.v+7.(740\ . 23. .1111. 1., ‘ .. I J 2...}. Ly .4 5 ‘ L- h 4...» IL}; I! .4 .11 2. (3.3!! . v9 . 0...!” Inuit," .9:I-v). )1 . Yn..lciatl.\' ?.!. 3,...»832‘tf ,flwn. 523: '1‘: .i. a $27I.~ ‘ . 2. ii. .. V) ..’hh. .t4néL.v:.Z A , . . 3...}. 3 tan . . frr L ..3:..\. ; .3131 . ‘ . , ‘v...5...\ (‘¢ , v tux. :w‘ ...J. . n\.l.r!v : 3.. : Ls».-. MICHIGAN STAT Mimi 1M 71m ‘IIWIHHI ) 43144 RARIES I! I iii/r» w! i. 3193009 This is to certify that the dissertation entitled Molecular Characterization of Oncogenic Marek's Disease Virus Attenuation: Genomic Mutations and their Affect on Viral Growth and Gene Expression. presented by Melinda R. Wilson has been accepted towards fulfillment of the requirements for Ph.D. degree in Animal Science LV— 4 Major professor Date 3/3‘I‘i3 MSU is an Affirmative Action/Equal Opportunity Institution 0 A 12771 v V LIBRARY Michigan State University \ J PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative Action/Equal Opportunity Institution cmmui MOLECULAR CHARACTERIZATION OF ONCOGENIC MAREK'S DISEASE VIRUS ATTENUATION: GENOMIC MUTATIONS AND THEIR AFFECT ON VIRAL GROWTH AND GENE EXPRESSION. By Melinda R. Wilson A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Animal Science 1 993 ABSTRACT MOLECULAR CHARACTERIZATION OF ONCOGENIC MAREK’S DISEASE VIRUS ATTENUATION: GENOMIC MUTATIONS AND THEIR AFFECT ON VIRAL GROWTH AND GENE EXPRESSION. BY Melinda R. Wilson Marek’s Disease virus (MDV), an oncogenic avian herpesvirus, produces malignant lymphomas and nerve demyelination in infected hosts. Serial passage of oncogenic Marek’s Disease virus in cultured avian cells results in attenuation of viral oncogenicity. Due to emergence of very virulent (v.v.) MDV in vaccinated flocks, analysis of viral attenuation is of considerable interest. However, the tightly cell-associated in vitro growth Characteristics of MDV present unique problems when attempting to purify, analyze, and manipulate MDV genomes. To facilitate molecular characterization of MDV, contour-clamped homogeneous electric fields electrophoresis (CHEF) was used to purify infectious MDV genomes. CHEF techniques were optimized for evaluation of total genome size and alterations in structure which occur during attenuation of oncogenic MDV. Using these techniques, a 200 bp deletion in attenuated genomes of the v.v. MDV. strain Mdll was identified. The 200 bp deletion, located in the BamHl-L fragment, is separate and distinct from the previously reported 132 bp amplified region. An attenuated variant which contains 132 bp amplified regions but has the wild type BamHl-L fragment, was utilized to determine if deletions in the BamHl-L fragment affect biological properties of attenuated Md11. Deletions in the BamHl-L region of attenuated Md11 were associated with increased viral growth and altered immediate early gene expression in vitro. In vivo studies indicate that the BamHl-L region does not directly contribute to MDV-induced oncogenicity. Analysis of RNA transcripts suggests that this region may be involved in latent infection of T- Iymphocytes by MDV. Another prominent characteristic of attenuated MDV is reduced expression of a predominant glycoprotein antigen, 90. To investigate the mechanism by which MDV gC expression is reduced during attenuation of the v.v. MDV strain Md11, protein levels, gene structure, steady state RNA levels, and transcription rates of MDV 90 from oncogenic and attenuated isolates of Md11 were determined. The 96 promoter, or MDV regulatory protein(s) that interact with the MDV gC promoter, are altered during attenuation of MDV strain Md11. Specific mutations within attenuated MDV genomes may be involved in separate steps of the pathway leading to MDV-induced oncogenicity. Copyrighted by Melinda Ricarda Wilson 1 993 To my husband, Michael F. Daeschlein for his love, support, and patience. and to my daughter, Amber M. Wilson-Daeschlein, for giving me the motivation to finish. ACKNOWLEDGMENTS A special thanks to my advisor, Paul M. Coussens, for his guidance, friendship, encouragement, financial support, and for creating a research environment in which I could grow creatively, scientifically, and intellectually. I am grateful to members of my guidance committee, Drs. S. Triezenberg, L. Velicer, J. Ireland, and M. Vandehaar for their helpful suggestions and discussions. I gratefully acknowledge all the people I have worked with in Dr. Coussens lab for their helpful conversations and technical prowess. I especially thank R. Southwick for his expertise with computers and excellent technical assistance. A sincere thanks to my parents Wanda Wilson and David Jordan; my sisters Wendy Middleton and Sandy Geiswite; and my mother-in-law Marge Daeschlein for there continuous encouragement. vi Table of Contents. Chapter 1. LITERATURE SURVEY ...................................... 1 I. HERPESVIRUSES .................................. 2 1. Introduction. .................................. 2 2. Herpesvirus classification. ........................ 5 3. Temporal Classification of herpesvirus genes. .......... 6 4. Herpesvirus proteins and their functions. ............. 6 a. HSV I IE proteins and their functions. ......... 6 b. VZV and EBV IE genes. ................... 7 c. Early and Late genes. ...................... 8 d. ”Non-essential" herpesvirus genes. ............ 9 II. MOLECULAR BASIS FOR VIRAL ATTENUATION. .......... 10 1. The poliovirus story. ........................... 10 2. Other RNA viruses. ............................ 14 3. Vaccinia virus. ................................ 16 4. Herpesviruses. ............................... 18 III. HISTORY OF VIRUS INDUCED CANCERS ................ 21 1. RNA viruses. ................................. 21 2. DNA tumor viruses. ............................ 22 vii ITI‘ 3. Herpesvimses. ................................ 23 IV. MAREK’S DISEASE VIRUS. .......................... 25 1. Pathology of Marek's Disease virus. ................ 25 2. Biology of MDV. ............................... 29 3. MDV genome structure. ......................... 31 4. MDV Proteins: general information. ................ 34 5. Molecular biology of MDV oncogenicity. .............. 35 a. Serotype comparisons. ..................... 37 b. MDV transformed cell lines ................... 38 6. MDV vaccines. ................................ 41 7. Attenuation of MDV. ............................ 45 a. Genome structure. ....................... 45 b. 90 expression. ........................... 47 C. Attenuation of v.v. MDV. ................... 48 Chapter 2. .............................................. 51 PURIFICATION AND CHARACTERIZATION OF INFECTIOUS MAREK'S DISEASE VIRUS GENOMES USING PULSED FIELD ELECTROPHORESIS. ABSTRACT ......................................... 52 I. INTRODUCTION. .................................. 53 II. MATERIALS AND METHODS ......................... 55 III RESULTS ....................................... 59 IV. DISCUSSION .................................... 63 IV. ACKNOWLEDGMENTS ............................. 65 viii Chapter 3. ............................................... 78 STRUCTURAL ALTERATIONS IN THE REPEAT REGIONS OF THE ATTENUATED MAREK’S DISEASE VIRUS GENOME AFFECTS GROWTH AND IMMEDIATE EARLY GENE TRANSCRIPTION IN VITRO. ABSTRACT ......................................... 79 I. INTRODUCTION. ................................... 80 II. MATERIALS AND METHODS .......................... 83 III. RESULTS ........................................ 87 IV. DISCUSSION ..................................... 95 ACKNOWLEDGMENTS ................................. 1 00 Chapter 4. ............................................... 119 MOLECULAR ANALYSIS OF THE GLYCOPROTEIN C-NEGATIVE PHENOTYPE OF ATTENUATED MAREK’S DISEASE VIRUS. ABSTRACT ......................................... 120 I. INTRODUCTION .................................... 121 II. MATERIALS AND METHODS .......................... 123 III. RESULTS ........................................ 128 IV. DISCUSSION ..................................... 133 ACKNOWLEDGMENTS ................................. 1 38 Chapter.5 ................................................ 150 DISCUSSION I. SUMMARY OF RESULTS AND CONCLUSIONS. ............ 150 II. THEORETICAL RELATIONSHIPS OF MDV ATTENUATION AND MDV-INDUCED ONCOGENICITY: WORKING HYPOTHESES TO BE TESTED. ..................... 155 III. FUTURE RESEARCH DIRECTIONS. .................... 158 ix Technique development. .......................... 158 MDV attenuation. ............................... 159 Vaccine development. ............................ 162 APPENDIX ............................................... 164 DESCRIPTION OF METHODS General information. ................................... 164 CHEF. ............................................. 164 Immunoprecipitation of proteins. ........................... 165 Nuclear run off transcription assay. ........................ 165 LIST OF REFERENCES ..................................... 171 LIST OF TABLES Chapter 2. TABLE 1. Genome Sizes of MDV Strains as Determined from High Resolution CHEF Gels. .................................. 73 Chapter 3. Table 1. Abundance of viral transcripts which hybridize to the MDV BamHI-L DNA fragment. ............................................ 1 12 Table 2. Temporal classification of viral transcripts which hybridize to the MDV BamHI-L DNA fragment. ..................................... 112 Table 3. Incidence of Marek’s Disease with oncogenic and attenuated Md11 and Md11p83-L. ........................................... 118 Chapter 4. Table 1. Abundance of viral transcripts in cells infected with LP and HP MDV strain Md11. ............................................. 146 xi LIST OF FIGURES Chapter 1. Figure 1. Pathway to MDV transformation of T-Iymphocytes ............ 27 Figure 2. Comparison of HVT and HSV genomic segment sizes to serotype 1 MDV genomic segment sizes. ................................. 33 Figure 3. Location of MDV genes on the MDV genome. .............. 36 Chapter 2. Figure 1. Separation of oncogenic and attenuated cell-free MDV genomes from chicken chromosomes using PFE. .............................. 68 Figure 2. Identification and localization of cell-free MDV genomes and chicken Chromosomes on CHEF gels. ................................. 70 Figure 3. Infectivity of cell-free MDV genomes after isolation from LMP agarose CHEF gels ................................................ 72 Figure 4. Cell-free genome size comparison of oncogenic, attenuated, and naturally occurring non-oncogenic MDV serotypes using high resolution CHEl-75 Figure 5. Characterization of alterations in attenuated MDV genome structure using CHEF purified call-free MDV genomes as hybridization probes. . . . . 77 Chapter 3. Figure 1. Identification of an attenuated MDV strain Md11 variant. ....... 102 Figure 2. Identification of the altered DNA fragment in the attenuated Md11 variant as BamHI L. ........................................ 104 xii Figure 3. Comparison of plaque morphology of oncogenic and attenuated Md11 to Md11p83-L . ........................................... 106 Figure 4. Panel A: comparison of plaque formation of oncogenic and attenuated Md11 to Md11p83-L in CEF. Panel B: comparison of viral DNA replication of oncogenic and attenuated Md11 to Md11p83-L in CEF. . . . . 108 Figure 5. Comparison of rates of plaque formation and DNA replication (day 1- 3) of oncogenic and attenuated Md11 to Md11p83-L in CEF. .......... 110 Figure 6. Restriction mapping to determine sub-localization of the 200 bp deletion in the attenuated Md11 genome. ........................ 114 Figure 7. Sub-localization of the 200 bp deletion in the BamHI-L fragment of attenuated Md11 genomes. ................................... 116 Chapter 4. Figure 1. Immunoprecipitation of 35S-labeled viral proteins from cell Iysates and media of DEF infected with oncogenic or attenuated MDV strain Md11. . . . 140 Figure 2. RFLP analysis of the MDV strain Md11 gene locus and flanking regions. ................................................. 142 Figure 3. MDV strain Md11 gC steady state RNA levels in DEF infected with oncogenic and attenuated isolates. ............................. 144 Figure 4. Transcription rates of MDV strain Md11 genes. ............. 148 Chapter 5. Figure 1. Hypothetical relationships between mutations of attenuated Md11 and the pathway to MDV transformation of T-Iymphocytes ................. 157 Appendix Figure 1. Clamped homogeneous electric fields electrophoresis. ........ 167 Figure 2. Resolution of large DNA molecules using CHEF. ............ 168 xiii Figure 3. Immunoprecipitation of proteins from MDV infected cells. ...... 169 Figure 4. Nuclear run off transcription assay. ...................... 170 xiv pCl/ml 281 MI/1 a.a. ANOVA DP CEF CHEF CHX cm CPM/ul DEF DNA EtBr fMoles GA 96 LIST OF ABBREVIATIONS microcuries pre milliliter serotype 2 MDV strain amino acid analysis of variance base pairs chick embryo fibroblasts clamped homogeneous electric fields cycloheximide centimeter counts per minute per microliter duck embryo fibroblasts deoxyribonucleic acid eany ethidium bromide femtamoles serotype 1 MDV strain glycoprotein B glycoprotein C XV HSV HP HVT ICP Ics IE IRL IR: JM kbp LMP LP MD Md1 1 MDNA MDV "'99 min nm nt. °C herpes simplex virus high passage herpesvirus of turkeys infected cell protein immune chicken sera immediate early internal repeat long internal repeat short serotype 1 MDV strain kilobase pairs late low melting point agarose low passage mole Marek’s disease Serotype 1 MDV strain Marek's disease nuclear antigen Marek’s disease virus Marek’s EcoRl 0 minute nanometer nucleotide degrees celsius xvi PAA PFU 9938 RaA RFLP RNA S.E.M. sa-1 8J1 SORF TBE TRL T33 uFD v.v. VIA VP phosphonoacetic acid plaque forming unit phosphoprotein of 38 molecular weight rabbit-anti-A (gC) restriction fragment length polymorphism analysis ribonucleic acid standard error of the mean serotype 2 MDV standard deviation specific open reading frame tris-borate-EDTA buffer terminal repeat long terminal repeat short microFarad microgram unique long unique short volt very virulent viral internal antigen virion protein xvii Chapter 1. LITERATURE SURVEY l. HERPESVIRUSES 1. Introduction. Herpesviruses are a large group of DNA viruses disseminated throughout the animal kingdom. These ubiquitous viruses cause disease in most organisms infected, thus they are intensely studied. In humans, herpesviruses are the causative agent of many common illnesses. Human herpesvirus infections cause cold sores (Herpes simplex virus I, HSV I), genital herpes (Herpes Simplex II, HSV ll), mononucleosis (Epstein barr virus, EBV), Chicken pox and shingles (Varicella- Zoster virus, VZV) (Volk, 1982). Many human herpesviruses can be life threatening to infants. Cytomegalovirus (CMV) is associated with congenital conditions such as mental retardation, microcephaly (small head), seizures, deafness, and jaundice (Volk, 1982). HSV II can be sub-Clinical or fatal to a newborn if an active infection occurs during childbirth. In some cases only mild Clinical symptoms occur in adults, with the exception of Herpes B virus. B virus is fatal in 75% of cases (Volk, 1982). Some herpesviruses have been implicated as the causative agent of some types of cancers. EBV is a suspect in human lymphomas. Lucke frog virus is associated with adenocarcinomas in frogs. HSV I is implicated in adeno- and fibrosarcomas. Herpesvims saimiri and H. Ateles are suspected as being the agent for lymphomas of monkeys. Marek’s disease virus (MDV) is the cause of lymphomas in chickens (Grodzicker and Hopkins, 1981). Herpesviruses that cause disease in cattle, sheep, goats, horses, swine, chickens, and turkeys have been identified (Review- FraenkeI-Conrat et al, 1988). 