMV6- XLS-hPKR, and the empty vector, pCMV6-XL5, were purchased from Origene (Rockville, MD). In general, as described previously in Chapter 2 (40), regular HepG2 cells (Figs. 5.3C, 5.5C) or cells with the FoxOl gene silenced by the siRNA of FoxOI (Fig. 5.5E) were subjected to forward transfections of the plasmids according to the Lipofectamine 2000 (Invitrogen) method. After 6 hours of trarnsfection, the cells were then cultured in regular media for 42 hours, and harvested (Fig. 5.5E) or treated with other chemicals (Figs. 5.3C, 5.5C) subsequently. Western blot analysis and immunoprecipitation— HepG2 cell lysate was obtained and Western Blot analysis was performed as previously described in Chapter 2 (40). Phospho site-specific anti-IRS1 (Tyr94l), PPPlA (Thr320), and anti-PPPlA antibodies were purchased from Abcam (Cambridge, MA), phospho site-specific anti-IKKa/B (Serl 76/ 180), FoxOl (Ser256), anti-biotin, anti-IKKB, and anti-FoxOl antibodies from Cell Signaling (Danvers, MA), phospho site-specific anti-IRS1 (Ser312), IRS2 (Ser731), PKR (Thr451), JNK (T183/Y185), anti-IRS], anti-IRS2, anti-PKR, anti-JNK, and anti- beta actin antibodies from Signa-Aldrich. Secondary anti-rabbit and anti-mouse antibodies were purchased from Pierce Biotechnology Inc. Real-time quantitative RT -PCR analysis—Total RNA was extracted from cells with the RNeasy mini kit (Qiagen, Valencia, CA) and depleted of contaminating DNA with RNase-free DNase (Qiagen). Equal amounts of total RNA (1 ug) were reverse- transcribed using an iScript cDNA synthesis kit (Bio-RAD). The first-strand cDNA was used as a template. The primers used for quantitative RT-PCR analyses of human IRS1 (5'-TCCACCTCGGATTGTCTCTT-3' and 5'-AGGGACTGGAGCCATACTCA-3'), 146 hmnan IRS2 (5'-CCACTCGGACAGCTI‘CTTCT-3' and 5'- AGGATGGTCTCGTGGATGTT-3 '), and human GAPDH (5'- AACTI'TGGTATCGTGGAAGGA-3' and 5'-CAGTAGAGGCAGGGATGATGT—3') were synthesized by Operon Biotechnologies, Inc. (Huntsville, AL). RT-PCR was performed as described previously in Chapter 2 (40), and normalized to GAPDH expression levels. PP2A phosphatase activity assay—The PP2A irnmunoprecipitation phosphatase assay kit purchased from Millipore (Ternecula, CA) was used to measure dephosphorylation of a phosphopeptide as an index of phosphatase activity. Briefly, the cells were lysed using tlne phosphatase extraction buffer specified by the assay kit, and the catalytic subunit of PP2A (PP2A/C) was immunoprecipitated with anti-PP2A-C supplied in the assay kit. Agarose-bound immune complexes were collected and resuspended in 80 ul of Ser/Thr buffer with 750 uM of phosphopeptide ((KRpTIRR), obtained from the kit). The reaction was conducted for 10 rrnin at 30°C in a shaking incubator. Supematants (25 ul) were transferred in a 96-well plate, and released phosphate was measured by adding 100 pl of malachite geen phosphate detection solution. Color was developed for 10 rrnin before reading the plate at 650 nm. The absorbance of the reactions was corrected by subtracting the absorbance in samples treated without Ab. Results were expressed as fold change of PP2A activity as compared with control cells. Statistical analysis—All experiments were performed at least three times, and representative results are shown. All data, unless specified, are shown as the mean :t SD. 147 for indicated number of experiments. One-way ANOVA with Student’s t-test were used to evaluate statistical significances between different treatment goups. Results Inhibiting PKR reversed phosphorylation of IRS1 at Ser312 induced by ceramide—Ceramide has been shown to inhibit insulin signaling and induce insulin resistance by up-regulating the Ser phosphorylation and blocking Tyr phosphorylation of IRS (22, 23). I confirmed cerarrnide promotes phosphorylation of IRS1 at Ser312 and suppresses phosphorylation at Tyn941 in HepG2 cells (Fig. 5.1A). Since cerarrnide also can activate PKR(26), I hypothesize that PKR mediates the role of ceramide in up- regulating the serine phosphorylation and down-regulating the tyrosine phosphorylation of IRS1. Inhibiting PKR in cerarrnide-treated cells decreased phosphorylation of IRS] at Ser312 and increased phosphorylation of IRS1 at Tyr94l (Fig. 5.1B). Anotlner inhibitor of PKR, 2-aminopurine (2-AP), had a similar effect on the phosphorylation of IRS1 (Fig. 5. l C). Taken together, the results of the chemical inhibitors of PKR suggest that the catalytic function of PKR is involved in inducing the inhibitory serine phosphorylation of IRS 1 , which suppresses tyrosine phosphorylation. However, considering the potential non-specific effects of chemical inhibitors, I further performed gene silencing studies using siRNA of PKR to confirm the role of PKR in regulating the phosphorylation of IRS1. Silencing PKR suppressed the serine phosphorylation and enhanced tyrosine phosphorylation of IRS1—siRNA targeting PKR markedly irnlnibited the gene (40) and 148 protein expressions of PKR (Fig. 5.2) (40). Silencing PKR with this siRNA in HepGZ cells significantly blocked the Ser312 and amplified the Tyr94l phosphorylation of IRSl induced by insulin (Fig. 5.2). A. p-IRS1 Tyr941 p-IRS1 Tyr941 p-lRS1 Ser312 p-lRS1 Ser312 IRS1 IRS1 p-PKR Thr451 p-PKR Thr451 PKR PKR beta Adi" beta Actin Ceramide (pM) Ceramide Ctrl GA 2AP B. p—lRS1 Tyr941 p-IRS1 Ser312 IRS1 p-PKR Thr451 PKR beta Actin Ceramide + + + + PKR Inhibitor (11M) 0 0.5 2 10 Figure 5.1. Effects of ceramide and PKR inhibitors on the phosphorylation of IRSl. HepGZ cells were exposed to different levels of ceramide for 12 hours (A). Pre-treated with 10 11M ceramide for 12 hours, HepG2 cells were exposed to different levels of PKR inhibitor dissolved in DMSO (control) (B) or 10 mM 2—AP dissolved in PBS:glacial acetic acid (200:1) (GA, control) (C) for another 12 hours. After treatment, the cells were harvested, and western blot analysis was performed to detect the level of beta Actin and the total and phosphorylated levels of PKR and IRS1. 149 Figure 5.2. Involvement of PKR in regulating the phosphorylation of [RSL Reverse transfection of suspended HepG2 cells were performed with scrambled siRNA (negative control) or siRNA of PKR for 24 hours and the transfected cells were cultured in regular media for another 24 hours. Cells were then treated with different concentrations of insulin for 15 minutes and harvested after the treatment. Western blot analysis was performed to detect the levels of beta actin and PKR, and the total and phosphorylated levels of IRS1. The phosphorylation levels of IRS1 at Ser312 and Tyr94l were quantified by normalizing to total IRS1 levels and are expressed as the average of three samples 3: SD fiom tlnree independent experiments. Student t-test was performed for analyzing the differences between samples transfected with siPKR and scrambled siRNA (negative control). *, significantly higher (Tyr94l) or lower (Ser312) than negative control (i.e., scrambled siRNA), p<0.01. 150 WB: siPKR Neg. Ctrl. P-IRS1Tyr941—> + - . ‘s 1‘: 3 .. ,-, 1:. :1. .r.. _.-.'. *r; w- "1'2: - t . 5,. 5:25:53 , .2. . l - ‘ ,.. -. .: ~ - r .. - . ~ -,:-,-.... p-IRS1 Ser312 IRS1 beta Actin 0.2 0.6 1.0 (15mins) 0.2 0.6 1.0 0 0 Insulin (nM) .. %//////x1 . . Meme ._ . l Scrambled 2 0 2.5 - .1 39.20 22. 5x. .59 a «5.8 $5-.. Insulin (nM) 151 Taken together, both the inhibition and gene silencing studies suggest PKR may be involved in regulating insulin signaling by inducing phosphorylation of IRS1 at Ser312 and suppressing phosphorylation at Tyr941. However, it is unclear how PKR regulates the phosphorylation of IRS1. Co-irnmunoprecipitation showed that PKR did not directly interact with IRS1 (Fig. 5.3A), suggesting other intermediate signaling molecules must mediate the effect of PKR on the phosphorylation of IRS1. IRS1 Ser kinases, which directly phosphorylate the serine residues of IRS], include IKK (32), mTOR (33), PKCl; (34), S6K1 (35) and JNK (36). PKR has been reported to positively activate the MAPKs, in particular JNK (40, 41), and IKK (39). Therefore, I hypothesize that PKR induces phosphorylation of IRS1 at Ser312 through IRS serine kinases, JNK, IKK, or both. PKR positively regulates JNK and IKK, both of which mediate the effect of PKR on the Ser phosphorylation of IRS1—I previously showed that PKR co- immunoprecipitates with JNK and activates JNK in HepG2 cells (40). Here, I show that IKK also is activated by PKR. Silencing PKR significantly reduced the phosphorylation of IKKo/B at Ser176/ 180, which indicates H(K activity (Fig. 5.38). To firrtlner confirm the involvement of IN K and IKK in mediating the effect of PKR on the phosphorylation of IRS1 at Ser312 and Tyr94l, I over-expressed PKR in HepG2 cells and inhibited the activity of JNK or IKK. Over-expressing PKR by transfecting the plasmid pCMV6-hPKR into HepG2 cells enhanced phosphorylation of IRSl at Ser312 and suppressed phosphorylation at Tyr94l (Fig. 5.3C), supporting our results (Figs. 5.1 and 5.2) that PKR induces serine phosphorylation of IRS1. Furthermore, in PKR over-expressed cells, inhibitors of JNK and IKK, SP600125 and $0514, respectively, reduced Ser312 phosphorylation and restored Tyr941 phosphorylation of IRS1 (Fig. 5.3C), suggesting 152 that both kinases, JNK and IKK, mediate the effect of PKR on the phosphorylation of IRSl at Ser312, which suppresses tyrosine phosphorylation of IRS1. Figure 3 A. re: IRS1 l8: IRS1 d lePKR *1 IBzPKR . IP: NC PKR lP: NC IRS1 B. C. pCMV pCMV-hPKR p-lKKor/B Ser176/180: .. _- ‘ ~ ~ p-IRS1Tyr941 IKKB .- .1 p-lRS1Ser312 p-JNK T183IY185 ’ RS, JNK p-PKR Thr451 i i i i NC siPKR PKR Figure 5.3. Involvement of JNK and IKK in regulating the phosphorylation of IRS]. Confluent HepG2 cells were harvested and immunoprecipitated with anti-PKR, anti-IRS1, or IgG as negative controls, and western blot analysis was performed to detect the protein level of IRS1 and PKR (A). Reverse transfection of suspended HepG2 cells was performed with scramble siRNA (negative control) or siRNA of PKR (siPKR) for 24 hours and the transfected cells were cultured in regular medium for another 24 hours (B). Cells were harvested, and western blot analysis was performed to detect the total and phosphorylated levels of JNK and IKK (B). In HepG2 cells, the forward transfection of empty vector pCMV6-XL5 (pCMV6) or plasmid containing PKR cDNA sequence (pCMV6-hPKR) was performed and the cells were then treated with the pharmaceutical inhibitor of JNK, SP600125 (25 uM), inhibitor of IKK, SC-514 (50 uM), or DMSO, vehicle of these two cherrnicals, for 1 hour (C). After treatments, cells were then harvested and western blot analysis was performed to detect the total and phosphorylated levels of IRS1 and PKR(C). 