. . abn_#-,,w%1€ n. a?" J i... / ._ _$?..mewhfl tank... a!" “59...... n. ‘ ‘55"? n_ a A A t 5.1 3.51, ; ‘ J" 0 ‘.E 6.1! Rd .anti . . gm» wagging ., :, Yvu. . n. ‘ .n ,‘ a. , A? l. . 5! I l THEELS Z LIBRARY 1000 Michigan State University This is to certify that the dissertation entitled MECHANISMS LEADING TO TRANSLATION ARREST IN AUTOGRAPHA CALIFORNICA MULTINUCLEOCAPSID NUCLEOPOLYHEDROVIRUS-INFECTED LYMANTRIA DISPAR CELLS presented by Christy Mecey has been accepted towards fulfillment of the requirements for the Doctoral degree in Genetics :4 m ' M 2 Major Professor’s Signature [77191 3 ‘ 2 O o é (j I Date MSU is an Affirmative Action/Equal Opportunity Institution on-a-o-a-o-o-I---n-n--I- .o-|-I-o-I-I---I-b-I-O-O-t-C-I-D-I-O-o-u- -- 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 l @9739qu 2/05 p:/C|RC/DateDue.indd-p.1 MECHANISMS LEADING TO TRANSLATION ARREST IN A UT OGRAPHA CALIF ORNICA MULTINUCLEOCAPSID NUCLEOPOLYHEDROVIRUS-INFECTED LYAMNTRIA DISPAR 652Y CELLS By Christy Mecey A DISSERTATION Submitted to Michigan State University in partial fulfillment for the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics Program 2006 ABSTRACT MECHANISMS LEADING TO TRANSLATION ARREST IN AUTOGRAPHA CALIFORNICA MULTINUCLEOCAPSID NUCLEOPOLYHEDROVIRUS-INFECTED LYMANTRIA DISPAR 652Y CELLS By Christy Mecey Infection of the Lymantria dispar 652Y (Ld652Y) cell line with the baculovirus Autographa calzfornica Multinuoleocapsid nucleopolyhedrovirus (AcMNPV) results in global translation arrest, initiation of apoptosis in the absence of a functional, virus- encoded, apoptotic suppressor and a non-productive infection. While AcMNPV enters Ld652Y cells and transcription of virus genes occurs, virus progeny are not produced. Expression of a single gene, host range factor-1 (hrf-I), isolated from Lymantria dispar Multinuoleocapsid Nucleopolyhedrovirus (LdMNPV), precludes translation arrest in AcMNPV-infected Ld652Y cells further allowing permissive replication of the virus. HRF-l also enables replication of Hyphantria cunea NPV, Bombyx mori NPV and Spodoptera exigua NPV in Ld652Y cells indicating that hrf-I is an essential host range factor for baculovirus infection in Ld652Y cells. Database and motif searches conducted with the hrf-I sequence provide no indication to its function. The goal of this study was to determine the mechanism of translation arrest in AcMNPV-infected Ld652Y cells to further identify possible functions for HRF-l. We Show that eIF20. phosphorylation at serine 51 occurs in AcMNPV-infected deSZY cells at 12 h.p.i. correlating to the onset of translation arrest. Ld652Y cells infected with vAchrf-l Show a reduced level of eIFZa phosphorylation for the full-length peIF20tserSl protein at 6, 12, 24 and 48 h.p.i. as compared to mock-infected cells. PK2, the AcMNPV eIF20t kinase inhibitor, was expressed in AcMNPV-infected Ld652Y cells but did not prevent eIF20t ser5 l phosphorylation. An apparent proteolytic cleavage of peIF20t was observed beginning at 12 h.p.i.; this cleavage event was investigated further. In Sf21 cells, a permissive cell line for AcMNPV replication, peIF20tscr51 was cleaved in both AcMNPV and vAchrf-l- infected cells indicating that HRF-l was not directly responsible for the cleavage. Cell 5ch l permeable protease inhibitors did not block cleavage of peIFZa . However, the proteasome inhibitor epoxomicin did reduce phosphorylation of eIF2a at serine 51, block cleavage of peIFZIIIschl and inhibit virus replication in AcMNPV and vAchrf-l-infected 8121 cells and vAchrf-l-infected Ld652Y cells. The proteasome inhibitor M0132 mimicked the effects of epoxomicin in vAchrf-l-infected Ld652Y cells. However, treatment with MG132 in AcMNPV and vAchrf-l-infected SfZl cells resulted in reduced virus titers and no block of peIFZItscr51 cleavage. This study suggests a role for the proteasome in enabling baculovirus replication. A cross-reactive antibody to the human eIF2 kinase, PKR, showed enhanced expression of a protein in vAchrf-l-infected cells at 12-24 h.p.i. Mass spectrometry analysis identified this protein as a 90 kDa heat shock protein (HSP90). Inhibition of HSP90 with geldanamycin resulted in a block of occluded virus production in baculovirus-infected $121 and Ld652Y cells at 48 h.p.i. Cleavage of occurs in productively-infected cells in the presence of geldanamycin indicating that cleavage is not dependent on HSP90 activity. The data suggest a possible role for HSP90 during the very late stages of baculovirus infection. Cepyright by Christy Mecey 2006 ACKNOWLEDGEMENTS I would like to thank my principal investigator, Dr. Suzanne Thiem. Her efforts have helped me to develop the critical thinking skills that are required to be an effective scientist. Her unwavering requirement for absolute fact in the interpretation of scientific data is demanding and valuable. Dr. Thiem has also taught me how to run a functional laboratory with limited resources and her determination in the face of set-back is inspirational. Many thanks to my remarkable committee: Dr. David Amosti, Dr. Rebecca Grumet, Dr. Steve Triezenberg and Dr. Suzanne Thiem. You were all chosen based not only on your tremendous contributions to science but on, your equal commitment to science education. Input from the committee was important, relevant and timely. I would like to especially thank Steve and Rebecca for their time and effort in coordinating the GAANN fellowship program. Your extended efforts enabled the funding of my graduate education and increased science literacy amongst educators in the state of Michigan. I would like to genuinely thank my husband Richard who is always kind even in the wrath of my wildest mood swings. Without his help I would never have been able to navigate through the daily challenges of being a mother and a graduate student. I would like to thank Montgomery for all of the weekends she had to spend playing in the lab instead of playing in the park. I am so fortunate that you are my daughter. I would like to thank my parents, who thought that getting a Ph.D. was the silliest idea I’ve ever had, for their continued support and practicality. Many thanks to the members of the Allison lab, past and present, Jamie, Cheryl, Meredith, David, Katie and Julia for their friendship and technical support. Thank you Cheryl Calhoun for teaching me that balance really is the most important aspect of life. I would also like to thank the members of the Thiem lab, past and present, Ale, J ian, Aubrey and Heather for their friendship, assistance and input. I have genuine, heartfelt appreciation for Jeannine Lee, the most talented program assistant at MSU. There was never a question she could not find an answer to or a problem that she could not solve. I would like to thank all of the friends that I have made throughout my graduate experience for their supportive discussions, their insight, their backgrounds and their vision for the future. A special thank you for Linda Danhof, master coordinator and technician for the PRL, there are so many technical procedures that Linda either taught me or found someone to teach me that it would be impossible to list them all as a part of this document. I would like to dedicate this effort to the family members that we have lost in the past year that are missed and remembered: Lee McIntosh, Florence Allison and my 34 year old sister, Jamie Dixon. There is not a day that passes that I don’t wish I could trade places with my sister and return her to her lost family. -vi- TABLE OF CONTENTS List of Figures ............................................................................................. x List of Tables ........................................................................................... xii List of Abbreviations .............................................................................. xiii Rationale for Study .................................................................................. 1 Chapter 1: Literature Review .................................................................................... 4 1.1. Baculovirus pathogenesis ........................................................................ 5 1.2. Baculovirus replication ........................................................................... 7 1.3. Baculovirus host range ......................................................................... 11 1.3.1. Invertebrate innate immunity to virus infection ................................. 12 1.3.2. Developmental resistance to baculovirus .......................................... 14 1.3.3. Species-mediated resistance to baculovirus ...................................... 14 1 .3 3. 1. Host cell factor-l (HCF-l) .............................................. 14 l .3 3. 2. p143... 15 1. 3. 3. 3. Apoptotic shppreSsors and host range .................................. 16 1..3 3. 4. Host range factor-1 (HRF- -1) ............................................ 18 1.3.3.5. Other baculovirus proteins associated with host range ............................................................. 20 1.4. Translation modifications in response to stress21 1.4.1. Translation initiation ............................................................... 22 1.4.2. Initiation stage translational modifications in response to stress ................................................................... 23 1.4.2.1. Phosphorylation of eIF20t ............................................... 24 1.4.2.2. 4E-BP and eIF4E ......................................................... 25 1.4.2.3. Proteolytic cleavage of eIF4GI, eIF4GII and other initiation factors. 26 1.4.3. Virus strategies for overcoming host-mediated translation arrest .................................................................... 27 1.4.4. Preferential translation of viral transcripts ...................................... 29 1.5. Stress response pathways and virus infection ............................................ 30 1.5. l. The unfolded protein response .................................................... 30 1.5.1.1. The unfolded protein response and virus infection ...................... 34 -vii- 1.5.1 . 1.1. Host-encoded molecular chaperone requirements for virus infection ............................................................... 35 1.6. Proteasome activation and virus infection ................................................... 36 1.6.1. The COP9 signalosome , protein degradation and kinase signaling ........ 38 Chapter 2: In baculovirus infected Lepidoptera cells, proteasome inhibitors reduce virus production and eIFZa phosphorylation at serine 51 ........................................ 40 2.1. Introduction ..................................................................................... 41 2.2. Materials and methods ......................................................................... 44 2.2.1. Viruses and cell lines ............................................................... 44 2.2.2. Cytoplasmic fraction extraction and polysome profiling ..................... 44 2.2.3. Cell stress induction assays ......................................................... 45 2.2.4. Protease and proteasome inhibitor assays ....................................... 46 2.2.5. Western blot analysis ............................................................... 46 2.2.6. Plaque assays ............... . ........................................................ 48 2.3. Results ....................................................................................... 48 2.3.1. Translation arrest in AcMNPV-infected Ld652Y cells occurs at the initiation stage ............................................. 48 2.3.2. Translation arrest in AcMNPV-infected Ld652Y cells correlates with eIF2a phosphorylation at serine 51 .............................................. 50 2.3.3. Proteolytic cleavage of peIF 20Lscr51 occurs in AcMNPV and vAchrf-l infected Sf21 cells ................................................................. 55 2.3.4. PK2 is expressed in AcMNPV-infected Ld652Y cells ....................... 56 2.3.5. Cleavage of peIF2ocsets l in Sf21 and Ld652Y cell lines is not induced by stress-inducing agents ............................................................. 57 2.3.6. Cleavage of peIF20tserSI is blocked by aphidicolin but not by protease inhibitors ............................................................................ 59 2.3.7. Proteasome inhibition reduces phosphorylation of eIF20t at serine 51 and virus production in lepidopteran cells productively-infected with baculovirus ......................................................................... 62 2.4. Discussion ................................................................................. 68 Chapter 3: HSP90 expression is enhanced in vAchrf-l-infected Ld652Y cells ...................... 75 3.1. Introduction ..................................................................................... 76 3.2. Materials and methods ......................................................................... 81 3.2.1. Viruses and cell lines ............................................................... 81 3.2.2. Irnmunoprecipitation and mass spectrometry analysis ........................ 81 - viii - 3.2.3. MALDI-TOF-MS .................................................................. 82 3.2.4. LC-ESF-MS ........................................................................ 82 3.2.5. Western blot analysis ............................................................... 83 3.3. Results ........................................................................................... 84 3.3.1. HSP90 is activated in Ld652Y cells that are productively-infected with vAchrf-l ....................................................................... 84 3.3.2. Inhibition of HSP90 with geldanamycin inhibits polyhedrin protein synthesis and occlusion body production in permissively-infected cells at 48 h.p.i. but has little effect on budded virus production. . . . ........86 3.3.3. Proteolytic cleavage of peIF2aser51 precedes and is independent of increased HSP90 expression in permissively-infected Ld652Y cells ...... 92 3.4. Discussion .................................................................................. 94 Chapter 4: Summary and Future Directions ............................................................... 97 4.1. Research Summary ........................................................................... 98 4.2. Future Directions .............................................. , ............................... 100 Appendix A: Supplementary Data ............................................................ 107 References .......................................................................................... 1 13 -ix- LIST OF FIGURES Figure 1. Global translation arrests and virus replication is blocked in AcMNPV-infected Ld652Y cells .......................................................................................... 2 Figure 2. Baculovirus infection cycle .............................................................. 6 Figure 3. Transcriptional regulation of baculovirus genes ...................................... 8 Figure 4. Early and late events that occur during AcMNPV infection of Ld652Y cells ....................................................................................... 20 Figure 5. Initiation of protein synthesis ......................................................... 23 Figure 6. Phosphorylation of eIF20t and viral mechanisms to overcome translation arrest that occur as a result of eIFZu phosphorylation ............................. 25 Figure 7. The unfolded protein response (UPR). . . . . . . . . . . . . . ................................ 34 Figure 8. Translation arrest occurs at translation initiation in AcMNPV-infected Ld652Y cells ..................................................................... 50 Figure 9. Phosphorylation of full length eukaryotic initiation factor 2a (eIF20t) at serine 51 is reduced in vAchrf-l-infected Ld652Y cells as compared to AcMNPV-infected Ld652Y cells ............................................................................................ 53 Figure 10. Phosphorylated eIF20tserSI is proteolytically cleaved in cell lines that are productively infected with AcMNPV or vAchrf-l ............................................... 54 Figure 11. PK2 is present at late time points post-infection in Ld652Y cells infected with AcMNPV or vAchrf-l ............................................................................... 57 Figure 12. Proteolytic cleavage of peIF20ts°'5 ' does not occur in Sf21 or Ld652Y cells that are treated with stress-inducing agents ....................................................... 58 Figure 13. Phosphorylation and proteolytic cleavage of pelFZoU"chI is inhibited in Lepidoptera cells that are productively-infected with baculovirus and treated with serSl - aphidicolin. Proteolytic cleavage of peIF20t 1s not blocked by protease inhibitors .............................................................................................. 60 Figure 14. Proteasome inhibitors, MG132 and epoxomicin, inhibit phosphorylation and proteolytic cleavage of peIF20tscr51 in vAchrf-l-infected Ld652Y cells ...................... 64 Figure 15. The very late baculovirus occlusion body protein, polyhedrin, is not produced in productively-infected cells that are treated with epoxomicin ............................... 65 Figure 16. Proteasome inhibitors prevent budded virus production in vAchrf-l-infected Ld652Y cells .......................................................................................... 