3 Two herpesviruses that infect cattle (BHV) cause rhinotracheitis, pustular vulvovaginitis, mammillitis, pseudolumpy skin disease and spontaneous abortion in cattle. Equine herpesvirus (EHV) can also cause abortion in horses and rhinopneumonoitis. Pseudorabies virus (PRV) infects swine and muses Aujeszky’s disease. MDV infects chickens and causes paralysis and lymphoma, commonly termed Marek's disease (MD) (Payne, 1985). Herpesviruses are relatively large in size compared to other viruses measuring 120 - 300 nm in diameter (Roizman, 1990). They contain a core in which the viral DNA resides. The core is encased by a capsid which is, in turn, surrounded by tegument. The entire virion is protected by an envelope. The envelope contains viral-encoded glycoprotein spikes protruding outward. Herpesviruses infect host cells by first attaching to the cell membrane. This initial step requires the interaction of viral glycoprotein spikes and the cell surface. After attachment, the virus penetrates the cell surface by a combination of phagocytosis and fusion of the virion envelope with the cell membrane, a step which also requires viral-encoded glycoproteins. The virus must then uncoat and get to the nucleus where, with the help of host cell machinery, it initiates transcription of viral genes and DNA replication (Review- Volk, 1982). All herpesviruses encode a variety of enzymes involved in nucleic acid metabolism, DNA synthesis, and transcriptional regulation (Roizman, 1990). Herpesvirus transcripts are transported to the cytoplasm where protein synthesis occurs (Volk, 1982). Viral glycoproteins are inserted into the nuclear membrane and the virion is assembled within the nucleus (Volk, 1982). Assembled virions then bud from 4 the nuclear membrane. The nuclear membrane containing virus encoded glycoproteins forms an envelope around the virus. Eventually, the cell dies, Iyses, and enveloped viruses are released to invade neighboring cells (Volk, 1982). During the course Of infection, herpesviruses "hide out" or become "latent". In latent herpesvirus infections, only a limited number of viral genes are expressed, and, in most cases, the genome exists as a Closed circular molecule. During a latent infection, no clinical signs of disease are evident. However, upon reactivation of latent herpesviruses, due to stress or immunosuppression, the virus embarks on a cytolytic life cycle described above. Due to the lack of viral protein expression during latency, the immune system is ineffective at ridding an organism of latent herpesviruses. Thus, in most cases, an organism that is infected with a herpesvirus is infected for life (Review— Roizman, 1990). The biological characteristics of a disease process caused by herpesviruses depends upon the interaction of virus encoded proteins with its host cell and host encoded proteins. To understand the disease process of herpesviruses, molecular biology techniques have been utilized to identify viral-encoded proteins and study their interaction with the host cell. The most intensely studied herpesvirus and guide for studying other rat-herpesviruses is HSV l. The following sections provide an overview of what is presently understood about HSV proteins at the molecular level. Other herpesviruses will be discussed when relevant. 2. Herpesvirus classification. Herpesviruses are classified based upon genome structure and biological properties. Genome structure classification is based on the location and 5 arrangement of repeated sequences within the genome. Six classes of genome structure exist within herpesviruses (Roizman, 1990). In most cases, genome structure Classification correlates with classification based on biological properties. Classification of herpesviruses based on biological properties has led to some controversy over classification of viruses which have genome structure similarities to one virus group but biological properties similar to another group. Such is the case with MDV. Previously, MDV was Classified as an y-herpesvirus. However, due to recent information on MDV genome structure, MDV is presently unclassified (Roizman, 1992). Classification based on biological properties is summarized below (Roizman, 1990): (II-herpesviruses are characterized by wide host range, short replication cycle, and usually latent infection of sensory root ganglia (HSV l, HSV II, VZV, BHV, EHV, PRV, and B virus). 8-herpesvlruses are characterized by relatively restricted host range, long replication cycle, and latency in secretory glands, lymphoreticular cells, kidney and other tissues (CMV). y-herpesvlruses are notable for infecting only T and B lymphocytes. Latency is confined to lymphoid tissues (EBV, H. Saimiri, and H. Ateles). 3. Temporal classification of herpesvirus genes. After herpes simplex virus infects a cell, expression of its genes occurs in a tightly regulated, temporal fashion (Review- Wagner, 1991). The virion carries 6 within it a protein called VP16. VP16 is a transactivator that, in conjunction with cellular proteins, initiates transcription of immediate early (IE) genes . IE genes encode regulatory proteins required for the synthesis of early (E) and late (L) genes. E genes primarily encode proteins involved in viral DNA replication, while most L genes are expressed only after the onset of viral DNA replication. However, DNA synthesis is not a strict requirement for L gene expression. L genes encode structural proteins for virion assembly and VP16. IE genes are expressed immediately upon infection, and IE gene transcripts will accumulate in the absence of viral protein synthesis (Roizman and Sears, 1990). E gene transcripts can be detected as short as 30 minutes after infection and will accumulate if DNA synthesis is blocked by chemical inhibitors (Wagner, 1991). Maximal expression of L genes can be detected at 3 hours post-infection (Wagner, 1991). Sequence analysis of the promoter regions of representative IE, E, and L genes indicate that, in most cases, each class has specific sequence and structural motifs that define the gene expression class. 4. Herpesvirus proteins and their functions. a. HSV l IE proteins and their functions. There are five known HSV I IE proteins: ICPO, ICP4, ICP27, ICP22, and ICP47 (Roizman and Sears, 1990). Of this group, the first four are known to affect herpesvirus gene expression. ICPO is non-essential in cell culture, is a promiscuous transactivator acting through another protein (Sandri-Goldin, 1991 ), and is hypothesized to be under antisense regulation during latency (Stevens et al, 1987). In addition, virus mutants which 7 do not express ICPO exhibit an larger particle to PFU ratio as compared to wild type viruses, suggesting that ICPO may confer a growth advantage to HSV l (Sandri-Goldin, 1991). ICP4 is essential for activation of E and L gene expression. In addition, ICP4 down-regulates its own promoter and other IE gene promoters. ICPO and ICP4 act synergistically to enhance E and L gene expression during transient assays (Sandri-Goldin, 1991). ICP27 is essential for viral growth, negatively regulates IE and E genes, is non-essential for DNA replication, and is required for L gene expression. ICP27 may be involved in the switch from E to L gene expression (Sandri-Goldin, 1991). ICP22 is essential for L gene expression in some cell types and may be required for repression of cellular proteins (Smiley et al, 1991). b. VZV and EBV IE genes. Regulation of VZV gene expression appears to be similar to HSV I gene regulation (Ruyechan et al, 1991). However, only two VZV regulatory proteins have been well characterized thus far. ORF62 is the functional homolog to HSV ICP4 and ORF61 is functionally analogous to HSV ICPO (Moriuchi, et al, 1992). EBV gene regulation during lytic infection has not been well characterized due to lack of appropriate in vitro culture systems (Hayward and Hardwick, 1991). However, three IE genes have been identified: Zta, Mta, and Rta. Zta (or ZEBRA) appears to be essential for induction of lytic cycle gene expression (Hayward and Hardwick, 1991). During latent EBV infection, EBNA 1 regulates at the level of DNA replication, and EBNA 2 is essential for growth transformation of lymphocytes (Raab-Traub and Gilligan, 1991). Since temporal classification of herpesvirus proteins are determined during 8 lytic infection, EBNA 1 and 2 are not Classified as IE proteins, but are considered regulatory proteins. EBNA 1 and 2 transcripts contain multiple splice sites. EBNA 2 has up to 11 intervening sequences spanning 25 kb (Raab-Traub and Gilligan, 1991) c. Early and Late genes. Early genes encode enzymes and proteins involved in DNA metabolism and replication. Their expression is dependent upon activation by IE proteins (Roizman, 1990). The early proteins which have been well characterized are viral ribonucleotide reductase, dUTPase, major DNA binding protein, thymidine kinase, and DNA polymerase (Roizman and Sears, 1990). Late genes encode structural components of the virion and envelope. As with early genes, late gene expression is dependent upon activation by IE proteins, especially ICP4, ICPO, and ICP27 (Blair and Snowden, 1991). The well characterized late genes are those which encode VP16 and glycoproteins. VP16 as described above, activates expression of IE genes and initiates the cascade of herpesvirus gene expression in lytic infections (Everett et al, 1991). The VP16 protein is produced late during infection and is bound to the tegument during virion assembly (Roizman, 1991). Glycoproteins encoded by L genes are the major antigenic determinants of herpesviruses. These are the molecules which are most often recognized by the host immune system. For the virus, glycoproteins function in virus attachment and entry into the cell. Six HSV glycoproteins have been well Characterized: 98, 9C, gD, gE, 9H, and gI (Roizman and Sears, 1990). Attempts to elucidate the role of glycoproteins during infection have yielded the following conclusions (Roizman and 9 Sears, 1990): 98 binds to heparin sulfate and functions during penetration and cell fusion. Neutralizing antibodies and cell-mediated immune response are produced against QB. 98 is the most highly conserved glycoprotein among a, B, and y herpesviruses. 90 also has an affinity for heparin sulfate and may function during attachment of herpesvirus to its host cell. The host will produce antibody against 90, however, it is unclear if anti-gC antibodies are protective. Oddly, 90 can bind the c3b component of complement and accelerate its decay. Functional significance of C3b binding by 9C is unclear. 9C is also conserved among cr- herpesviruses. Glycoprotein D functions during penetration and cell fusion and is the most potent inducer of neutralizing antibodies among herpesvirus glycoproteins. When complexed with gl, gE forms a high affinity receptor for host IgG. As with gC-c3b interactions, the functional significance of gE-gI-lgG binding is unclear. gE is not, itself, a potent inducer of humoral immune response. gH functions to stabilize attachment of virions to the host cell during penetration and cell fusion and can induce neutralizing antibodies. gH is also highly conserved among herpesviruses. d. "Non-essential" herpesvirus genes. Experiments utilizing herpesvirus mutants and antibody blocking studies have indicated that some HSV genes are not required for growth In vlfro. Some of these non-essential genes encode: gE, gl, gC, ribonucleotide reductase, dUTPase, and ICPO (Roizman and Sears, 1990). Although they are not required in cell culture (a basic system), and therefore are often considered non-essential, they may, in fact, function in a cell-specific or tissue-specific manner in the host organism. In addition, these proteins may play 10 a role in the characteristic manifestation of a specific disease process. ll. MOLECULAR BASIS FOR VIRAL ATTENUATION. Viral attenuation leads to reduced ability of viruses to induce disease in their natural host. The most notable attenuating mutations affect the interactions of viral-encoded proteins with host cells (Review- Tyler and Fields, 1990). In this section, literature related to analysis of the molecular mechanisms of attenuation of viral pathogenesis will be reviewed. Discussion will start with poliovirus, one of the smallest viruses known to cause disease, and end with vaccinia virus and herpes virus, the largest viruses known to cause disease. 1. The poliovirus story. Paralytic poliomyelitis has afflicted humankind for centuries. The earliest known record of poliomyelitis is found on an Egyptian stone carving dating 1500 BC (Melnick, 1982). From ancient times to the end of the 19'" century, polio was reported as sporadic cases in endemic proportions (Melnick, 1982). During the late 19‘" and early 20'" centuries, urbanized industrial centers of Europe and the United States experienced polio epidemics of increasing frequency and severity. Although concern for this disease was great, research on polio advanced slowly, with isolation of the causative agent in 1909 leading to development of effective vaccines in the mid 1950s (Salk, 1953; Sabin, 1957). During the twelve years after introduction of the inactivated (Salk) and live (Sabin) polio vaccines, a 99% decrease in reported poliomyelitis was observed (Melnick, 1982). However, polio cases continued to be reported and 17% of reported cases were in children who 11 had received the vaccine (Melnick, 1982). Presently, a world wide residual level of poliomyelitis of 0.84 cases/million of population/year continues to exist, and is associated with vaccinated populations (Melnick, 1982). The causative agent of paralytic poliomyelitis is poliovirus, a small virus containing a single stranded RNA genome of 7450 nucleotides (Putnak and Phillips, 1981). The RNA genome is translated as a single polycistronic protein. Extensive splicing and proteolytic cleavage occurs after transcription. The poliovirus genome encodes four structural proteins (VP1-VP4), a protease, a polymerase, and VPg (protein primer for RNA synthesis) (Phillips, 1981). There are three distinct serotypes of poliovirus, and each serotype causes identical disease symptoms. Serotype 3 poliovirus is associated with vaccine breaks at a higher frequency than serotypes 1 or 2 (Collingham et al., 1978). The residual level of poliomyelitis and sporadic episodes of polio associated with vaccinated populations prompted attempts to' understand the molecular mechanisms of attenuation of the Sabin polio vaccine. In 1984, Stanway et al. reported on the comparison of the complete nucleotide sequence of serotype 3 wild type polio and its attenuated vaccine derivative. Their data shows that only 10 point mutations occurred during attenuation and the 3’ and gained a single base (G). 