153 Figure 5.4. Involvement of PKR in regulating IRSZ. Confluent HepG2 cells were treated with 5 uM PKR inhibitor (P1) or its analogue as a negative control (NC) or 10 mM 2-AP dissolved in PBS:glacial acetic acid (200:1) (GA) for 12 hours (A, C). Reverse transfection of suspended HepG2 cells were performed with scrambled siRNA (negative control) or siRNA of PKR for 24 hours and the transfected cells were cultured in regular media for another 24 hours. Cells were then harvested (D) or treated with different concentrations of insulin for 15 minutes and harvested after the treatment (B). Western blot analysis was performed to detect the total and phosphorylated levels of IRS2 at Ser731 (A, B). RT-PCR was performed to detect the gene expression levels of IRSl and IRS2 in response to the PKR inhibitors (C) or siRNA of PKR (D). Gene expression data were expressed as the average of nine samples 3: SD from three independent experiments. The protein levels of IRS2 were quantified by normalizing to beta actin, and the phosphorylation levels of IRS2 at Ser731 were quantified by normalizing to total IRS2 levels. Botln the protein and phosphorylation levels of IRS2 are expressed as the average of three samples :t SD fiom three independent experiments. Student t-test was performed for analyzing the differences between siPKR and scrambled siRNA (negative control) (B, D) or PKR inhibitors and negative controls. *, significantly lower than negative control (i.e., scrambled siRNA (B, D) or chemical analogue of the PKR inhibitor (C)), p<0.05. **, significantly lower than control (GA, solvent of 2-AP), p<0.01. 154 WB: p-lRS2 Ser731 IRSZ beta Actin NC PI GA 2AP WB: siPKR 11131!!! p-IRS2 Ser731 | R32 beta Actin 0.2 0.6 1.0 (15mins) 0.2 0.6 1.0 0 0 Insulin (nM) ar/ .0 b b m m C S I its: 1100 39.20 28. «mm. Insulin (nM) l Scrambled // 2.5 ~ _ m 5 4| 1 $955 2o: 3”: .50: Stew «9:8 2_ a 0.5~ Insulin (nM) 155 Figure 5.4 continued C- r). 1.8 - 1 2 _ I IRS1 ‘ llRS1IlR82 '5 2 5 E 1 ‘ '7 c f; -' A .2 O) c 0 0.8 - r: o D r g -- s a 0 § 5 0 6 - x '0 '- ur a a 2 g 2: ru ,3 0.4 — E net 0.2 — 0 _ Neg. Ctrt. PKR GA 2-AP NC siPKR inhibitor Thus far, I have identified a novel function of PKR in regulating the phosphorylation of IRS1. In addition to IRS1, I also investigated the potential effect of PKR on another major IRS family protein, IRSZ, which also mediates insulin signaling in the liver (3). PKR up-regulates the protein expression level of IRS2—Both PKR irnhibitors, PKR inhibitor and 2-AP, down-regulated the protein level of IRS2 (Fig. 5.4A), but not IRS1 (Fig. 5.1), suggesting that PKR is required for cells to maintain proper protein expression level of IRSZ. Similarly, silencing the gene expression of PKR also down- regulated the protein level of IRSZ (Fig. 5.4B), further supporting the effect of PKR on the protein expression level of IRSZ. Notably, upon inhibiting or silencing PKR, the phosphorylation of IRSZ at Ser731 varied proportionally to its total protein level (Fig. 5.4A and B), suggesting that PKR is not affecting the phosphorylation of IRS2 at Ser731. To determine whether PKR transcriptionally regulates the protein level of IRS2, I measured the mRNA expression levels of IRS1 and IRSZ upon PKR inhibition and gene silencing. Both the PKR inhibitors and the siRNA of PKR down-regulated the mRNA 156 expression of IRSZ, but not of IRS1 (Fig. 540, suggesting PKR transcriptionally regulates IRS2. PKR rap-regulates the protein level of IRSZ through the transcription factor FoxOI—PKR, a protein kinase, activates several transcription factors, such as IRF-l , p53, and NF-ch (52, 53), but these have not been shown to regulate the transcription of IRS2. However, the transcription of IRSZ has been shown to be dependent on the transcription factor FoxOl in the liver (42, 43) or CREB in pancreatic beta-cells (44). It is not known whether PKR interacts with either of tlnese two transcription factors. I found that PKR had no effect on the activity or translocation of CREB in HepG2 cells (not shown). However, silencing the PKR gene significantly increased the phosphorylation of FoxOl at Ser256 (Fig. 5.5A). Phosphorylation at Ser256 inhibits the DNA-binding activity of F 0x01 and its nuclear import by suppressing the nuclear targeting signal on its DNA binding domain (54, 55). Our results suggest, for the first time, that PKR reduces the phosphorylation of FoxOl and thereby activates it. As a protein kinase, PKR does not have the phosphatase activity that is required to dephosphorylate FoxOl . However, PKR is known to phosphorylate 85611, the regulatory subunit of protein phosphatase 2A (PP2A), which then activates the catalytic subunit of PP2A (56). PP2A, in turn, is known to dephosphorylate FoxOl at Ser256 (57). Indeed, silencing PKR significantly suppressed the activity of PP2A (Fig. 5.5B), thereby confirming that PKR activates PP2A in HepG2. To further investigate the potential involvement of PP2A in mediating the dephosphorylation of Ser256 on FoxOl by PKR, I over-expressed PKR in HepG2 cells and inhibited the activity of PP2A with okadaic acid (OA). Over-expressing PKR by trarnsfecting the plasmid pCMV6-hPKR into HepG2 cells reduced the phosphorylation of 157 FoxOl at Ser256 (Fig. 5.5C), supporting our results that PKR induces dephosphorylation of FoxOl at Ser256 (Fig. 5.5A). More importantly, 0A, a specific PP2A irnhibitor, restored the serine phosphorylation level of FoxO] in PKR over-expressed cells (Fig. 5.5C). Taken together, PKR dephosphorylates and activates FoxOl , mediated by PP2A. To confirm the positive effect of FoxOl on the expression of IRS2 in HepG2 cells, I performed gene silencing study of FoxOl. Silencing the gene expression of FoxO] significantly reduced the protein level of IRS2, but not IRS1 (Fig. 5.5D), suggesting that FoxO] controls the expression of IRS2 in HepG2 cells. To firrther confirm that FoxOl mediates the effect of PKR on IRS2 protein level, I over-expressed PKR in HepG2 cells as well as silenced FoxOl. Over-expressing PKR in HepGZ cells increased the protein level of IRS2 (lanes 1 vs. 2 in Fig. 5.5E), whereas silencing FoxOl in control and PKR over-expressed cells significantly reduced the protein level of IRS2 (Fig. 5.5E), confirming that PKR up-regulates the protein level of IRS2 through the transcription factor FoxOl. In summary, our results suggest two different pathways by which PKR regulates IRS proteins and thereby insulin signaling (Fig. 5.9). First, PKR induces phosphorylation of IRS] at Ser312, and suppresses tyrosine phosphorylation of IRS], mediated by IRS kinases, JNK and IKK. Second, PKR activates a transcription factor, FoxOl, which up- regulates the gene expression of IRS2. In addition, the activity of PKR was recently shown to be suppressed by insulin or insulin-like gowth factor 1 (IGFl) in muscle cells (45). I confirmed a similar effect of insulin on the activity of PKR in HepG2 cells. Therefore I propose that PKR is a novel downstream target of insulin signaling. 158 Figure 5.5. Involvement of Fox01 in mediating the effect of PKR on IRS2. Reverse transfection of suspended HepG2 cells were performed with scrambled siRNA (control) or siRNA of PKR (A, B) or siRNA of Fox01 (D) for 24 hours and the transfected cells were cultured in regular media for another 24 hours. The forward transfection of empty vector pCMV6-XL5 (pCMV6) or plasmid containing PKR cDNA sequence (pCMV6- hPKR) was performed and the cells were then treated with OA (2 nM) or its vehicle, ethanol, as a control, for 1 hour (C). Reverse transfection of scramble siRNA (negative control, lanes 1 and 2) or siRNA of Fox01 (siPKR, lanes 3 and 4) was performed, followed by the forward transfection of empty vector pCMV6-XL5 (pCMV, lanes 1 and 3) or the plasmid containing PKR cDNA sequence (hPKR, lanes 2 and 4) (B). After treatment, the cells were harvested, and western blot analysis was performed to detect the total and phosphorylated levels of Fox01 (A, C) or the total levels of IRS], IRS2, Fox01 and beta actin (D, E). PP2A activity assay was performed to detect the phosphatase activity of PP2A (B). Student t-test was performed and p values were calculated for analyzing the differences between the indicated samples. 159 C. 1" we- we: p-Fox01 $256 _’ p-FoxO1 $256 —> was: r" . > 1 , Fox01 { 3 f Fox01 a It — 1. NC siPKR CMV hPKR hPKR+OA B. A 1.2 “ D. 0 O C 2 IRSZ 0 'U E IRS1 > . ' RKO1 5 . O. NC siFoxO1 NC siPKR IRSZ IRS1 PKR an...- , .j. .H' m, d beta Actin NC siFoxO1 Insulin down-regulates the activity of PKR in HepGZ cells, mediated by IRS— Physiological concentrations of insulin decreased the phosphorylation of PKR at Thr451in HepG2 cells, in a time and dose dependent manner (Fig. 5.6 A and B), suggesting that PKR may be a downstream target of the insulin signaling pathway. R85, 160 including IRS] and IRS2 in liver cells, have been identified as central transmitters of insulin signaling (1, 2). To determine whether IRS mediates the suppressive effect of insulin on PKR phosphorylation, I silenced the expression of IRS] (Fig. 5.7A) or IRSZ (Fig. 5.7B) in HepG2 cells. The negative effect of insulin on the phosphorylation of PKR, which was observed in control cells, was abolished in both IRSI- and IRS2-silenced cells (Fig. 5.7), indicating that IRS] and IRS2 mediate the effect of insulin on the activity of PKR. As a tyrosine kinase, IRS does not have Ser/Thr kinase or phosphatase activity. Therefore an intermediate protein must mediate the dephosphorylation of PKR upon simulation of IRS by insulin. It is well established that in the liver, insulin up-regulates the glycogen-targeting subunits of protein phosphatase ] (PP1), i.e., GL and R5 (58, 59), which activate the catalytic subunit of PP] (PPlc) (60). This process, the up-regulation of the regulatory subunits and activation of PPlc by insulin, has been suggested to be mediated by the pathway IRS-PI-3 kinase-Akt (6], 62). As a major eukaryotic protein serine/threonine phosphatase, PP1 binds to PKR and dephosphorylates PKR at the serine and threonine residues (63). Therefore, I hypothesize that PP] mediates the dephosphorylation of PKR induced by insulin. To test this hypothesis, I measured the phosphorylation of PPlc in response to insulin and studied the inhibition of PPlc. 161 A. WB: p-lRS1Ser312 [ W . p beta A011" W Insulin (nM) 0 0.1 0.2 0.4 1.0 (15 mins) B. p-IRS1 Ser312 p-IRS1 Tyr941 —> f IRS1 p-PKR Thr451 —> . PKR beta Actin Insulin (min) 0 5 10 15 30 60 (1.0 nM) Figure 5.6. Effect of insan on the activity of PKR. HepG2 cells were exposed to different levels of insulin for 15 rrninutes (A) or lnM of insulin for 5, 10, 15, 30, or 60 rrninutes (B). After treatment, the cells were harvested, and western blot analysis was performed to detect the total and phosphorylated levels of PKR and IRS]. 