67 Figure 17. Epoxomicin prevents AcMNPV or vAchrf-l budded virus production in 891 cells, M0132 reduces budded virus production in those same infected cells ................... 68 Figure 18. HSP90 levels increase in vAchrf-l-infected Ld652Y cells from 12-24 h.p.i ................................................................................................... 85 Figure 19. Geldanamycin delays or prevents occlusion body formation in lepidopteran cells permissively-infected with baculovirus .................................................... 89 Figure 20. In cells treated with the HSP90 inhibitor, geldanamycin, polyhedrin protein levels are reduced in Sf21 and Ld652Y cells productively-infected with baculovirus ........................................................... . ................................. 90 Figure 21. Geldanamycin reduces budded virus production in vAchrf-l-infected Ld652Y cells ................................................................................................... 92 Figure 22. Proteolytic cleavage of peIF20tscrsl occurs in Ld652Y and Sf21 cells productively-infected with AcMNPV or vAchrf-l .............................................. 93 Figure 23. Proposed model of post-translational modification of eIF20t in response to baculovirus infection ............................................................................... 101 Figure 24. The interaction of eIF2a with eIFZB ............................................. 104 Figure 25. Proposed model for HRF-l function in vAchrf-l-infected Ld652Y cells. . . .106 Figure 26. AcMNPV infection in Ld652Y cells does not impair the m7G cap-binding ability of eIF4E .................................................................................... 128 Figure 27. 4E-BP is not up-regulated in response to AcMNPV infection in Ld652Y cells .................................................................................................. 130 Figure 28. The lepidopteran kinase, 86K, is expressed in AcMNPV-infected Ld652Y cells at a time that correlates to the onset of global translation arrest ...................... 131 Figure 29. Anti-PKR immunoprecipitation gels for mass spectrometry analysis... . . ....132 -xi- LIST OF TABLES Table l. Sequest summary of mass spectrometry analysis for immunoprecipitation with anti-PKR antibodies in vAchrf-l infected Ld652Y cells at 12 h.p.i .................................................................. 86 -xii- 4E-BP AcMNPV ATF ATP BeK BiP BmN BmNPV BV Cf-203 CflVINPV COP9 / CSN Dm DNA Dnapol dsRNA EDEM EDTA KEY TO SYMBOLS OR ABBREVIATIONS eIF4E binding protein amino acid Autographa califomica Multicapsid nuclear Polyhedrosis Virus alkaline nuclease activating transcription factor adenosine triphosphate Bombyx mori eIF2a kinase glucose regulated protein 78 Bombyx mori N cell line Bombyx mori Multi Nucleocapsid Polyhedrovirus budded virus ' Choristoneurafumiferana 203 cell line Choristoneurafumiferana Multi Nucleocapsid Polyhedrovirus COP9 signalosome Drosophila melanogaster deoxyribonucleic acid DNA polymerase double-stranded RNA ER degradation—enhancing a-mannosidase-like protein (ethylenedinitrilo) tetraacetic acid - xiii - eIF ER ERAD ERSE GCN GDP GTP HCF HCMV HCV HHV HIV HPI HPPV HRI HSC HSP HSV IAP eukaryotic initiation factor endoplasmic reticulum endoplasmic reticulum-associated degradation endOplasmic reticulum stress responsive element general control non-derepressible guanosine diphosphate guanosine triphosphate host cell factor human cytomegalovirus hepatitis C virus human herpes virus human immunodeficiency virus hours post-infection human papillomavirus hours post-treatment homologous repeat host range factor heme-regulated inhibitor of protein synthesis heat shock cognate protein heat shock protein herpes simplex virus inhibitor of apoptosis ~xiv- IE IRE IRES JEV JN K LEF Ld652Y LdMNPV Met MOI mRNA mTOR ODV OpMNPV OV P35 P39 PABP PERK/PEK PK2 immediate-early gene inositol-requiring enzyme internal ribosomal entry site Japanese encephalitis virus c-Jun kinase late expression factor Lymantria dispar 652Y cell line Lymantria dispar Multi Nucleocapsid Polyhedrovirus methionine multiplicity of infection messenger RNA mammalian target of raparnycin nuclear factor-kappa B occlusion derived virus Orygia pseudotsugata Multi Nucleocapsid Polyhedrovirus occluded virus baculovirus 35 kD apoptotic suppressor protein baculovirus 39 kD capsid protein poly-A binding protein RNA-dependent-like endoplasmic reticulum kinase AcMNPV inhibitor of eIF2a kinases ribonucleic acid -xv- RNAi RNA Pol. II SeMNPV St‘9 Sf21 Tn368 TNF TRAF tRNA UPR UTR vAcAp35 VLF vRNA Pol XBP RNA interference RNA polymerase II Spodoptera exigua Multi Nucleocapsid Polyhedrovirus Spodopterafiugiperda 9 cell line Spodopterafrugiperda 21 cell line Trichoplusia m' 368 cell line tumor necrosis factor TNF-receptor kinase transfer RNA unfolded protein response untranslated region AcMNPV strain with the apoptotic suppressor, p35, deleted very late expression factor viral RNA polymerase X-box binding protein -xvi- Rationale for Study In 2001, annual international insecticide sales totaled over 9 billion dollars (Donaldson et al., 2004). Chemical insecticides are often effective but can act broadly and harm non-target species. Many chemical insecticides have deleterious effects on the environment and have been removed from the commercial market or relicensed to limit use (Donaldson et al., 2004). Biopesticides and biological control agents are living organisms or products from living organisms used to suppress pest populations. Biopesticides offer an alternative to chemical-based insecticides and generally affect a narrower range of target species (Thiem, 1997). Baculoviruses primarily infect Lepidopteran insects during the larval stage of development and are a biological control agent used commercially to control soybean pests. Narrow host range is the primary limiting factor to the further utilization and subsequent development of baculoviruses as biopesticides (Maeda, 1995; Moscardi, 1989; Moscardi, 1999; Thiem, 1997). Unlike bacterial and fungal host-pathogen systems, the molecular mechanisms that govern invertebrate host-specificity to virus infection are not understood. In cell culture, many baculoviruses are able to enter non-host invertebrate cells, virus replication is initiated but then terminated during the late stages of infection (Du and Thiem, 1997a; Lu and Miller, 1996). The Gypsy Moth, Lymantria dispar, is a non-permissive host to Autographa caliform'ca Multinucleocapsid Nucleopolyhedrovirus (AcMNPV), the type virus for the Baculoviridae family (McClintock, Dougherty, and Weiner, 1986). Virus replication begins in Lymantria dispar 652Y (Ld652Y) cells that are infected with AcMNPV but global translation arrest is initiated by 9 h.p.i. correlating to the late stages of infection and no virus progeny are produced (Fig. 1A) (Guzo et al., 1992). Expression of a Single gene, host range factor-I (HRF-l), isolated from Lymantria dispar Multinucleocapsid nucleopolyhedrovirus (LdMNPV), blocks translation arrest and enables the production of virus progeny in AcMNPV-infected Ld652Y cells and L. dispar larvae (Fig. 13) (Chen et al., 1998; Du and Thiem, 1997a; Du and Thiem, 1997b; Thiem et al., 1996). Homology and motif searches conducted with the hrf-I sequence provide no indications to its function. AcMNPV Infection in Ld652Y cells 8. HRF-1 Late Stages of Infection Global , ‘3 Translation . ' / Arrest Ld652Y Cells N_o Progeny Virus Produced Virus enters cell, viral transcripts are made, translation of early transcribed genes OCCUI’S . Ld652Y Cells | I B ' @I % 0 . I I AcMNPV ' ‘- - 0 + 7- .. _ g or w o n . 0 . H RF-‘I A Productive Virus (LdMNPV) Lymantria dispar Infection larvae Figure 1. Global translation arrests and virus replication is blocked in AcMNPV- infected Ld652Y cells (A). Expression of the LdMNPV gene, host range factor-1 (HRF-l), precludes translation arrest in AcMNPV-infected Ld652Y cells and Lymantria dispar larvae further enabling the production of virus progeny (B). (Image Presented in Color) It has not been determined whether global translation arrest is the direct mechanism leading to a block in the production of virus progeny in AcMNPV-infected Ld652Y cells or simply a correlative event resulting from infection. Global translation arrest may be a host-mediated innate immune response to control virus replication or a host-induced response to stress. In mammalian systems, viruses use a variety of mechanisms to block translation of host-encoded genes as a means of ensuring preferential translation of viral transcripts (Jang et al., 1988; Joachims, Van Breugel, and Lloyd, 1999; Kerekatte et al., 1999; Pelletier and Sonenberg, 1988). This is unlikely in baculovirus-infected cells as host macromolecular synthesis appears to be controlled at the transcriptional level (Nobiron, O'Reilly, and Olszewski, 2003; Ooi and Miller, 1988). In this study, my goal was to identify the mechanism(s) that leads to translation arrest in AcMNPV-infected Ld652Y cells. I hypothesized that identifying this mechanism might provide some indication to the function of HRF-l. Furthermore, we hoped to gain insight into invertebrate host responses that are initiated during non- permissive infections to further understand the components that determine baculovirus host range. Chapter 1 Literature Review Introduction 1.1 Baculovirus Pathogenesis Baculoviruses are large, double-stranded DNA viruses that primarily infect lepdiopteran insects during the larval grth stage. Autographa californica Multicapsid Nuclear Polyhedrosis virus (AcMNPV) is the type virus for the family Baculoviridae. Baculoviruses have a biphasic replication cycle which results in rod-shaped virions being packaged into a budded and occluded form (Theilmann et al., 2004). Budded virus, produced from a primary infection, is enveloped at the plasma membrane and infects neighboring cells by adsorptive endocytosis (Blissard and Wenz, 1992; Volkman and Goldsmith, 1988) resulting in a systemic infection. Synthesis of budded virus is reduced by 24 hours post-infection (h.p.i.) and there is a switch to occluded virus production (V olkman and Knudson, 1986). The occluded form of the virus (0V) is environmentally stable and consists of multiple, enveloped nucleocapsids that are packaged into a polyhedrin protein-rich matrix. Virions condensed in a polyhedrin matrix are wrapped in a virus-derived, carbohydrate-rich, calyx which forms a unilamellar structure with spike- like projections, forming the occlusion body. Caterpillars ingest the occluded form of the virus, the polyhedra dissolve rapidly within the alkaline pH of the midgut (Summers, 1971; Summers and Amott, 1969) and the virions pass through the peritrophic membrane of the insect. The virus enters cells via fusion of virion envelope with the microvillar membrane (Granados, Lawler, and Burand, 1981). The virus then travels along the microvillus aided by microtubules (Charlton and Volkman, 1993; Granados, Lawler, and Burand, 1981) and enters the nucleus where it uncoats at the nuclear pore or just inside the nuclear membrane (Bilimoria, 1991) (For review of the baculovirus infection cycle see Fig. 2) (Blissard and Rohrmann, 1990). Advanced infection of the larvae results in complete cellular destruction and a condition known as “melted” (Federici, 1997). Nuclear Membrane Nuclear Pore Virogenic Stroma O Occlusion Body Primary . . ’ Infectiom . Budded Vlrus i lngested by caterpillar 21'3“" vaI'I‘ Secondary lnfection- é ' gm ° “50W" End°°Yt°sis é Polyhedra solubilized = .‘ ______ , in the midgut-released when insect tissues deteriorate Figure 2. Schematic diagram of the baculovirus infection cycle adapted from Blissard and Rohrmann 1990 (Image Presented in Color) 1.2 Baculovirus Replication Within the nucleus of the midgut epithelial cells the infection process begins with immediate and early gene transcription of the virus followed by DNA replication, late gene expression, the production and release of budded virus, very late gene expression and the synthesis of occluded virus (See Fig. 3). Baculoviruses encode a DNA polymerase gene that is transcribed early in the infection process and is essential to replication of the virus (V anarsdall, Okano, and Rohrmann, 2005). Additional viral- encoded genes required for virus replication include: immediate-early gene-1 (ieI), late expression factor flef) I and 3, alkaline nuclease (an), p143 and p35 (Ito et al., 2004; LaCount and Friesen, 1997; McDougal and Guarino, 2000; Mikhailov and Rohrmann, 2002; Stewart et al., 2005; Todd, Passarelli, and Miller, 1995). lnfectlon Beglns I Early Late Very Late 3 fl : fl :fl Transcription § ie1, ieO, RNA Pol", § 5 i Mediated by: HR repeats, host : E Transcription factors, Virus-encoded 5 other host factors 5 alpha-atnantin resistant I RNA polymerase Virus Genes ie1, ieO, let-1, lef-2, , . Transcribed: § lef-3, lei-7, hcf-1, § Structural Proteins. § P°'Yh°d"" p143, p35, ie2, DNA — E pol. and pe-38 ‘ ‘ DNA Replication : (Genes required leI, Ief-1, Ief-3, p143,§ for replication) 5 p35, DNA 5 """"""""""""" ’ polymerase I Most RNA Pol. ll Transcription Downregulated Budded Virus Production Occluded Virus I I I— I Production ; : : 5 Figure 3. Schematic diagram of transcriptional regulation that occurs during the baculovirus replication cycle adapted from Rohrman 1992 Transcription of early expressed baculovirus genes is regulated by host RNA polymerase II and IE1. IE1 has a splicing variant known as IE0. Individually either IE1 or IEO can support virus replication but both are required to achieve a wildtype infection (Stewart et al., 2005). Homologous repeat (HR) sequences that serve as transcriptional enhancer regions can be found in eight locations of the AcMNPV genome (Rasmussen et al., 1996) and HR regions have been identified in all sequenced baculoviruses. IE1 regulates early virus transcription in cis by binding a 28-mer palindromic site in the HR regions and host transcription factors also bind within the HR region (Landais et al., 2005). LEF—1 is a DNA primase (Mikhailov and Rohrmann, 2002) and p143 a DNA helicase (McDougal and Guarino, 2000) that associates with LEF-3 (Evans et al., 1999), a single-stranded DNA binding protein, for nuclear transport and localization (Chen and Carstens, 2005). It has been proposed that the HR sites also serve as origins of DNA replication, indeed HR sites are essential for virus DNA replication (Leisy et al., 1995). Immunoprecipitation followed by crosslinking has shown that IE-l , P143 and LEF-3 bind viral chromatin at closely linked sites, suggesting that replication complexes form at the homologous repeat regions (Ito et al., 2004). Baculoviruses also encode an early expressed alkaline nuclease (AN) which is essential for virus replication. AN has 5’-to-3’ endonuclease activity on single-stranded DNA and 5’-to-3’ exonuclease activity when associated with LEF-3. The data suggests that AN plays a role in DNA recombination and the resolution of replication intermediates (Mikhailov, Okano, and Rohrmann, 2004). The apoptotic suppressor, p35 (Clem, Fechheimer, and Miller, 1991) directly binds caspases, blocking substrate proteolysis. P35 functions at early and late timepoints post- infection (LaCount and Friesen, 1997). Baculovirus IE2, LEF-7, HCF-1 and P38 stimulate viral replication but are not essential to a productive infection (Miller, 1997). The late stage events of baculovirus infection are defined as those that occur after DNA replication of the virus. The production of budded virus temporally correlates to a time immediately following DNA replication and the production of occluded virus is the result of the late stages of infection in cells infected with baculovirus (Friesen and Miller, 1986; Rice and Miller, 1986). Characterized late genes are transcribed at six to twenty- four hours post-infection (h.p.i.), the corresponding gene products are primarily structural in nature and include the VP39 capsid protein (Thiem and Miller, 1989) and P69 core protein (Wilson et al., 1987). Very late genes are transcribed in a burst of expression from eighteen to seventy-two h.p.i. and include the polyhedrin occlusion body protein (Rohrmann, 1986) and p10, which effects nuclear disintegration upon exit of the occluded virus (Rohrmann, 1992). Regulation of the late stages of infection is primarily transcriptional. The segue of early-to-late gene transcription is coupled to a decrease in host mRNA levels at six to eight h.p.i. (Nobiron, O'Reilly, and Olszewski, 2003; Ooi and Miller, 1988) and the induction of a virus-encoded (Guarino et al., 1998), alpha-amantin resistant RNA polymerase (vRNApol) (Beniya et al., 1996; Fuchs, 1983; Grula, 1981a; Grula, 1981b; Huh and Weaver, 1990; Yang, Stetler, and Weaver, 1991). By 18 h.p.i. transcription is exclusive to vRNApol (Fuchs, 1983; Huh and Weaver, 1990) (For a review of baculovirus temporal gene regulation, see Fig. 2). The viral RNA polymerase specifically binds the virus template at a consensus sequence of TAAG found in the promoters of late and very late genes (Morris and Miller, 1994; Ooi, Rankin, and Miller, 1989; Rankin, Ooi, and Miller, 1988; Thiem and Miller, 1989; Thiem and Miller, 1990). However, host and virus transcription factors, including IE1, IEO and host GATA factors, bind the virus DNA template in concert with the vRNApol (Chen, Tsai, and Chen, 2001; Hefferon, 2004; Krappa et al., 1992; Landais et al., 2005) and the HR enhancer elements play a critical role in gene regulation even at late times during the infection process (Hefferon, 2004). Twenty viral early or constitutively expressed genes are required for late and very late baculovirus gene expression, these include: ie-I, ie-2, dnapol, p143, p47, p35, 39K and the late expression factors (LEFS) 1-12 (Hefferon, 2004; Li, Harwood, and 10 Rohrmann, 1999; Lu and Miller, 1995b). The baculovirus-encoded LEF-8 and LEF-9 are homologous to subunits of RNA polymerase II. LEF-4 has guanyltransferase activity and can be purified as a component of the vRNApol complex (Guarino, Jin, and Dong, 1998). The very late expression factor 1 (V LF -1) is required for transcription from the polyhedrin promoter (McLachlin and Miller, 1994; Mistretta and Guarino, 2005). 