01 these ten differences, 7 do not result in amino acid (a.a.) changes. The 3 Changes which did result in a.a. substitutions were located within VP3, VP1, and the non- structural protein. Two mutations reside within the 5’ noncoding region of the RNA genome. Based on sequence comparisons, these authors concluded that only mutations resulting in a.a. substitutions were responsible for attenuation of 12 serotype 3 poliovirus neurovirulence. Further clarification of mutations responsible for attenuation of neurovirulence came from comparative studies of the genome sequences of serotype 3 wild type, its vaccine derivative, and a vaccine revertant (Pa/119) recovered from a patient with a fatal case of vaccine-associated poliomyelitis (Cann et al., 1984). P3/119 differed from the vaccine strain at 7 nucleotide locations. One change in the 5' non-coding (nt. 472) was a true reversion to the wild type nucleotide, three Changes resulted in a.a. substitutions in VP1 and VP2, and one change resulted in loss of two nucleotides at the 3’ end just prior to the polyadenylation tract. Evidence of true reversion at nt. 472 to wild type prompted studies on partial sequence characterization of vaccine strains recovered from infants after primary vaccination (Evans et al., 1985). In all vaccine vims isolates tested, nt. 472 reverted to the wild type nt. 48 hours after viral replication in the gut of infants that received the live vaccine. Direct evidence for the involvement of nt. 472 in attenuation of neurovirulence came from studies of Westrop et al. (1989). They constructed infectious Chimeric cDNAs between wild type and vaccine strains or wild type, revertant and vaccine strains. After transfection in cell culture to produce infectious vims, virus was injected into monkeys to determine the level of neurovirulence for each poliovirus recombinant. Mutations strongly correlated with attenuation of neurovimlence were located in the 5’ non-coding region (nt. 472) and a single a.a. substitution in VP3 (nt. 2034), which results in a serine to phenylalanine substitution. However, each mutation alone did not have a pronounced affect on attenuation, thus these mutations were additive. To elucidate the biological significance of nt. 472, recombinant viruses 13 were tested fortheirtranslational efficiency in a cell-free translation system (Svitkin et al., 1990). Recombinant viruses containing the attenuated nt. 472 displayed reduced translational efficiency compared to the wild type parent. In addition, recombinants containing the wild type base at nt. 472 in attenuated background displayed higher translational efficiency compared to attenuated virus. Translational efficiencies directly correlated with level of neurovirulence in monkey tests. During analysis of a protease-negative mutant, functional significance of the 5’ non-coding region was elucidated. Andino et al. (1990) determined that interaction of stem-loop structures with 30pro (protease Cleavage product of the nonstructural protein) plays an essential role in RNA replication. Disruption of stem loop structures encompassing the attenuating nt. 472 results in increased temperature-sensitivity (a marker for attenuated poliovirus) indicating that RNA secondary structure in the 5’ noncoding region is important for poliovirus attenuation (Macadam et al., 1992). Biological significance of the single a.a. change in VP3 is less well understood. The same wild type/vaccine strain recombinants described by Westrop et al. (1989) were tested for their ability to induce III-interferon in human leukocytes in vitro (Hovi et al, 1991). Their results indicate that attenuated serotype 3 poliovirus induces less interferon than its wild type counterpart. In addition, induction of or-interferon is directly associated with the VP3 capsid protein. Similar studies were carried out for attenuated serotypes 1 and 2 poliovirus. In contrast to attenuated serotype 3 poliovirus, attenuated serotype 1 and 2 14 poliovirus contain 56 and 22 nucleotide changes, respectively (Christodoulou et al., 1990; Moss et al., 1989). For serotype 1, nucleotide changes correlated with attenuation of neurovirulence are located in the 5’ noncoding region (nt. 480), polymerase, and the capsid region (Christodoulou et al., 1990). For serotype 2, nucleotide changes associated with attenuation of neurovirulence are located within the 5' noncoding region (nt. 481) and the capsid protein VP1 (Moss et al., 1989). Differences in cell culture and in viva passage histories leading to the attenuated phenotype (Sabin and Boulger, 1974), may be related to the differences in the attenuating nucleotides within the coding regions of all three serotypes. The consistency of the attenuating nucleotide in the 5' noncoding region would be expected as this region has the highest degree of homology among all three serotypes (Toyoda et al., 1984). In summary, studies on poliovirus indicate that attenuation of a viral induced disease state can occur by perturbation of viral transcriptional or translational mechanisms and alteration of interactions between capsid proteins and the poliovirus receptor (Moss and Racaniello, 1991). 2. Other RNA viruses. A survey of other RNA viruses containing polycistronic RNA genomes indicates that attenuation by mutations in the 5’ non-coding region is a common theme. Two separate groups have reported sequence data comparisons of virulent and attenuated Japanese encephalitis virus (Aihara, et al, 1991; Nitayaphan at al, 1990). Although their results differ in location of changes within the coding region, both groups found mutations within the 5’ non-coding region of 15 attenuated isolates to be important for reduced virulence. In addition, Pritchard et al.(1992) has reported that a single nucleotide Change located in the 5’ non-coding region of Theiler’s virus attenuates neurovirulence in mice. The Change alters predicted stem loop structures in the 5’ non-coding region. Kuhn et al. (1992) reported that defined mutations in the 5’ and 3’ non- coding regions of Sindbis virus results in attenuation of virulence. They also demonstrated that growth properties of mutant viruses were different in mouse cells compared to chicken or mosquito cells. They hypothesize that cell-specific host proteins which interact with the non-coding regions are different among cell types, resulting in altered growth properties. From in vivo studies, Kuhn et al. concluded that changes in the 5’ and 3’ regulatory regions may have tissue- specific as well as host-specific affects. Aside from alterations in regulatory regions of attenuated RNA viruses, a few mechanisms of attenuation affecting glycoprotein genes have been reported. Sequence analysis of an attenuated Rift valley fever virus indicates that a new AUG codon, upstream of that which initiates translation of a viral glycoprotein precursor, was evident. Translation initiation at this new AUG codon may alter glycoprotein structure (T akehara et al., 1989). Lobigs et al. (1990) attenuated Murry valley encephalitis virus by 10 passages in human cells. Sequence analysis of the E glyCOprotein gene of 6 attenuated isolates revealed an alteration in the RGD sequence (RGD: arginine-glycine—asparginine sequence found in the leader peptide of adhesion molecules) of ng in 5 of the 6 isolates. Since the RGD sequence is highly conserved among flaviviruses, the authors conclude that the 16 RGD sequence is an important general determinant for flavivirus pathogenicity. Thus, there may be a variety of molecular mechanisms of viral attenuation. Most notable are alterations in regulatory regions which may affect RNA transcription or replication and alter the interaction of tissue or cell-specific proteins with these regulatory regions. Mutations within the coding sequence of glycoprotein genes can alter their structure affecting initial virus-host interactions. 3. Vaccinia virus. As with poliovirus, poxviruses are ancient with a 3000 year recorded history of human disease (Volk, 1982). Poxvirus is the largest and most complex animal virus and was the first and only virus to be seen in the light microscope. Human poxvirus causes smallpox, a deadly disfiguring disease characterized by large pox- Iike lesions on the skin. Smallpox vaccines were the first vaccines to be developed. In 1878, E. Jenner used poxvirus isolated from cows to produce a mild form of disease in humans. This mild infection resulted in effective immunization against human smallpox (Review— Fraenkel-Conrat et al., 1988). Since 1977, smallpox is considered extinct due to massive immunization programs using the related vaccinia virus strain directed by the World Health Organization (Volk, 1982). Poxviruses share many common features with herpesviruses: large size, genome structure, and type of genes encoded. Poxviruses contain a linear double-stranded genome of 130 to 300 kbp (Moss, 1990). As with herpesviruses, the poxvirus genome encodes genes for macromolecular synthesis. In contrast to herpesviruses however, poxvirus transcription and virus replication occur in the 17 cytoplasm. The poxvirus genome consists of a unique long region with ends flanked by inverted terminal repeats. Although the unique long segment is highly conserved among poxviruses, the size of terminal repeats can vary from 2.3 to 12.5 kbp (Moss, 1990). Development of poxviruses as vectors for foreign genes and recombinant vaccines has raised concern about retention of virulence within poxvirus vectors, especially vaccinia virus. This concern has prompted researchers to identify virulence factors and attenuating regions within the vaccinia virus genome. Attenuation of vaccinia vims by serial passage in chick embryo cells has two major effects on poxvirus genome structure. As a result of intermolecular recombination, large deletions occur within the unique long region, near the ends (Review-Moss, 1990). In addition, 70 bp tandem repeats within the terminal repeat regions become amplified by tandem reiteration. Amplification by tandem reiteration is thought to occur by homologous recombination between flanking short repeats (Moss, 1981). Presently no phenotypic alteration has been assigned to amplification within terminal repeat regions. Site-directed deletions within the unique region of the vaccinia virus genome have been characterized with respect to attenuation of pock formation in mice (Lee et al., 1992). Most notable, mutations or deletions in the ribonucleotide reductase gene or the hemagglutinin gene resulted in attenuated phenotypes. Deletions in the thymidine kinase gene did not affect pock formation with some deletion mutants, but other mutants showed decreased ability to form pocks. This discrepancy could have been due to strain differences among mutants. An 11 kb ‘1. l 18 deletion, located between the host range determinant gene and the ribonucleotide reductase gene, resulted in the complete absence of pock formation. However, the mutant virus grew to parental virus titers, indicating a highly attenuated phenotype. In addition, a 20 kb deletion located at the left end of the unique long region greatly reduced pock formation. Genes encoded within the 11 kb and 20 kb deletions are presently unknown (Lee et al, 1992). Virulence determinants of vaccinia virus, proposed by other investigators have been located in the VVGF (a protein that binds epidermal growth factor), thymidine kinase, and a serine protease inhibitor homolog (located within the right inverted terminal repeat region) (Tyler and Fields, 1990). Although poxviruses were the first viruses identified and used for vaccines, the molecular mechanisms of attenuation are difficult to determine due to its extremely large size. 4. Herpesviruses. As with vaccinia viruses, study of herpesvirus attenuation is difficult due to their large genome size and the many genes they encode. Presently, studies related to attenuation of herpesvirus virulence are carried out using defined mutants (HSV, PRV) and attenuated vaccine viruses (PRV). Some recent findings will be discussed in this section. Most reports, with one exception, are at the level of identifying genes involved in virulence. One paper (Robbins et al, 1984) reports results which hint at a possible mechanism of attenuation. Pseudorabies virus, an alphaherpesvirus which infects swine, causes an infection similar to that of Herpes simplex virus. Due to its economic importance, 19 an attenuated PRV was developed and is currently used for vaccination of swine in Europe (Card at al., 1992). Virulent PRV are neurotropic and produce lytic Infection in the central nervous system of swine. Attenuated PRV is also neurotropic, but does not cause pronounced lytic infection. The mechanism of attenuation remains to be elucidated, but evidence indicates that alterations of glycoproteins in the viral envelope may be important. The genome of attenuated PRV contains mutations in the UL and Us regions. A large deletion in the Us has removed coding sequence for gl and gp63 genes. These two glycoproteins form a complex required for neurovirulence. Mutations in the UL region effect glll and the BamHl-4 fragment which encodes 4 genes thought to be involved in capsid assembly. Functional significance of these mutations has been investigated. Mettenleiter et al. (1987a, 1987b) reported that absence of 91 affects release of virus from rabbit kidney cells and plays a major role in neurovirulence. Robbins et al. (1989) investigated the significance of mutations in the gill gene of attenuated PRV genomes. To eliminate the contributions of other deletions in attenuated PRV, they constructed recombinants with the glll gene from attenuated PRV in a wild type background. They showed that, although steady state RNA levels were the same with attenuated glll as compared to wild type, protein levels from in vitro translation and purified virions were greatly reduced. However, levels of glll from immunoprecipitations of infected cell lysates were the same for attenuated glll as compared to wild type. Sequence analysis indicated that a nucleotide change in the attenuated glll signal sequence caused a change from leucine to proline. The authors conclude that mutation in the signal sequence 20 resulted in improper protein localization, and that glll accumulates in the cytoplasm of infected cells instead of being inserted into the virus envelope. Card at al. (1992) constructed a series of recombinant viruses containing deletions in the glll plus gl genes, glll alone, or gl alone. These mutants were then