162 A. WB: P-PKRThr451—P -. ,..... Wc eat“ an- PKR "bu—renowned.» IRS2 “wufimum” IRs1 ‘ u h fl a... |nsulin(nM) 0 0.2 0.6 1.0 0 0.2 0.6 1.0 (15mins) Neg. Ctrl. siIRS1 WB: p-PKRThr451—> " PKR IRSZ IRS1 beta Actin |nsulin(nM) 0 0.2 0.6 1.0 0 0.2 0.6 1.0 (15mins) Neg. Ctrl. silRSZ Figure 5.7. Involvement of IRSl and IRSZ in mediating the effect of insulin on the phosphorylation of PKR. Reverse transfection of suspended HepG2 cells was performed with scrambled siRNA (negative control) or siRNA of IRS] (A) or siRNA of IRS2 (B) for 24 hours and the transfected cells were cultured in regular media for another 24 hours. Cells were then treated with different concentrations of insulin for 15 minutes and harvested after the treatment. Western blot analysis was performed to detect the total and phosphorylated level of PKR, and the levels of IRS], IRSZ, and beta Actin. PP1 mediates the inhibitory effects of insulin on PKR—Insulin suppressed phosphorylation of PPlc at Thr320, a key residue tlnat upon phosphorylation inhibits the activity of PPlc (64) (Fig. 5.8). To determine if PP1 mediates the effect of insulin on PKR, I inhibited PPlc activity with Tautomycetin (TMC), a specific inhibitor of PPlc 163 (48). The phosphorylation of PKR no longer decreases in response to insulin (Fig. 5.8), suggesting that PP1 mediates the down-regulation of PKR phosphorylation by insulin. Therefore, I propose that insulin down-regulates the phosphorylation of PKR, mediated by the pathway from IRS to PP1 (Fig. 5.9). p-PKR Thr451 —> PKR p-PP1c Thr320 PP1c beta Actin Insulin (nM) 0 0.6 1.0 0 0.6 1.0 (15 mins) Neg. Ctrl. TMC Figure 5.8. Involvement of PP1 in mediating the effect of insulin on the phosphorylation of PKR. Confluent HepG2 cells were pre-treated with 5 uM Tautomycetin (TMC) for 4 hours and exposed to different levels of insulin for 15 nninutes. After treatment, the cells were harvested, and western blot analysis was performed to detect the total and the phosphorylated levels of PKR and PPlc, and the level of beta Actin. In summary, 1 found that PKR is irnhibited by insulin through the IRS-PP1 pathway, and serves as a downstream substrate of insulin signaling. More importantly, PKR regulates the central transnnitters of intracellular insulin signaling, the IRS proteins, through distinct mechanisms. PKR up-regulates phosphorylation of IRS] at Ser312, through JNK and IKK, thereby suppressing tyrosine phosphorylation of IRS]. Concomitantly, PKR activates the transcription factor, Fox01, up-regulating IRS2 expression. Taken together, these results suggest PKR is involved in insulin signaling through a feedback mechanism (Fig. 5.9). 164 I IR IRS1: p-SerI I lRSZ: protein level ,..—.1 IRS1 IRSZ( I Akt I IKK JNK PP1 t t I . + PKR I IRS 2 PP2A -> Fox01 Figure 5.9. Proposed signaling pathways through which PKR is involved in insulin signaling network in HepG2 cells. Insulin activates insulin signaling by IR and IRS, leading to the activation of PP], which in turn dephosphorylates PKR at Thr451. PKR induces the phosphorylation of IRS] at Ser312 through two other kinases, JNK and IKK. In addition, by activating PP2A, PKR dephosphorylates a transcription factor, Fox01, which up-regulates the gene expression of IRS2. Discussion In this chapter, I identified novel effects of PKR on 2 major IRS proteins, IRS] and IRS2, in HepG2 cells. First, PKR up-regulates the phosphorylation of IRS] at Ser312, which in turn suppresses the tyrosine phosphorylation of IRS]. This effect of PKR is mediated by JNK and IKK (Fig. 5.3). It is well known that PKR stimulates the transcription factor NF-ch by activating HéK (39), and this process does not require the catalytic activity of PKR. Instead, The N-terminus of PKR is responsible for the activation of IKK (65). As discussed previously (40), PKR has been reported also to play 165 a role in the phosphorylation of the 3 rrnitogen-activated protein kinases (MAPK) (JN K, ERK, and p38) in the rank order of JNK>p38>ERK (41). Among the three MAPKs, JNK has been suggested to play a central role in inducing the inhibitory serine phosphorylation of IRS] (21, 36, 37). I did not test the effects of PKR on the other two less responsive MAPK proteins, ERK and p3 8, which were also suggested to induce phosphorylation of IRS] at Ser residues (66, 67). Thus, although JNK may be an important intermediate, it is not likely the only one involved in mediating the signaling patlnway from PKR to IRS] phosphorylation. Currently, it remains unclear how PKR interacts with MAPKs, e. g. JNK. PKR was shown to interact with apoptosis signal-regulating kinase ] (ASKl) (68), a MAPK kinase kinase (MAPKKK), which phosphorylates SEKl/MKK4 or MKK3/MKK6 and in turn activates JNK or p38 MAPK, respectively (69). In addition, it is believed that the activation of the MAPK cascade requires a scaffold protein that assembles the MAPKKK, MAPKK, and MAPK proteins together into a signaling module (70). Therefore, the activation of JNK by PKR may require the recruitment of other proteins, including a scaffold protein that assembles JNK and its upstream activators in the MAPK signaling pathway. In other words, the PKR—JN K pathway 1 identified does not preclude other possible intermediates between PKR and JNK. The phosphorylation of IRS] at Ser312 by PKR provides a potential mechanism through which ceramide, an activator of PKR and inducer of insulin resistance, promotes the inhibitory serine phosphorylation of IRS]. Indeed, other activators of PKR have also been identified to induce insulin resistance and type 2 diabetes. For example, HCV core protein, which directly binds and activates PKR (27), has been shown to induce insulin resistance in liver cells by inducing the phosphorylation of IRS] at Ser312 (25). The 166 ability of PKR to promote serine phosphorylation of IRS] provides a possible mechanism by which PKR mediates HCV infection and the development of insulin resistance and ultimately type 2 diabetes (24, 25). Another novel function of PKR is regulating the protein level of IRS2 through the transcription factor Fox01. Fox01 directs the expression of genes involved in a wide variety of cellular responses including myogenic gowtln, differentiation, apoptosis, and stress resistance (71, 72). In liver, FoxOl is regulated by insulin (73) and plays a significant role in regulating glucose homeostasis and energy metabolism (74). Through the IRS1-P13K-Akt pathway, insulin induces the phosphorylation of Fox01, thereby inhibiting its activity (73). In healthy states, the low insulin level during fasting sustains the activity of Fox01, which facilitates the transcription of key enzymes involved in gluconeogenesis (75). Fox01 also can up-regulate IRSZ gene expression through a feedback loop (42, 43), which I also found to occur in HepG2 cells (Fig. 5.5). However in insulin resistance models, the persistent activation of Fox01, due to disruption of the IRS-P13 K-Akt pathway, contributes to the development of hyperglycemia and glucose intolerance (76, 77). Therefore, Fox01, which becomes activated during irnsulin resistance, is believed to serve as a dominant regulator of hepatic gene expression. The roles of Fox01 in regulating fasting glucose homeostasis and enhancing hyperglycemia and glucose intolerance in insulin resistance models begs the question whether PKR activation is involved in the regulation of glucose metabolism mediated by Fox01. Here, I identified that PKR regulates IRS] and IRS2 by different mechanisms. Although the protein structures of the IRS proteins are highly conserved, both animal and cell studies indicate that IRS] and IRS2 serve complementary, rather than redundant, 167 in: "I" roles in insulin signaling (78-81). It has been suggested that hepatic IRS] and IRS2 control different aspects of hepatic metabolism, with IRS] more closely related to glucose homeostasis and IRS2 more closely related to lipid metabolism (82). However more recently, using specific knockouts of liver IRS1 or IRS2, researchers demonstrated that the two proteins may overlap in their insulin action (7 7 ). Nevertheless, investigators have claimed that IRS] is stably expressed and frrnctions in the postprandial state, while IRS2 is highly expressed and activated in the fasted state (77 , 83), which could be due to elevated activation of Fox01. Taken together, IRS] and IRS2, nevertheless, are believed to function independently despite these possible overlaps, and the varying effects of PKR on these two proteins may serve as a potential mecharnism by which the two IRS proteins are differentially regulated. I also found that in HepG2 cells the phosphorylation of PKR is down-regulated by insulin (Fig. 5.6). It is interesting that this down-regulation was abolished upon longer treatment with insulin (Fig. 5.6B), suggesting, possibly, a negative feedback is regulating the effect of insulin on PKR. Indeed, insulin signaling is tuned by various negative feedbacks, e. g., degadation of IR induced by insulin binding (84), negative regulation of insulin signaling activity by MAPKs (85), mTOR (86), PI-3 kinase (87), etc. These negative feedbacks modulate the downstream signaling events induced by long-term treatment of insulin (16). In summary, I identified that PKR, which is down-regulated by insulin through the IRS to PP1 pathway, may fimction as a key regulator of insulin signaling (Fig. 5.9). PKR may inhibit insulin signaling and potentially induce insulin resistance by promoting the phosphorylation of IRS at Ser312. This firnction of PKR explains perhaps how 168 several PKR activators (e. g., ceramide) induce the inhibitory phosphorylation of IRS], leading to insulin resistance. On the other hand, PKR activates the transcription factor, Fox01, which up-regulates the protein expression level of IRS2, with possible implications on the regulation of glucose homeostasis. Taken together, PKR appears to be a novel player in the regulation of insulin signaling tlnrough the major transmitters of insulin signaling, IRS] and IRS2. 