1.3 Baculovirus Host Range Baculoviruses have been commercially developed as eukaryotic protein expression vectors and as biopesticides. Hundreds of different baculoviruses have been isolated from hundreds of species of infected lepidopteran larvae and it is predicted that thousands more exist (Federici, 1997; Theilrnann et al., 2004). A small percentage of characterized baculoviruses have the potential to be useful in controlling insect pests (Moscardi, 1999; Yearian, 1982). The use of baculoviruses as insect control agents has been successful for controlling a lepidopteran pest of soybeans (Moscardi, 1989). The limited commercial success is due, in part, to the narrow host range of baculoviruses. AcMNPV has the broadest host range and can permissively infect six Lepidopteran species and semi-permissively infect another twenty-six (Federici, 1997). Permissive hosts are denoted as those being highly susceptible to a particular virus infection while semi-permissive are those that are less susceptible (Bishop, 1995). Environmentally, the limited host range of baculoviruses is beneficial because effects on non-target species are minimal to non-existent. Economically, the limited host range of baculoviruses is detrimental because agricultural producers require a single broad-range insecticide to maintain cost effectiveness. Substantial research has been devoted to expanding host range and/or enhancing insecticidal properties through the development of recombinant 11 baculoviruses that express host range factors, insect regulatory proteins and/or general insecticidal toxins (Black, 1997; Maeda, 1995). In mammalian cells, virus host range is frequently determined by the interaction of a viral protein with a host-encoded receptor. This does not appear to be true for baculoviruses. In a number of non-permissive and semi-pennissive baculovirus infections the virus enters the cell nucleus, viral DNA replication begins but infection is blocked during the late stages of infection (Miller, 1997; Simon et al., 2004; Thiem, 1997; Yanase, Yasunaga, and Kawarabata, 1998). As an example, AcMNPV infects, has some late gene expression but does not productively replicate in cell lines from Bombyx mori, Choristoneura firmiferana, Mamestra brassicae and Lymantria dispar. AcMNPV infects Drosophila melanogaster cell lines but late gene expression does not occur (Morris and Miller, 1992; Morris and Miller, 1993). AcMNPV can enter and evoke an immune response in human cells (Abe et al., 2005) but viral DNA is not transcribed because viral promoters are not recognized by host machinery (Thiem, 1997). Host immune response also limits virus infection in mammals. In contrast, little is known about immune response to virus infection in invertebrates. 1.3.1. Invertebrate Innate Immunity to Virus Infection Due to a lack of a molecularly well-characterized virus pathogen and invertebrate host model, the pathways leading to induction of invertebrate innate immune response in response to virus infection have not been clearly elucidated. Recent evidence shows that Toll pathways may contribute to antiviral response in Drosophila melanogaster. Adult Drosophila dif mutants lacking a Toll pathway-activated NF-KB transcription factor were found to have a decreased tolerance to Drosophila X virus (Zambon et al., 2005). 12 Microarray data from the midgut of Aedes aegypti vectoring the Sindbis virus indicate a role for Toll and c-Jun amino terminal cascades in the infection process (Sanders et al., 2005). The RNAi pathway in Anopheles gambiae can be silenced with dsRNA of the ArgonauteZ gene. These Ag02 knock-down mosquitos are also more vulnerable to infection with O’nyong-nyong virus, an RNA alphavirus, as compared to untreated animals, which leads to the conclusion that RNAi is an integral antiviral response in A. gambiae (Keene et al., 2004). Heliothis virescens larvae are a fully permissive host to AcMNPV while Helicoverpa zea larvae, a closely-related species, are semi-permissive, showing a 1000- fold difference in susceptibility (Allen, 1969; Vail, 1978; Vail, 1982). However, larvae of both species that were orally fed AcMNPV ODV show no difference in their susceptibility during primary infection of the midgut cells or secondary infection of the tracheal cells. However, by 48 h.p.i in Helicoverpa zea, hemocytes encapsulate melanized infection foci in the trachea and clear the virus. Helicoverpa zea hemocytes are resistant to AcMNPV infection while hemocytes of Heliothis virescens are not. Encapsulation of infection foci does not occur in AcMNPV-infected Heliothis virescens. Originally, virus encapsulation and clearing were believed to be an innate immune response to baculovirus infection in Lepidoptera, however, further study indicates that swelling and tracheal damage occur in response to AcMNPV infection in Helicoverpa zea, making melanization and encapsulation a probable response to wounding that occurs from virus-damaged tissues (Trudeau, Washburn, and Volkman, 2001; Washburn, 1996) 13 1.3.2. Developmental Resistance to Baculovirus There appears to be some level of developmental resistance to baculovirus infection. As larvae age they become more resistant to oral-ingested baculovirus infection. This phenomena appears to be the result of midgut cell sloughing during the molting process. Larvae in the temporal late stages of instar development, immediately preceeding or during a molt, are less likely to establish a productive AcMNPV infection than those at the early instar stage. This is in contrast to H. virescens or T richoplusia m' larvae that are hemocoel-injected with one plaque forming unit of AcMNPV BV, which results in 59-89% mortality. It would appear that infected midgut cells are shed during the molting process before a systemic baculovirus infection can be established, thus “clearing” the virus from the insect. This decreased susceptibility is mimicked in H. virescens larvae that are fed on cotton where a plant-encoded foliar peroxidase initiates midgut sloughing in the insect (Trudeau, Washburn, and Volkman, 2001; Washburn, Kirkpatrick, and Volkman, 1995; Washburn, 1996). 1.3.3. Species-Mediated Resistance to Baculovirus Based on the current data, many different mechanisms appear to mediate or contribute to the species-specific host range of baculoviruses. In cell culture, host range mechanisms that have been investigated in detail commonly show impaired baculovirus replication and a block during the late stages of infection (Miller, 1997; Thiem, 1997). 1.3.3.1 Host Cell Factor-1 (HCF-l) In Sf21 and TN368 cells, both of which are permissive to AcMNPV infection, transient expression assays were conducted that combined a late viral promoter, a reporter gene and expression of eighteen AcMNPV LEFs. In Sf21 cells, expression of the 14 eighteen characterized LEFs was sufficient to initiate transcription of the reporter gene. However, in Tn368 cells transcription from a late and very late viral promoter required the expression of an additional baculovirus gene, host cell factor-1 (hcf-I) (Lu and Miller, 1995a), which is not homologous with mammalian hcf-I. Infection with AcMNPV hcf-I deletion mutants had no effect on virulence, virus DNA replication or temporal gene expression in SfZl cells. However, in T richoplusia m' 368 (Tn368) ovarian cells infected with hcf-I null mutants, DNA replication and late gene expression of the virus did not occur and an arrest of host and viral protein synthesis occurred by 18 h.p.i. (Lu and Miller, 1996). Infection with hcf-I null mutants in Hi-5 cells, a cell line derived from Trichoplusia ni eggs, and in T. ni larvae resulted in a reduced virulence of the virus. A 50-fold decrease in infectivity was observed for the budded form of the mutant virus. However, there was no reduction in oral infectivity for larvae that were fed polyhedra lacking the hcf-l gene. Thus, HCF-1 has species and tissue-specific effects (Lu and Miller, 1996). HCF-1 has been characterized as an early protein that localizes to the nucleus in infected Tn368 cells and has been shown to interact with itself in coprecipitation and yeast two-hybrid studies (Hefferon, 2004). 1.3.3.2. p143 Bombyx mori N (BmN) cells are permissive to Bombyx mori Multi Nucleocapsid Polyhedrovirus (BmNPV) but Tn368, and the Spodoptera frugiperda Sf21 and CLS-79 cell lines are non-permissive to BmNPV infection. Conversely, BmN cells are non- permissive to AcMNPV infection (Summers, 1978). Co-infection of BmNPV and AcMNPV yielded a variant that was capable of replicating in all three cell types that were non-permissive to BmNPV infection (Kondo and Maeda, 1991). Using construction of 15 recombinant BmNPV variants carrying AcMNPV genome fragments, the gene responsible for the expanded host range was identified as p143-a putative DNA helicase (Maeda, Kamita, and Kondo, 1993). The region within p143 associated with expanding the host range of AcMNPV was narrowed'to 3 amino acids that allowed replication of AcMNPV in B. mori larvae (Croizier et al., 1994). Eventually a single amino acid (AA) was identified as being sufficient to expand the host range of AcMNPV (Karnita and Maeda, 1997). At a high multiplicity of infection (MOI), AcMNPV variants carrying this single AA substitution were capable of infecting Sf9 and BmN cell lines (Karnita and Maeda, 1996). Since the original host range studies involving p143 were conducted, p143 has been confirmed as an in-vitro DNA helicase (McDougal and Guarino, 2000) with ATP-binding (McDougal and Guarino, 2001) activity that can be found associated with DNA in crosslinking studies (Ito et al., 2004). Suggestions as to how a DNA helicase might alter host range have not been offered. However, it is worth noting that the DNA polymerase from Lymantria dispar MNPV (LdMNPV) can functionally replace that of AcMNPV in transient DNA replication assays in Sf21 cells; however, p143 from LdMNPV can not replace AcMNPV p143 in the same assay (Ahrens and Rohrmann, 1996). The host specificity of the helicase indicates a potential role for its interaction with host-encoded proteins within the DNA replication complex of the virus. 1.3.3.3. Apoptotic Suppressors and Host Range There are multiple hypotheses to explain why apoptosis occurs in virus-infected cells. The predominant hypothesis for invertebrates is that apoptosis results in a premature cell lysis further preventing virus replication and subsequent infection of neighboring host cells. Paradoxically, apoptosis may also function as a means for virus 16 dissemination and has thus, co-evolved with permissive virus infections (Friesen and Miller, 1996). It is important to consider that invertebrates are not equipped with an adaptive immune system and thus, must rely on innate immune responses to guard against pathogen infection. While mammals also rely on innate immunity for protection from pathogens, they are also equipped with immunological memory resulting in antibody-based recognition of epitopes from invading pathogens. The apoptotic suppressor, p3 5, is non-essential for AcMNPV replication in Tn368 cells (Clem, Fechheimer, and Miller, 1991; Clem and Miller, 1993; Clem and Miller, 1994). Virus replication occurs in Trichoplusia ni larvae infected with AcMNPV p35 deleted (vAcAp35) virus but there is a reduction in the production of occluded virus, with less melting at the end of infection as compared to a wildtype infection (Clem and Miller, 1993; Clem and Miller, 1994). Spodoptera fiugiperda 21 (Sf21) cells that are infected with vAcAp35 show severely impaired virus replication and a non-efficient shut-off of host protein synthesis (Clem and Miller, 1993; Hershberger, Dickson, and Friesen, 1992). In Sf21 cells infected with vAcAp35, p35 can be replaced with an inhibitor of apoptosis (lap) gene from other baculoviruses and wildtype infection is restored (Bimbaum, Clem, and Miller, 1994). Spodoptera fiugiperda larvae that are infected with vAcAp35 show a 25-fold decrease in rate of infection as compared to wildtype AcMNPV (Clem and Miller, 1993; Clem and Miller, 1994). Spodoptera Iittoralis-Z (SI-2) (Chejanovsky and Gershburg, 1995) and Charistoneura fitmrferana-203 (Cf-203) cells (Palli et al., 1996) infected with AcMNPV undergo apoptosis, resulting in a decrease in budded virus production and no occluded virus production. Both cell types are permissive to other baculovirus infections. P35 may be insufficient to block apoptotic response in these cell 17 lines, as transcriptional analysis revealed that p35 expression is delayed in AcMNPV- infected Cf-203 cells (Palli et al., 1996). Thus, baculovirus apoptotic suppressors are also species and tissue-specific regulators of virus host range. 1.3.3.4. Host Range Factor-l (Inf-I) AcMNPV enters but does not productively infect the Lymantria dispar 652Y (Ld652Y) cell line (McClintock, Dougherty, and Weiner, 1986). Global translation arrest begins in AcMNPV-infected Ld652Y cells by 8 h.p.i. correlating with the onset of DNA replication and the late stages of virus infection. All temporal classes of AcMNPV mRNA are synthesized and transported to the cytoplasm in infected Ld652Y cells and these mRNAs can be translated in vitro (Guzo et al., 1992). Expression of single gene, host range factor-1 (hrf-I ), isolated from Lymantria dispar Multinucleocapsid Nucleopolyhedrovirus (LdMNPV) prevents translation arrest in AcMNPV-infected Ld652Y cells and enables permissive replication of the virus in cell culture and infected larvae (Chen et al., 1998; Du and Thiem, 1997b; Thiem et al., 1996). HRF-l also expands the host range of the nonperrnissive Hyphantria cunea Multinucleocapsid Nucleopolyhedrovirus (HcMNPV) and the semi-permissive BmNPV and Spodoptera exigua Multinucleocapsid Nucleopolyhedrovirus (SeMNPV) (Ishikawa et al., 2004). These data suggest that HRF-l is an essential host range factor for baculovirus infection of Ld652Y cells. Motif and homology searches provide no indications as to the functional role of HRF-l. A truncated conceptual protein of 78 amino acids with homology to HRF-l exists in the genome of Orygia pseudotsugata Multinucleocapsid Nucleopolyhedrovirus (OpMNPV) but does not complement for LdMNPV HRF-l function in AcMNPV-infected Ld652Y cells (Ikeda, Reimbold, and Thiem, 2005). HRF- 18 1 contains a domain that is predicted to be highly acidic, mutations in this domain result in varied changes in HRF-l enabled AcMNPV replication in Ld652Y cells suggesting a role for the acidic domain in HRF-l function (Ikeda, Reimbold, and Thiem, 2005). In the absence of p35 and HRF-l , translation arrest and apoptosis are still initiated during the early stages of infection in AcMNPV-infected Ld652Y cells, apoptosis is enhanced and global translation arrest occurs at a time that correlates to the onset of DNA replication and the late stages of infection (Thiem and Chejanovsky, 2004). Blocking DNA replication and the late phases of virus replication with aphidicolin precludes global translation arrest in AcMNPV-infected Ld652Y cells, thus associating the arrest of protein synthesis with the late stages of virus infection (Thiem and Chejanovsky, 2004). Events correlating to the late stages of infection include the onset of DNA replication of the virus, the expression of late viral proteins and the Shut-off of host RNA Polymerase II-mediated transcription (Fig. 4) (Friesen and Miller, 1986; Thiem and Miller, 1989). Global translation arrest and apoptosis appear to be linked in the response because infecting Ld652Y cells with an AcMNPV mutant carrying a non-fimctional version of the apoptotic suppressor p35 does not result in global translation arrest (Thiem and Chejanovsky, 2004). The global translation arrest infection phenotype is restored when the same AcMNPV mutant expresses Opiap, Cpiap or p49, apoptotic suppressors isolated from other baculoviruses that have different modes of action as compared to p35 (Thiem and Chejanovsky, 2004). Treatment of vAcAp3S-infected Ld652Y cells with peptide caspase inhibitors does not result in global translation arrest and supports high levels of budded virus production, suggesting that apoptosis must be initiated for global translation arrest to occur (Thiem and Chejanovsky, 2004). 19 Earl Events Durin Infection Late Events Durin Infection Blocked by p35 Blocked by p35 Apoptosis Enhanced Stages Initiated? Of A tosis / \ DNA / 9°” suffix: Replication Of the Virus In Ld652Y _ Cells \ Translation / \ Global Translation Arrest . Arrest Initiated Sensor leading to late stage events: DNA Recombination? Host Transcription Shut-off? Late Gene Expression of Virus? Blocked by PK2? Blocked by HRF-1 Figure 4. Events that occur during AcMNPV-infection of Ld652Y cells (Image Presented in Color) 1.3.3.5. Other Baculovirus Proteins Associated with Virus Host Range The late expression factor, LEF-7, has homology to the herpes-virus UL29 family of single-stranded binding proteins (Lu and Miller, 1995b). LEF-7 has been identified as being an enhancer of viral DNA replication in Sf-21 cells (Chen and Thiem, 1997; Lu and Miller, 1995a) and a contributor to homologous recombination in plasmid transfected insect cells when in the presence of a hill complement of viral proteins (Crouch and Passarelli, 2002). Viral replication was reduced to ten percent of wildtype AcMNPV levels in Sf21 and SElc cells infected with a lef-7 deletion mutant. The same mutation had no effect on virus replication in Tn368 cells indicating a species-specific role for host range effects of LEF-7 (Chen and Thiem, 1997). 