tested in viva using the rat visual neuronal system. Patterns of infection and specific cell types infected were able to be followed in this system. glll-negative mutants did not attenuate neurovirulence, however gl-negative mutants were substantially attenuated for neurovirulence. glll-negative plus gI-negative mutants were more attenuated than the single gene mutants. In addition, they determined that gl-negative mutants were unable to infect dorsal root ganglia but were able to infect retinal ganglia. Also, glll was not required for infection of either cell type. The authors conclude that gl has a significant impact on neurovirulence and PRV glycoproteins confer cell type tropism to the virus. gl, glll and gp63 are functional homologues of gE, g0 and gl, respectively of HSV. Interestingly, these proteins are non-essential for growth in vitra, but seem to be major virulence determinants in viva, at least for PRV. Studies to define genes responsible for HSV virulence have used defined mutants and clinical isolates. The thymidine kinase gene was considered to be involved in neurovirulence and latency reactivation. However, some investigators argue that thymidine kinase mutants still retain neurovirulence. (Review— Fields, 1990). In the mouse model, additional genes which attenuate neurovirulence are gE, gG, IE22, ribonucleotide reductase, and protein kinase. Recently, Robertson et al. (1992) have reported on a neurovirulence gene located in the terminal 21 portion of the long repeat region. A mutant containing a 759 bp deletion in this gene fails to grow in dorsal root ganglia, but is capable of replicating in the foot pads of mice. These results suggest that the NV gene confers cell or tissue specificity. In addition, the NV-negative mutant was able to establish latency but was impaired for its ability to reactivate. Another recent report has implicated the nan-essential gene, dUPTase in being a neurovirulence determinant in the mouse model. dUTPase mutants were able to replicate in the peripheral nervous system and establish latency, but reactivated from latency with reduced efficiency. (Pyles et al, 1992). As with PRV, HSV non-essential genes in vitra seem to be important for virulence in viva. Ill. HISTORY OF VIRUS INDUCED CANCERS. To prove the etiology of a disease, rules were adapted from postulates of the late 19th century scientist, Robert Koch. Koch's postulates are summarized below (Review— Volk, 1982): 1. The same organism must be found in all cases of a given disease. 2. The organism must be isolated and grown in pure culture from the infected patient. 3. The organism from the pure culture must reproduce the disease when inoculated into a susceptible animal. 4. The organism must then be isolated in pure culture from the experimentally infected animal. 1. RNA viruses. The role of viruses in the etiology of cancer was not appreciated until Rous sarcoma virus (RSV), a single stranded RNA virus was shown to cause cancer in experimental chickens. In 1911, Peyton Rous was able to induce tumors in 22 Chickens by injecting filtered extracts of tumor tissue isolated from a cancerous chicken, thus establishing that a virus was the etiologic agent. After Rous’s original report, many other RNA viruses were similarly suspected of playing a role In cancer (Review— Weiss, 1984). The discovery that a virus can induce cancer lead to speculation that the virus may encode a gene responsible for tumor induction. It was not until 1970 that Martin reported that there was a genetic basis for tumor induction by RSV. This viral genetic component was found to be a single gene, src, encoded within RSV. Since then, many viral oncogenes have been identified in RNA viruses (Review— Bishop, 1987). Most of these oncogenes are structurally and functionally similar to transcription factors, receptors, hormones, or metabolic messengers (Bishop, 1987). The final outcome of an active oncogene in a cell is stimulation of cells to divide indefinitely. 2. DNA tumor viruses. The involvement of DNA viruses in the etiology of cancer began with the demonstration that papillama virus, using similar techniques of Rous, is the musative agent of warts (Shop, 1932). Since then other DNA viruses (palyoma virus, simian virus 40, adenoviruses, JC, and BK) have been blamed as the culprits of cell transformation. The viral proteins implicated in cell transformation by polyoma and SV40 are involved in stimulating viral DNA replication (T -antigens) (Grodzicker and Hopkins, 1981). Similarly, regions of viral DNA associated with transformation have been identified in adenovirus (Flint, 1981). 23 3. Herpesviruses. The involvement of herpesvirus in cancer was difficult to establish. The first hint that herpesviruses were associated with tumors came from studies by Fawcett in 1956. Using electron microscopy, Fawcett was able to prove the existence of herpesvirus (Lucké virus) particles in tumor tissue of frogs injected with filtered extracts of naturally occurring tumors (zurHausen, 1981 ). However, some tumors were free of any signs of herpesviruses. Thus, attempts to satisfy Koch's first postulate were unsuccessful. Herpes simplex virus I and II are associated with cervical carcinomas, although the evidence is indirect. Many women suffering from cervical carcinoma have antibody to HSV ll. However, virus particles or viral DNA in tumors have not been reproducibly detected (Spear and Roizman, 1981). The ability of a certain restriction endonuclease fragment to transform cells in vitra has been reported (Camacho and Spear, 1978). CMV is also associated with human cancers in a similar fashion as HSV. Patients suffering Kaposi's sarcoma have high antibody titers to CMV, however CMV virions or DNA have never been detected in tumor tissue (zurHausen, 1981). Due to lack of an appropriate in viva model system for human herpesviruses, testing of Koch’s postulates is impossible. Other herpesviruses which infect primates and are considered oncogenic are H. Saimiri and H. Ateles. However these viruses are innocuous to their natural host, but highly oncogenic to other primate species (zurHausen, 1981). The human herpesvirus which is most strongly implicated as a causative 24 agent in cancer, and has prompted the most concern, is Epstein-Barr virus (EBV). EBV was initially isolated from cultured Iymphoblastoid cells which originated from a Burkitt’s lymphoma patient (Epstein et al, 1964). In addition, high antibody titers to EBV are found in Burkitt’s lymphoma patients. Burkitt's lymphoma is a relatively rare cancer occurring in children 5-12 years old only in certain regions of East Africa (Burkitt, 1962) and New Guinea. Data which prompted concern was that antibodies to EBV were found in 80% of healthy adults all over the world (Henle and Henle, 1966). Although many investigators have attempted to find direct evidence for the link between EBV and human cancers, the data remains only circumstantial. However this Circumstantial evidence is strong: human cells can be transformed by EBV in vitra; EBV can induce lymphoid tumors in non-human primates; Most but not all Burkitt's lymphoma and nasopharyngeal carcinoma cells carry EBV DNA. Again, attempts to satisfy Koch’s postulates failed. Thus far, all herpesviruses discussed are associated with some form of cancer or tumor formation, however testing of Koch's postulates has failed or cannot be done in humans. Recovery of herpesvirus from only some cancer cells may be due to coincident infection with oncogenic viruses. Failure to find virus particles or viral DNA in some human tumor cells may be related to the characteristics of herpesvirus latency. Herpesvirus particles are absent during latency, and very few copies (undetectable) of herpesvims DNA are present in latently infected cells. In addition, some herpesviruses may transform a cell by a "hit and run" mechanism (zurHausen, 1981). The idea that herpesviruses can cause cancer was proved by three studies. In 1967, Churchill and Biggs isolated 25 a herpes-like virus in cell culture and subsequently injected the virus infected cells into Chickens. The experimental chickens developed lymphoid tumors (acute Marek's disease). In 1970, development of Marek’s disease from injection of purified cell-free MDV virions into experimental chickens also lead to MD (Nazerian and Witter, 1970). Additional proof came with the development of effective vaccines against Marek's disease. These were pivotal observations for the involvement of herpesviruses in the etiology of cancer. IV. MAREK’S DISEASE VIRUS. 1. Pathology of Marek’s Disease virus. Clinical signs of Marek's Disease (MD) were firstidescribed by Joseph Marek in 1907 (Payne, 1985). There are two distinct pathological forms of MD: classical and acute (Payne, 1985). The classical form of MD is characterized by impairment of neural function and cytolytic infection. The acute form of MD is marked by lymphoproliferation and tumor formation and includes classical form symptoms. A Chicken infected with MDV displays visible signs of nerve function impairment, beginning with partial and ending with complete paralysis. A common characteristic of Classical MD is uneven gait or one leg stretched forward and the other leg stretched backward. Closed eyelids, lameness, and droopy wings are also characteristic of nerve involvement (Purchase, 1985). This Classical form of MD was considered the most common in the field, historically. Presently, there is a higher prevalence for the acute form of MD (Payne, 1985). In some cases the bird recovers, however in most instances, MDV infects lymphoid tissue and I“- . '0', 2.1 a. In. 5‘ I.» I 26 lymphocyte proliferation leads to the acute form of MD (Purchase, 1985). In addition, the virus becomes latent in T-Iymphocytes, persisting throughout the bird's lifetime (Schat, 1985). The cellular pathway to MDV transformation of T- lymphocytes is summarized in Figure 1 and described in the following paragraph. MDV is contracted from the environment via the respiratory system. MDV travels from lungs to lymphoid tissue by a yet undetermined mechanism. The first evidence (3-6 days post inoculation) of virus infection is in lymphoid organs (spleen, thymus, and bursa) (Payne, 1985). Primarily B-lymphocytes, and a few T-lymphocytes, are infected (Calnek et al, 1984). Initial infection of lymphoid tissue is cytolytic in Character but does not result in the release of cell-free virions. This "semi-productive” infection is marked by the presence of viral internal antigens (VIA) and naked virus particles (without envelopes) within lymphoid tissues (Payne, 1985). Tissue Changes which accompany this initial infection are infiltration of macrophages and granulocytes and reticular cell hyperplasia. The net effect is enlargement of the Spleen and atrOphy of the bursa and thymus (Payne, 1985). 27 MDV INFECTED CELL . A. ' 2 - CYTOLYTIC PHASE LATENT PHASE fa" a-T :- Al" “'12 (3:: Figure 1. Pathway to MDV transformation of T-lymphocytes. 28 From 5 to 7 days post infection, the type of infection changes from cytolytic infection of primarily B-lymphocytes to latent infection of predominantly T- lymphocytes. In cells latently infected with MDV, VIA are absent, Marek's associated tumor surface antigen (MATSA: activated T-Cell marker) is present, and virus can be recovered only after co-cultivation with permissive cells (McCall et al, 1987; Powell et al, 1974). Evidence for antibody and cell-mediated immune response is found coincident with the switch from predominately cytolytic to predominantly latent MDV infection (Sharma, 1985). About 9 days post infection, transient immunosuppression occurs, but after a 2 week recovery period, permanent immunosuppression ensues (Payne, 1985). Latently infected T- Iymphocytes are disseminated via the Circulatory system and localized cytolytic lesions are present in various organ and tissue systems during the immunosuppression recovery period (9 days to 3 weeks) (Payne, 1985). Infiltration of macrophages and lymphocytes in peripheral nerves ultimately leads to demyelination and impairment of nerve function. Only in feather follicle epithelium does the cytolytic infection result in release of cell-free virions (Witter, 1970). The final outcome of acute MDV infection is transformation of T-lymphocytes and infiltration of transformed cells into visceral organs, skin, muscle and nerves, establishing tumorous masses which disrupt organ function. Severity and tissue distribution of tumors of acute MD depends upon MDV strain and genetic susceptibility, age, sex, and immune status of the host chicken (Payne, 1985). Infection of turkeys with MDV also results in lymphoproliferative lesions, with similar pathology as chickens. However, tumor cells are predominately of the B- 29 cell origin (Nazerian and Silva, 1988) and only some strains of serotype 1 MDV are able to establish infection (Elmubarak, 1981). 2. Biology of MDV. The herpesvirus responsible for development of MD was initially isolated in chick kidney cell cultures (Churchill and Biggs, 1967). They described a highly cell-associated herpes-like virus particle in cells displaying cytopathic effects (CPE: cellrounding and syncytia formation). Evidence from electron microscopy studies by Nazerian and Bunnester (1968) showed icosahedral, herpes-like particles 95- 100 um in diameter within the nucleus. However, enveloped viruses within the cytoplasm were found at low frequency and not detected in culture media. The cell-associated nature of MDV in vitra is most likely related to the absence of enveloped virions within cells or growth fluid. MDV is classified into 3 serotypes based primarily an antigenic similarities between virus isolatesand ability to form lymphoid tumors in susceptible Chickens. Serotype 1 MDV strains are oncogenic whereas serotypes 2 and 3 are non-oncogenic (Payne-review, 1982). Serotype 3 MDV or herpesvirus of turkeys (HVT) was originally isolated as an innocuous virus from turkeys (Witter, 1970). Serotypes 2, 3 and attenuated serotype 1 strains have been used as effective vaccines against oncogenic serotype 1 MDV (Payne, 1982; Witter, 1983). Nucleotide sequence analysis and comparison of banding patterns generated by restriction endonuclease digestions of MDV DNA indicates that MDV genome structure differs among serotypes and within each serotype (Coussens et al., 1989; 30 Hirai et al., 1979; 1981; 1984; lgarashi et al., 1987; Ross et al., 1984; Silva and Witter, 1985). Serotype 1 strains are further subdivided based upon severity of disease they cause. Vegy highly oncogenic strai_ns_ (very virulent: v.v.) are oncogenic in genetically resistant and HVT- vaccinated birds. This group also causes acute lethal cytolytic infections in MD susceptible birds. Highly oncogenic m (acute) cause a high incidence of visceral and neural tumors in genetically susceptible but low incidence of tumors in genetically resistant birds. Moderately to mildly oncogenic strains (classic) induce low incidence of mainly neural MD in susceptible birds. Minimally oncogenic strains cause minimal lesions only in very susceptible birds. Serotype 1 strains of varying pathotype are not distinguishable using conventional or monoclonal antibody methods. Serotype 1 pathotypes have been distinguished by their ability to induce immunosuppression: highly oncogenic strains are more immunosuppressive than less oncogenic strains (Witter et al, 1980). In addition, cell lines established from less oncogenic strains contain more viral antigen than cell lines derived from highly oncogenic strains (Review- Schat, 1985) The tightly cell-associated in vitra growth characteristics of MDV (Schat, 1985) have hampered progress on the molecular analysis of MDV genomes. Separation of MDV DNA from host cell DNA by density-gradient centrifugation results in extremely low yields of pure MDV DNA (0.1 to 10 ug MDV DNA per 10" infected cells). Often, these preparations are contaminated with host cell DNA (Kato and Hirai, 1985), as buoyant density of MDV DNA (1.705 g/cm’) (Lee et al., 31 1971) is similar to that of host cell genomic DNA (1.700 g/cm"). In addition, physical integrity of MDV genomes is often compromised during conventional isolation procedures. The cloning and restriction mapping of MDV and HVT DNA (Fukuchi el at, 1987: lgrashi et al, 1988) have greatly enhanced our ability to analyze the molecular biology of MDV and HVT. The BamHl restriction maps of MDV and HVT are presently used as standard references for location of genes on the MDV or HVT genomes (Figure 3). In addition, adaptation of pulsed field gel electrophoresis technology to the study of MDV has facilitated detection of novel Changes in genome structure of various MDV isolates (lsfort et al, 1990; This Dissertation Ch. 3). 