169 10. References Saltiel, A. R., and Kahn, C. R. (2001) Insulin signalling and the regulation of glucose and lipid metabolism. Nature 414, 799-806 Patti, M. E., and Kahn, C. R. (1998) The insulin receptor--a critical link in glucose homeostasis and insulin action. J Basic Clin Physiol Pharmacol 9, 89-109 Thirone, A. C., Huang, C., and Klip, A. (2006) Tissue-specific roles of IRS proteins in insulin signaling and glucose transport. Trends Endocrinol Metab 17, 72-78 White, M. F. (1998) The IRS-signalling system: a network of docking proteins that mediate insulin action. Mol Cell Biochem 182, 3-11 Lowenstein, E. J ., Daly, R. J., Batzer, A. G., Li, W., Margolis, B., Lammers, R., Ullrich, A., Skolrnik, E. Y., Bar-Sagi, D., and Schlessinger, J. (1992) The SH2 and SH3 domain-containing protein GRB2 links receptor tyrosine kinases to ras signaling. Cell 70, 431-442 Skolrnik, E. Y., Lee, C. H., Batzer, A., Vicentini, L. M., Zhou, M., Daly, R., Myers, M. J., Jr., Backer, J. M., Ullrich, A., White, M. F., and et a1. (1993) The SH2/SH3 domain-containing protein GRB2 interacts with tyrosine-phosphorylated IRS] and She: implications for insulin control of ras signalling. EMBO J 12, 1929-1936 Backer, J. M., Schroeder, G. G., Kahn, C. R., Myers, M. G., Jr., Wilden, P. A., Calnill, D. A., and White, M. F. (1992) Insulin stimulation of phosphatidylinositol 3-kinase activity maps to insulin receptor regions required for endogenous substrate phosphorylation. J Biol Chem 267 , 1367-1374 Alessi, D. R., James, S. R., Downes, C. P., Holmes, A. B., Gaffrney, P. R., Reese, C. B., and Cohen, P. (1997) Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and activates protein kinase Balpha. Curr Biol 7, 261-269 Standaert, M. L., Galloway, L., Kamam, P., Bandyopadhyay, G., Moscat, J., and F arese, R. V. (1997) Protein kinase C-zeta as a downstream effector of phosphatidylinositol 3-kinase during insulin stimulation in rat adipocytes. Potential role in glucose transport. J Biol Chem 272, 30075-30082 Sone, H., Suzuki, H., Takahashi, A., and Yarnada, N. (2001) Disease model: hyperinsulinemia and insulin resistance. Part A-targeted disruption of insulin signaling or glucose transport. Trends Mol Med 7, 320-322 170 ll. 12. l3. 14. 15. 16. 17. l8. 19. 20. 21. Kahn, B. B. (1998) Type 2 diabetes: when insulin secretion fails to compensate for insulin resistance. Cell 92, 593-596 Reaven, G. M. (1988) Banting lecture 1988. Role of insulin resistance in human disease. Diabetes 37, 1595-1607 Irnamura, T., Takata, Y., Sasaoka, T., Takada, Y., Morioka, H., Haruta, T., Sawa, T., Iwanishi, M., Hu, Y. G., Suzuki, Y., and et al. (1994) Two naturally occurring mutations in the kinase domain of insulin receptor accelerate degadation of the insulin receptor and impair the kinase activity. J Biol Chem 269, 31019-31027 McElduff, A., Hedo, J. A., Taylor, S. I., Roth, J ., and Gorden, P. (1984) Insulin receptor degadation is accelerated in cultured lymphocytes from patients with genetic syndromes of extreme insulin resistance. J Clin Invest 74, 1366-1374 Zick, Y. (2005) Ser/Thr phosphorylation of IRS proteins: a molecular basis for irnsulin resistance. Sci SIKE 2005, pe4 Herschkovitz, A., Liu, Y. F., Ilan, E., Ronen, D., Boura-Halfon, S., and Zick, Y. (2007) Common inhibitory serine sites phosphorylated by IRS-1 kinases, triggered by insulin and inducers of insulin resistance. J Biol Chem 282, 18018-18027 Taylor, S. I., and Arioglu, E. (1998) Syndromes associated with insulin resistance and acanthosis nigicans. J Basic Clin Physiol Pharmacol 9, 419-439 Paz, K., Hemi, R., LeRoith, D., Karasik, A., Elhanany, E., Kanety, H., and Zick, Y. (1997) A molecular basis for insulin resistance. Elevated serine/threonine phosphorylation of IRS-l and IRS-2 inhibits their binding to the juxtamembrane region of the insulin receptor and impairs their ability to undergo insulin-induced tyrosine phosphorylation. J Biol Chem 272, 29911-29918 Hotamisligil, G. S., Peraldi, P., Budavari, A., Ellis, R., White, M. F., and Spiegelrrnarn, B. M. (1996) IRS-l-mediated inhibition of insulin receptor tyrosine kinase activity in TNF-alpha- and obesity-induced insulin resistance. Science 271, 665-668 Pederson, T. M., Kramer, D. L., and Rondinone, C. M. (2001) Serine/threonine phosphorylation of IRS-l triggers its degadation: possible regulation by tyrosine phosphorylation. Diabetes 50, 24-31 Hirosumi, J., Tuncman, G., Chang, L., Gorgun, C. Z., Uysal, K. T., Maeda, K., Karin, M., and Hotamisligil, G. S. (2002) A central role for JNK in obesity and insulin resistance. Nature 420, 333-336 171 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Miura, A., Kajita, K., Ishizawa, M., Kanoh, Y., Kawai, Y., Natsume, Y., Sakuma, H ., Yamamoto, Y., Yasuda, K., and Ishizuka, T. (2003) Inhibitory effect of ceramide on insulin-induced protein kinase Czeta translocation in rat adipocytes. Metabolism 52, 19-24 Kanety, H., Hemi, R., Papa, M. Z., and Karasik, A. (1996) Sphingomyelinase and ceramide suppress insulin-induced tyrosine phosphorylation of the insulin receptor substrate-1. J Biol Chem 271, 9895-9897 Aytug, S., Reich, D., Sapiro, L. E., Berrnsteirr, D., and Begum, N. (2003) Impaired IRS-1/PI3-kinase signaling in patients with HCV: a mechanism for increased prevalence of type 2 diabetes. Hepatology 38, 1384-1392 Banerjee, S., Saito, K., Ait-Goughoulte, M., Meyer, K., Ray, R. B., and Ray, R. (2008) Hepatitis C virus core protein upregulates serine phosphorylation of insulin receptor substrate-1 and impairs the downstream akt/protein kinase B signaling pathway for insulin resistance. J Virol 82, 2606-2612 Ruvolo, P. P., Gao, F., Blalock, W. L., Deng, X., and May, W. S. (2001) Ceramide regulates protein syntlnesis by a novel mechanism involving the cellular PKR activator RAX. J Biol Chem 276, 11754-11758 Yarn, X. B., Battaglia, S., Boucreux, D., Chen, Z., Brechot, C., and Pavio, N. (2007) Mapping of the interacting domains of hepatitis C virus core protein and the double-stranded RNA-activated protein kinase PKR. Virus Res 125, 79-87 Taylor, S. S., Haste, N. M., and Ghosh, G. (2005) PKR and eIF2alpha: integation of kinase dimerization, activation, and substrate docking. Cell 122, 823-825 Stark, G. R., Kerr, I. M., Williams, B. R., Silverrnan, R. H., and Schreiber, R. D. (1998) How cells respond to interferons. Annu Rev Biochem 67, 227-264 Williams, B. R. (2001) Signal integation via PKR. Sci ST KE 2001, RE2 Barber, G. N. (2005) The dsRNA-dependent protein kinase, PKR and cell death. Cell Death Difler 12, 563-570 Gao, Z., Hwang, D., Bataille, F., Lefevre, M., York, D., Quon, M. J., and Ye, J. (2002) Serine phosphorylation of insulin receptor substrate 1 by inhibitor kappa B kinase complex. J Biol Chem 277, 48115-48121 Ozes, O. N., Akca, H., Mayo, L. D., Gustin, J. A., Maeharna, T., Dixon, J. E., and Donner, D. B. (2001) A phosphatidylinositol 3-kinase/Akt/mTOR patlnway 172 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. mediates and PTEN antagonizes tumor necrosis factor inhibition of insulin signaling through insulin receptor substrate-1. Proc Natl Acad Sci U S A 98, 4640- 4645 Ravichandran, L. V., Esposito, D. L., Chen, J., and Quon, M. J. (2001) Protein kinase C-zeta phosphorylates irnsulin receptor substrate-1 and impairs its ability to activate phosphatidylinositol 3-kinase in response to insulin. J Biol Chem 276, 3543-3 549 Um, S. H., Frigerio, F., Watanabe, M., Picard, F., Joaquin, M., Sticker, M., Fumagalli, S., Allegini, P. R., Kozrna, S. C., Auwerx, J., and Thomas, G. (2004) Absence of S6Kl protects against age- and diet-induced obesity while enhancing insulin sensitivity. Nature 431, 200-205 Aguirre, V., Uchida, T., Yenush, L., Davis, R., and White, M. F. (2000) The c-Jun NH(2)-terminal kinase promotes insulin resistance druing association with insulin receptor substrate-l and phosphorylation of Ser(3 07). J Biol Chem 275, 9047-9054 Lee, Y. H., Giraud, J., Davis, R. J., and White, M. F. (2003) c-Jun N-terminal kinase (JNK) mediates feedback inhibition of the insulin signaling cascade. J Biol Chem 278, 2896-2902 Zick, Y. (2004) Uncoupling insulin signalling by serine/threonine phosphorylation: a molecular basis for insulin resistance. Biochem Soc Trans 32, 812-816 Bonnet, M. C., Wei], R., Dam, B., Hovanessian, A. G., and Meurs, E. F. (2000) PKR stimulates NF -kappaB irrespective of its kinase function by interacting with the IkappaB kinase complex. Mol Cell Biol 20, 4532-4542 Yang, X., and Chan, C. (2009) Repression of PKR mediates palmitate-induced apoptosis in HepG2 cells through regulation of Bcl-2. Cell Res Zhou, H. R., Lau, A. S., and Pestka, J. J. (2003) Role of double-stranded RNA- activated protein kinase R (PKR) in deoxynivalenol-induced ribotoxic stress response. T oxicol Sci 74, 335-344 Ide, T., Shirnano, H., Yahagi, N., Matsuzaka, T., Nakakuki, M., Yamamoto, T., Nakagawa, Y., Takahashi, A., Suzuki, H., Sone, H., Toyoshima, H., Fukarnizu, A., and Yarnada, N. (2004) SREBPs suppress IRS-2-mediated insulin signalling in the liver. Nat Cell Biol 6, 351-357 Zhang, J ., Ou, J ., Bashmakov, Y., Horton, J. D., Brown, M. S., and Goldstein, J. L. (2001) Insulin inhibits transcription of IRS-2 gene in rat liver through an insulin 173 45. 46. 47. 48. 49. 50. 51. 