20 When larvae of Spodoptera frugr'perda and Trichoplusia m' are fed occluded AcMNPV which lack the immediate early gene, ie2, virus infectivity is reduced 1000 to 100 fold, respectively. Changes in infectivity are attributed to a lack of virions packed in the ODV of the mutant virus. However, based on this change oral infectivity, IE2 has been denoted as a possible determinant of virus host range. Larvae infected via intrahemocoelic injection with the same mutant had infection rates similar to wildtype AcMNPV infections (Prikhod'ko et al., 1999). Like LEF-7, the immediate early gene, ie2, of BmNPV and AcMNPV also enhances virus DNA replication of BmNPV in BmN cells (Gomi et al., 1997) and AcMNPV in Sf21 cells (Lu and Miller, 1995a), respectively. IE2 has been shown to transactivate IE1 in AcMNPV-infected Sf21 cells (Yoo and Guarino, 1994a; Y00 and Guarino, 1994b), to block cell cycle progression in cell lines from Spodoptera fiugiperda and Trichoplusia ni (Prikhod'ko and Miller, 1998) and IE2 from BmNPV has E3 ubiquitin ligase activity (Irnai et al., 2003). The diverse firnctional activity of IE2 in different virus-host relationships make it diffith to assess its possible role in determining virus host range. 1.4. Translational Modifications in Response to Stress Global translation is reduced in most, if not all, types of cellular stress (Holcik and Sonenberg, 2005). Control of translation provides the cell with the ability to immediately respond to a broad range of stress-induced signals. There is increasing evidence, based on disproportionate amounts of mRN A as compared to coordinate protein levels, which indicate an essential role for translational regulation and protein stability in mediating the cellular response to stress (Gygi et al., 1999; Holcik and Sonenberg, 2005; Rajasekhar and Holland, 2004; Rajasekhar et al., 2003). Translation 21 can be separated into three distinct phases: initiation, elongation and termination (Mathews, 2000). The majority of examples that currently exist show translational control occurring at the initiation stage as a cessation of translation (Holcik and Sonenberg, 2005). 1.4.1. Translation Initiation The initiation of translation is a complex process coordinated by numerous protein subunits that join at the 5’ and 3’ ends of an mRNA transcript. In the cytoplasm, ternary complexes are formed which consist of a GTP molecule, a methionine initiator tRN A and the eukaryotic initiation factor 2 complex. The ternary structure joins with the 43S pre-initiation complex which consists of the 40S ribosomal subunit and eIF3, eIF 1, eIFlA and eIF5. Binding of the pre-initiation complex to the5’ cap structure is believed to require bridging interactions between eIF 3 and the eIF 4F protein complex which is bound through eIF4E, the cap binding protein, at the m7GpppN cap structure. Other subunits within the eIF4F complex include eIF4A, the dead-box RNA helicase, and eIF4G which acts as a scaffold protein between eIF4E, eIF4A and eIF 3. The subunit eIF4G also interacts with the poly-A binding protein (PABP) and this interaction between proteins bound at the 5’ and 3’ end of the mRNA transcript, respectively, results in a circularization of the template (See Fig. 5). After recruitment to the mRNA, it is believed that the 43S initiation complex scans the template for an AUG translation initiation codon (Pestova and Kolupaeva, 2002). The methionine initiator tRNA binds to the AUG start site and the 43S ribosomal subunit is joined by the 60S ribosomal subunit to form the 808 ribosome. Ribosome formation is followed by the release of initiation factors for cellular recycling (See Fig. 4) (Gebauer and Hentze, 2004). 22 2 < > 2 < E .- —- ...—.---.. -___~37kDa - --~26kDa "" ~11kDa Figure 10. Phosphorylated eIF211’c'I5l is proteolytically cleaved in cell lines that are productively infected with AcMNPV or vAchrf-l. Proteolytic cleavage of eIF2u was assayed at 24 h.p.i. in mock, AcMNPV and vAchrf-l-infected Ld652Y and Sf21 cells with antibodies that specifically recognize the phosphorylated or total endogenous forms of the protein. Proteins from whole cell lysates were separated on 15% SDS-PAGE gels and analyzed with immunoblotting. The same blot for Ld652Y cells was probed with antibodies to peIF2oIschI (AB: peIFZawSI) (lanes 4-6) then stripped and reprobed with antibodies to eIF2a (AB: eIF20r) (lanes 1-3). Cell lines and antibodies used are shown above the panels and the molecular sizes of immune reactive bands are indicated to the right of the panels. In AcMNPV-infected Ld652Y cells, incubation of cell lysates with m7GTP-bound sepharose followed by western blot analysis showed that the cap-binding ability of eIF4E was not impaired (Fig. 26, Supplementary Data). In addition, western blots generated 54 with antibodies raised against Drosophila melanogaster 4E-BP (kindly provided by Mathieu Miron) did not show an increase in 4E-BP protein levels in AcMNPV-infected Ld652Y cells (Fig. 27, Supplementary Data). The results suggest that modification or sequestering of eIF4E is not the primary mechanism leading to global translation arrest in AcMNPV-infected Ld652Y cells. To show that eIF2or phosphorylation was the mechanism leading to global translation arrest in AcMNPV-infected Ld652Y cells, experiments were conducted to silence endogenous eIF201 expression in invertebrate cell lines using RNAi technology. Silenced Ld652Y cells were infected with a recombinant AcMNPV that constitutively expressed the Spodopterafiugiperda eIF2a gene with an alanine substitution at serine 51 (vAcHAeIFZaaIISI). Isotopic labeling of protein synthesis in Ld652Y cells silenced for eIF201 expression and infected with the vAcHAeIF 201aIaSI recombinant were inconclusive (data not shown). 2.3.3. Proteolytic cleavage of peIFZamSI occurs in AcMNPV and vAchrf-l-infected 8121 cells. serSl Cleavage of peIF20I in vAchrf-l-infected cells suggested a possible mechanism for HRF -1 to overcome translation arrest in AcMNPV-infected Ld652Y cells. serS 1 One hypothesis was that cleavage of peIF201 could prevent its binding of the guanosine nucleotide exchange factor eIFZB thus allowing protein synthesis to continue. To determine whether cleavage of peIF20tser51 was mediated by HRF-l we compared peIF2as°III cleavage in cell lysates of vAchrf-l- and AcMNPV-infected Sf21 cells, using western blots probed with antibodies to peIF201selrs I. In contrast to Ld652Y cells, Sf21 55 “'5 I occurred in both cells are permissive for AcMNPV replication. Cleavage of peIF2a AcMNPV and vAchrf-l-infected Sf21 cells by 24 h.p.i. (Fig. 10, lanes 8 and 9) and not in mock-infected cells (Fig. 10, lane 7) suggesting that cleavage is not directly-mediated by sets 1 HRF-l. In the cell lines studied, cleavage of peIF20t occurs only in those that are “'5‘ may be directly or productively-infected with baculovirus. Cleavage of pelF2a indirectly mediated by HRF-l in Ld652Y cells as a host protein may be functionally homologous to HRF -1 during AcMNPV infection in Sf21 cells. 2.3.4. PK2 is expressed in AcMNPV-infected Ld652Y cells. PK2 was expected to inhibit eIF2a kinases, yet similar levels of phosphorylation were observed in cells infected with AcMNPV and vAcApk2 (Fig. 9, lanes 6 and 10 as compared to lanes 8 and 12). To determine if PK2 was expressed in AcMNPV-infected Ld652Y cells, we performed western blot analysis using antibodies to PK2 (R. Clem from L.K. Miller collection). Although PK2 levels decreased over time, it was still present in Ld652Y cells infected with AcMNPV and Achrf-l at 6, 12 and 24 hours post- infection (Fig. 11, lanes 3, 4, 7, 8, 11 and 12). Thus, it is likely that concentrations of PK2 are too low to block eIF2u kinase activity at later time points. Alternatively, it is possible that the eIFZot kinase activated by AcMNPV infection in Ld652Y cells is not inhibited by PK2. There is a non-specific band of ~31 kD that is evident in all samples at 6 h.p.i. (Fig. 11, lanes 1, 2, 3 and 4) and has previously been observed with this antibody (Li and Miller 1995). 56 ohm lznm 24hm EEE EEE SEE gaésgsés‘éEE-s ...E__§ < i E31: g 6-55.59? “I: rang.-. ’I-ir-r' 344369” 123456789101112 Figure 11. PK2 is present at late time points post-infection in Ld652Y cells infected with AcMNPV and Achrf-l. Infected cells were harvested at 6, 12 and 24 hours post- infection, subjected to SDS-PAGE on 15% polyacrylamide gels and immunoblotted with antibodies to PK2. 2.3.5. Cleavage of peIFZaI'I"51 in 8121 and Ld652Y cell lines is not induced by stress- inducing agents. To test if peIF201scr51 cleavage resulted from a general stress response, we treated ser5 1 Sf21 and Ld652Y cells with stress-inducing agents and assayed for peIF20t cleavage by western blot analysis. To screen for the effects of genotoxic stress, uninfected cells were treated with methyl methanesulfonate. Nutritional stress was induced with rapamycin, oxidative stress with hydrogen peroxide, proteasome inhibition with epoxomicin and heat shock by incubation at 39° C. ER—stress and the unfolded protein response (UPR) were induced using thapsigargin and tunicamycin. Assays were “'5 I was not conducted at 6 and 24 hours post-treatment (h.p.t.). Cleavage of peIF2a observed in either Ld652Y or Sf21 cells, in response to any of these treatments (Fig. 12). The reduction in eIFZot phosphorylation, in both Ld652Y and Sf21 cells treated with 57 hydrogen peroxide at 24 h.p.t. (Fig.12A, lane 8 and 12B, lane 8), may be the result of cell damage initiated by reactive oxygen species. A. Ld652Y Cells Mock Epox. HS H202 MMS Rap. Thap. Tun. vAchrf-1 252:2:E=E=EEE3E33 omomomo§o§io§o§o§§ I W~W~peIFZG - 1234567891011121314151617 4'8"} . B. Sf21 Cells Mock Epox. HS H202 MMS Rap. Thap. Tun. AcMNPV age§s=ssess=EcE£= V V V 0N3N¢ONONON©§¢D§IO§§ - ‘I234567891011121314151617 Figure 12. Proteolytic cleavage of peIFZa’”51 does not occur in 8121 or Ld652Y cells that are treated with stress-inducing agents. Ld652Y and Sf21 cells were treated with stress inducing agents as indicated above each panel. Cells were harvested at 6 and 24 hours post-treatment. Proteins from whole cell lysates were fractionated on 15% SDSSl- SCI PAGE gels and analyzed on western blots probed with antibodies against peIF20t . Cell lines and treatments are shown above the panels. 58 2.3.6. Cleavage of peIFZa’M’l is blocked by aphidicolin but not by protease inhibitors. Late events in the baculovirus replication cycle enhance apoptosis in AcMNPV- infected Sf21 cells in the absence of the virally encoded apoptotic suppressor, p35 (LaCount and Friesen, 1997), and stimulate translation arrest in AcMNPV-infected Ld652Y cells (Thiem and Chejanovsky, 2004). To test if peIF20Ise'5l cleavage correlated with the onset of late viral gene expression, we treated vAchrf-l-infected Ld652Y cells serS 1 with the DNA polymerase inhibitor aphidicolin and scored for peIF20t cleavage at 24 h.p.i. on western blots. Cleavage was not observed and eIF20t phosphorylation was reduced in aphidicolin-treated, vAchrf-l-infected Ld652Y cells and AcMNPV-infected Sf21 cells (Fig. 13, lanes 1 and 6) indicating that phosphorylation and cleavage of peIF201I’c'5 I required a late event in the AcMNPV replication cycle. 59 Ld652Y/vAchrf-1 Sf21/AcMNPV n C =5 5:5 ,5; .38 938 E Esngm23n->- >. >> +ee§e .;gessw: >->- coco coco §§§§33 §§:§33%% 3335?? 3355??55 :‘zz‘gtgxzz‘g‘gzz 8§%%2§ 08%%22%% §E<<>> §§<<>><< ..., Polyhedrin : - - 1 2 3 4 5 6 7 8 91011121314 Figure 20. In cells treated with the HSP90 inhibitor, geldanamycin, polyhedrin protein levels are reduced in 8121 and Ld652Y cells productively infected with baculovirus. Polyhedrin protein levels were assayed by using anti-polyhedrin antibodies Total cell lysates were prepared from Ld652Y cells that were mock-infected or infected with AcMNPV or vAchrf-l or AcMNPV-infected Sf21 cells. Infected cells were harvested at 24 and 48 hours post-infection. Proteins from whole cell lysates were separated on 15% SDS-PAGE gels and analyzed with immunoblotting. 90 A. Geldanamycin in Infected Ld652Y cells 1.ooe+os , - . .. . . mAchrM : - ~‘ ~ flvAchrf-1IGeld. IAcMNPV 1.00E+07 = EACMNPVIGeld. rt. n' N: in“ f H.511“ .. 1 , 7 (1).“: 34253112; 2‘:th gaggeegaéset» <> <> <>§<§ --...— ...-sun- ...-.- ~37 kDa d all. -- --zskoa 123456789101112 Figure 22. Proteolytic cleavage of peIFZamfl occurs in Ld652Y and 8121 cells permissively-infected with vAchrf-l and AcMNPV in the presence of geldanamycin. Western blots showing proteolytic cleavage of peIF201ser51 in Ld652Y cells infected with AcMNPV and vAchrf-l that were untreated (lanes 2 and 3) or in the presence of geldanamycin (lanes 5 and 6) at 24 h.p.i. Proteolytic cleavage of peIF201$ch1 also occurs in Sf21 cells infected with AcMNPV or vAchrf-l that were untreated (lanes 8 and 9) or treated with geldanamycin (lanes 11 and 12). 93 3.4. Discussion Infection of the Lymantria dispar 652Y cell line with the baculovirus AcMNPV results in global translation arrest and a non-productive infection (Guzo et al., 1992; McClintock, Dougherty, and Weiner, 1986). Expression of a single gene, host range factor-I, isolated from LdMNPV precludes translation arrest in AcMNPV-infected Ld652Y cells and results in virus replication (Chen et al., 1998; Du and Thiem, 1997b; Thiem et al., 1996). Apoptosis is enhanced and global translation arrest is initiated during the late stages of AcMNPV infection in Ld652Y cells (Thiem and Chejanovsky, 2004). During a productive infection, events correlating to the late phases of virus replication are DNA replication of the virus, expression of late viral genes and a virus-mediated shut-off of host transcription. In this study we identify an increased expression of the chaperone protein HSP90 that occurs in cells permissively-infected with vAchrf-l during the late stages of infection (Fig. 18). Further, we determine that geldanamycin-mediated inhibition of HSP90 delays or inhibits the production of the very late baculovirus polyhedrin protein and reduces or eliminates the production of occluded virus in productively-infected cells at 48 h.p.i., a time when occluded virus is normally present (Fig. 19 and 21). Inhibition of HSP90 with geldanamycin has no inhibitory effect on the proteasome-dependent cleavage of peIF201s¢r5 1, indicating that HSP90 activation occurs independently or downstream of eIF2a phosphorylation at serine 51 (Fig. 22). 94 Other groups have shown that chaperone proteins may be required during permissive baculovirus infection. HSC70, also known as BiP, was shown to be transcriptionally up-regulated in Sf9 cells permissively-infected with AcMNPV (N obiron, O'Reilly, and Olszewski, 2003). HSP90 was not identified in the same study as being transcriptionally activated; however, the same group shows the majority of host transcription being shut-off between 12 and 18 h.p.i. during a permissive infection. The shut-off of host transcription suggests a potential role for translational regulation of chaperone transcripts during a permissive infection. Indeed, HSP90 levels increase from 12 to 24 h.p.i. in Ld652Y cells that are permissively-infected with vAchrf-l and HSP90 appears to play a role in the production of late viral proteins and occluded virus (Fig. 19, 20 and 21). In the non-productive infection of AcMNPV in Ld652Y cells, the virus enters the nucleus and executes the early stages of virus replication. During these early stages of infection, host and virus transcription and translation continue (Guzo et al., 1992). Previously, we have shown that eIF20 phosphorylation at serine 51 occurs during AcMNPV infection of Ld652Y cells at a time that correlates to the onset of global translation arrest. Additionally, we have shown that both eIF20t phosphorylation at serine 51 and cleavage of peIF2aser51 are enabled by the proteasome (Mecey-manuscript in progress). The additional load of viral transcripts produced during the early stages of infection may result in stress on the protein processing capacity of the endoplasmic reticulum and its localized chaperone machinery cumulating in the induction of an unfolded protein response. 95 The unfolded protein response (U PR) has been characterized in C. elegans and plays a regulatory role in the development of the nematode (Shen et al., 2001). Many homologous components of the mammalian UPR pathway have been identified and characterized in other invertebrates (Olsen et al., 1998; Santoyo et al., 1997) including the lepidopteran, Bombyx mori (Goo et al., 2004). Hallmarks of the unfolded protein response include eIF20t phosphorylation at serine 51 resulting in a global arrest of cap- dependent protein synthesis and a diverted pathway resulting in either apoptosis and cell destruction, or cell recovery as mediated by an increased synthesis of chaperone proteins and ERAD degradation of unfolded/unprocessed proteins (Ma and Hendershot, 2004; Rutkowski and Kaufinan, 2004). The proteasome has recently been identified as a key regulator in the UPR by facilitating ERAD degradation (N g, Spear, and Walter, 2000). Although our host range system has many of the hallmarks of a UPR stress response pathway, one must consider that this pathway has not been fully characterized for Lepidopteran insects and that the comparisons made are based on information obtained from mammalian systems. Different stress response pathways such as those resulting from DNA or oxidative damage may also be induced in response to infection. Investigation into kinase signaling leading to eIF20. phosphorylation in AcMNPV- infected Ld652Y cells may help to define initiated pathways during non-permissive infection. Additonal information should also be gathered with regards to the expression of other chaperone proteins during baculovirus infection. 