3. MDV genome structure. Total size of the MDV genome, derived from contour mapping electron microscopy or sedimentation values of MDV DNA, was determined to be 180 kb for serotype 1 MDV and 160 kb for HVT (Cebrian et al, 1982; Hirai et al., 1979; Lee et al., 1971). Structure of the MDV genome was elucidated using electron microscopy of partially denatured DNA (Cebrian et al .1982). These studies revealed a structure closely related to HSV I (type E genome structure), containing a long unique region (UL), and a short unique region (Us). Both UL and Us are flanked by terminal repeat (T RL and TRs) and internal inverted repeat sequences (IRL and ms) (Figure 2). Cebrian et al (1982) also compared the unit length of MDV DNA regions with that of HSV and HVT. In all regions, MDV DNA was larger than HSV or HVT DNA, with the exception of the Us. MDV Us is 3 kb smaller than the Us of HSV, but 3 kb larger than the Us of HVT. The MDV TRL is 7.6 kb and 32 6.1 kb larger than HVT and HSV, respectively. The MDV UL is also 7.6 kb larger than UL of HVT or HSV. In addition, the MDV TRs is 1.5 kb and 6.1 kb larger than the TRs of HVT and HSV, respectively (Figure 2). Evidence for gene co-Iinearity on the MDV genome compared to a-herpesvirus genes was supplied by Buckmaster et al. (1988). Using random sequence analysis, Buckmaster et al. localized 35 putative VZV homolog genes in a co-linear array on the MDV genome. In addition, the entire Us region was completely sequenced by Brunovskis and Velicer (1992) and they have determined that MDV Us genes which are homologous to HSV genes are arranged in a colinear fashion with the HSV genome. Brunovskis and Velicer also identified 3 open reading frames (SORF 1-4) which are unique to MDV and show no homology to existing herpesvirus Baum 3958 0.88% >92 P 8296.... 2 moufi EcEmom 0.89.8 >w... can ._.>_.. .o comcmano a 2.6... 9. to- 9. a.a.. 9. ms- 9. 5. >9. 9. m. 7 9. Rm- 9. ms- 9. ms- 5... HUI—“HI I >92 mm» m: a... .m. .2 .5. 09 m2. amp m9 our mop om mm om me 9. o 34 sequences in current databases. They also identified a fowlpox virus homolog (SORF 2) (Brunovskis and Velicer, 1992). Total size of the 3 SORFs and the fowlpox homolog nearly equals the difference in size of the MDV Us compared to the HVT Us. Proteins encoded within these SORFs may be important in the Characteristic pathological differences between MDV and HVT. 4. MDV Proteins: general Information. There are at least six antigenically active viral proteins found in all three serotypes of MDV (lkuta et al., 1983; Silva and Lee, 1984). The genes and protein products of two of these antigens, MDV A (90) and B (98) have been extensively Characterized. MDV 96 and 98 are homologous to herpes simplex virus-1 (HSV- 1) glycoproteins C and B, respectively (Binns and Ross, 1989; Buckmaster et al., 1988; Chen and Velicer, 1991b; Coussens et al., 1990; Ross et al., 1989). In HSV-1, 90 is non-essential for viral replication in vitra whereas gB, highly conserved among alphaherpesviruses, is required for virus attachment to host cells (Roizman, 1991). MDV 90, an abundant viral glycoprotein antigen, is found in sera of infected birds, infected cell culture fluids, and to a lesser extent associated with the plasma membrane of infected cells (lsfort et al., 1986). Evidence from pulse-chase experiments and tunicamycin (inhibits protein glycosylation) treated cells suggests that MDV 9C is synthesized on the endoplasmic reticulum and processed through the Golgi via two pathways. Secretion of N-linked glycosylated MDV 90 occurs through a major pathway while an unglycosylated form is secreted through a minor pathway (lsfort et al., 1986). Approximately 95% of serotype 1 MDV go is secreted while the remaining 5% is associated with plasma membranes 35 of infected cells (lsfort et al., 1986). In comparison, MDV 9B is found primarily in infected cell Iysates and the glycoprotein processing pathway is similar to that of other herpesvirus gB homologues (Chen and Velicer, 1991b; Ross et al., 1989; Velicer et al., 1978). The gene for one MDV IE protein, MDV ICP4, has been localized to the MDV genome (Anderson et al 1992). The most extensively characterized genes are illustrated in Figure 3. Regions of the MDV genome implicated in playing a role in MDV pathogenesis are all, with one exception (90), located within the IR and TR, the most divergent region of the MDV genome as compared to or- herpesvirus genomes. These genes will be discussed in the next section. 5. Molecular biology of MDV oncogenicity. A variety of approaches have been utilized to identify regions of the MDV genome which may encode genes important in MDV-induced disease processes. One approach is comparison of proteins and gene sequences from non-oncogenic serotypes 2 and 3 to oncogenic serotype 1. Another approach is analysis of transcripts and their products in MDV transformed lymphoblastoid cell lines. Yet another approach is comparison of proteins, genome structure, and transcription patterns of oncogenic strains to their attenuated, non-oncogenic counterparts (the subject of section 7). Serotype comparisons and MDV transformed cell lines will be addressed in separate sections below. .mEocoo >05. 2.. co mecca >92 3 5.88.. .n 2:2“. Q... I a.E-.§.3=§9_3 we: can: . _ _ . I . .zz _ _ _ Soonefi s.\. I, N a 33' \‘\\ X. o _ 2.... ll \\.\ I \ I \\ I \\ 6 II iii 3 II \.\\ I .\\.. I 0. ” \\.\ llmll .. . I am; 3 mm. .m. a .E. 2.0. or; On an Iv IXIr Ivhv p __ _ : _ b L__:: :_ _ __ _ _ b. _ _ : b : . _ __ q 2 d _ «_~d: :4 _ __ _ d _ d A q : _ ._J t .n. .{< .7... z v.2_z/ m i..:/m _/~v. n. o o._ 0 ~__. \ 3...... «o m. kmm .O A.... a. «.3. o No \ ll. \ I’ll. —:Eam .. 8:8 m: \S III}! S ll AAA A A v v A A A v mo .8 on 3E8 v... «m: numOm 2m: 5: «ED» {mow 37 a. Serotype comparisons. 96. Although 90 is antigenically related among serotypes, serotype 2 9C is lower molecular weight than serotypes 1 or 3 and serotype 3 9C is more alkaline than serotype 1 90 (vanZaane et al, 1982). Sequence data comparison of serotype 1 and 3 90 genes indicates overall 66% and 81 % homologies at the DNA and protein levels, respectively. However, the 5’ coding and flanking regions are quite divergent (Coussens et al, 1990). Differences in the 5' regions between serotypes 1 and 3 would suggest a possible role in different pathologies of Serotype 1 and 3. Thus, it would be of interest to further investigate the divergent 5’ regions in relation to MDV pathogenesis. pp38. By comparison of proteins among serotypes that are immunoprecipitated with a panel of monoclonal antibodies, Silva and Lee (1984) identified serotype 1 specific phosoproteins of 44, 38, and 24 kd. Subsequently, the gene encoding the 38 kd phosphoprotein (pp38) was localized to the UleL junction of the MDV genome (Chen and Velicer, 1992; Cui et al, 1991). pp38 is found predominantly in the cytoplasm of infected cells and is not considered a DNA binding protein (Cui et al, 1991). In addition, pp38 produces only non- neutralizing antibodies in viva (Nazerian et al, 1992). Also, pp38 may be an IE or E protein based upon its expression in the presence of phosphonoacetic acid (Chen and Velicer, 1992). Data by Pratt et al (1992) suggests that pp38 may be transactivated by the MDV ICP4. Because pp38 is expressed in MDV transformed cell lines and MDV-induced tumor lesions in viva, pp38 is implicated as playing a role in MDV-induced oncogenicity (Chen and Velicer, 1992; Cui et al, 1991). . an- 38 However, pp38 is also expressed in cells lytically infected with oncogenic as well as attenuated MDV strains (Chen and Velicer, 1992; Silva and Lee, 1984). In addition, the original pp38 monoclonal antibody, H19, was generated from spleens of mice that were injected with the attenuated Md11 strain Md11/750 (Lee et al, 1984). Thus, pp38 may not be the transforming protein of MDV, but may be involved in a multi-step process leading to transformation. Clearly, further investigation on the functional role of pp38 is required. b. MDV transformed cell lines. MDV tumor cell lines have been established by cultivation of MDV lymphoma cells in vitra. Most cell lines characterized thus far are activated T- lymphocytes (Schat et al, 1991). Virus can be rescued from most cell lines by co- cultivation with susceptible primary cultures. After prolonged in vitra cultivation, the ability to rescue virus is lost. These cell lines are Classified as non-producer cell lines. Viral antigens are not detectable in non-producer cell lines. Producer cell lines are defined as cells from which virus can be rescued after co-cultivation. Producer lines are further subdivided as expression or non-expression based on amount of viral antigen produced. 5-iodo-2-deoxyuridine (IUDR) treatment of non- expression cell lines induces expression of viral antigen (Review-Schat, 1989). MDNA. A 28 kd nuclear protein (MDNA) was found to bind to specific sequences located in the TRs and Us regions of the MDV genome (Wen et al, 1988). MDNA was not expressed in lytically infected cells, but was expressed in two producer cell lines (an expression and a non-expression cell line). Although 39 the authors conclude that MDNA is analogous to EBNA-1 of EBV (EBV nuclear antigen that binds near the origin of replication in latently infected cells), two important experiments are lacking. First, binding specificity was not conclusively shown. Second, no evidence was given that MDNA is a viral-encoded protein. In addition, it is questionable whether the cell lines used were truly latently infected. Perhaps MDNA is a lymphocyte specific factor, expressed in activated lymphocytes. meq. Analysis of viral specific RNA expression in producer and non- producer cell lines revealed abundantly expressed transcripts located within the IRL of the MDV genome (Tillotson et al, 1988). In addition, these transcripts were not expressed in lytic infections of attenuated MDV strain Md11/75C but were expressed at low levels in virulent MDV strain JM infected cells (T illotson at al, 1988; Jones et al, 1992). The IRL transcripts were sub-localized to the EcoRI Q fragment (spans the BamHl-I2 and BamHl-Q2 fragments). Sequence analysis of the EcoRI O fragment revealed a putative 40 kd protein containing a leucine zipper repeat and an upstream domain rich in basic amino acids, suggesting a protein similar to the fos/jun family of transcriptional activators. Antiserum raised against a synthetic peptide corresponding to the leucine zipper region was used to detect a 40 Rd protein in MDV transformed cell lines, but not lytically infected MDV strain GA cells (Jones et al, 1992). Further work is required to determine if meq (Marek's EcoRl O) is a direct acting oncogene, as its physical structure would suggest. Transcriptional activity In MDV transformed cell lines. Because MDV 40 cell lines are established directly from tumor cells , transcriptional activity in MDV transformed cell lines is of interest. Three groups have investigated transcription patterns in various MDV transformed cell lines. Maray et al (1988) found transcriptional activity dispersed throughout various regions of the MDV genome in MSB-1 cells (a producer, expression line). Schat et al (1989) compared transcriptional activity in a variety of MDV transformed cell lines. This group also found transcriptional activity dispersed throughout various regions of the MDV genome in an expression (producer) cell line, but non-producer and non- expression cell lines expressed only a limited number of immediate early transcripts from discrete genomic regions. These IE transcripts are clustered in the IRL (BamHI-Is and -L fragments), le (BamHl-A fragment) and the unique short (non-expression only, BamHI-P1). After treatment with IUDR, transcripts which hybridize to a BamHl-H probe (also in IRL) were detected in non-producer cell lines. Similar treatment with IUDR of non-expression cell lines yielded transcripts located in BamHl-H as well as other transcripts dispersed throughout the MDV genome. In addition, a 0.6 kb IE transcript located within BamHI-L is truncated by 0.3 kb in the non-producer cell line. Sugaya et al (1990) also investigated transcription patterns in MDV transformed cell lines. Using a non-expression cell line, they found transcripts in the same regions as Schat et al (1989). However, in this particular cell line, they detected additional transcripts located within the BamHI-H region. The presence of MDV transcripts originating from a limited region of the MDV genome is reminiscent of transcription patterns in transformed cells latently infected with EBV 41 (Schat et al, 1989). Comparison of transcription patterns in various cell lines suggests that MDV infection is frozen in different stages of latency. Thus, non- producer cell lines may be considered latently infected, containing an aberration in MDVs ability to reactivate. Whereas, MDV in expression and non-expression cell lines may be frozen in earlier stages of latency. Gene products from the IR._ and IRS may play a role in the initiation and maintenance of MDV transformation, and/or initiation, maintenance, and reactivation from latency. 