52. response element (IRE) that resembles IRES of other insulin-repressed genes. Proc Natl Acad Sci U S A 98, 3756-3761 Jhala, U. S., Canettieri, G., Screaton, R. A., Kulkarni, R. N., Krajewski, S., Reed, J ., Walker, J ., Lin, X., Wlnite, M., and Montrniny, M. (2003) cAMP promotes pancreatic beta-cell survival via CREB-mediated induction of IRS2. Genes Dev 17, 1575-1580 Eley, H. L., Russell, S. T., and Tisdale, M. J. (2008) Role of the dsRNA-dependent protein kinase (PKR) in the attenuation of protein loss from muscle by insulin and insulin-like gowth factor-I (IGF-I). Mol Cell Biochem Srivastava, S., and Chan, C. (2006) Hydrogen peroxide and hydroxyl radicals mediate palmitate-induced cytotoxicity to hepatoma cells: Relation to mitochondrial permeability transition. Free Radic Res 41, 38-49 Gual, P., Gonzalez, T., Gremeaux, T., Barres, R., Le Marchand-Brustel, Y., and Tanti, J. F. (2003) Hyperosmotic stress inhibits insulin receptor substrate-l function by distinct mechanisms in 3T3-L1 adipocytes. J Biol Chem 278, 26550-26557 Mitsuhashi, S., Shima, H., Tanuma, N., Matsuura, N., Takekawa, M., Urano, T., Kataoka, T., Ubukata, M., and Kikuchi, K. (2003) Usage of tautomycetin, a novel inhibitor of protein phosphatase 1 (PP1), reveals that PP1 is a positive regulator of Raf-l in vivo. J Biol Chem 278, 82-88 Shim, J. H., Lee, H. K., Chang, E. J., Chae, W. J., Han, J. H., Han, D. J ., Morio, T., Yang, J. J ., Bothwell, A., and Lee, S. K. (2002) Irnmunosuppressive effects of tautomycetin in vivo and in vitro via T cell-specific apoptosis induction. Proc Natl Acad Sci USA 99,10617-10622 Kishore, N., Sommers, C., Matlnialagarn, S., Guzova, J., Yao, M., Hauser, S., Huynh, K., Bonar, S., Mielke, C., Albee, L., Weier, R., Graneto, M., Hanau, C., Perry, T., and Tripp, C. S. (2003) A selective H(K-2 inhibitor blocks NF-kappa B-dependent gene expression in interleukin-1 beta-stimulated synovial fibroblasts. J Biol Chem 278, 32861-32871 Muranyi, A., Gergely, P., Nagy, G. M., and Fekete, M. I. (1997) The possible role of protein phosphatase 2A in the sodium sensitivity of the receptor binding of opiate antagonists naloxone and naltrindole. Brain Res Bull 44, 273-279 Kumar, A., Yang, Y. L., Flati, V., Der, S., Kadereit, S., Deb, A., Haque, J ., Reis, L., Weissmann, C., and Williams, B. R. (1997) Deficient cytokine signaling in mouse embryo fibroblasts with a targeted deletion in the PKR gene: role of IRF-1 and NF- kappaB. EMBO J 16, 406-416 174 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. Cuddihy, A. R., Li, S., Tam, N. W., Wong, A. H., Taya, Y., Abraham, N., Bell, J. C., and Koromilas, A. E. (1999) Double-stranded-RNA-activated protein kinase PKR enhances transcriptional activation by tumor suppressor p53. Mol Cell Biol 19, 2475-2484 Rena, G., Woods, Y. L., Prescott, A. R., Peggie, M., Unterman, T. G., Williams, M. R., and Cohen, P. (2002) Two novel phosphorylation sites on FKHR that are critical for its nuclear exclusion. EMBO J21, 2263-2271 Rena, G., Prescott, A. R., Guo, S., Cohen, P., and Unterman, T. G. (2001) Roles of the forkhead in rhabdomyosarcoma (F KHR) phosphorylation sites in regulating 14- 3-3 binding, transactivation and nuclear targetting. Biochem J 354, 605-612 Xu, Z., and Williams, B. R. (2000) The B56alpha regulatory subunit of protein phosphatase 2A is a target for regulation by double-stranded RN A-dependent protein kinase PKR. Mol Cell Biol 20, 5285-5299 Yan, L., Lavin, V. A., Moser, L. R., Cui, Q., Kanies, C., and Yang, E. (2008) PP2A regulates the pro-apoptotic activity of FOXO]. J Biol Chem 283, 7411-7420 Browne, G. J ., Delibegovic, M., Keppens, S., Stalmans, W., and Cohen, P. T. (2001) The level of the glycogen targetting regulatory subunit R5 of protein phosphatase 1 is decreased in the livers of insulin-dependent diabetic rats and starved rats. Biochem J 360, 449-459 Doherty, M. J., Cadefau, J., Stalmans, W., Bollen, M., and Cohen, P. T. (1998) Loss of the hepatic glycogen-binding subunit (GL) of protein phosphatase 1 underlies deficient glycogen synthesis in insulin-dependent diabetic rats and in adrenalectonnized starved rats. Biochem J 333 ( Pt 2), 253-257 Cohen, P. T. (2002) Protein phosphatase 1--targeted in many directions. J Cell Sci 115, 241-256 Bollen, M., Keppens, S., and Stalmans, W. (1998) Specific features of glycogen metabolism in the liver. Biochem J 336 ( Pt 1), 19-31 Wada, T., Sasaoka, T., Funaki, M., Hori, H., Murakarni, S., Ishiki, M., Haruta, T., Asano, T., Ogawa, W., Ishihara, H., and Kobayashi, M. (2001) Overexpression of SH2-containing inositol phosphatase 2 results in negative regulation of insulin- induced metabolic actions in 3T3-L1 adipocytes via its 5'-phosphatase catalytic activity. Mol Cell Biol 21, l 63 3-1 646 175 63. 65. 66. 67. 68. 69. 70. 71. 72. 73. Tan, S. L., Tareen, S. U., Melville, M. W., Blakely, C. M., and Katze, M. G. (2002) The direct binding of the catalytic subunit of protein phosphatase 1 to the PKR protein kinase is necessary but not sufficient for inactivation and disruption of enzyme dimer formation. J Biol Chem 277, 36109-36117 Dohadwala, M., da Cruz e Silva, E. R, Hall, F. L., Williams, R. T., Carbonaro-Hall, D. A., Nairrn, A. C., Greengard, P., and Berndt, N. (1994) Phosphorylation and inactivation of protein phosphatase 1 by cyclin-dependent kinases. Proc Natl Acad Sci USA 91, 6408-6412 Bonnet, M. C., Daurat, C., Ottone, C., and Meurs, E. F. (2006) The N-terminus of PKR is responsible for the activation of the NF-kappaB signaling pathway by interacting with the IKK complex. Cell Signal 18, 1865-1875 Rui, L., Aguirre, V., Kim, J. K., Shulmarn, G. 1., Lee, A., Corbould, A., Dunaif, A., and White, M. F. (2001) Insulin/IGF-l and TNF-alpha stimulate phosphorylation of IRS-l at inhibitory Ser307 via distinct patlnways. J Clin Invest 107, 181-189 Fujishiro, M., Gotoh, Y., Katagiri, H., Sakoda, H., Ogihara, T., Anai, M., Onishi, Y., Ono, H., Abe, M., Shojima, N., Fukushima, Y., Kikuchi, M., Oka, Y., and Asano, T. (2003) Three mitogen-activated protein kinases inhibit insulin signaling by different mechanisms in 3T3-L1 adipocytes. Mol Endocrinol 17, 487-497 Takizawa, T., Tatematsu, C., and Nakanishi, Y. (2002) Double-stranded RNA- activated protein kinase interacts with apoptosis signal-regulating kinase 1. Implications for apoptosis signaling pathways. Eur J Biochem 269, 6126-6132 Ichijo, H., Nishida, E., Irie, K., ten Dijke, P., Saitoh, M., Moriguchi, T., Takagi, M., Matsumoto, K., Miyazono, K., and Gotoh, Y. (1997) Induction of apoptosis by ASK], a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275, 90-94 Chang, L., and Karin, M. (2001) Mammalian MAP kinase signalling cascades. Nature 410, 37-40 Buteau, J ., and Accili, D. (2007) Regulation of pancreatic beta-cell function by the forkhead protein FoxOl. Diabetes Obes Metab 9 Supp12, 140-146 Huang, H., and Tindall, D. J. (2007) Dynamic F oxO transcription factors. J Cell Sci 120, 2479-2487 Barthel, A., Schmoll, D., and Unterman, T. G. (2005) FoxO proteins in insulin action and metabolism. Trends Endocrinol Metab 16, 183-189 176 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. Gross, D. N., van den Heuvel, A. P., and Birnbaum, M. J. (2008) The role of FoxO in the regulation of metabolism. Oncogene 27, 2320-2336 Puigserver, P., Rhee, J., Donovan, J., Walkey, C. J., Yoon, J. C., Oriente, F., Kitarnura, Y., Altomonte, J ., Dong, H., Accili, D., and Spiegehnan, B. M. (2003) Insulin-regulated hepatic gluconeogenesis through FOXO] -PGC-1 alpha interaction. Nature 423, 550-555 Samuel, V. T., Choi, C. S., Phillips, T. G., Romanelli, A. J., Geisler, J. G., Bhanot, S., McKay, R., Monia, B., Shutter, J. R., Lindberg, R. A., Shulrnarn, G. 1., and Veniant, M. M. (2006) Targeting foxol in mice using antisense oligonucleotide improves hepatic and peripheral insulin action. Diabetes 55, 2042-2050 Dong, X. C., Copps, K D., Guo, S., Li, Y., Kollipara, R., DePinho, R. A., and White, M. F. (2008) Inactivation of hepatic Foxol by insulin signaling is required for adaptive nutrient homeostasis and endocrine gowth regulation. Cell Metab 8, 65-76 Tarnemoto, H., Kadowaki, T., Tobe, K., Yagi, T., Sakura, H., Hayakawa, T., Terauchi, Y., Ueki, K., Kaburagi, Y., Satoh, S., and et al. (1994) Insulin resistance and gowth retardation in rrnice lacking insulin receptor substrate-l. Nature 372, 182-186 Araki, B., Lipes, M. A., Patti, M. B., Bruning, J. C., Haag, B., 3rd, Johnson, R. S., and Kahn, C. R. (1994) Alternative pathway of insulin signalling in mice with targeted disruption of the IRS-1 gene. Nature 372, 186-190 Kido, Y., Burks, D. J ., Withers, D., Bruning, J. C., Kahn, C. R., White, M. F., and Accili, D. (2000) Tissue-specific insulin resistance in mice with mutations in the insulin receptor, IRS-1, and IRS-2. J Clin Invest 105, 199-205 Withers, D. J., Gutierrez, J. S., Towery, H., Burks, D. J., Ren, J. M., Previs, S., Zhang, Y., Bemal, D., Pons, S., Shulman, G. I., Bonner-Weir, S., and White, M. F. (1998) Disruption of IRS-2 causes type 2 diabetes in mice. Nature 391, 900-904 Tarniguchi, C. M., Ueki, K., and Kahn, R. (2005) Complementary roles of IRS-1 and IRS-2 in the hepatic regulation of metabolism. J Clin Invest 115, 718-727 Kubota, N., Kubota, T., Itoh, S., Kumagai, H., Kozono, H., Takarnoto, 1., Mineyama, T., Ogata, H., Tokuyama, K., Ohsugi, M., Sasako, T., Moroi, M., Sugi, K., Kakuta, S., Iwakura, Y., Noda, T., Ohnishi, S., Nagai, R., Tobe, K., Terauchi, Y., Ueki, K., and Kadowaki, T. (2008) Dynamic functional relay between insulin receptor substrate 1 and 2 in hepatic insulin signaling during fasting and feeding. Cell Metab 8, 49-64 177 84. 85. 86. 87. Opara, E., Kutlu, M., van Haeften, T., Kennedy, F. P., and Gerich, J. E. (1988) Effect of internalization and degadation of insulin on rat adipocyte insulin receptor binding kinetics. Clin Physiol Biochem 6, 1-11 Kusari, A. B., Byon, J ., Bandyopadhyay, D., Kenner, K. A., and Kusari, J. (1997) Insulin-induced rrnitogen-activated protein (MAP) kinase phosphatase-1 (MKP-l) attenuates insulin-stimulated MAP kinase activity: a mechanism for the feedback inhibition of insulin signaling. Mol Endocrinol 11, 1532-1543 Tremblay, F., and Marette, A. (2001) Arrnino acid and insulin signaling via the mTOR/p70 S6 kinase pathway. A negative feedback mechanism leading to insulin resistance in skeletal muscle cells. J Biol Chem 276, 38052-3 8060 Martin, K. A., Merenick, B. L., Ding, M., Fetalvero, K. M., Rzucidlo, E. M., Kozul, C. D., Brown, D. J., Chiu, H. Y., Shyu, M., Drapeau, B. L., Wagner, R. J., and Powell, R. J. (2007) Raparnycin promotes vascular smooth muscle cell differentiation through insulin receptor substrate-l/phosphatidylinositol 3- kinase/Akt2 feedback signaling. J Biol Chem 282, 36112-36120 178 CHAPTER 6: Conclusion and Discussion In this thesis, my research focused on the major hepatic cellular activities, including insulin signaling activity and the saturated FFA-induced cytotoxicity and apoptosis. Human hepatoblastoma cells (HepG2/C3A) were used for the study. These human-ori gin cells offer the advantages of ease of culture and experimentation, and have been shown to retain many hepatospecific functions and tlnerefore suggested as a good model cell system for hepatocellular functions, such as lipid metabolism (1) and fatty acid transport (2). Cell lines offer an advantage over primary hepatocytes in that they are more amenable to genetic manipulation. In the saturated F FA-induced cytotoxicity and apoptosis studies, two different types of F FAs were employed, i.e. saturated and monounsaturated fatty acids, corresponding to the major dietary fractions. Palmitate was chosen as the representative saturated fatty acid and oleate as the monounsaturated fatty acid. They are the major fatty acids of their classes found in serum/plasma. While the total concentration of FFAs in plasma may reach rrnillirnolar range under pathological conditions (3), most studies of obese/type 2 diabetic patients have reported fatty acid concentration at about 0.7 mM. We therefore used 0.7 mM or less as the concentrations of palmitate and oleate. Using the research techniques from molecular and cellular biology, a