96 Chapter 4 Summary and Future Directions 97 4.1. Research Summary a.) Research conducted by other laboratory members show that translation arrests gradually in AcMNPV-infected Ld652Y cells with a complete arrest of protein synthesis occm'ring by 13-14 h.p.i. (work conducted by X. Du). In those same infected cells, polysome profiles show a monosome peak occurring at 9 h.p.i. indicating that translation arrests at the initiation stage (work conducted by W.A. Williams). Possible effects occurring from the inhibition of translation elongation were not investigated in this study. b.) Phosphorylation of eIF20t at serine 51 occurs in Ld652Y cells that are infected with AcMNPV at 12 h.p.i. In Ld652Y cells that are infected with vAchrf-l, eIF20t appears to be protected from phosphorylation as compared to mock and AcMNPV-infected cells. Expression of the proposed AcMNPV-encoded, dominant-negative inhibitor of eIF20t kinases, PK2, continued throughout the infection but PK2 was insufficient at blocking phosphorylation of eIF20t in AcMNPV-infected Ld652Y cells at late times post-infection. $ch 1 c.) Proteolytic cleavage of peIF2a appears to occur in Ld652Y cells permissively- infected with vAchrf-l and Sf21 cells permissively-infected with AcMNPV and vAchrf- ser5 1 1. Cleavage of peIF20r. was not induced by stress-inducing agents and was not blocked with cell permeable protease inhibitors. (1.) Treatment with the proteasome inhibitor epoxomicin in Sf21 and Ld652Y cell lines infected with AcMNPV and vAchrf-l or vAchrf-l, respectively, resulted in prevention of ser5 l eIF201 phosphorylation at serine 51, a block of peIF20t cleavage and no virus replication. The proteasome inhibitor, MG132, yielded similar results in vAchrf-l- infected Ld652Y cells but only reduced virus titers while not blocking peIF201scr51 98 cleavage in infected Sf21 cells. The conflicting result observed for both proteasome inhibitors may result from a varying potency of each inhibitor or its specific mode of action. Both results indicate that the proteasome is involved during baculovirus infection. 0.) Heat shock protein 90 expression is enhanced in Ld652Y cells that are permissively- infected with vAchrf-l. Inhibition of HSP90 with geldanamycin resulted in a delay or reduction of very late stage baculovirus protein expression and a block of occlusion body ser5 1 production at 48 h.p.i. Cleavage of pelF201 occurs independently of enhanced HSP90 expression. These results suggest that enhanced expression of HSP90 accompanies productive baculovirus infection and is integral to the production of occluded virus. 99 4.2. Future Directions The current data indicates that proteasome-enabled cleavage of peIFZasm' accompanies productive virus infection in both Ld652Y and Sf21 cells infected with vAchrf-l and AcMNPV and vAchrf-l, respectively. Additional data produced in this study demonstrates that the inhibition of the proteasome with epoxomicin or MG132 blocks phosphorylation of eIF20t at serine 51 and inhibits or reduces the production of virus progeny in those same infected, lepidopteran cell lines (See Fig. 23). The primary questions that remain are: is baculovirus replication enabled by the proteolytic cleavage of peIF20ts¢r5 l and, if so, does HRF-l participate in that process? 100 b eIFZo Phosphorylation No Virus Blocked Replication Proteasome m Infection Activity _—>a <:l ————> E Virus Replication peIFZo‘m‘ Proteolytically ' "1 Blocked? Ieaved 1 Virus Replication Enabled Figure 23. Proposed model of post-translational modification of eIF20t in response to baculovirus infection. The current data shows eIF2a becoming phosphorylated in response to AcMNPV-infection in Ld652Y cells. In Sf21 and Ld652Y cells that are productively-infected with AcMNPV and vAchrf-l or vAchrf-l, respectively, a (Fig. 23 caption, cont.) proteasome-enabled cleavage of peIF201§cT51 occurs by 24 h.p.i. In those same infected cells, when treated with proteasome inhibitors, eIF20t is protected from phosphorylation at serine 51 and virus replication is inhibited. (Image Presented in Color) 101 ser5 1 Indeed, if cleavage of peIF20t enables baculovirus replication by preventing global translation arrest, then one can envision two scenarios as being mechanistically significant. The first and most probable mechanism involves disabling the inhibitory ser5 1 capabilities of peIF20t on eIF2B, the second involves the proteolytic degradation of peIF20ts°m coupled to an increased synthesis of eIF201. The nucleotide exchange factor eIFZB is a limited factor in the cell, phosphorylation of eIF2a at serine 51 results in the binding and sequestering of eIF2B. This impoundment of eIF2B prevents the exchange of GDP to GTP required for the formation of an active eIF2 ternary complex resulting in translation arrest at the initiation stage (See Fig. 24A and 24B). Proteolytic cleavage of peIF20tse'51 may render two partial eIF20t fragments that are incapable of inhibiting eIF2B but capable of forming a functional ternary complex with other subunits (See Fig. 24C). One method to test this hypothesis is to express the predicted partial eIF2a fi'agments individually from a recombinant AcMNPV in Ld652Y cells and then conduct plaque assays to determine budded virus titers. Alternatively, engineered plasmid constructs with truncated forms of eIF2oc could be transfected into Ld652Y cells followed by AcMNPV infection and plaque assay. While Figure 24C does not reflect it, functional ternary complex formation could result from interaction with either partial eIF2a component so, both would have to be expressed and evaluated. If budded virus titers increased as compared to those produced at 0 hours post-infection, one could conclude that AcMNPV replication had been enabled by the expression of a truncated eIF20t sequence. It is possible that phosphorylation of eIF201 at serine 51 occurs as a general response to stress in lepidopteran insects. While experiments utilizing general stress induction agents in 102 both Ld652Y and Sf21 cells did not result in cleavage of peIF2asc'5‘, levels of full-length peIF20tscr51 may show differential regulation in response to stress (Fig. 12). More experiments need to be conducted to determine the effects of stress on eIF2or phosphorylation in lepidopteran cells. Proteasome-enabled degradation of peIF201‘cr5 l may occur as a means to overcome translation arrest while levels of eIF20t recover. If budded virus titers do not increase in Ld652Y cells infected with recombinant AcMNPV expressing either eIF20t truncated protein (vAccleIF201), protein synthesis levels should be evaluated to determine if global translation arrest occurs. If translation arrest is prevented in vAccleIF20t-infected Ld652Y cells and a productive infection is not established, it would indicate that global translation arrest is not the mechanism preventing AcMNPV replication in that cell line. Other groups have shown that expression of a C-terminal, truncated form of eIF20t, was not sequestered by eIFZB in mammalian SAOS-2 cells. Cells expressing this mutant were able to overcome translational repression induced by PKR even though the truncated eIF201 was phosphorylated by the kinase (Satoh, et. a1. 1999). 103 0; Met Jo" 12111+ [u Figure 24. The interaction of eIF20t with eIFZB. A.) The nucleotide exchange factor eIF2B coordinates the exchange of GDP to GTP to restore functional ternary complex formation required for the initiation of protein synthesis. B.) When eIF20L is phosphorylated at serine 51 it binds eIFZB preventing nucleotide exchange, active ternary complexes can not form and global translation arrest is initiated. C.) A proposed model of how a cleaved fragment of pelF201§cr51 may interact with eIFZB and still fimction in an active ternary complex to initiate protein synthesis. (Image Presented in Color) The second proposed mechanism involves the proteasome—enabled cleavage of peIF2as°r51 as a means to remove the phosphorylated protein from the cell as the cell recovers from a particular stress. Eukaryotic cells are dependent on eIF20L as a required component of the initiation of protein synthesis so, this proteolytic degradation would have to be accompanied by an increased synthesis of eIF20t which is best evaluated at the transcriptional level using northern blots or real-time PCR. Current data do not suggest that eIF20t levels increase in productively-infected cells but that is difficult to determine by observing protein levels as host transcription is being shut-off. In either model as it relates to vAchrf-l-infected Ld652Y cells, a viral protein may mediate the proteasome- ser5 1 enabled cleavage of peIF20t to initiate recovery fiom stress resulting from virus 104 infection, thus restoring protein synthesis and possibly enabling permissive virus infection. As a part of an initiated stress response pathway, proteasome activity can lead to to eIF20t phosphorylation at serine 51 (Jiang and Wek, 2005). One can envision a scenario where once proteasome-dependent eIF2a phosphorylation is initiated in baculovirus-infected cells, a viral protein, like HRF -1, may mediate the degradation of peIF2orsc'5‘ by directing proteasome activity (Fig. 25). The question of HRF-l involvement in enabling baculovirus replication in Lymantria dispar and its correlation to “'51 are best answered by searching for proteins that the proteolytic cleavage of peIF201 directly interact with HRF-l . Yeast two-hybrid analysis or immunoprecipitation studies coupled to mass spectrometry analysis may identify HRF-l interacting proteins and further provide indications to the level of function for HRF-l. Another protein(s) that functions in a similar manner to HRF-l may be encoded by the Spodoptera fi-ugiperda genome thus facilitating AcMNPV replication in Sf21 cells. This would offer an “'5‘ occurs in Sf21 cells during infection With explanation of why cleavage of peIF2a AcMNPV. Alternatively, Ld652Y cells may physiologically differ from 8121 cells to such an extreme, that a HRF-l-type protein is not required to promote AcMNPV replication in Sf21 cells. 105 Proteasome } elF2r1 is ACtIVItY phosphorylated v Viral protein (HRF-1?) directs degradation of pelFZa“r51 via proteasome activity? Infection 4 Figure 25. Proposed model for HRF-l function in vAchrf-l-infected Ld652Y cells. AcMNPV-infection of Ld652Y cells results in eIFZa phosphorylation at serine 51 and global translation arrest. This study shows that proteasome inhibition blocks the phosphorylation of eIF 201, cleavage of pelF 211W51 and blocks the production of virus progeny in baculovirus-infected, invertebrate cell lines. In the proposed model, HRF-l would act downstream of proteasome activation to coordinate cleavage of peIFZaws 1. 106 Appendix A Supplementary Data 107 eIF4E is available to bind cap substrates in AcMNPV-infected Ld652Y cells. Protein synthesis can be attenuated through the formation of an inhibitory eIF4E- 4E binding protein (4E-BP) complex. To exclude the possibility that mechanisms other than eIF20t phosphorylation might be contributing to translation arrest we isolated total cell lysates from mock, AcMNPV and Achrf-l-infected Ld652Y cells and precipitated against m7G-bound sepharose (Amersham-Pharmacia). Western blots of precipitated proteins were probed with Lepidopteran eIF4E antibodies. Western blots show equal amounts of eIF4E bound to the m7G cap in mock, AcMNPV and Achrf-l infected cells at 3 and 20 hpi (Fig. 15, lanes 1, 2, 3, 4, 5 and 6). This indicates that eIF4E is not prevented from binding the cap and by inference, that it was not sequestered by 4E-BP. 3hpi 20hpi ~ AcMNPV Cap-bound eIF4E Figure 26. AcMNPV infection in Ld652Y cells does not impair the m7G cap-binding ability of eIF4E. Total cell lysates from mock-, AcMNPV-, and Achrf-l-infected Ld652Y cells were immunoprecipitated with m7GTP-bound sepharose. Proteins were boiled off sepharose beads, subjected to SDS-PAGE and western blotting with lepidopteran eIF 4E antibodies. 108 4E-BP is not over-expressed in AcMNPV-infected Ld652Y cells. To further investigate the possibility of 4E-BP contributing to global translation arrest in AcMNPV-infected Ld652Y cells, we probed western blots of total cell lysates isolated from Mock, AcMNPV and Achrf-l-infected cells at 3, 12 and 20 hpi with antibodies raised against Drosophila melanogaster 4E-BP (kindly provided by Mathieu Miron). There was no detectable increase in 4E-BP levels in infected cells over mock- infected (Fig. 16, lanes 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11 and 12) suggesting that, unlike translation arrest induced in Drosophila in response to bacterial infection and wounding (Bemal and Kimbrell, 2000), over-expression of 4E-BP does not contribute to translation arrest in AcMNPV-infected Ld652Y cells. These experiments do not rule out possible changes in 4E-BP phosphorylation. However, the combined eIF4E and 4E-BP data indicates that translation arrest is not controlled through inhibition of eIF4E. 109 3' 12. N A 3' '2. AcMNPV vAch ri-1 VACApk2 E 'r 2! s a a E 2 s s E ~‘a’\ 5.-.. ..-“ 4E-BP ... 1234 5678 9101112 Figure 27. 4E-BP is not up-regulated in response to AcMNPV-infection in Ld652Y cells. Total cell lysates were prepared from Ld652Y cells that were mock-infected or (Caption continued from p. ) infected with AcMNPV, vAchrf-l or vAcApk2. Lysates were collected at 12 h and 24 h post-infection. The same blot for each time point was stripped and reprobed sequentially with antibodies to Drosophila 4E-BP and a-tubulin. The lepidoteran eIFZa kinase, BeK, is activated in AcMNPV-infected Ld652Y cells. A novel lepidopteran eIF20 kinase, BeK, was identified in Bombyx mori and characterized as being activated in response to heat shock and osmotic stress in transformed Drosophila 82 cells but not in response to immune, endoplasmic reticulum or oxidative stress. The same group showed that PK2 was capable of reducing BeK activity suggesting that BeK may act in an anti-viral response (Prasad et al., 2003). Using western blots and antibodies raised against BeK (P. Brey), we screened infected Ld652Y cells for BeK induction. Western blots of infected cell lysates show expression of a 65 kDa protein that interacts with anti-BeK antibodies at 12 hours post-infection in AcMNPV-infected Ld652Y cells (Fig. 14), a time that correlates to eIF20t phosphorylation in AcMNPV-infected Ld652Y cells. Western blots produced from 110 mock-infected Ld652Y cells or those infected with vAchrf-l from 3-24 hours post- infection are devoid of any antibody interaction when probed with anti-BeK antibodies (Fig. 14). An additional band of ~40 kDa is also seen in AcMNPV-infected Ld652Y cells at 12 h.p.i. Results from this western blot indicates that BeK is expressed at a time that correlates to eIF20t phosphorylation and global translation arrest in AcMNPV- infected Ld652Y cells. Ld652Y AcMNPV vAchrf-1 +Mi3612244836122448 A "" 3:: BeK Sf21 AcMNPV vAchrf-1 +Mi36122448361224 B a —..-' ---—-- - ___,__.___.--- BeK .“ -- .- ...—.--” Figure 28. The lepidopteran kinase, BeK, is expressed in AcMNPV-infected Ld652Y cells at a time that correlates to the onset of global translation arrest. Total cell lysates were collected at 3, 6, 12, 24 and 48 hours post-infection, subjected to SDS- PAGE and immunoblotted with polyclonal antibodies to BeK. lll “Qt—J" 4—— PKR Fusion Protein -r J. ‘J .o i Exclsed Bands for MS Excised Band .... . I! I u ’— PKR Mb" :- ‘ Protein 8' Western Blot from IP Gel Above Figure 29. Polyacrylamide gel of proteins immunoprecipitated with human PKR antibodies in vAchrf-l-infected Ld652Y cells. A. Arrows indicate the bands that were excised for mass spectrometry analysis and co-migration of the PKR control protein. Infected cell lysates were immunoprecipitated against sepharose bound antibody overnight, then boiled and separated using SDS-PAGE, gel was stained with Coomassie blue stain and destained for 5 hours. B. One lane from the IP gel was excised and transferred to western blot along with the PKR control protein lane. Western blots were incubated with anti-PKR antibodies; HRP-linked, anti-rabbit secondary antibodies and bands were detected using ECL detection kit. 112 References 113 Abe, T., Hemmi, H., Miyamoto, H., Moriishi, K., Tamura, S., Takaku, H., Akira, S., and Matsuura, Y. (2005). Involvement of the Toll-like receptor 9 signaling pathway in the induction of innate immunity by baculovirus. J Viral 79(5), 2847-58. Ahrens, C. H., and Rohrmann, G. F. (1996). The DNA polymerase and helicase genes of a baculovirus of Orgyia pseudosugata. J Gen Virol 77 ( Pt 5), 825-37. Allen, G. E. a. C. M. I. (1969). The nucleopolyhedrosis virus of Heliothis quantitative in viva estimates of virulence. J. Invertebr. Pathol. 