6. MDV vacclnes. Prior to the development of vaccines against MDV, 20 to 30% mortality rates in commercial flocks were common (Pattison, 1985). In 1970, vaccination against MDV using HVT was introduced (Witter, 1970). During the 12 year period following HVT vaccination, from 1970 to 1982, national condemnation of young chickens due to MD declined by 95% (PurChase, 1985). Total cost to the US. poultry industry for vaccination is $44.4 million dollars/year. Although MD vaccines have been effective at reducing mortality and condemnation rates, there continues to be occasional outbreaks or vaccine breaks, which contributes further cost to the poultry industry. Economic loss in the US. due to mortality, condemnations, and loss of egg production is $115.9 million year. In 1984, total worldwide cost of MD was estimated at $943 million a year. Thus, control of MD is of economic importance to the poultry industry (Purchase, 1985). The most widely used MDV vaccines in the us. and other countries are cell-associated and cell-free HVT preparations (Pattison, 1985). HVT produces no 42 clinical symptoms in chickens or turkeys (Payne, 1985). In turkeys, HVT is highly contagious. ln chickens, however, spread of HVT by contact is minimal (Payne, 1985). Antigenic cross-reactivity of HVT and MDV antigens prompted its development as a MD vaccine (Pattison, 1985). Attenuated mildly virulent serotype 1 vaccines have been widely used in The Netherlands (Witter, 1992a). This type of vaccine readily spreads by contact, but spread does not occur rapidly enough to provide passive immunization of hatchmates (Witter, 1992b). Attenuated mildly virulent serotype 1 vaccines have not been used extensively due to their residual pathogenicity. Serotype 2 vaccines have also been used as vaccines against MD (Witter, 1992a). The serotype 2 strain in current use, SB—1, spreads readily by contact and is efficacious against most virulent MD strains (Review- Witter, 1985). Recombinant MDV vaccines are presently being developed and two vaccine candidates are worth noting. First, a fowlpox vaccine expressing MDV 98 was 100% effective against challenge with GA and M05 (virulent and v.v. MDV strains, respectively), but only 90% protective against challenge with v.v. MDV strain RBIB (Nazerian et al, 1992). Second, a bivalent recombinant vaccine against MDV and Newcastle disease virus (NDV) was developed by Morgan et al. (1992). These HVT recombinants expressing two vimlence determinants of NDV gave 100% protection against NDV. However the vaccine was only 64-76% protective against challenge with v.v. MDV strain RBIB (Morgan et al, 1992). Since 1988, the application of a bivalent vaccine, a combination of 88-1 and HVT, has increased in the US (Witter,1992). The bivalent vaccine efficacy rivals 43 or exceeds that of HVT alone (Pruthi et al, 1989). Although this vaccine is extremely efficacious, concern was raised when Bacon et al. (1989) reported that SB-1 could augment lymphoid Ieukosis (avian retrovirus induced disease) in white leghom chickens. The mechanism behind this possible interaction between MDV and retroviruses was suggested to occur via MDV transactivator proteins (T ieber et al., 1990). 88-1 can enhance retroviral replication in vitro and retroviral transformed B-cells can harbor SB-1 DNA after co-infections in vivo (Pulaski et al., 1992; Fynan et al, 1992). In addition, retroviral integration into the MDV genome has been shown to occur in vitro (lsfort et al., 1992). Presently, there is little concern for these interactions occurring under field conditions due to the low level of lymphoid Ieukosis in commercial flocks (Witter, 1992a). However, published reports on the interaction of two viruses producing an undesirable outcome in vitro and in vivo are extensive. Possible interference with vaccine efficacy is illustrated by reports on PRV. Recombination of two PRV vaccine strains during replication in swine resulted in vimses with restored virulence (Katz et al., 1990; Henderson et al, 1991). During the course of vaccination against Marek’s Disease, some flocks experience the emergence of MD characterized by a more severe pathotype. These v.v. MDV strains are associated with, and considered to be the cause of, MD outbreaks in vaccinated flocks (\Mtter, 1992a). During the first quarter of 1992, Delaware experienced a 75% increase in incidence of MD over the previous year. Assuming vaccination and management practices have not changed, this increase may be due to the emergence of v.v. MDV. 44 The increasing virulence of v.v. MDV has prompted attempts to develop vaccines against v.v. MDV (Witter, 1992). These attempts have met with limited success. Residual pathogenicity is associated with attenuated, mildly virulent, serotype 1 vaccines. One vaccine, CV1988/Rispens, was found to be the most effective vaccine against v.v. MDV with efficacy of only 80-85% (Witter, 1992b). CV1988/Rispens has been used with success in many countries, but is not currently licensed in the US. The possibility of providing better protection against v.v. MDV strains prompted development of an attenuated vaccine using v.v. MDV strain Md11 (Witter et al, 1982). MDV strain Md11 was originally isolated from an HVT- vaccinated commercial flock experiencing vaccine breaks (Witter, 1983). Md11 is characterized by inducing severe early bursa and thymus atrophy, lymphoid cell necrosis, and early death in the absence of lymphomas or nerve enlargement (Payne, 1985). The Md11/750 vaccine provided 69% protection in chickens that did not have homologous maternal antibodies. However, vaccine protection in chickens with homologous maternal antibodies was reduced to 14% (Witter, 1984). A trivalent test vaccine containing HVT, 88-1, and Md11/75C was 100% effective against challenge with vimlent and v.v. MDV (Witter, 1984). Practical considerations on use of this trivalent vaccine, such as growth and preparation of trivalent vaccine mixtures and the residual pathogenicity associated with the Md11/75C component, have most likely prevented its use by the poultry industry. Due to the increasing prevalence of v.v. MDVs isolated from vaccinated flocks, development of vaccines against v.v. MDV is a priority concern (Witter, 45 1992a). However, basic understanding of the attenuation process, and genes involved in virulence would aid in designing effective strategies for vaccine development. Different virus strains become attenuated at different rates or passages (Schat, 1985b). MDV strain Md11 is only partially attenuated at passage 52, further passage reduces in vivo lesions, but attenuated Md11 has an associated residual pathogenicity of approximately 2% (Ch. 4, this dissertation). This may be due to the presence of oncogenic sub-populations. The continued and increasing frequency of vaccine breaks makes understanding virus attenuation a priority. 7. Attenuation of MDV. Attenuation of serotype 1 MDV occurs after 30 - 55 serial in vitro passages (Schat et al., 1985b). Coincident with loss of oncogenicity in viva, attenuated MDV fails to induce early cytolytic infection, have reduced ability to infect or replicate in lymphocytes, and lose the ability to spread by contact, as compared to oncogenic MDV (Schat et al., 1985; Witter,1984). in vitro studies indicate that attenuated MDV grow faster in cell culture, are able to infect a wider variety of cultured cells, display reduced expression of glycoprotein C expression, and contain altered genome sthcture and sizes compared to oncogenic MDV (Nazerian, 1979; Schat, 1985; Review— Ross, 1985; Wilson and Coussens, 1991). a. Genome structure. Comparison of genome structure of attenuated isolates to their oncogenic counterparts revealed gross changes in limited regions of the MDV genome (Hirai 46 et al., 1979; 1981; 1984; lgarashi et al., 1987; Ross et al., 1984; Silva and Witter, 1985). These changes were attributed to expansion, by tandem reiteration, of a 132 bp repeat region located within the TRL and IR, (BamHI-D and -H fragments) of the serotype 1 MDV genome (Maotani et al, 1986). Although a protein product encoded within this region has yet to be identified, transcripts encompassing the 132 bp repeat region have been intensely studied. Comparison of transcripts located near the 132 bp repeats in oncogenic and attenuated MDV indicated that a 1.8 kb transcript family becomes tmncated as a result of in vitro attenuation (Bradley et al, 1989). Nucleotide sequence analysis and generation of cDNAs from the BamHl-H fragment by several investigators have yielded diverse results. Although exact location of transcripts differ among investigators, it is apparent that a family of spliced and unspliced RNAs can be detected (Bradley et al, 1989;, Chen and Velicer, 1991; Peng et al, 1992; lwata et al, 1992). Predicted amino acid sequence homologies with various regions of the BamHl-H region to gag (retrovirus antigen), avian fps (kinase related transforming protein), v-fms (epidermal growth factor-like oncogene), mouse TLM oncogene, and fes/fps oncogenes have been found (Bradley et al, 1989; Peng et al, 1992; lwata et al, 1992). In addition, transcripts are initiated at multiple sites in a bi-directional manner (Chen and Velicer, 1991). In some transcripts, the 132 bp repeats are contained within introns (Peng et al, 1992). Interestingly, the 132 bp repeats are located downstream of a bi-directional promoter containing a putative origin of replication, CAAT and Sp1 binding site motifs (Bradley et al, 1989). Kawamura et al (1991) have shown that an antisense oligonucleotide complementary to a splice 47 donor sequence in one of the BamHl-H transcripts will inhibit proliferation of an MDV transformed cell line (producer, expression line). Although the BamHl-H region continues to be the subject of intensive investigation, functional significance of this region remains to be elucidated. b. 90 expresslon. In all MDV serotypes tested, expression of 9C is greatly reduced after 30-50 serial passages in vitro (Bulow and Biggs, 1975; Churchill et al., 1969; lkuta et al., 1983a; lkuta et al., 1983b; Nazerian, 1980; Purchase at al., 1971; Silva and Lee, 1984). Coincident with reduced gC expression, MDV serotypes 2 and 3 lose their vaccine efficacy and serotype 1 becomes attenuated with respect to oncogenicity (Calnek and Witter, 1991; Churchill et al., 1969; Kato and Hirai, 1985; Silva and Witter, 1985). M DV gB expression does not change significantly after serial in vitro passage (Silva and Lee, 1984). Despite an apparent correlation between MDV gC expression and MDV oncogenicity, isolation of naturally occurring non-oncogenic (serotype 2) MDV which express 90 and clonal isolates of oncogenic (serotype 1) strains which do not express gC (Bulow and Biggs, 1975a; Bulow and Biggs, 1975b; Glaubiger et al, 1983; Purchase et al., 1971), suggests that gC does not directly affect the oncogenic potential of MDV. However, in vivo data indicates that attenuation of MDV reduces the efficiency of lymphocyte infection (Schat et al., 1985). Therefore, MDV 90 may play a role in lymphocyte targeting. Regulation of MDV gC may have important implications for vims-host interactions, and have important consequences in the process of viral attenuation. 48 Although the gC-negative phenotype of attenuated MDV has been documented for more than 20 years, the molecular basis for this alteration has remained obscure. The inability of attenuated MDV to produce 90 may be due to mutations or deletions of the MDV 90 gene or promoter. Alternatively, decreased 90 expression may result from mutations in MDV encoded or induced trans-acting factors which interact with the QC promoter. Analysis of the mechanism for reduced expression of 9C is presented in this dissertation (Ch. 2). c. Attenuation of v.v. MDV. Due to the emergence of v.v. MDV in vaccinated flocks, and limited success with developing effective vaccines against v.v. MDV, analysis of the attenuation process of v.v. MDV strains is of interest. One such v.v. MDV, strain Md11, is characterized by causing acute cytolytic infection leading to early mortality (Witter, 1984). After 57 serial in vitro passages, v.v. MDV strain Md11 became attenuated with respect to oncogenicity. However, microscopic nerve lesions were detectable (Witter, 1982). After passage 75, nerve lesions were not detected. Thus, complete attenuation was not achieved until passage 75. Also, attenuated Md1 1 p75 was shown to induce considerably less interferon in chick embryos than that of oncogenic strains (Sharma, 1989). Decreased expression of glycoprotein C in attenuated Md11 has also been established (Witter and Lee, 1984) and the molecular basis of reduced 90 expression was investigated in detail (Wilson et al., submitted; this Dissertation Ch. 2). In addition, Tieber et al. (1989) have determined that attenuated Md11 is less able to transactivate an exogenous RSV promoter in vitro as compared to oncogenic Md11. Genome stmcture studies 49 indicate that amplification of the 132 bp repeats also occurs during attenuation of Md11 (Silva and Witter, 1985). Recently, our laboratory has identified an additional change in the genome of attenuated Md11(Wilson and Coussens, 1991; this Dissertation Ch. 3). Using cell-free MDV genomic probes, we identified a 200 bp deletion, also located within the IRL and TRL of the attenuated Md11 genome. This change, however is separate and distinct from the 132 bp amplified region. 132 bp amplification occurs in the BamHl-D and -H fragments (Silva and Witter, 1985), and the 200 bp deletion was mapped to the BamHI-L fragment (Wilson and Coussens, 1991). Further characterization of the 200 bp deletion in BamHl-L is presented in Chapter 4 of this Dissertation. The subject of this dissertation is analysis of the attenuation mechanisms of MDV. Research focus was directed toward two observations of attenuated MDV isolates: genome structure alterations and reduced expression of a major glycoprotein antigen. During preliminary investigations, it became clear that development of techniques to isolate cell-free MDV genomes would aid in the analysis of attenuated MDV isolates. Development of these techniques and characterization of attenuated MDV genomes is the subject of chapter 2. During the course of developing these techniques, novel changes in the genome of attenuated isolates of the v. v. MDV strain Md11 were identified and further investigated as described in chapter 3. The attenuation of this particular virus strain is of interest because of its highly oncogenic nature and potential for vaccine development of its attenuated counterpart. Thus, molecular analysis of the 50 reduced expression of a major glycoprotein antigen in v. v. MDV strain Md11 is presented in chapter 4. Attenuation of MDV as it relates to oncogenicity and vaccine development will be discussed in chapter 5. Chapter 2. PURIFICATION AND CHARACTERIZATION OF INFECTIOUS MAREK'S DISEASE VIRUS GENOMES USING PULSED FIELD ELECTROPHORESIS. Melinda R. Wilson and Paul M. Coussens Molecular Virology Laboratory, Department of Animal Science Michigan State University East Lansing, Michigan 48824 Virology. 