key protein, PKR, was identified to be highly involved in both palmitate-induced apoptosis and the regulation of insulin signaling. PKR in Palmitate-Induced Apoptosis—Multiple mecharnisms have been proposed for palmitate-induced cytotoxicity and apoptosis. Chapter 2 of this thesis 179 discusses how PKR, found to be anti-apoptotic in HepG2 cells, plays a central role in mediating the palrrnitate-initiated pathways that led to hepatic apoptosis. PKR is best krnown for its pro-apoptotic role by phosphorylating eIF-20. and thereby inhibiting general protein synthesis (4, 5). In contrast, more recent studies suggest that PKR has an anti- apoptotic role in certain conditions (6-9). Elevated levels of PKR protein and activity were observed in human breast cancer cells (6), melanoma cells (7, 8), and hepatitis C virus (HCV)-related hepatocellular carcinoma (9). However, the mechanism is still unclear. Research suggests that PKR may suppress apoptosis by activating the NF-ch pathway before phosphorylating eIF-20t, tlnereby initially inducing cell survival and, subsequently, cell death in NIH3T3 cells expressing PKR (10). Chapter 2 discusses the two potential pathways by which PKR mediates the anti-apoptotic signals through Bcl-2. These results provide a potential mechanism of palrrnitate-induced apoptosis, by suppressing PKR, in HepG2 cells. In the first pathway, PKR is involved in controlling the transcription of Bel-2 in HepG2 cells, mediated by the transcription factor NF-ch. As discussed in Chapter 2, NF- KB is commonly considered an anti-apoptotic transcription factor (11, 12), which is consistent with the anti-apoptotic role of PKR in HepG2 cells. However the second anti- apoptotic patlnway, namely, that PKR up-regulates the phosphorylation of Bcl-2 at Ser70, is mediated by JNK, a multi-functional kinase considered a “double-edged sword” (13) in regulating apoptosis. Most of the studies on JNK used a specific chemical inhibitor, SP600125. As discussed in Chapter 2, the concentrations or treatment lengths of this inhibitor in palnnitate-related studies play an unclear but important role in determirning the apoptosis level of the cells, thereby raising concerns of using this inhibitor in the studies 180 of the palmitate-induced apoptosis. Furthermore, this particular JNK inhibitor was found to ineffectively inhibit the activity of JNK in the palnnitate media. Since SP600125 is hydrophobic, it is possible that the chemical inhibitor may aggegate with palrrnitate, which is also hydrophobic, or even BSA, the vehicle used to dissolve palmitate, thereby losing its inhibitory function on JNK. In Liedtke et al. (14), an adenol-viral vector expressing dominant-negative TAK], which was identified to specifically inhibit JNK activity, was shown to induce strong apoptosis in Huh7 and HepG2 cells, supporting the result that JNK is anti-apoptotic in HepG2 cells. The results in Chapter 2 show for the first time that palmitate treatment decreases the activity of PKR; however, the mechanism by which palnnitate induces dephosphorylation of PKR is still unclear. The auto-phosphorylation of PKR requires its dimerization (15, 16) and binding of ATP to its ATP binding site on the catalytic domain (17). Another PhD student in our lab is currently investigating the potential interaction between the free palmitate molecules and the ATP binding site of PKR. Interestingly, both computational and experimental results indicate that palmitate binds to the ATP binding site on PKR and thereby disrupts the auto-phosphorylation of PKR. In summary, an anti-apoptotic role of PKR, mediated by the protein expression level and phosphorylation status of Bel-2, was identified in HepG2 cells. The transcription factor NF-ch and the MAP kinase JNK appear to be involved in mediating the effects of PKR on the protein level and the phosphorylation of Bcl-2, respectively. Furthermore, palrrnitate was found to suppress these two pathways by inhibiting PKR and tlnereby attenuate the anti-apoptotic machinery of HepG2 cells. 181 PKR in Regulating Insulin Signaling—As a serine-threonine protein kinase that is ubiquitously expressed, PKR is actively involved in multiple signaling pathways by regulating the activities of transcription factors, protein kinases, and phosphatases (] 8). Therefore, PKR serves as a signal integator, regulating various cellular activities, such as protein synthesis and degadation, stress and immune responses, and apoptosis (18, 19). In Chapter 5 of this thesis, a novel function of PKR in regulating insulin signaling pathway was identified for HepG2 cells. In detail, PKR differentially regulates the two major IRS proteins, IRS] and IRSZ, through post-translational modification and transcriptional regulation, respectively. The involvement of PKR in regulating insulin signaling may connect the effectors of PKR with insulin signaling. For example, according to the results in Chapter 5, the inhibitory serine phosphorylation of IRS] induced by ceramide is attributed to the activation of PKR by cerannide. In addition to ceramide, PKR is also responsive to many other factors, e.g., HCV infection (20), endoplasmic reticulum (ER) stress (2] , 22), cytokines such as tumor necrosis factor (TNF)-01 (23) and interleukin (IL)-1 (24), deoxyrnivalenol (DON, or vorrnitoxin) (25) and lipopolysaccharide (24). Whether and how PKR is involved in regulating insulin signaling in response to these factors is worth further investigation. The function of PKR in regulating the protein level of IRS2 involves the transcription factor Fox01. Fox01 directs the expression of genes involved in a wide variety of cellular responses. The gene expression progarn regulated by Fox01 is considered to be cell-protective and anti-apoptotic (26). For example, Fox01 facilitate DNA repair by up-regulating GADD45 and DDBl and suppresses cell-cycle by increasing expression of p27'dp, thereby protecting cells from DNA damage (26). F 0x01 182 also suppresses the rate of aging by up-regulating superoxide dismutase (27), which repairs oxidative damage (27). The effect of PKR on Fox01 has not been studied previously. Chapter 5 shows that PKR dephosphorylates and activates FoxOl mediated by the protein phosphatase PP2A. PKR is best known for its pro-apoptotic role by phosphorylating eIF-2a and thereby inhibiting general protein synthesis (4, 5). In contrast, more recent studies including the work in Chapter 2 suggest that PKR has an anti- ap0ptotic role (8, 28). Due to the cytoprotective and anti-apoptotic roles of Fox01, the connection identified between PKR and Fox01 also supports the anti-apoptotic function of PKR, thereby suggesting another interesting research target for PKR-regulated apoptosis. In addition to the regulatory function of PKR on insulin signaling, PKR was also identified as a potential downstream substrate of the insulin signaling pathway. Irnitially known as a virus responsive protein, PKR has been shown to be involved in regulating a variety of important cellular activities, such as immune response, general protein syntlnesis, protein degadation, apoptosis, cell gowtln and proliferation (29), and therefore associated with diseases including virus infection, e.g., HCV infection (30), cancer (31), neuron degenerative diseases, e.g., Alzheimer's (32), etc. Therefore, by revealing the involvement of PKR in insulin signaling pathways, I propose that PKR has a central position that could correlate insulin (or IGF-l) signaling with other cellular activities, such as immune response, apoptosis, protein syntlnesis and degadation, some of which have been already identified to be related to insulin action (33-37). The results in Chapter 5 therefore suggest a promising research direction to identify potential interactions between insulin signaling and these PKR-related cellular activities. 183 Chapter 5 also showed the abolition of the down-regulation in PKR phosphorylation by insulin upon longer treatment with insulin (Fig. 5.6B), suggesting a possible negative feedback in the regulation of PKR by insulin. Indeed, insulin signaling is tuned by various negative feedbacks, e.g., degadation of IR induced by insulin binding (38), negative regulation of insulin signaling activity by MAPKs (39), mTOR (40), PI-3 kinase (41), etc. These negative feedbacks modulate the downstream signaling events induced by long-term treatment of insulin (42). Given that many signaling components of the feedback pathways in the insulin signaling system have the potential to cross-talk with PKR, the contribution of each to the dynamic profile of PKR in the insulin network system is unclear. This complex feedback system represents a good target for mathematical modeling. Indeed, there have been many mathematical models proposed to model insulin action, most of which, however, focus on the binding kinetics of insulin with its receptor, IR (43, 44), and GLUT4 trafficking (45, 46). There are very few models that describe the kinetics of insulin signaling, incorporating only a limited number of feedback pathways (47, 48). This is due to the difficulty in quantitatively measuring all the biochemical and kinetic parameters of the signaling pathways. Thus, an alterrnative modeling method is needed, which infers the network dynamics in the absence of the detail kinetic parameters. Currently a Boolean network model is beirng built in our goup to simulate the insulin signaling network. It is a discrete logic model that solely relies on the network structure. Preliminary results show that this model, in spite of its simplicity, is able to reproduce some of the key dynamic features of the insulin signaling system, thereby suggesting promising applications of this methodology in providing valuable insights into the execution and regulation of insulin signaling. 