13, 378-381. Aragon, T., de la Luna, 8., Novoa, 1., Carrasco, L., Ortin, J., and Nieto, A. (2000). Eukaryotic translation initiation factor 4G] is a cellular target for N81 protein, a translational activator of influenza virus. Mol Cell Biol 20(17), 6259—68. Attardo, G. M., Hansen, I. A., and Raikhel, A. S. (2005). Nutritional regulation of vitellogenesis in mosquitoes: implications for anautogeny. Insect Biochem Mol Biol 35(7), 661-75. Attardo, G. M., Higgs, S., Klingler, K. A., Vanlandingham, D. L., and Raikhel, A. S. (2003). RNA interference-mediated knockdown of a GATA factor reveals a link to anautogeny in the mosquito Aedes aegypti. Proc Natl Acad Sci U S A 100(23), 13374-9. Awasthi, N., and Wagner, B. J. (2005). Upregulation of heat shock protein expression by proteasome inhibition: an antiapoptotic mechanism in the lens. Invest Ophthalmol Vis Sci 46(6), 2082-91. Azevedo, C., Sadanandom, A., Kitagawa, K., Freialdenhoven, A., Shirasu, K., and Schulze-Lefert, P. (2002). The RARl interactor SGTl, an essential component of R gene-triggered disease resistance. Science 295(5562), 2073-6. Baltzis, D., On, L. K., Papadopoulou, S., Blais, J. D., Bell, J. C., Sonenberg, N., and Koromilas, A. E. (2004). Resistance to vesicular stomatitis virus infection requires a functional cross talk between the eukaryotic translation initiation factor 2alpha kinases PERK and PKR. J Viral 78(23), 12747-61. Beattie, E., Denzler, K. L., Tartaglia, J., Perkus, M. E., Paoletti, E., and Jacobs, B. L. (1995). Reversal of the interferon-sensitive phenotype of a vaccinia virus lacking E3L by expression of the reovirus S4 gene. J Viral 69(1), 499-505. Bech-Otschir, D., Seeger, M., and Dubiel, W. (2002). The COP9 signalosome: at the interface between signal transduction and ubiquitin-dependent proteolysis. J Cell Sci 115(Pt 3), 467-73. 114 Beere, H. M. (2001). Stressed to death: regulation of apoptotic signaling pathways by the heat shock proteins. Sci ST KE 2001(93), REl. Belsham, G. J., McInemey, G. M., and Ross-Smith, N. (2000). Foot-and-mouth disease virus 3C protease induces cleavage of translation initiation factors eIF4A and eIF4G within infected cells. J Viral 74(1), 272-80. Beniya, H., Funk, C. J., Rohrmann, G. F., and Weaver, R. F. (1996). Purification of a virus-induced RNA polymerase from Autographa californica nuclear polyhedrosis virus-infected Spodoptera frugiperda cells that accurately initiates late and very late transcription in vitro. Virology 216(1), 12-9. Bernal, A., and Kimbrell, D. A. (2000). Drosophila Thor participates in host immune defense and connects a translational regulator with innate immunity. Proc Natl Acad Sci U S A 97(11), 6019-24. Bertolotti, A., Zhang, Y., Hendershot, L. M., Harding, H. P., and Ron, D. (2000). Dynamic interaction of BiP and ER stress transducers in the unfolded- protein response. Nat Cell Biol 2(6), 326-32. Birnbaum, M. J., Clem, R. J., and Miller, L. K. (1994). Anapoptosis-inhibiting gene from a nuclear polyhedrosis virus encoding a polypeptide with Cys/His sequence motifs. J Viral 68(4), 2521-8. Bishop, D. H. L., Hirst, M.L., Possee, R.D., and Cory, J.S. (1995). Genetic engineering of microbes: Virus insecticides-A case study. In "50 Years of Microbials" (P. A. H. G.K. Darby, and A.D. Russell, Ed.), pp. 249-277 Cambridge University Press, Cambridge, England. Black, B. C., Brennan, L.A., Dierks, RM. and LE. Gard (1997). Commercialization of Baculoviral Insecticides. In "The Baculoviruses" (L. K. Miller, Ed.), pp. 341-387, Plenum Press, New York. Blissard, G. W., and Rohrmann, G. F. (1990). Baculovirus diversity and molecular biology. Annu Rev Entomol 35, 127-55. Blissard, G. W., and Wenz, J. R. (1992). Baculovirus gp64 envelope glycoprotein is sufficient to mediate pH-dependent membrane fusion. J Virol 66(11), 6829- 35. Brecht, M., and Parsons, M. (1998). Changes in polysome profiles accompany trypanosome development. Mol Biochem Parasitol 97(1-2), 189-98. Brewer, J. W., and Diehl, J. A. (2000). PERK mediates cell-cycle exit during the mammalian unfolded protein response. Proc Natl Acad Sci U S A 97(23), 12625-30. 115 Busch, 8., Eckert, S. E., Krappmann, S., and Braus, G. H. (2003). The COP9 signalosome is an essential regulator of development in the filamentous fungus Aspergillus nidulans. Mol Microbiol 49(3), 717-30. Bushell, M., Poncet, D., Marissen, W. E., Flotow, H., Lloyd, R. E., Clemens, M. J., and Morley, S. J. (2000). Cleavage of polypeptide chain initiation factor eIF4G] during apoptosis in lymphoma cells: characterisation of an internal fragment generated by caspase-3-mediated cleavage. Cell Death Differ 7(7), 628-36. Buttle, D. J., Murata, M., Knight, C. G., and Barrett, A. J. (1992). CA074 methyl ester: a proinhibitor for intracellular cathepsin B. Arch Biochem Biophys 299(2), 377-80. Calfon, M., Zeng, H., Urano, F., Till, J. H., Hubbard, S. R., Harding, H. P., Clark, S. G., and Ron, D. (2002). IREl couples endoplasmic reticulum load to secretory capacity by processing the XBP-l mRNA. Nature 415(6867), 92-6. Carroll, K., Elroy-Stein, 0., Moss, B., and Jagus, R. (1993). Recombinant vaccinia virus K3L gene product prevents activation of double-stranded RNA- dependent, initiation factor 2 alpha-specific protein kinase. J Biol Chem 268(17), 12837-42. Cassady, K. A., and Gross, M. (2002). The herpes simplex virus type 1 U(S)11 protein interacts with protein kinase R in infected cells and requires a 30- amino-acid sequence adjacent to a kinase substrate domain. J Viral 76(5), 2029-35. Charlton, C. A., and Volkman, L. E. (1993). Penetration of Autographa californica nuclear polyhedrosis virus nucleocapsids into IPLB Sf 21 cells induces actin cable formation. Virology 197(1), 245-54. Chejanovsky, N., and Gershburg, E. (1995). The wild-type Autographa californica nuclear polyhedrosis virus induces apoptosis of Spodoptera littoralis cells. Virology 209(2), 519-25. Chen, C. J., Quentin, M. E., Brennan, L. A., Kukel, C., and Thiem, S. M. (1998). Lymantria dispar nucleopolyhedrovirus hrf-l expands the larval host range of Autographa californica nucleopolyhedrovirus. J Viral 72(3), 2526-31. Chen, C. J., and Thiem, S. M. (1997). Differential infectivity of two Autographa californica nucleopolyhedrovirus mutants on three permissive cell lines is the result of lef-7 deletion. Virology 227(1), 88-95. 116 Chen, H. H., Tsai, F. Y., and Chen, C. T. (2001). Negative regulatory regions of the PAT] promoter of Hz-l virus contain GATA elements which associate with cellular factors and regulate promoter activity. J Gen Vir0182(Pt 2), 313-20. Chen, J. J. (2000). Heme-regulated eIF2alpha kinase. In "Translational Control of Gene Expression" (N. J. Sonenberg, Hershey, J.W.B., Mathews M.B., Ed.), pp. 529-546. Cold Spring Harbor Laboratory Press, New York. Chen, J. J., and London, I. M. (1995). Regulation of protein synthesis by heme- regulated eIF-2 alpha kinase. Trends Biochem Sci 20(3), 105-8. Chen, Z., and Carstens, E. B. (2005). Identification of domains in Autographa californica multiple nucleopolyhedrovirus late expression factor 3 required for nuclear transport of P143. J Viral 79(17), 10915-22. Cheng, G., Feng, Z., and He, B. (2005). Herpes simplex virus 1 infection activates the endoplasmic reticulum resident kinase PERK and mediates eIF-2alpha dephosphorylation by the gamma(l)34.5 protein. J Viral 79(3), 1379-88. Chillaron, J., and Haas, I. G. (2000). Dissociation from BiP and retrotranslocation of unassembled immunoglobulin light chains _ are tightly coupled to proteasome activity. Mol Biol Cell 11(1), 217-26. Cioca, D. P., Aoki, Y., and Kiyosawa, K. (2003). RNA interference is a functional pathway with therapeutic potential in human myeloid leukemia cell lines. Cancer Gene T her 10(2), 125-33. Clem, R. J., Fechheimer, M., and Miller, L. K. (1991). Prevention of apoptosis by a baculovirus gene during infection of insect cells. Science 254(5036), 1388-90. Clem, R. J., and Miller, L. K. (1993). Apoptosis reduces both the in vitro replication and the in vivo infectivity of a baculovirus. J Viral 67(7), 3730-8. Clem, R. J., and Miller, L. K. (1994). Control of programmed cell death by the baculovirus genes p35 and iap. Mol Cell Biol 14(8), 5212-22. Clemens, M. J. (1993). The small RNAs of Epstein-Barr virus. Mol Biol Rep 17(2), 81-92. Clemens, M. J. (2005). Translational control in virus-infected cells: models for cellular stress responses. Semin Cell Dev Biol 16(1), 13-20. Clemens, M. J., Bushell, M., Jeffrey, I. W., Pain, V. M., and Morley, S. J. (2000). Translation initiation factor modifications and the regulation of protein synthesis in apoptotic cells. Cell Death Ditfer 7(7), 603-15. 117 Clemens, M. J., Bushell, M., and Morley, S. J. (1998). Degradation of eukaryotic polypeptide chain initiation factor (eIF) 4G in response to induction of apoptosis in human lymphoma cell lines. Oncogene 17(22), 2921-31. Crevel, G., Bates, H., Huikeshoven, H., and Cotterill, S. (2001). The Drosophila Dpit47 protein is a nuclear Hsp90 co-chaperone that interacts with DNA polymerase alpha. J Cell Sci 114(Pt 11), 2015-25. Crews, C. M. (2003). Feeding the machine: mechanisms of proteasome-catalyzed degradation of ubiquitinated proteins. Curr Opin Chem Biol 7(5), 534-9. Croizier, G., Croizier, L., Argaud, 0., and Poudevigne, D. (1994). Extension of Autographa californica nuclear polyhedrosis virus host range by interspecific replacement of a short DNA sequence in the p143 helicase gene. Proc Natl Acad Sci U S A 91(1), 48-52. Crouch, E. A., and Passarelli, A. L. (2002). Genetic requirements for homologous recombination in Autographa californica nucleopolyhedrovirus. J Viral 76(18), 9323-34. Dever, T. E., Sripriya, R., McLachlin, J. R., Lu, J., Fabian, J. R., Kimball, S. R., and Miller, L. K. (1998). Disruption of cellular translational control by a viral truncated eukaryotic translation initiation factor 2alpha kinase homolog. Proc Natl Acad Sci U S A 95(8), 4164-9. Donaldson, D., Kiely, T. and Grube, A. (2004) 2000-2001 Pesticide market estimates. Environmental Protection Agency, Washington, D.C. Du, X., and Thiem, S. M. (1997a). Responses of insect cells to baculovirus infection: protein synthesis shutdown and apoptosis. J Viral 71(10), 7866-72. Du, X., and Thiem, S. M. (1997b). Characterization of host range factor 1 (hrf-l) expression in Lymantria dispar M nucleopolyhedrovirus- and recombinant Autographa californica M nucleopolyhedrovirus-infected IPLB-Ld652Y cells. Virology 227(2), 420-30. Duan, L., Reddi, A. L., Ghosh, A., Dimri, M., and Band, H. (2004). The Cbl family and other ubiquitin ligases: destructive forces in control of antigen receptor signaling. Immunity 21(1), 7-17. Evans, J. T., Rosenblatt, G. S., Leisy, D. J., and Rohrmann, G. F. (1999). Characterization of the interaction between the baculovirus ssDNA-binding protein (LEF-3) and putative helicase (Pl43). J Gen Viral 80 ( Pt 2), 493-500. 118 Everett, R. D., Orr, A., and Preston, C. M. (1998). A viral activator of gene expression functions via the ubiquitin-proteasome pathway. Emba J 17(24), 7161-9. Evers, D. L., Wang, X., and Huang, E. S. (2004). Cellular stress and signal transduction responses to human cytomegalovirus infection. Microbes Infect 6(12), 1084-93. Federici, B. (1997). "Baculovirus Pathogenesis." The Baculoviruses (L. K. Miller, Ed.) Plenum New York. Feigenblum, D., and Schneider, R. J. (1993). Modification of eukaryotic initiation factor 4F during infection by influenza virus. J Viral 67(6), 3027-35. Freilich, S., Oron, E., Kapp, Y., Nevo-Caspi, Y., Orgad, S., Segal, D., and Chamovitz, D. A. (1999). The COP9 signalosome is essential for development of Drosophila melanogaster. Curr Biol 9(20), 1187-90. Friesen, P. D., and Miller, L. K. (1986). The regulation of baculovirus gene expression. Curr T op Microbiol Immunol 131, 31-49. Fuchs, L. Y., Woods, M.S., and RF. Weaver (1983). Viral transcription during Autographa californica nuclear polyhedrosis virus infection: a novel RNA polymerase induced in infected Spodoptera/rugiperda cells. J Virol 48, 641. Gale, M., J r., Tan, S. L., and Katze, M. G. (2000). Translational control of viral gene expression in eukaryotes. Microbial Mal Biol Rev 64(2), 239-80. Gebauer, F., and Hentze, M. W. (2004). Molecular mechanisms of translational control. Nat Rev Mol Cell Biol 5(10), 827-35. Gomi, 8., Zhou, C. E., Yih, W., Majima, K., and Maeda, S. (1997). Deletion analysis of four of eighteen late gene expression factor gene homologues of the baculovirus, BmNPV. Virology 230(1), 35-47. Goo, T. W., Yun, E. Y., Choi, K. H., Kim, S. H., Nho, S. K., Kang, S. W., and Kwon, O. Y. (2004). ATFC is a novel transducer for the unfolded protein response in Bombyx mori BMS cells. Biochem Biophys Res Commun 325(2), 626-31. Goodwin, R. H., Tompkins, G. J., and McCawley, P. (1978). Gypsy moth cell line divergent in viral susceptibility In Vitra 14, 485-494. Gradi, A., Svitkin, Y. V., Imataka, H., and Sonenberg, N. (1998). Proteolysis of human eukaryotic translation initiation factor eIF4GII, but not eIF4GI, coincides with the shutoff of host protein synthesis after poliovirus infection. Proc Natl Acad Sci U S A 95(19), 11089-94. ll9 Granados, R. R., Lawler, K. A., and Burand, J. P. (1981). Replication of Heliothis zea baculovirus in an insect cell line. Interviralogy 16(2), 71-9. Gray, E. S., and Tsai, R. W. (1994). Characterization of striped bass growth hormone receptors by disulfide-bond reduction and cross-linking studies. J Exp Zool 268(6), 428-35. Green, M. C., Monser, K. P., and Clem, R. J. (2004). Ubiquitin protein ligase activity of the anti-apoptotic baculovirus protein Op-IAP3. Virus Res 105(1), 89-96. Greene, J. C., Whitworth, A. J., Kuo, I., Andrews, L. A., Feany, M. B., and Pallanck, L. J. (2003). Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proc Natl Acad Sci U S A 100(7), 4078-83. Grula, M. A., Buller, P.L. and Weaver, R.F. (1981a). alpha-amanitin resistant viral RNA synthesis in nuclei isolated from nuclear polyhedrosis virus-infected Heliothis zea larvae and Spadapterafrugiperda cells. J Virol 38, 916. Grula, M. A. a. W., R.F. (1981b). An improved method for isolation of Heliothis zea DNA-dependent RNA polymerase: separation and characterization of a form 111 RNA polymerase activity. Insect Biochem 11, 149. Grundmann, 0., Mosch, H. U., and Braus, G. H. (2001). Repression of GCN4 mRNA translation by nitrogen starvation in Saccharomyces cerevisiae. J Biol Chem 276(28), 25661-71. Guarino, L. A., Jin, J., and Dong, W. (1998). Guanylyltransferase activity of the LEF-4 subunit of baculovirus RNA polymerase. J Viral 72(12), 10003-10. Guarino, L. A., Smith, G., and Doug, W. (1995). Ubiquitin is attached to membranes of baculovirus particles by a novel type of phospholipid anchor. Cell 80(2), 301-9. Guarino, L. A., Xu, B., Jin, J., and Dong, W. (1998). A virus-encoded RNA polymerase purified from baculovirus-infected cells. J Viral 72(10), 7985-91. Gunnery, 8., Rice, A. P., Robertson, H. D., and Mathews, M. B. (1990). Tat- responsive region RNA of human immunodeficiency virus 1 can prevent activation of the double-stranded-RNA-activated protein kinase. Proc Natl Acad Sci U S A 87(22), 8687-91. Guzo, D., Rathburn, H., Guthrie, K., and Dougherty, E. (1992). Viral and host cellular transcription in Autographa californica nuclear polyhedrosis virus- infected gypsy moth cell lines. J Viral 66(5), 2966-72. 120 Gygi, S. P., Rochon, Y., Franza, B. R., and Aebersold, R. (1999). Correlation between protein and mRNA abundance in yeast. Mol Cell Biol 19(3), 1720-30. Hansen, 1. A., Attardo, G. M., Park, J. H., Peng, Q., and Raikhel, A. S. (2004). Target of rapamycin-mediated amino acid signaling in mosquito anautogeny. Proc Natl Acad Sci U SA 101(29), 10626-31. Harari-Steinberg, 0., and Chamovitz, D. A. (2004). The COP9 signalosome: mediating between kinase signaling and protein degradation. Curr Protein Pept Sci 5(3), 185-9. Harding, H. P., Novoa, 1., Zhang, Y., Zeng, H., Wek, R., Schapira, M., and Ron, D. (2000a). Regulated translation initiation controls stress-induced gene expression in mammalian cells. Mol Cell 6(5), 1099-108. Harding, H. P., Zhang, Y., Bertolotti, A., Zeng, H., and Ron, D. (2000b). Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol Cell 5(5), 897-904. Harding, H. F., Zhang, Y., and Ron, D. (1999). Protein translation and folding are coupled by an endoplasmic-reticulum-resident kinase. Nature 397(6716), 271- 4. Harding, H. P., Zhang, Y., Zeng, H., Novoa, 1., Lu, P. D., Calfon, M., Sadri, N., Yun, C., Popko, B., Paules, R., Stojdl, D. F., Bell, J. C., Hettmann, T., Leiden, J. M., and Ron, D. (2003). An integrated stress response regulates amino acid metabolism and resistance to oxidative stress. Mol Cell 11(3), 619-33. He, B., Gross, M., and Roizman, B. (1997). The gamma(l)34.5 protein of herpes simplex virus 1 complexes with protein phosphatase lalpha to dephosphorylate the alpha subunit of the eukaryotic translation initiation factor 2 and preclude the shutoff of protein synthesis by double-stranded RNA-activated protein kinase. Proc Natl Acad Sci U S A 94(3), 843-8. Hefferon, K. L. (2004). Baculovirus late expression factors. J Mol Microbiol Biotechnol 7(3), 89-101. Hellen, C. U., and Sarnow, P. (2001). Internal ribosome entry sites in eukaryotic mRNA molecules. Genes Dev 15(13), 1593-612. Hershberger, P. A., Dickson, J. A., and Friesen, P. D. (1992). Site-specific mutagenesis of the 35-kilodalton protein gene encoded by Autographa californica nuclear polyhedrosis virus: cell line-specific effects on virus replication. J Viral 66(9), 5525-33. 121 Hershey, J. W. B., and Mathews, M. B. (1996). "Translational Control." (N. Sonenberg, Ed.) Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Hirai, M., Terenius, 0., Li, W., and Faye, I. (2004). Baculovirus and dsRNA induce Hemolin, but no antibacterial activity, in Antheraea pernyi. Insect Mol Biol 13(4), 399-405. Hoege, C., Pfander, B., Moldovan, G. L., Pyrowolakis, G., and Jentsch, S. (2002). RAD6-dependent DNA repair is linked to modification of PCNA by ubiquitin and SUMO. Nature 419(6903), 135-41. Holcik, M., and Sonenberg, N. (2005). Translational control in stress and apoptosis. Nat Rev Mol Cell Biol 6(4), 318-27. Hu, J., and Seeger, C. (1996). Hsp90 is required for the activity of a hepatitis B virus reverse transcriptase. Proc Natl Acad Sci U S A 93(3), 1060-4. Huh, N. E., and Weaver, R. F. (1990). Identifying the RNA polymerases that synthesize specific transcripts of the Autographa californica nuclear polyhedrosis virus. J Gen Viral 71 (Pt 1), 195-201. . Hung, J. J., Chung, C. S., and Chang, W. (2002). Molecular chaperone Hsp90 is important for vaccinia virus growth in cells. J Viral 76(3), 1379-90. Ikeda, M., Reimbold, E. A., and Thiem, S. M. (2005). Functional analysis of the baculovirus host range gene, hrf-l. Virology 332(2), 602-13. Imai, N., Matsuda, N., Tanaka, K., Nakano, A., Matsumoto, S., and Kang, W. (2003). Ubiquitin ligase activities of Bombyx mori nucleopolyhedrovirus RING finger proteins. J Viral 77(2), 923-30. Ishikawa, H., Ikeda, M., Alves, C. A., Thiem, S. M., and Kobayashi, M. (2004). Host range factor 1 from Lymantria dispar Nucleopolyhedrovirus (NPV) is an essential viral factor required for productive infection of NPVs in IPLB- Ld652Y cells derived from L. dispar. J Viral 78(22), 12703-8. Isler, J. A., Skalet, A. H., and Alwine, J. C. (2005). Human cytomegalovirus infection activates and regulates the unfolded protein response. J Viral 79(1 1), 6890-9. Ito, E., Sahri, D., Knippers, R., and Carstens, E. B. (2004). Baculovirus proteins IE- 1, LEF-3, and P143 interact with DNA in vivo: 11 formaldehyde cross-linking study. Virology 329(2), 337-47. 122 Janas, R. M., Marks, D. L., and LaRusso, N. F. (1994). Purification and partial characterization of a heat-resistant, cytosolic neuropeptidase from rat liver. Biochem Biophys Res Commun 198(2), 574-81. Jang, S. K., Krausslich, H. G., Nicklin, M. J., Duke, G. M., Palmenberg, A. C., and Wimmer, E. (1988). A segment of the 5' nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J Viral 62(8), 2636-43. Jentsch, S., and Schlenker, S. (1995). Selective protein degradation: a journey's end within the proteasome. Cell 82(6), 881-4. Jiang, H. Y., and Wek, R. C. (2005). Phosphorylation of the alpha-subunit of the eukaryotic initiation factor-2 (eIF2alpha) reduces protein synthesis and enhances apoptosis in response to proteasome inhibition. J Biol Chem 280(14), 14189-202. Joachims, M., Van Breugel, P. C., and Lloyd, R. E. (1999). Cleavage of poly(A)- binding protein by enterovirus proteases concurrent with inhibition of translation in vitro. J Viral 73(1), 718-27. Kalejta, R. F., Bechtel, J. T., and Shenk, T. (2003). Human cytomegalovirus pp7l stimulates cell cycle progression by inducing the proteasome-dependent degradation of the retinoblastoma family of tumor suppressors. Mol Cell Biol 23(6), 1885-95. Kalejta, R. F., and Shenk, T. (2003). Proteasome-dependent, ubiquitin-independent degradation of the Rb family of tumor suppressors by the human cytomegalovirus pp71 protein. Proc Natl Acad Sci U S A 100(6), 3263-8. Kamita, S. G., and Maeda, S. (1996). Abortive infection of the baculovirus Autographa californica nuclear polyhedrosis virus in Sf-9 cells after mutation of the putative DNA helicase gene. J Viral 70(9), 6244-50. Kamita, S. G., and Maeda, S. (1997). Sequencing of the putative DNA helicase- encoding gene of the Bombyx mori nuclear polyhedrosis virus and fine- mapping of a region involved in host range expansion. Gene 190(1), 173-9. Kampmueller, K. M., and Miller, D. J. (2005). The cellular chaperone heat shock protein 90 facilitates Flock House virus RNA replication in Drosophila cells. J Viral 79(11), 6827-37. Kaufman, R. J. (2000). In "Translational Control of Gene Expression" (N. Sonenberg, Hershey, J.W.B. and Mathews, M.B., Ed.), pp. 503-527. Cold Spring Harbor Laboratory Press, Cold Spring Harbor. 123 Keene, K. M., Foy, B. D., Sanchez-Vargas, I., Beaty, B. J., Blair, C. D., and Olson, K. E. (2004). RNA interference acts as a natural antiviral response to O'nyong-nyong virus (Alphavirus; Togaviridae) infection of Anopheles gambiae. Proc Natl Acad Sci U SA 101(49), 17240-5. Kerekatte, V., Keiper, B. D., Badorff, C., Cai, A., Knowlton, K. U., and Rhoads, R. E. (1999). Cleavage of Poly(A)-binding protein by coxsackievirus 2A protease in vitro and in vivo: another mechanism for host protein synthesis shutoff? J Viral 73(1), 709-17. Khoo, D., Perez, C., and Mohr, I. (2002). Characterization of RNA determinants recognized by the arginine- and proline-rich region of Us] 1, a herpes simplex virus type l-encoded double-stranded RNA binding protein that prevents PKR activation. J Viral 76(23), 11971-81. Kim, T., Hofmann, K., von Arnim, A. G., and Chamovitz, D. A. (2001). PCI complexes: pretty complex interactions in diverse signaling pathways. Trends Plant Sci 6(8), 379-86. Kimball, S. R. (1999). Eukaryotic initiation factor eIF2. Int J Biochem Cell Biol 31(1), 25-9. Kondo, A., and Maeda, S. (1991). Host range expansion by recombination of the baculoviruses Bombyx mori nuclear polyhedrosis virus and Autographa californica nuclear polyhedrosis virus. J Viral 65(7), 3625-32. Krappa, R., Behn-Krappa, A., Jahnel, F., Doerfler, W., and Knebel-Morsdorf, D. (1992). Differential factor binding at the promoter of early baculovirus gene PE38 during viral infection: GATA motif is recognized by an insect protein. J Viral 66(6), 3494-503. Kumar, R., Azam, S., Sullivan, J. M., Owen, C., Cavener, D. R., Zhang, R, Ron, D., Harding, H. P., Chen, J. J., Han, A., White, B. C., Krause, G. S., and DeGracia, D. J. (2001). Brain ischemia and reperfusion activates the eukaryotic initiation factor 2alpha kinase, PERK. J Neurochem 77(5), 1418- 21. LaCount, D. J., and Friesen, P. D. (1997). Role of early and late replication events in induction of apoptosis by baculoviruses. J Viral 71(2), 1530-7. Laemmli, U. K. (1970). Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 227(5259), 680-5. Lamphear, B. J., Yan, R., Yang, F., Waters, D., Liebig, H. D., Klump, H., Kuechler, E., Skern, T., and Rhoads, R. E. (1993). Mapping the cleavage site in protein 124 synthesis initiation factor eIF—4 gamma of the 2A proteases from human Coxsackievirus and rhinovirus. J Biol Chem 268(26), 19200-3. Landais, 1., Vincent, R., Bouton, M., Devauchelle, G., Duonor-Cerutti, M., and Ogliastro, M. (2005). Functional analysis of evolutionary conserved clustering of bZIP binding sites in the baculovirus homologous regions (hrs) suggests a cooperativity between host and viral transcription factors. Virology. Lee, H. H., and Miller, L. K. (1978). Isolation of genotypic variants of Autographa californica nuclear polyhedrosis virus. Journal of Virology 27, 754-767. Lee, K., Tirasophon, W., Shen, X., Michalak, M., Prywes, R., Okada, T., Yoshida, H., Mori, K., and Kaufman, R. J. (2002). IREl-mediated unconventional mRNA splicing and SZP-mediated ATF6 cleavage merge to regulate XBPl in signaling the unfolded protein response. Genes Dev 16(4), 452-66. Leisy, D. J., Rasmussen, C., Kim, H. T., and Rohrmann, G. F. (1995). The Autographa californica nuclear polyhedrosis virus homologous region 1a: identical sequences are essential for DNA replication activity and transcriptional enhancer function. Virology 208(2), 742-52. Li, L., Harwood, S. H., and Rohrmann, G. F. (1999). Identification of additional genes that influence baculovirus late gene expression. Virology 255(1), 9-19. Li, S., and Koromilas, A. E. (2001). Dominant negative function by an alternatively spliced form of the interferon-inducible protein kinase PKR. J Biol Chem 276(17), 13881-90. Li, Y., and Miller, L. K. (1995). Expression and functional analysis of a baculovirus gene encoding a truncated protein kinase homolog. Virology 206(1), 314-23. Lin, W., Choe, W. H., Hiasa, Y., Kamegaya, Y., Blackard, J. T., Schmidt, E. V., and Chung, R. T. (2005). Hepatitis C virus expression suppresses interferon signaling by degrading STAT]. Gastroenteralagv 128(4), 1034-41. Liu, C. Y., Schroder, M., and Kaufman, R. J. (2000). Ligand-independent dimerization activates the stress response kinases IREl and PERK in the lumen of the endoplasmic reticulum. J Biol Chem 275(32), 24881-5. Liu, Y., Schiff, M., Serino, G., Deng, X. W., and Dinesh-Kumar, S. P. (2002). Role of SCF ubiquitin-ligase and the COP9 signalosome in the N gene-mediated resistance response to Tobacco mosaic virus. Plant Cell 14(7), 1483-96. Lloyd, R. M., and Shatkin, A. J. (1992). Translational stimulation by reovirus polypeptide sigma 3: substitution for VAI RNA and inhibition of 125 phosphorylation of the alpha subunit of eukaryotic initiation factor 2. J Viral 66(12), 6878-84. Lowe, J., Stock, D., Jap, B., Zwickl, P., Baumeister, W., and Huber, R. (1995). Crystal structure of the 208 proteasome from the archaeon T. acidophilum at 3.4 A resolution. Science 268(5210), 533-9. Lu, A., and Miller, L. K. (1995a). Differential requirements for baculovirus late expression factor genes in two cell lines. J Viral 69(10), 6265-72. Lu, A., and Miller, L. K. (1995b). The roles of eighteen baculovirus late expression factor genes in transcription and DNA replication. J Viral 69(2), 975-82. Lu, A., and Miller, L. K. (1996). Species-specific effects of the hcf-l gene on baculovirus virulence. J Viral 70(8), 5123-30. Ma, Y., and Hendershot, L. M. (2004). ER chaperone functions during normal and stress conditions. J Chem Neuraanat 28(1-2), 51-65. Maeda, S. (1995). Further development of recombinant baculovirus insecticides. Curr Opin Biotechnal 6(3), 313-9. Maeda, S., Kamita, S. G., and Kondo, A. (1993). Host range expansion of Autographa californica nuclear polyhedrosis virus (NPV) following recombination of a 0.6-kilobase-pair DNA fragment originating from Bombyx mori NPV. J Viral 67(10), 6234-8. Maitra, R. K., McMillan, N. A., Desai, S., McSwiggen, J., Hovanessian, A. G., Sen, G., Williams, B. R., and Silverman, R. H. (1994). HIV-1 TAR RNA has an intrinsic ability to activate interferon-inducible enzymes. Virology 204(2), 823-7. Marissen, W. E., Guo, Y., Thomas, A. A., Matts, R. L., and Lloyd, R. E. (2000). Identification of caspase 3-mediated cleavage and functional alteration of eukaryotic initiation factor 2alpha in apoptosis. J Biol Chem 275(13), 9314- 23. Mathews, M. B., Sonenberg, N., and Hershey J.W.B. (2000). In "Translational Control of Gene Expression" (N. Sonenberg, Hershey, J.W.B. and Mathews, M.B., Ed.), pp. 1-32. Cold Spring Harbor Laboratory Press, Cold Spring Harbor. McClintock, J. T., Daugherty, E. M., and Weiner, R. M. (1986). Semipermissive replication of a nuclear polyhedrosis virus of Autographa californica in a gypsy moth cell line. Journal of Virology 57, 197-204. 126 McDougal, V. V., and Guarino, L. A. (2000). The Autographa californica nuclear polyhedrosis virus p143 gene encodes a DNA helicase. J Viral 74(11), 5273-9. McDougal, V. V., and Guarino, L. A. (2001). DNA and ATP binding activities of the baculovirus DNA helicase P143. J Viral 75(15), 7206-9. McLachlin, J. R., and Miller, L. K. (1994). Identification and characterization of vlf- 1, a baculovirus gene involved in very late gene expression. J Viral 68(12), 7746-56. Means, J. C., Muro, I., and Clem, R. J. (2003). Silencing of the baculovirus Op-iap3 gene by RNA interference reveals that it is required for prevention of apoptosis during Orgyia pseudotsugata M nucleopolyhedrovirus infection of Ld652Y cells. J Viral 77(8), 4481-8. Meng, L., Mohan, R., Kwok, B. H., Elofsson, M., Sin, N., and Crews, C. M. (1999). Epoxomicin, a potent and selective proteasome inhibitor, exhibits in vivo antiinflammatory activity. Proc Natl Acad Sci U S A 96(18), 10403-8. Meusser, B., Hirsch, C., Jarosch, E., and Sommer, T. (2005). ERAD: the long road to destruction. Nat Cell Biol 7(8), 766-72. Mikhailov, V. S., Okano, K., and Rohrmann, G. F. (2004). Specificity of the endonuclease activity of the baculovirus alkaline nuclease for single-stranded DNA. J Biol Chem 279(15), 14734-45. Mikhailov, V. S., and Rohrmann, G. F. (2002). Baculovirus replication factor LEF-1 is a DNA primase. J Viral 76(5), 2287-97. Miller, L. K. a. A. L. (1997). The Molecular Basis of Baculovirus Host Range. In "The Baculoviruses" (L. K. Miller, Ed.), pp. 217-235, Plenum Press, New York. Miron, M., Verdu, J., Lachance, P. E., Birnbaum, M. J., Lasko, P. F., and Sonenberg, N. (2001). The translational inhibitor 4E-BP is an effector of PI(3)K/Akt signalling and cell growth in Drosophila. Nat Cell Biol 3(6), 596- 601. Mistretta, T. A., and Guarino, L. A. (2005). Transcriptional activity of baculovirus very late factor 1. J Viral 79(3), 1958-60. Miyata, Y., and Yahara, I. (2000). p53-independent association between SV40 large T antigen and the major cytosolic heat shock protein, HSP90. Oncogene 19(11), 1477-84. 127 Mori, K., Ogawa, N., Kawahara, T., Yanagi, H., and Yura, T. (1998). Palindrome with spacer of one nucleotide is characteristic of the cis-acting unfolded protein response element in Saccharomyces cerevisiae. J Biol Chem 273(16), 9912-20. Morley, S. J., Coldwell, M. J., and Clemens, M. J. (2005). Initiation factor modifications in the preapoptotic phase. Cell Death Differ 12(6), 571-84. Morris, T. D., and Miller, L. K. (1992). Promoter influence on baculovirus-mediated gene expression in permissive and nonpermissive insect cell lines. J Viral 66(12), 7397-405. Morris, T. D., and Miller, L. K. (1993). Characterization of productive and non- productive AcMNPV infection in selected insect cell lines. Virology 197(1), 339-48. Morris, T. D., and Miller, L. K. (1994). Mutational analysis of a baculovirus major late promoter. Gene 140(2), 147-53. Moscardi, F. (1989). Use of viruses for pest control in Brazil: The case of the nuclear polyhedrosis virus of the soybean caterpillar Anticarsia gemmatalis. Memorias do Instituta Oswaldo Cruz Rio de Janeira 84, 51-56. Moscardi, F. (1999). Assessment of the application of baculoviruses for control of lepidoptera. Annu. Rev. Entomal. 44, 257-89 Mulvey, M., Poppers, J., Ladd, A., and Mohr, l. (1999). A herpesvirus ribosome- associated, RNA-binding protein confers a growth advantage upon mutants deficient in a GADD34-related function. J Viral 73(4), 3375-85. Mundt, K. E., Liu, C., and Carr, A. M. (2002). Deletion mutants in COP9/signalosome subunits in fission yeast Schizosaccharomyces pombe display distinct phenotypes. Mol Biol Cell 13(2), 493-502. Mundt, K. E., Porte, J., Murray, J. M., Brikos, C., Christensen, P. U., Caspari, T., Hagan, I. M., Millar, J. B., Simanis, V., Hofmann, K., and Carr, A. M. (1999). The COP9/signalosome complex is conserved in fission yeast and has a role in S phase. Curr Biol 9(23), 1427-30. Nakagawa, T., Zhu, H., Morishima, N., Li, E., Xu, J., Yankner, B. A., and Yuan, J. (2000). Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-beta. Nature 403(6765), 98-103. Ng, D. T., Spear, E. D., and Walter, P. (2000). The unfolded protein response regulates multiple aspects of secretory and membrane protein biogenesis and endoplasmic reticulum quality control. J Cell Biol 150(1), 77-88. 128 Nishio, M., Tsurudome, M., Ito, M., Garcin, D., Kolakofsky, D., and Ito, Y. (2005). Identification of paramyxovirus V protein residues essential for STAT protein degradation and promotion of virus replication. J Viral 79(13), 8591- 601. Nobiron, I., O'Reilly, D. R., and Olszewski, J. A. (2003). Autographa californica nucleopolyhedrovirus infection of Spodoptera frugiperda cells: a global analysis of host gene regulation during infection, using a differential display approach. J Gen Vir0184(Pt 11), 3029-39. O'Reilly, D. R., Miller, L. K., and Luckow, V. A. (1994). "Baculovirus Expression Vectors, A Laboratory Manual." Oxford University Press, New York, Oxford. Ohlmann, T., Prevot, D., Decimo, D., Roux, F., Garin, J., Morley, S. J., and Darlix, J. L. (2002). In vitro cleavage of eIF4GI but not eIF4GII by HIV-1 protease and its effects on translation in the rabbit reticulocyte lysate system. J Mol Biol 318(1), 9-20. Okada, T., Yoshida, H., Akazawa, R., Negishi, M., and Mori, K. (2002). Distinct roles of activating transcription factor 6 (ATF6) and double-stranded RNA- activated protein kinase-like endoplasmic reticulum kinase (PERK) in transcription during the mammalian unfolded protein response. Biochem J 366(Pt 2), 585-94. Olsen, D. S., Jordan, B., Chen, D., Wek, R. C., and Cavener, D. R. (1998). Isolation of the gene encoding the Drosophila melanogaster homolog of the Saccharomyces cerevisiae GCN2 eIF-2alpha kinase. Genetics 149(3), 1495- 509. 001, B. G., and Miller, L. K. (1988). Regulation of host RNA levels during baculovirus infection. Virology 166(2), 515-23. Ooi, B. G., Rankin, C., and Miller, 1.. K. (1989). Downstream sequences augment transcription from the essential initiation site of a baculovirus polyhedrin gene. J Mol Biol 210(4), 721-36. Palli, S. R., Caputo, G. F., Sohi, S. S., Brownwright, A. J., Ladd, T. R., Cook, B. J., Primavera, M., Arif, B. M., and Retnakaran, A. (1996). CfMNPV blocks AcMNPV-induced apoptosis in a continuous midgut cell line. Virology 222(1), 201-13. Park, Y. W., Wilusz, J., and Katze, M. G. (1999). Regulation of eukaryotic protein synthesis: selective influenza viral mRNA translation is mediated by the cellular RNA-binding protein GRSF-l. Proc Natl Acad Sci U S A 96(12), 6694-9. 129 Pataer, A., Vorburger, S. A., Barber, G. N., Chada, S., Mhashilkar, A. M., Zou- Yang, H., Stewart, A. L., Balachandran, S., Roth, J. A., Hunt, K. K., and Swisher, S. G. (2002). Adenoviral transfer of the melanoma differentiation- associated gene 7 (mda7) induces apoptosis of lung cancer cells via up- regulation of the double-stranded RNA-dependent protein kinase (PKR). Cancer Res 62(8), 2239-43. Patel, C. V., Handy, I., 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(48), 37993-8. Patil, C. K., Li, H., and Walter, P. (2004). Gcn4p and novel upstream activating sequences regulate targets of the unfolded protein response. PLoS Biol 2(8), E246. Pelletier, J., and Sonenberg, N. (1988). Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature 334(6180), 320-5. Peng, Z., Serino, G., and Deng, X. W. (2001a). Molecular characterization of subunit 6 of the COP9 signalosome and its role in multifaceted developmental processes in Arabidopsis. Plant Cell 13(11), 2393-407. Peng, Z., Serino, G., and Deng, X. W. (2001b). A role of Arabidopsis COP9 signalosome in multifaceted developmental processes revealed by the characterization of its subunit 3. Development 128(21), 4277-88. Pestova, T. V., and Kolupaeva, V. G. (2002). The roles of individual eukaryotic translation initiation factors in ribosomal scanning and initiation codon selection. Genes Dev 16(22), 2906-22. Pestova, T. V., Kolupaeva, V. G., Lomakin, I. B., Pilipenko, E. V., Shatsky, I. N., Agol, V. 1., and Hellen, C. U. (2001). Molecular mechanisms of translation initiation in eukaryotes. Proc Natl Acad Sci U SA 98(13), 7029-36. Peters, G. A., Khoo, D., Mohr, 1., and Sen, G. C. (2002). Inhibition of PACT- mediated activation of PKR by the herpes simplex virus type 1 Usll protein. J Viral 76(21), 11054-64. Piron, M., Delaunay, T., Grosclaude, J., and Poncet, D. (1999). Identification of the RNA-binding, dimerization, and eIF4GI-binding domains of rotavirus nonstructural protein NSP3. J Viral 73(7), 5411-21. Prasad, M. D., Han, S. J., Nagaraju, J., Lee, W. J., and Brey, P. T. (2003). Cloning and characterization of an eukaryotic initiation factor-2alpha kinase from the silkworm, Bombyx mori. Biochim Biophys Acta 1628(1), 56-63. 130 Pratt, W. B., and Toft, D. O. (2003). Regulation of signaling protein function and trafficking by the hsp90/hsp70-based chaperone machinery. Exp Biol Med (Maywaad) 228(2), 111-33. Prikhod'ko, E. A., Lu, A., Wilson, J. A., and Miller, L. K. (1999). In vivo and in vitro analysis of baculovirus ie-2 mutants. J Viral 73(3), 2460-8. Prikhod'ko, E. A., and Miller, L. K. (1998). Role of baculovirus IE2 and its RING finger in cell cycle arrest. J Viral 72(1), 684-92. Prosch, S., Priemer, C., Hoflich, C., Liebenthaf, C., Babel, N., Kruger, D. H., and Volk, H. D. (2003). Proteasome inhibitors: a novel tool to suppress human cytomegalovirus replication and virus-induced immune modulation. Antivir T her 8(6), 555-67. Rajasekhar, V. K., and Holland, E. C. (2004). Postgenomic global analysis of translational control induced by oncogenic signaling. Oncogene 23(18), 3248- 64. Rajasekhar, V. K., Viale, A., Socci, N. D., Wiedmann, M., Hu, X., and Holland, E. C. (2003). Oncogenic Ras and Akt signaling contribute to glioblastoma formation by differential recruitment of existing mRNAs to polysomes. Mol Cell 12(4), 889-901. Rankin, C., Ooi, B. G., and Miller, L. K. (1988). Eight base pairs encompassing the transcriptional start point are the major determinant for baculovirus polyhedrin gene expression. Gene 70(1), 39-49. Rao, R. V., Castro-Obregon, S., Frankowski, H., Schuler, M., Stoka, V., del Rio, G., Bredesen, D. E., and Ellerby, H. M. (2002). Coupling endoplasmic reticulum stress to the cell death program. An Apaf-l-independent intrinsic pathway. J Biol Chem 277(24), 21836-42. Rasmussen, C., Leisy, D. J., Ho, P. S., and Rohrmann, G. F. (1996). Structure- function analysis of the Autographa californica multinucleocapsid nuclear polyhedrosis virus homologous region palindromes. Virology 224(1), 235-45. Reilly, L. M., and Guarino, L. A. (1996). The viral ubiquitin gene of Autographa californica nuclear polyhedrosis virus is not essential for viral replication. Virology 218(1), 243-7. Reyes-Del Valle, J., Chavez-Salinas, S., Medina, F., and Del Angel, R. M. (2005). Heat shock protein 90 and heat shock protein 70 are components of dengue virus receptor complex in human cells. J Viral 79(8), 4557-67. 131 Rice, W. C., and Miller, L. K. (1986). Baculovirus transcription in the presence of inhibitors and in nonpermissive Drosophila cells. Virus Res 6(2), 155-72. Richardson, H., and Kumar, S. (2002). Death to flies: Drosophila as a model system to study programmed cell death. J Immunol Methods 265(1-2), 21-38. Rohrmann, G. F. (1986). Polyhedrin structure. J Gen Viral 67 ( Pt 8), 1499-513. Rohrmann, G. F. (1992). Baculovirus structural proteins. J Gen Virol 73 (Pt 4), 749- 61. Rutkowski, D. T., and Kaufman, R. J. (2004). A trip to the ER: coping with stress. Trends Cell Biol 14(1), 20-8. Salvatore, M., Basler, C. F., Parisien, J. P., Horvath, C. M., Bourmakina, S., Zheng, H., Muster, T., Palese, P., and Garcia-Sastre, A. (2002). Effects of influenza A virus NSl protein on protein expression: the NS] protein enhances translation and is not required for shutoff of host protein synthesis. J Viral 76(3), 1206-12. Sanders, H. R., Foy, B. D., Evans, A. M., Ross, L. S., Beaty, B. J., Olson, K. E., and Gill, S. S. (2005). Sindbis virus induces transport processes and alters expression of innate immunity pathway genes in the midgut of the disease vector, Aedes aegypti. Insect Biochem Mol Biol 35(11), 1293-307. Santoyo, J., Alcalde, J., Mendez, R., Pulido, D., and de Haro, C. (1997). Cloning and characterization of a cDNA encoding a protein synthesis initiation factor- 2alpha (eIF-2alpha) kinase from Drosophila melanogaster. Homology To yeast GCN2 protein kinase. J Biol Chem 272(19), 12544-50. Sarmento, L. M., Huang, H., Limon, A., Gordon, W., Fernandes, J., Tavares, M. J., Miele, L., Cardoso, A. A., Classon, M., and Carlesso, N. (2005). Notchl modulates timing of Gl-S progression by inducing SKP2 transcription and p27Kipl degradation. J Exp Med 202(1), 157-68. Satoh, S., Hijikata, M., Handa, H., and Shimotohno, K. (1999). Caspase-mediated cleavage of eukaryotic translation initiation factor subunit 2alpha. Biochem J 342 (Pt 1), 65-70. Schneider, R. J. (2000). Adenovirus inhibition of cellular protein synthesis and preferential translation of viral mRNAs. In "Translational Regulation" (M. B. Mathews, Sonenberg, N., Hershey, J.W.B., Ed.), pp. 901-914. Cold Spring Harbor Laboratory Press, Cold Spring Harbor. Schneider, R. J., and Mohr, I. (2003). Translation initiation and viral tricks. Trends Biochem Sci 28(3), 130-6. 132 Schwechheimer, C., and Deng, X. W. (2001). COP9 signalosome revisited: a novel mediator of protein degradation. Trends Cell Biol 11(10), 420-6. Schwechheimer, C., Serino, G., Callis, J., Crosby, W. L., Lyapina, S., Deshaies, R. J., Gray, W. M., Estelle, M., and Deng, X. W. (2001). Interactions of the COP9 signalosome with the E3 ubiquitin ligase SCFTIRI in mediating auxin response. Science 292(5520), 1379-82. Seeger, M., Kraft, R., Ferrell, K., Bech-Otschir, D., Dumdey, R., Schade, R., Gordon, C., Naumann, M., and Dubiel, W. (1998). A novel protein complex involved in signal transduction possessing similarities to 26$ proteasome subunits. F aseb J 12(6), 469-78. Sharpless, N. E., and DePinho, R. A. (2002). p53: good cop/bad cop. Cell 110(1), 9- 12. Shen, X., Ellis, R. E., Lee, K., Liu, C. Y., Yang, K., Solomon, A., Yoshida, H., Morimoto, R., Kumit, D. M., Mori, K., and Kaufman, R. J. (2001). Complementary signaling pathways regulate the unfolded protein response and are required for C. elegans development. Cell 107(7), 893-903. Simon, 0., Williams, T., Lopez-Ferber, M., and Caballero, P. (2004). Virus entry or the primary infection cycle are not the principal determinants of host specificity of Spodoptera spp. nucleopolyhedroviruses. J Gen Virol 85(Pt 10), 2845-55. Sin, N., Kim, K. B., Elofsson, M., Meng, L., Auth, H., Kwok, B. H., and Crews, C. M. (1999). Total synthesis of the potent proteasome inhibitor epoxomicin: a useful tool for understanding proteasome biology. Bioorg Med Chem Lett 9(15), 2283-8. Skowronek, M. H., Hendershot, L. M., and Haas, I. G. (1998). The variable domain of nonassembled lg light chains determines both their half-life and binding to the chaperone BiP. Proc Natl Acad Sci U S A 95(4), 1574-8. Soboloff, J., and Berger, S. A. (2002). Sustained ER Ca2+ depletion suppresses protein synthesis and induces activation-enhanced cell death in mast cells. J Biol Chem 277(16), 13812-20. Sood, R., Porter, A. C., Ma, K., Quilliam, L. A., and Wek, R. C. (2000). Pancreatic eukaryotic initiation factor-2alpha kinase (PEK) homologues in humans, Drosophila melanogaster and Caenorhabditis elegans that mediate translational control in response to endoplasmic reticulum stress. Biochem J 346 Pt 2, 281-93. 133 Spankuch, B., and Strebhardt, K. (2005). RNA interference-based gene silencing in mice: the development of a novel therapeutical strategy. Curr Pharm Des 11(26), 3405-19. Steinhilb, M. L., Turner, R. S., and Gaut, J. R. (2001). The protease inhibitor, MG132, blocks maturation of the amyloid precursor protein Swedish mutant preventing cleavage by beta-Secretase. J Biol Chem 276(6), 4476-84. Stewart, D., Kazemi, S., Li, S., Massimi, P., Banks, L., Koromilas, A. E., and Matlashewski, G. (2004). Ubiquitination and proteasome degradation of the E6 proteins of human papillomavirus types 11 and 18. J Gen Viral 85(Pt 6), 1419-26. Stewart, T. M., Huijskens, I., Willis, L. G., and Theilmann, D. A. (2005). The Autographa californica multiple nucleopolyhedrovirus ie0-ie1 gene complex is essential for wild-type virus replication, but either IEO or IE1 can support virus growth. J Viral 79(8), 4619-29. Su, H. L., Liao, C. L., and Lin, Y. L. (2002). Japanese encephalitis virus infection initiates endoplasmic reticulum stress and an unfolded protein response. J Viral 76(9), 4162-71. Summers, M. D. (1971). Electron microscopic observations on granulosis virus entry, uncoating and replication processes during infection of the midgut cells of Trichoplusia ni. J Ultrastruct Res 35(5), 606-25. Summers, M. D., and Amott, H. J. (1969). Ultrastructural studies on inclusion formation and virus occlusion in nuclear polyhedrosis and granulosis virus- infected cells of Trichoplusia ni (Hubner). J Ultrastruct Res 28(5), 462-80. Summers, M. D., Volkman, LE. and Hseih, C-H. (1978). Immunoperoxidase detection of baculovirus antigens in insect cells. J. Gen Virol. 40, 545 Svitkin, Y. V., Gradi, A., Imataka, H., Morino, S., and Sonenberg, N. (1999). Eukaryotic initiation factor 4GII (eIF4GII), but not eIF4GI, cleavage correlates with inhibition of host cell protein synthesis after human rhinovirus infection. J Viral 73(4), 3467-72. Tamai, M., Yokoo, C., Murata, M., Oguma, K., Sota, K., Sato, E., and Kanaoka, Y. (1987). Efficient synthetic method for ethyl (+)-(ZS,38)-3-[(S)-3-methyl- 1-(3- methylbutylcarbamoyl)butylcarbamoyl]-2-oxiranecarb oxylate (EST), a new inhibitor of cysteine proteinases. Chem Pharm Bull (T okyo) 35(3), 1098-104. Tardif, K. D., Mori, K., Kaufman, R. J., and Siddiqui, A. (2004). Hepatitis C virus suppresses the IREl-XBPl pathway of the unfolded protein response. J Biol Chem 279(17), 17158-64. 134 Tee, A. R., and Proud, C. G. (2002). Caspase cleavage of initiation factor 4E-binding protein 1 yields a dominant inhibitor of cap-dependent translation and reveals a novel regulatory motif. Mol Cell Biol 22(6), 1674-83. Theilmann, D., Blissard, G., Banning, B., Jehle, J. A., O'Reilly, D. R., Rohrmann, G., Thiem, S. M., and vlak, J. M. (2004). Baculoviridae. In "Virus Taxonomy. Eighth Report of the International Commitee on Taxonomy of Viruses", pp. 177-185. Thiem, S. M. (1997). Prospects for altering host range for baculovirus bioinsecticides. Curr Opin Biotechnol 8(3), 317-22. Thiem, S. M., and Chejanovsky, N. (2004). The role of baculovirus apoptotic suppressors in AcMNPV-mediated translation arrest in Ld652Y cells. Virology 319(2), 292-305. Thiem, S. M., Du, X., Quentin, M. E., and Berner, M. M. (1996). Identification of baculovirus gene that promotes Autographa californica nuclear polyhedrosis virus replication in a nonpermissive insect cell line. J Viral 70(4), 2221-9. Thiem, S. M., and Miller, L. K. (1989). Identification, sequence, and transcriptional mapping of the major capsid protein gene of the baculovirus Autographa californica nuclear polyhedrosis virus. J Viral 63(5), 2008-18. Thiem, S. M., and Miller, L. K. (1990). Differential gene expression mediated by late, very late and hybrid baculovirus promoters. Gene 91(1), 87-94. Todd, J. W., Passarelli, A. L., and Miller, L. K. (1995). Eighteen baculovirus genes, including lef-ll, p35, 39K, and p47, support late gene expression. J Viral 69(2), 968-74. Trudeau, D., Washburn, J. 0., and Volkman, L. E. (2001). Central role of hemocytes in Autographa californica M nucleopolyhedrovirus pathogenesis in Heliothis virescens and Helicoverpa zea. J Viral 75(2), 996-1003. Vail, P. V., Jay, D.L., Stewart, F.D., Martinez, A.J. and H.T. Dolmage (1978). Comparitive susceptibility of Heliothis virescens and H. zea to the nuclear polyhedrosis virus isolated from Autographa californica. J. Econ. Entomol. 71, 293-296. Vail, P. V. a. S. S. C. (1982). Comparative replication, mortality and inclusion body production of Autographa californica nuclear polyhedrosis virus in Heliothis sp. Ann. Entamal. Soc. Am. 75, 376-382. 135 van Oers, M. M., Doitsidou, M., Thomas, A. A., de Maagd, R. A., and Vlak, J. M. (2003). Translation of both S'TOP and non-TOP host mRNAs continues into the late phase of Baculovirus infection. Insect Mol Biol 12(1), 75-84. Vanarsdall, A. L., Okano, K., and Rohrmann, G. F. (2005). Characterization of the replication of a baculovirus mutant lacking the DNA polymerase gene. Virology 331(1), 175-80. Vaughn, J. L., Goodwin, R. H., Tompkins, G. J., and McCawley, P. (1977). The establishment of two cell lines from the insect Spodoptera frugiperda (Lepidoptera; Noctuidae). In Vitra 13, 213-217. Volkman, L. E., and Goldsmith, P. A. (1988). Resistance of the 64K protein of budded Autographa californica nuclear polyhedrosis virus to functional inactivation by proteolysis. Virology 166(1), 285-9. Washburn, J. 0., Kirkpatrick, B. A., and Volkman, L. E. (1995). Comparative pathogenesis of Autographa californica M nuclear polyhedrosis virus in larvae of Trichoplusia ni and Heliothis virescens. Virology 209(2), 561-8. Washburn, J. O., Kirkpatrick, B.A. and L.E. Volkman (1996). Insect protection against viruses. Nature 383, 767 Wei, N., and Deng, X. W. (1998). Characterization and purification of the mammalian COP9 complex, a conserved nuclear regulator initially identified as a repressor of photomorphogenesis in higher plants. Phatachem Photobial 68(2), 237-41. Williams, B. R. (1999). PKR; a sentinel kinase for cellular stress. Oncogene 18(45), 6112-20. Wilson, M. E., Mainprize, T. H., Friesen, P. D., and Miller, L. K. (1987). Location, transcription, and sequence of a baculovirus gene encoding a small arginine- rich polypeptide. J Viral 61(3), 661-6. Yamagishi, J., Isobe, R., Takebuchi, T., and Bando, H. (2003). DNA microarrays of baculovirus genomes: differential expression of viral genes in two susceptible insect cell lines. Arch Virol 148(3), 587-97. Yanase, T., Yasunaga, C., and Kawarabata, T. (1998). Replication of Spodoptera exigua nucleopolyhedrovirus in permissive and non-permissive lepidopteran cell lines. Acta Virol 42(5), 293-8. Yang, C. L., Stetler, D. A., and Weaver, R. F. (1991). Structural comparison of the Autographa californica nuclear polyhedrosis virus-induced RNA polymerase 136 and the three nuclear RNA polymerases from the host, Spodoptera frugiperda. Virus Res 20(3), 251-64. Ye, J., Rawson, R. B., Komuro, R., Chen, X., Dave, U. P., Prywes, R., Brown, M. S., and Goldstein, J. L. (2000). ER stress induces cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Mol Cell 6(6), 1355-64. Yearian, W. C. a. S. Y. Y. (1982). Control of insect pests of agricultural importance by viral insecticides. In "Microbial and viral pesticides" (E. Kurstak, Ed.), pp. 387-423, Marcel Dekker, New York. Yoneda, T., Imaizumi, K., Oono, K., Yui, D., Gomi, F., Katayama, T., and Tohyama, M. (2001). Activation of caspase-12, an endoplastic reticulum (ER) resident caspase, through tumor necrosis factor receptor-associated factor 2- dependent mechanism in response to the ER stress. J Biol Chem 276(17), 13935-40. Yoo, S., and Guarino, L. A. (1994a). The Autographa californica nuclear polyhedrosis virus ie2 gene encodes a transcriptional regulator. Virology 202(2), 746-53. Yoo, S., and Guarino, L. A. (1994b). Functional dissection of the ie2 gene product of the baculovirus Autographa californica nuclear polyhedrosis virus. Virology 202(1), 164-72. Yoshida, H., Haze, K., Yanagi, H., Yura, T., and Mori, K. (1998). Identification of the cis-acting endoplasmic reticulum stress response element responsible for transcriptional induction of mammalian glucose-regulated proteins. Involvement of basic leucine zipper transcription factors. J Biol Chem 273(50), 33741-9. Yoshida, H., Matsui, T., Hosokawa, N., Kaufman, R. J., Nagata, K., and Mori, K. (2003). A time-dependent phase shift in the mammalian unfolded protein response. Dev Cell 4(2), 265-71. Yoshida, H., Matsui, T., Yamamoto, A., Okada, T., and Mori, K. (2001). XBPl mRNA is induced by ATF6 and spliced by [RBI in response to ER stress to produce a highly active transcription factor. Cell 107(7), 881-91. Young, J. C., Barral, J. M., and Ulrich Hartl, F. (2003). More than folding: localized functions of cytosolic chaperones. Trends Biochem Sci 28(10), 541-7. Yu, G. Y., and Lai, M. M. (2005). The ubiquitin-proteasome system facilitates the transfer of murine coronavirus from endosome to cytoplasm during virus entry. J Viral 79(1), 644-8. 137 Yu, Y., Wang, S. E., and Hayward, G. S. (2005). The KSHV immediate-early transcription factor RTA encodes ubiquitin E3 ligase activity that targets IRF7 for proteosome-mediated degradation. Immunity 22(1), 59-70. Zambon, R. A., Nandakumar, M., Vakharia, V. N., and Wu, L. P. (2005). The Toll pathway is important for an antiviral response in Drosophila. Proc Natl Acad Sci U SA 102(20), 7257-62. Zheng, T. S., Hunot, S., Kuida, K., and Flavell, R. A. (1999). Caspase knockouts: matters of life and death. Cell Death Differ 6(1 1), 1043-53. 138 uiiiiiiiiiiiiii"