185(2):673-680 51 if! G. 1h «in ) fr. .I ,{ ABSTRACT Marek’s Disease virus (MDV) is an acutely oncogenic avian herpesvirus. The tightly cell-associated in vitro growth characteristics of MDV present unique problems when attempting to purify, analyze, and manipulate MDV genomes. To facilitate molecular characterization of MDV, contour-clamped homogeneous electric fields electrophoresis (CHEF) was used to purify infectious M DV genomes. CHEF techniques were optimized for evaluation of total genome size and alterations in structure which occur during in vitro attenuation of oncogenic MDV. Our results indicated that genomes of attenuated serotype 1 MDV strain JM may contain deletions totaling 15 kbp while high passage serotype 2 non-oncogenic MDV strains 88-1 and 281Ml/1 were 5 kbp and 3 kbp larger, respectively, than their low passage counterparts. Using cell-free CHEF purified MDV genomes as hybridization probes, we identified a 200 bp deletion in attenuated genomes of the very virulent MDV strain Md11. Presently, it is unclear if this 200 bp deletion is related to mutations which lead to loss of oncogenicity or pathogenicity in Md11. This study is the first report which describes procedures for purification of infectious herpesvirus genomes from pulsed field gels. Our results demonstrate that pulsed field purified viral DNA will facilitate molecular characterization of M DV and other cell-associated herpesviruses. 52 I. INTRODUCTION. Marek’s Disease virus (MDV), an acutely oncogenic gammaherpesvirus, causes lymphoproliferation and nerve demyelination in domestic chickens (Payne-review, 1982). MDV is classified into 3 serotypes based primarily on antigenic similarities between virus isolates and ability to form lymphoid tumors in susceptible chickens. Serotype 1 MDV strains are oncogenic whereas serotypes 2 and 3 are non-oncogenic (Payne- review, 1982). The MDV genome, co-Iinear with alphaherpesvirus genomes, consists of unique long (UL) and unique short (Us) regions flanked by terminal repeats (T RL, TRs) and internal repeats (IRL, IRS) (Buckmaster et al, 1988; Cebrian et al., 1982). Nucleotide sequence analysis and comparison of banding patterns generated by restriction endonuclease digestions of MDV DNA indicates that MDV genome structure differs among serotypes, within each serotype, and after serial in vitro passage of MDV (Coussens et al., 1989; Hirai et al., 1979; 1981; 1984; lgarashi et al., 1987; Ross et al., 1984; Silva and Witter, 1985). A prominent change in attenuated serotype 1 MDV strains is amplification by tandem reiterations of a 132 bp sequence located in the IRL of MDV genomes (Maotani et al., 1986; Bradley et al., 1989; Chen and Velicer, 1991). Progress on the molecular analysis of MDV genomes is hampered by the tightly cell-associated in vitro growth characteristic of MDV (Schat, 1985). Separation of MDV DNA from host cell DNA by density-gradient centrifugation results in extremely low yields of pure MDV DNA (0.1 to 10 ug MDV DNA per 10° infected cells). Often, these preparations are contaminated with host cell DNA 53 I'I f K .r I . mm [mm by Infl. .MV int NW .II he C If r. . «r5 {:1 VI. u: 5 .V 54 (Kato and Hirai, 1985), as buoyant density of MDV DNA (1.705 g/cm3) (Lee et al., 1971) is similar to that of host cell genomic DNA (1.700 g/cm"). In addition, physical integrity of MDV genomes is often compromised during conventional isolation procedures. Thus, development of alternative procedures to obtain purified, intact, and infectious MDV DNA is of critical importance to the study of MDV molecular biology. In 1984, pulsed field electrophoresis (PFE) technology was introduced to analyze large DNA molecules (Schwartz and Cantor, 1984). Separation of large viral DNA genomes from host cell chromosomes by PFE has been used for Chlorella vims (Rhozinski et al., 1989), poxvirus (Bostock, 1988), Epstein-Barr virus (Harris and Bentley, 1988), and more recently for MDV (lsfort et al., 1990). Although these viral DNAs were separated from cellular DNA using PFE, infectivity of intact viral genomes isolated from pulsed field gels has not been demonstrated in any viral system. In this report, we present procedures developed to recover infectious cell-free MDV genomes using contour-clamped homogeneous electric fields (CHEF; Chu et al., 1986) electrophoresis. MDV DNA, separated from chicken chromosomes by PFE, was given the designation "cell-free" MDV genomes. CHEF was used to investigate genome structure differences among and within MDV serotypes, and after serial in vitro passage of various MDV strains. Our results indicate that oncogenic serotype 1 cell- free MDV genomes are 17-18 Kbp larger than non-oncogenic serotype 2 and 3 cell-free genomes and that the genome of attenuated serotype 1 MDV strain JM is 10 Kbp smaller than its 55 oncogenic counterpart. Using full length, cell-free MDV genomes as hybridization probes, we have identified a 200 bp deletion in attenuated genomes of the very virulent (v.v.) MDV strain Md11. Techniques and results presented in this report will contribute to our understanding of MDV pathogenicity and are directly applicable to studies of other large genome viruses. II. MATERIALS AND METHODS Cells and Viruses. Preparation, maintenance, and infection of duck and chick embryo fibroblasts (DEF and CEF, respectively) were performed as described (Glaubiger at al., 1983). MDV strains GA, JM, Md11, SB-1,281Ml/1 and HVT strain F0126, have also been described (Coussens et al., 1989; Tieber et al., 1990; Carter and Silva, 1990). Low passage (GAp5, HVTp5, Md11 p10, JMp14, SB-1p20, and 281Ml/1p15) and high passage (Md11p80, JMp213, SB-1p100, and 281 MI/1 p96) MDV isolates were kindly provided by the USDA-Avian Diseases and Oncology Laboratory, East Lansing, Michigan, USA. Preparation of MDV infected cells for CHEF. MDV infections of CEF and DEF were allowed to progress until cell monolayers appeared fully infected and cells began to detach from the substratum (5-7 days). Infected cells were washed from culture plates in phosphate buffered saline (Sambrook et al., 1989), pelleted, and resuspended in L Buffer (0.1 M EDTA, 0.01 M Tris-HCI pH 7.6, 0.02 M NaCl) (Sambrook et al., 1989) at an nah-:3 h 56 approximate concentration of 1x107 cells per ml. Cells in L Buffer were warmed to 42°C and an equal volume of 42°C 1.2% low melting point (LMP) agarose (Bethesda Research Laboratories, Gaithersburg, MD.) was mixed with cell suspensions. Cell/agarose mixtures were pipetted into sample plug molds (Bio-Rad, Inc., Richmond, CA.) (250 Lil per plug) and allowed to harden for 5 min at -20°C or 15 min at 4°C. Agarose embedded cells were lysed for 48 hr in 2 changes of L buffer (3 agarose plug volumes each) containing 1% (v/v) Sarkosyl and 0.1% (w/v) proteinase K at 50°C. Following Iysis, agarose plugs were washed three times for 15 min each in TE (10 mM tris-HCI, pH 8.0, 1 mM EDTA) at 50°C and stored in 0.5 M EDTA at 4°C until used. Agarose blocks were stored in this manner for six months with minimal degradation of DNA. Pulsed Field Electrophoresis. Lysed cells were analyzed by CHEF using the hexagonal electrode array developed by Chu et al. (1986). Agarose and LMP agarose gels were prepared in 0.5x Tris-Borate-EDTA buffer (TBE) (Sambrook et al. 1989). Prior to electrophoresis, excess EDTA in agarose plugs was removed by soaking plugs in 10 volumes of TE, ph 8.0 at 22°C for 20 min. Agarose plugs were placed in agarose gel sample wells and sealed with 1% LMP agarose. Yeast chromosomes (Bio-Rad, Inc., Richmond, CA.) or concatermized lambda DNA (Sambrook et al., 1989) were used as DNA size markers during electrophoresis. Electrophoresis was carried out in a CHEF-Dry II system (Bio-Rad, Inc., Richmond, CA.) gel chamber at 16°C in recirculating 0.5x TBE. Gels were stained 57 for 30 min in 0.5 ug per ml ethidium bromide (EtBr) in 0.5x TBE followed by destaining for 30 min in the same buffer. DNA bands were visualized and photographed with a UV light source. Sizes of DNA bands were determined from standard curves based on either yeast chromosome or lambda DNA concatemer standards. Preparation of cloned MDV fragment and cell-free MDV genomic probes. To prepare full length MDV genomic probes, gel slices containing cell-free MDV genomes were excised from LMP agarose CHEF gels and DNA was isolated by melting LMP agarose followed by phenol extractions as described (Sambrook et al., 1989). Concentrations of viral DNA recovered from CHEF gels were determined spectrophotometrically. Yields of cell-free MDV DNA varied from 130 ng for strain GAp7 to 1 ug for strain SB-1p101 per 10° infected cells. Cell-free MDV genomes were labeled with [alpha-”P1dCTP using a random primed labeling kit (Boehringer Mannheim Biochemicals, Indianapolis, IND.) as per manufacturer’s specifications. Due to probe complexity, incubation times for labeling reactions were one hour. Excess unincorporated nucleotides were removed using Sephadex G-50 spin columns as described (Sambrook et al., 1989). MDV fragment specific probes were prepared from a MDV strain GA DNA clone library (a generous gift of M. Nonoyama, Showa University Research Institute for Biomedicine in Florida, St. Petersberg, FL). BamHl-A, -H, -l2, -L, and -P, MDV insert DNA fragments were gel purified and labeled with [or-32P1dCTP as described above. r“ t) I F 453 58 Southern Blots. DNA in EtBr stained agarose gels was nicked by irradiation with UV light for 3-5 min, transferred to nylon membranes (Zeta- Probe, Bio-Rad, Inc., Richmond, CA.) under alkaline conditions (0.4 N NaOH), and hybridized to radiolabeled DNA probes as described (Budowle and Baechtel, 1990). CHEF gels containing DNA in the megabase size range were allowed to transfer for 24 hours. Conventional agarose gels containing DNA in the 50 to 0.1 kilobase pair (kbp) size range were transferred for 16 hours. Transfection of CHEF-purified MDV DNA. CHEF-purified, cell-free MDV genomes were introduced into CEF cells by electroporation (Neumann et al., 1982). Typically, 50-75 ng of CHEF-purified MDV DNA were used per 60 mm culture dish. Primary CEF cells were trypsinized, counted and pelleted. Cells were resuspended in Leibovitz-McCoy media supplemented with 0.3 M sucrose at a concentration of 5x10° cells per 0.8 ml., transferred to electroporation cuvettes (Bio-Rad, Inc., Richmond, CA), and warmed to 37° for 5 min. LMP agarose slices containing cell-free MDV genomes were melted at 65°, cooled to 37°, mixed with cell suspensions, and incubated at 37° for 10 min. Electroporation was carried out using a Gene Pulsar and capacitance extender (Bio-Rad, Inc., Richmond, CA.). Different combinations of voltage and capacitance settings were evaluated. In one set of experiments, voltage was held constant at 450v while capacitance settings were varied (125, 250, 500, and 960 uFD). In 59 another set of experiments, capacitance was held constant at 960 uFD and voltages of 150v, 250v, 350v, and 450v were used. Following electroporation, samples were incubated for 10 min at 37° and transferred to 4 ml Leibovitz-McCoy media containing 4% calf serum in 60 mm culture dishes. Electroporated cells were maintained at 37°C in an atmosphere supplemented with 5% 002. Culture media was changed to Leibovitz-McCoy supplemented with 1% calf serum, 48 hr after electroporation. III. RESULTS Separation and Identification of MDV genomes and chicken chromosomes using CHEF. On EtBr stained CHEF gels containing samples from high and low passage Md11 infected cells, four DNA bands of 2,100, 1,400, 660, and 180 Kbp were visible (Figure 1). A broad smear of less than 180 kbp was observed in all lanes. Lanes containing uninfected CEF displayed all EtBr stained bands found in infected cell lanes except the 180 kbp band, suggesting this band was of viral origin (Figure 1). Southern transfer analysis was used to determine whether particular bands in CHEF gels of MDV infected cells were of viral or cellular origin. Radiolabeled cloned MDV DNA fragments hybridized primarily to the 180 kbp band (Figure 2, A), indicating that this band was viral DNA. Hybridization of MDV probes to DNA retained in sample walls was also evident (Figure 2, A), and addressed previously (lsfort et al, 1990). In lanes containing Md11p80 DNA, an additional DNA band 60 of 125 kbp which hybridized to the MDV DNA probe was evident below the more intense 180 kbp band (Figure 2, A). Hybridization of the 2,100, 1,400, and 660 kbp bands to CEF DNA probes was observed in uninfected CEF, Md11p10, and Md11p80 (after longer exposure to X-ray film ) sample lanes (Figure 2, B), suggesting that these bands were chicken chromosomes which begin to resolve under CHEF conditions. To distinguish between cell-associated MDV DNA retained in sample wells and MDV DNA which migrated at 180 kbp on CHEF gels, we designated the 180 kbp band as "cell-free" MDV DNA. Infectivity of cell-free CHEF purified MDV DNA. Although analysis of CHEF gels indicated that a 180 kbp band was the full length MDV genome, we were not certain this DNA band retained biological activity. To assess infectivity and physical integrity of cell-free CHEF purified MDV genomes, electroporation of CHEF-purified MDV DNA into CEF cells was performed as described in Materials and Methods. Typical MDV viral plaque formations were evident 6 days post electroporation on CEF cells electroporated with 450v or 250v at 960uFD (Figure 3, A 3). Resultant viral plaques were purified and expanded 2 passages to obtain sufficient quantities of DNA for further studies. Analysis of BamHl restriction digests of viral DNA obtained from cells transfected with CHEF- purified MDV DNA verified that viral plaque formation resulted from electroporation of CHEF purified MDV DNA. No major genome rearrangements were visible in viral DNA of cells 61 transfected with CHEF purified MDV DNA (Figure 3, B). Transfection efficiency (20 PFU/ug viral DNA) from electroporation of cell-free MDV genomes was similar to transfection efficiencies of total DNA isolated from MDV infected cells using the polybrene/DMSO procedure described by Kawai and Nishizawa (1984). However, transfection efficiencies from electroporation of cell-free MDV DNA were 5 times lower than that obtained by electroporation of total DNA isolated from MDV infected cells (100 PFU/ug viral DNA). Detection of Genome Size Differences Among MDV Serotypes. To determine if serotype classification could be correlated to cell-free MDV genome size differences among MDV strains, and to evaluate use of CHEF as a diagnostic aid to MDV serotyping, cells infected with MDV serotypes 1 (GAp7, Md11p11, Md11p81, JMp15, and JMp214), 2 (SB-1p21, SB-1p101, 281MI/1p16, and 281 MIN p97), and 3 (HVTp7 strain F0126) were analyzed on high resolution CHEFgms Faster migration of low passage MDV serotypes 2 and 3 cell-free genomes indicated lower size genomes of SB-t, 281Ml/1, and HVT as compared to low passage serotype 1 strains (T able 1). Results for serotypes 1 and 3 are consistent with previously reported genome sizes derived from sedimentation values of MDV DNA from serotype 1 (180 kbp) and 3 (167 kbp) (Hirai