184 NAFLD, NASH, and Insulin Resistance—The hepatic cellular activities that are studied in this thesis, including insulin signaling and saturated FFA-induced cytotoxicity and apoptosis, are directly related to the initiation and development of the liver disorders such as NAFLD, NASH, and hepatic insulin resistance. Insulin resistance is very common in patients who have NAFLD (49), and is believed to play an important role in the pathogenesis of NAFLD by impairing mitochondrial [ii-oxidation of FFAs and promoting fat accumulation in liver (50). Indeed, obesity and type 2 diabetes, in which insulin resistance is usually prevalent, are the two most common risk factors for developing NAFLD. Although the development of NASH, the more severe, second stage of NAFLD, has not been fully understood, several models have been proposed for the pathogenesis of NASH from NAFLD (51). The prevailing theory is the two-lnit model (52), with insulin resistance accounting for the first hit, inducing fatty liver, and FFA-induced cytotoxicity accounting for the second hit, inducing inflammation and cell death (52). Therefore, it is well-recognized that in the liver, insulin resistance is the cause of fat accumulation, NAFLD and eventually NASH, rather than the consequence of NAF LD. In this thesis, a regulatory role of PKR in irnsulin signaling was identified (Chapter 5). Furtlnermore, saturated FFA, i.e., palrrnitate, decreased the phosphorylation of PKR (Chapter 2). Taken together, these findings raise an interesting possibility for further study. Namely, can palmitate regulate insulin signaling through PKR? Palmitate has been identified as an inducer of insulin resistance at the whole-body level (53). In hepatocytes, palmitate inhibits insulin action by both receptor and post-receptor events (54, 55). However, it has been recently shown that irnlnibiting insulin signaling by palnritate in 185 hepatoma cells is dependent upon oxidation of fatty acyl-CoA species and requires intact insulin receptor expression (55). Therefore, the palmitate-induced insulin resistance is likely due to the combination of different events, in which the down-regulation of IR expression may play a major role, although the exact mechanism is still not clear. It may be worthwhile to evaluate the potential involvement of PKR in mediating the induction of hepatic insulin resistance by palrrnitate. In summary, a part of the research in this thesis focused on the regulation of insulin signaling and potentially induction of insulin resistance, which play roles in the development of type 2 diabetes, as well as fat accrrrnulation in the liver, which is a big risk factor of NAFLD. The other half of the research was on saturated FFA-induced cytotoxicity and apoptosis, which are known to contribute to the development of NASH in NAFLD patients (56, 57). Therefore, the detailed mechanistic information involved in the regulation of insulin signaling and palmitate-induced cytotoxicity would enhance our understanding of the pathology of liver-related diseases such as NAFLD, NASH, hepatic insulin resistance, and type 2 diabetes, and may further identify potential research and drug targets. Research Tools in Systems Biology—In addition to the molecular biology studies on the saturated FFA-induced apoptosis and cytotoxicity, the principles of systems biology was also applied in this thesis. In collaboration with mathematicians and computational experts, systems biology methodologies were applied to gain a systems- level view on lipotoxicity. Phenotype-specific gene networks were reconstructed and analyzed. The rationale for applying systems biology in studying the cellular activities is that reconstructing gene networks fiom the gene- gene interactions, which could be 186 extracted fi'om gene expression profiles, could shed light into the possible mechanism involved in context-specific cellular activities. The assumption behind this is that only a subset of the genes are responsible for a certain phenotype and the mechanistic information is enriched in the phenotype-specific gene network. Two strategies (Chapters 3 and 4) were developed for reconstruction of the phenotype-specific gene network. Both strategies consist of two phases, 1) selection of the phenotype-related genes and 2) reconstruction of the gene network that is specific to the phenotype. For each phase, we applied two different approaches to achieve our goal. For the selection of phenotype-related genes, it is well recognized that regular statistical feature selection methods are usually computationally expensive and susceptible to the quality of data (58), thereby requiring integation of additional information. In the two methods discussed in Chapters 3 and 4, the information incorporated to facilitate the gene-selection process was obtained from two different sources, metabolic data (Chapter 3) and prior knnowledge (Chapter 4). Botln approaches are adaptable to different phenotypes. However, while the first approach integates metabolic information, which may be more biologically meaningful since it reflects the real biological process that induces the phenotype, it is more labor-intensive in obtaining the metabolic information. On the other hand, the second approach, which incorporates prior knowledge instead of experimental data, provides valuable information in selecting the genes, but at the same time, the results of this approach largely depend on the source and quality of the prior knowledge. In addition, some prior knowledge may not be valid under all biological systems and experimental conditions. Therefore, selecting the source 187 of prior krnowledge and optimizing the weight of the prior knowledge in the gene- selection process are critical to this method. Similarly, we viewed the reconstruction of the phenotype-specific gene network from two different standpoints gounded in different understandings of the biological system. In Chapter 3, the gene network was composed of gene pairs with high synergy scores. This network therefore captured genes that interacted synergistically (cooperatively or antagonistically) on the phenotype. Since the concept of synergy is intuitively based upon the phenotype, we obtained a single phenotype-specific network from the gene expression profiles of the different conditions. Therefore, this method does not require the comparison of different networks across the experimental conditions, thereby eliminating the noise due to the artificial differences, such as the sample size and noise level differences, between the experimental conditions. However, this method was not designed to recover the direct physical interactions between genes and proteins, thereby limiting the direct experimental validation of synergistic gene pairs in the synergy network. Nevertheless, this method was able to uncover some novel mecharnistic information that was not revealed by other methods. In Chapter 4, gene modules and module interactions, instead of pair-wised gene interactions, were identified. This method takes advantage of the prior knowledge and can tune the amount of weight or value to place on the prior knowledge in obtaining optimal results. Although this module network is not capable of uncovering specific gene- gene interactions, it successfully recovered functional gene modules, in which genes involved in certain cellular functions or activities were significantly enriched. Therefore, this module-interaction network captured the organized modularity in biological network 188 systems. However, this method requires building and comparing the networks under different conditions, making it susceptible to sampling differences across experimental conditions. Therefore, a proper normalization process is needed to reduce artifactual noise. Thus, both system biology methodologies described in this thesis were able to capture a systems-level view of the phenotype of interest and provided valuable insiglnts into potential mechanisms involved in the induction of the different phenotypes. Both methods suggested some similar findings, which was encouraging. These findings include the potential involvement in palmitate-induced cytotoxicity of endoplasmic reticulum (ER) stress, and intracellular metabolism regulation, etc., some of which are supported by the literature (59-61) and experiments. Interestingly, the gene module approach also suggested that PKR is involved in regulating apoptosis that contributes to palmitate-induced cytotoxicity. This finding is consistent with the molecular biology study of PKR in palnnitate-induced apoptosis (Chapter 2). Summary In this thesis, I investigated the hepatocellular activities, such as regulation of insulin signaling and saturated FFA-induced cytotoxicity, which are krnown to play central roles in the initiation and development of multiple liver disorders including NAFLD, NASH, and hepatic insulin resistance. Using molecular and cellular biology techniques, I identified that 1), PKR mediates the palmitate-induced apoptosis by regulating Bcl-2, and 2), PKR, affected by insulin, is involved in insulin signaling by differentially regulating the major IRS proteins. In general, PKR exploits a role as a 189 signaling hub that regulates multiple hepatocellular activities through different signaling events. In addition, in collaboration with other researchers, I also applied systems biology methodologies to obtain furtlner understandings into saturated FFA-induced lipotoxicity. Both strategies, based on different aspects of the gene network systems, provided valuable mechanistic information on the phenotype of interest, thereby serving as powerful research tools that complement the molecular biology study, and suggested novel research targets and possible mechanisms. 190 10. References Cianflone, K. ,Vu, H. ,Zhang, Z, and Sniderrnan, A. D. (1994) Effects of albumin on lipid synthesis, apo B- 100 secretion, and LDL catabolism in HepG2 cells. Atherosclerosis 107, 125-135 Guo, W., Huang, N., Cai, J., Xie, W., and Hamilton, J. A. (2006) Fatty acid transport and metabolism in HepG2 cells. Am J Physiol Gastrointest Liver Physiol 290, G528-534 Artwohl, M., Roden, M., Waldhausl, W., Freudenthaler, A., and Baumgartrner- Parzer, S. M. (2004) Free fatty acids trigger apoptosis and inhibit cell cycle progession in human vascular endothelial cells. FASEBJ. 18, 146-148 Taylor, S. S., Haste, N. M., and Ghosh, G. (2005) PKR and eIF2alpha: integation of kinase dimerization, activation, and substrate docking. Cell 122, 823-825 Stark, G. R., Kerr, 1. M., Williams, B. R., Silvermarn, R. H., and Schreiber, R. D. (1998) How cells respond to interferons. Annu Rev Biochem 67, 227-264 Kim, S. H., Forman, A. P., Mathews, M. B., and Gunnery, S. (2000) Human breast cancer cells contain elevated levels and activity of the protein kinase, PKR. Oncogene 19, 3086-3094 Kim, S. H., Gunnery, S., Choc, J. K., and Mathews, M. B. (2002) Neoplastic progession in melanoma and colon cancer is associated with increased expression and activity of the interferon-inducible protein kinase, PKR. Oncogene 21, 8741- 8748 Delgado Andre, N., and De Lucca, F. L. (2007) Knockdown of PKR expression by RNAi reduces pulmonary metastatic potential of Bl6-F10 melanoma cells in rrnice: Possible role of NF-kappaB. Cancer Lett 258, 118-125 Hiasa, Y., Karnegaya, Y., Nuriya, H., Onji, M., Kohara, M., Schmidt, E. V., and Chung, R. T. (2003) Protein kinase R is increased and is firnctional in hepatitis C virus-related hepatocellular carcinoma. Am J Gastroenterol 98, 2528-2534 Donze, O., Deng, J ., Curran, J ., Sladek, R., Picard, D., and Sonenberg, N. (2004) The protein kinase PKR: a molecular clock that sequentially activates survival and death progarns. EMBO J 23, 564-571 191 ll. 12. l3. 14. 15. 16. l7. 18. 19. 20. 21. 22. Kucharczak, J ., Simmons, M. J ., Fan, Y., and Gelinas, C. (2003) To be, or not to be: NF-kappaB is the answer--role of Rel/NF-kappaB in the regulation of apoptosis. Oncogene 22, 8961-8982 Li, M., Shillinglaw, W., Henzel, W. J., and Beg, A. A. (200]) The Rela(p65) subunit of NF-kappaB is essential for irnhibiting double-stranded RNA-induced cytotoxicity. J Biol Chem 276, 1185-] 194 Liu, J ., and Lin, A. (2005) Role of JNK activation in apoptosis: a double-edged sword. Cell Res 15, 36-42 Liedtke, C., Plumpe, J ., Kubicka, S., Bradham, C. A., Manns, M. P., Brenner, D. A., and Trautwein, C. (2002) Jun kinase modulates tumor necrosis factor-dependent apoptosis in liver cells. Hepatology 36, 315-325 Dey, M., Cao, C., Dar, A. C., Tamura, T., Ozato, K., Sicheri, F., and Dever, T. E. (2005) Mechanistic link between PKR dimerization, autophosphorylation, and eIF2alpha substrate recognition. Cell 122, 901-913 Lemaire, P. A., Lary, J., and Cole, J. L. (2005) Mechanism of PKR activation: dimerization and kinase activation in the absence of double-stranded RNA. J Mol Biol 345, 81-90 Dar, A. C., Dever, T. E., and Sicheri, F. (2005) Higher-order substrate recognition of eIF2alpha by the RNA-dependent protein kinase PKR. Cell 122, 887-900 Williams, B. R. (2001) Signal integation via PKR. Sci STKE 2001, RE2 Barber, G. N. (2005) The dsRNA-dependent protein kinase, PKR and cell death. Cell Death Difi‘er 12, 563-570 Yan, X. B., Battaglia, S., Boucreux, D., Chen, Z., Brechot, C., and Pavio, N. (2007) Mapping of the interacting domains of hepatitis C virus core protein and the double-stranded RNA-activated protein kinase PKR. Virus Res 125, 79-87 Onuki, R., Bando, Y., Suyama, E., Katayarna, T., Kawasaki, H., Baba, T., Tohyama, M., and Taira, K. (2004) An RNA-dependent protein kinase is involved in tunicamycin-induced apoptosis and Alzheimer's disease. EMBO J23, 959-968 Langland, J. 0., Cameron, J. M., Heck, M. C., Jancovich, J. K., and Jacobs, B. L. (2006) Inhibition of PKR by RNA and DNA viruses. Virus Res 119, 100-110 192 rm 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. Meusel, T. R., Kehoe, K. B., and Imani, F. (2002) Protein kinase R regulates double-stranded RNA induction of TNF-alpha but not IL-1 beta mRN A in human epithelial cells. J Immunol 168, 6429-6435 Goh, K. C., deVeer, M. J ., and Williams, B. R. (2000) The protein kinase PKR is required for p38 MAPK activation and the innate immune response to bacterial endotoxin. EMBO J 19, 4292-4297 Zhou, H. R., Lau, A. S., and Pestka, J. J. (2003) Role of double-stranded RNA- activated protein kinase R (PKR) in deoxynivalenol-induced ribotoxic stress response. T oxicol Sci 74, 335-344 van der Horst, A., and Burgering, B. M. (2007) Stressing the role of FoxO proteins in lifespan and disease. Nat Rev Mol Cell Biol 8, 440-450 Murphy, C. T., McCarroll, S. A., Bargnann, C. 1., Fraser, A., Kamatln, R. S., Ahringer, J., Li, H., and Kenyon, C. (2003) Genes that act downstream of DAF-l6 to influence the lifespan of Caenorhabditis elegans. Nature 424, 277-283 Yang, X., and Charn, C. (2009) Repression of PKR mediates palmitate-induced apoptosis in HepG2 cells through regulation of Bcl-2. Cell Res Williams, B. R. (1997) Role of the double-stranded RNA-activated protein kinase (PKR) in cell regulation. Biochem Soc Trans 25, 509-513 Korth, M. J ., and Katze, M. G. (2000) Evading the interferon response: hepatitis C virus and the interferon-induced protein kinase, PKR. Curr T op Microbiol Immunol 242, 197-224 Shir, A., Friedrich, I., and Levitzki, A. (2003) Tumor specific activation of PKR as a non-toxic modality of cancer treatment. Semin Cancer Biol 13, 309-314 Peel, A. L. (2004) PKR activation in neurodegenerative disease. J Neuropathol Exp Neurol 63, 97-105 Grinspoon, S. (2003) Mechanisms and strategies for insulin resistance in acquired immune deficiency syndrome. Clin Infect Dis 37 Suppl 2, S85-90 Curi, R., Homem-de-Bittencourt, P. 1., Jr., Costa-Rosa, L. F., Femandes, L. C., el- Razi, S., Seelaender, M. C., and Carpinelli, A. R. (1995) Insulin and prostaglandins as signals between tumor and the host immune system. Braz J Med Biol Res 28, 773-779 193 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. Keku, T. 0., Lund, P. K., Galanko, J., Simmons, J. G., Woosley, J. T., and Sandler, R. S. (2005) Insulin resistance, apoptosis, and colorectal adenoma risk. Cancer Epidemiol Biomarkers Prev 14, 2076-2081 Bolster, D. R., Jefferson, L. S., and Kimball, S. R. (2004) Regulation of protein synthesis associated with skeletal muscle hypertrophy by insulin-, amino acid- and exercise-induced signalling. Proc Nutr Soc 63, 351-3 56 Eley, H. L., Russell, S. T., and Tisdale, M. J. (2008) Role of the dsRNA-dependent protein kinase (PKR) in the attenuation of protein loss from muscle by insulin and insulin-like gowtln factor-I (IGF-I). Mol Cell Biochem 313, 63-69 Opara, E., Kutlu, M., van Haeften, T., Kennedy, F. P., and Gerich, J. E. (1988) Effect of internalization and degadation of insulin on rat adipocyte insulin receptor binding kinetics. Clin Physiol Biochem 6, l-ll Kusari, A. B., Byon, J., Bandyopadhyay, D., Kenner, K. A., and Kusari, J. (1997) Insulin-induced nnitogen-activated protein (MAP) kinase phosphatase-1 (MKP-l) attenuates insulin-stimulated MAP kinase activity: a mechanism for the feedback inhibition of insulin signaling. Mol Endocrinol 11, 1532-1543 Tremblay, F ., and Marette, A. (2001) Amino acid and insulin signaling via the mTOR/p70 S6 kinase pathway. A negative feedback mecharnism leading to insulin resistance in skeletal muscle cells. J Biol Chem 276, 38052-3 8060 Martin, K. A., Merenick, B. L., Dirng, M., Fetalvero, K. M., Rzucidlo, E. M., Kozul, C. D., Brown, D. J., Chiu, H. Y., Shyu, M., Drapeau, B. L., Wagner, R. J., and Powell, R. J. (2007) Raparnycin promotes vascular smooth muscle cell differentiation through insulin receptor substrate-1/phosphatidylinositol 3- kinase/Akt2 feedback signaling. J Biol Chem 282, 36112-36120 Herschkovitz, A., Liu, Y. F., Ilan, B., Ronen, D., Boura-Halfon, S., and Zick, Y. (2007) Common inhibitory serine sites phosphorylated by IRS-l kinases, triggered by insulin and inducers of insulin resistance. J Biol Chem 282, 18018-18027 Wanant, S., and Quon, M. J. (2000) Insulin receptor binding kinetics: modeling and simulation studies. J Theor Biol 205, 355-364 Quon, M. J ., and Carnpfield, L. A. (1991) A mathematical model and computer simulation study of insulin receptor regulation. J Theor Biol 150, 59-72 194 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. Holman, G. D., Lo Leggio, L., and Orshman, S. W. (1994) Insulirn-stimulated GLUT4 glucose transporter recycling. A problem in membrane protein subcellular trafficking tlnrough multiple pools. J Biol Chem 269, 17516-17524 Quon, M. J. (1994) Advances in kinetic analysis of insulin-stimulated GLUT-4 translocation in adipose cells. Am J Physiol 266, El44-150 Sogard, P., Harlen, M., Svensson, L. T., Zieratln, J. R., and Nilsson, P. (2005) Integation of mathematical and experimental approaches to resolve insulin signalling. Acta Physiol Scand 183, 125-126 Sedaghat, A. R., Sherman, A., and Quon, M. J. (2002) A mathematical model of metabolic insulin signaling pathways. Am J Physiol Endocrinol Metab 283, E1084- 1 101 Hamaguchi, M., Kojima, T., Takeda, N., Nakagawa, T., Tarriguchi, H., Fujii, K., Omatsu, T., Nakajima, T., Sarui, H., Shimazaki, M., Kato, T., Okuda, J ., and Ida, K. (2005) The metabolic syndrome as a predictor of nonalcoholic fatty liver disease. Ann Intern Med 143, 722-728 Salt, W. B., 2nd (2004) Nonalcoholic fatty liver disease (NAFLD): a comprehensive review. J Insur Med 36, 27-41 Khashab, M., and Chalasani, N. (2007) Use of insulin sensitizers in NASH. Endocrinol Metab Clin North Am 36, 1067-1087; xi Sanyal, A. J., Carnpbell-Sargent, C., Mirshahi, F., Rizzo, W. B., Contos, M. J., Sterling, R. K., Luketic, V. A., Shiffrnan, M. L., and Clore, J. N. (2001) Nonalcoholic steatohepatitis: association of insulin resistance and mitochondrial abnormalities. Gastroenterology 120, 1183-1192 Thrush, A. B., Heigenhauser, G. J ., Mullen, K. L., Wright, D. C., and Dyck, D. J. (2008) Palnnitate acutely induces insulin resistance in isolated muscle from obese but not lean humans. Am J Physiol Regul Integr Comp Physiol 294, R1205-1212 Mordier, S., and Iynedjian, P. B. (2007) Activation of mammalian target of rapamycin complex 1 and insulin resistance induced by palrrnitate in hepatocytes. Biochem Biophys Res Commun 362, 206-211 Ruddock, M. W., Stein, A., Landaker, E., Park, J., Cooksey, R. C., McClain, D., and Patti, M. E. (2008) Saturated Fatty Acids Inhibit Hepatic Insulin Action by Modulating Insulin Receptor Expression and Post-Receptor Signalling. J Biochem 195 56. 57. 58. 59. 60. 61. Wieckowska, A., Papouchado, B. G., Li, Z., Lopez, R., Zein, N. N., and Feldstein, A. E. (2008) Irncreased hepatic and circulating interleukin-6 levels in human nonalcoholic steatohepatitis. Am J Gastroenterol 103, 1372-1379 Nehra, V., Angulo, P., Buchmarn, A. L., and Lindor, K D. (2001) Nutritional and metabolic considerations in the etiology of nonalcoholic steatohepatitis. Dig Dis Sci 46, 2347-23 52 Le Phillip, P., Balnl, A., and Ungar, L. H. (2004) Using prior knowledge to improve genetic network reconstruction from rrnicroarray data. In Silica Biol 4, 335-353 Guo, W., Wong, S., Xie, W., Lei, T., and Luo, Z. (2007) Palmitate modulates intracellular signaling, induces endoplasmic reticulum stress, and causes apoptosis in mouse 3T3-L1 and rat primary preadipocytes. Am J Physiol Endocrinol Metab 293, E576-586 Karaskov, E., Scott, C., Zhang, L., Teodoro, T., Ravazzola, M., and Volchuk, A. (2006) Chronic palmitate but not oleate exposure induces endoplasmic reticulum stress, which may contribute to INS-1 pancreatic beta-cell apoptosis. Endocrinology 147, 3398-3407 Listenberger, L. L., Han, X., Lewis, S. E., Cases, S., Farese, R. V., Jr., Ory, D. S., and Schaffer, J. E. (2003) Triglyceride accumulation protects against fatty acid- induced lipotoxicity. Proc Natl Acad Sci U S A 100, 3077-3082 196 M71111117111111!flilllllllllliilll“