et al., 1979; Lee et al., 1971). An estimate of MDV serotype 2 (281Ml/1) genome size, derived from the sum of Neil restriction fragment sizes, is 161.8 kbp (D. Reilly, personal 62 communication), consistent with results from CHEF gels (Table 1). To determine if repeated passage in culture resulted in altered mobility of cell-free MDV genomes on CHEF gels, we compared low and high passage pairs of serotype 1 and 2 samples on high resolution CHEF gels (Figure 4, Table 1). High passage serotype 1 strains were either smaller (JMp214) or remained the same (Md11p81) in cell-free genome size as compared to low passage counterparts (Figure 4, Table 1). High passage serotype 2 strains (SB-1 p101 and 281/Mlp97) were larger in cell-free genome size as compared to low passage counterparts (Figure 4, Table 1). Characterization of alterations in attenuated MDV genomes using full length cell-tree MDV genomes as a hybridization probe. Attenuated v.v. MDV strain Md11 contains 0.6 to 5.4 Kbp of 132 bp repeat units in tandem reiteration (Silva and Witter, 1985). However, results from high resolution CHEF gels indicated that size of cell-free Md11 genomes remained unchanged after attenuation (Figure 4, Table 1). To investigate the lack of genome size difference between oncogenic and attenuated MDV strain Md11, we performed Southern transfer analysis on gels containing restriction digests of total DNA isolated from cells infected with oncogenic (Md11p11) and attenuated (Md11p81) MDV isolates. Using CHEF purified, cell- free Md11 p10 genomes as a hybridization probe viral bands (contained within the mixture of cellular and viral DNA immobilized on membranes) were detected (Figure 5, A). BamHI and Bgll restriction fragments of viral DNA _——---u_ 63 were assigned letter designations according to published maps of the MDV genome (Fukuchi et al., 1984). Amplified 132 bp repeat sequences located within BamHl-D, -H, and Bgll-F viral DNA fragments were evident (Figure 5, A). These results are consistent with previous findings (Hirai et al., 1981; Maotani et al., 1986; Silva and Witter, 1985). In addition, BamHI-L and Bgll-J viral DNA fragments were absent in attenuated Md11 DNA (Figure 5A, lane H). Instead, attenuated Md11 DNA contained fragments 200 bp smallerthan the BamHl-L and Bgll J fragments of oncogenic Md11 DNA (Figure 5, A). According to published maps (Fukuchi et al., 1984), BamHl-L and Bgll J are overlapping fragments located within the IRL region near the IRL/ IRs junction. To verify changes assigned to specific MDV BamHl fragments using CHEF purified cell-free Md11p10 genomic probes, membranes were stripped and hybridized to cloned BamHl-L (Figure 5, B) and BamHl-H (Figure 5, C) MDV DNA fragments. Both probes hybridized to identical bands identified using CHEF purified cell-free Md11p10 genomic probes (Figure 5, B and C). IV. DISCUSSION Difficulties intrinsic to preparation of cell-free MDV and its cell-associated nature during in vitro cultivation have hampered efforts to characterize MDV at the molecular level. As described in this report, these difficulties can be circumvented by using PFE to isolate infectious cell-free MDV genomes. Correlation of genome size to serotype classification of MDV strains was evaluated by comparing mobilities of cell-free MDV genomes of each serotype on 64 CHEF gels. Oncogenic serotype 1 cell-tree MDV genomes were 17-18 kbp larger than non-oncogenic serotypes 2 and 3 MDV DNA. Extra genetic material in the larger genomes of serotype 1 may code for factors involved in MDV-induced lymphoid proliferation. Comparison of attenuated MDV serotype 1 JMp214 relative to its low passage, oncogenic counterpart (JMpt 4) supports this hypothesis. MDV strain JMp214 (an attenuated, non-oncogenic MDV isolate) contains at least 40 copies of the 132 bp tandem repeat (M. R. Wilson, unpublished results). These repeats would increase the JMp214 genome by 5.28 kbp. However, cell-free JMp214 genomes purified by CHEF are approximately 10 kbp smaller than genomes of their oncogenic counterparts (JMp14). Thus, total size of deleted genetic material in the JMp214 genome may exceed 15 kbp. It appears that deletions and expansion of tandem repeats may occur concurrently during attenuation of oncogenic MDV strains. Deletions, which may be similar in size to expanded 132 bp repeats, may account for the lack of detectable cell-free genome size difference between oncogenic and attenuated v.v. MDV strain Md11. Minor differences in genome sizes may go undetected due to resolution limits of PF gels. Using CHEF purified cell- free Md11p10 genomic probes, we identified a 200 bp deletion located in the IRL region of attenuated Md11 DNA. Presently, it is unclear if this deletion is related to mutations which lead to loss of oncogenicity in attenuated MDV strain Md11. While MDV strain Md11 lost the ability to form lymphoid tumors after 57 serial in vitro passages, it retained the ability to produce typical MD microscopic nerve lesions in chickens. At passage 75, this ability is lost (Witter and Lee, 1984). It is possible that the 200 65 bp deletion in the BamHI-L fragment effects neuroinvasive mechanisms of MDV. Further comparison of this region isolated from oncogenic and attenuated Md11 strains is in progress. In contrast to serotype 1 MDV, high passage serotype 2 cell-free genomes increased in size by 5 kbp (SB-1) and 3 kbp (281Ml/1) relative to their low passage counterparts. Increased copy number of tandem repeats may contribute to increased size of serotype 2 genomes after serial in vitro passage. Expansion of approximately 15 repeats is observed in SB-1p100 (M. R. Wilson, unpublished results). However, similar expansion of repeat sequences in the 281Ml/1 p96 genome has not been observed (H. Camp, unpublished results). Infectivity of viral genomes, isolated from pulsed field gels, has not been assessed previously. Results presented in this report represent the first example of procedures for recovery of purified infectious viral DNA from pulsed field gels. Procedures described in this report may be applied to other large DNA viruses for which isolation of pure intact viral genomes is difficult. IV. ACKNOWLEDGMENTS We thank Dr. Robert Bull and Dr. John Gerlach for use of a pulsed field apparatus employed in early investigations and for many helpful discussions. We also thank Dr. L. F. Velicer and Dr. S. Triezenberg for critical review of the manuscript and Mr. Ronald Southwick for excellent technical assistance. This work was supported by the Research Excellence Fund, State of Michigan, by the 66 Michigan Agricultural Experiment Station and Grants no. 88-37266-3983 and 90-34116-5329 awarded to P. M. Coussens under the Competitive and Special Research Grants Programs, respectively, administered by the U. S. Department of Agriculture. M. R. Wilson was supported by a Michigan State University Biotechnology Research Fellowship award. 67 Figure 1. Separation of oncogenic and attenuated cell-free MDV genomes from chicken chromosomes using PFE. Cells infected with oncogenic (Md11p10) and attenuated (Md11p80) serotype 1 MDV isolates were lysed in agarose plugs and electrophoresed on a 1% agarose CHEF gel for 24 hours at 50-90 sec. ramped pulse times, followed by staining with EtBr. CEF: uninfected chicken embryo fibroblasts. Yeast chromosomes and lambda concatemers were used as size markers. Sizes, in kilobase pairs, are indicated on the right. 68 Figure 1. 4660 0.5), and only slightly higher than oncogenic Md11 viral DNAfinput viral DNA (0.10.5). Rates of viral DNA replication from 27 to 69 hours post infection in CEF were compared for all 3 strains. Md11p83-L rate of viral DNA replication was intermediate to oncogenic and attenuated Md11, consistent with data for rates of plaque formation (Figure 5). Rate of viral DNA replication for attenuated Md11 91 was slightly higher than that of Md11p83-L (p<0.05). Rate of viral DNA replication for oncogenic Md1 1 was significantly lowerthan that of attenuated Md1 1 (p<0.001) or Md11p83-L (p<0.001). Transcriptional activity and expression of major glycoprotein antigens. Attenuation of MDV strain Md11 results in a severe reduction of 9C expression, where as expression of other major MDV antigens (98, and p79) remains unchanged (Wilson et al., submitted). To determine if the 200 bp deletion within the BamHl-L fragment affected the expression of major MDV antigens, immunoprecipitations were performed on oncogenic and attenuated Md11, and Md11p83-L. Using convalescent chick sera (immunoprecipitates QB, 90 and p79) or gC specific rabbit sera, there were no differences observed between amounts of -viral antigens expressed in cells infected with attenuated Md11 or Md11p83-L (data not shown). Thus, presence of the wild type BamHl-L fragment in the Md11p83-L genome did not restore gC expression levels to that of oncogenic Md11. These results suggest that the 200 bp deletion in the BamHI-L fragment did not significantly alter the expression of the major antigens of MDV. To determine if transcripts within or partially within the BamHI-L fragment were affected by the 200 bp deletion in attenuated Md11, northern blot analysis of total infected cell RNA was performed. Hybridization to a MDV BamHl-L fragment probe was used to compare transcriptional activity in the BamHI-L region of oncogenic and attenuated Md11, and Md11p83-L. Three major transcripts, which hybridized to the BamHl-L probe, of 2.8, 1.7, and 0.8 Kb are present in all 92 strains (T able 1). In addition, a 10 kb transcript present only in CHX treated groups was detectable (data not shown). These data were similar to previous reports using RNA from other oncogenic MDV strains (Schat et al, 1989; Sugaya et al, 1990). In attenuated Md11, the 0.8 kb transcript was expressed at higher levels than that of oncogenic Md11 or Md11p83-L (Table 1). Relative abundance of specific transcripts varied among strains and transcriptional class. Data for transcripts expressed in oncogenic Md11 indicates that the 2.8, 1.7, and 0.8 kb transcripts were expressed as IE genes (Table 2). In attenuated Md11 infected cells treated with CHX, the 1.7 and 0.8 kb transcripts were not detected (T able 2) even though these transcripts were expressed in the untreated and FAA treated groups. In Md11p83-L infected cells treated with CHX, the 1.7 kb transcript was not detected, but was expressed in the untreated and PAA treated groups. However, the 0.8 kb transcript in Md11p83-L infected cells was expressed in a similar fashion to that in oncogenic Md11 infected cells (T able 2). These data indicate that patterns of expression of the 0.8 kb IE transcript was affected by the 200 bp deletion in the BamHl-L fragment of attenuated Md11. Data also indicates that the 0.8 kb IE transcript was expressed as an E transcript in attenuated Md11. In addition, attenuation of Md11 affects patterns of expression of the 1.7 kb IE transcript. As with the 0.8 kb IE transcript, data indicates that the 1.7 kb IE transcript is expressed as an E transcript in attenuated Md11. However, deletion in the BamHl-L fragment did not have a significant affect an altered expression of the 1.7 kb IE transcript, as expression patterns for the 1.7 kb IE transcript in Md11p83-L were similar to that of attenuated Md11. 93 To confirm that only IE and E genes were detected, and CHX or PAA treatment effectively blocked late gene expression, membranes were stripped and re-hybridized with a late gene (MDV gC) specific probe. Only untreated lanes containing oncogenic Md11 hybridized to the 9C probe (data not shown). Sub-localization of the 200 bp deletion within the BamHI-L fragment. Previous reports indicate that 0.6-0.8 kb immediate early (IE) transcripts which hybridize to the BamHl-L fragment probe are expressed in MDV transformed lymphocytes (Schat et al, 1989; Ohashi and Schat, 1992). Recently the same group generated a cDNA (termed L,) derived from RNA isolated from MDV transformed lymphocytes. L1 hybridizes to IE transcripts encoded within or partially within the BamHl-L fragment (Ohashi and Schat, 1992). Physical mapping studies were initiated to determine where in the BamHl-L fragment the 200 bp deletion occurred in attenuated Md11 DNA and if this deletion was within or near the L1 cDNA region. DNA isolated from oncogenic, attenuated, and Md1 1p83-L infected cells was double digested with BamHI and Hpall, Haelll, or Hinil and subjected to Southern transfer analysis. Initially, membranes were probed with a full length BamHI-L DNA fragment. 1.13 and 0.36 kb Haelll, 1.05 and 0.67 kb Hpall, and 0.95 kb Hinll BamHl-L sub-fragments in attenuated Md11 DNA were decreased in size as compared to oncogenic Md11 and Md11p83-L DNA (Figure 6A). Fragment sizes in oncogenic Md11 and Md11p83-L DNA were identical. Fragment sizes in oncogenic Md11 and Md11p83-L DNA corresponding to fragments affected by the 94 deletion in attenuated Md11 DNA were 1.25 and 0.37 kb Haelll, 1.15 and 0.745 kb Hpall and 1.05 kb Hintl sub fragments of BamHl-L. L, cDNA, which was previously mapped to the 5' end of MDV BamHl-L fragment (Ohashi, 1992), was used as a probe for stripped membranes of panel A (Figure 68). L, hybridized strongly to the 1.13 kb Haelll and 1.05 kb Hpall fragments in attenuated Md11 lanes (Figure 6B). Very weak hybridization was visible with the 0.95 kb Hintl fragment. These results indicated that the 200 bp deletion was located within the 0.95 kb Hintl fragment. This fragment was adjacent to the 3’ end of L, cDNA (Figure 7). To confirm this observation, membranes were stripped and probed with the 1.0 kb Hintl sub-fragment of the MDV strain GA clone. As predicted, the 1.0 kb Hintl probe hybridized strongly to the 1.13 and 0.36 kb Haelll fragments, 1.05 and 0.67 kb Hpall fragments, and the 0.95 kb Hintl fragment of attenuated Md11 DNA (Figure 60). Weak hybridization was evident in bands which were not altered in attenuated Md11 DNA. Sub-localization of the deletion in relation to location of L, cDNA is illustrated in Figure 7. In oncogenic Md11 DNA, minor populations of DNA fragments which co-migrated with DNA fragments containing the 200 bp deletion of attenuated Md11 DNA were evident (Figure 6A, B, and C). Conversely, in attenuated Md11 DNA, minor populations of DNA fragments which co-migrated with DNA fragments of oncogenic Md11 DNA were also evident, but was not detected in Md11p83-L DNA. These data indicate the presence of minor sub-populations in oncogenic and attenuated Md11 DNA. 95 Assessment of oncogenicity in viva. Physical mapping data of the deletion in the BamHI-L fragment indicated that the BamHI-L fragment in Md11p83-L was identical in structure to that of oncogenic Md11. Therefore it was of interest to determine if the BamHI-L region contributed directly to oncogenicity. MD susceptible chicks were inoculated with oncogenic, attenuated Md11 or Md11p83-L. As expected, a high incidence of MD (95%) was found in oncogenic Md11 (T able 3). In the attenuated Md11 group, only 1 birds displayed visible signs of MD, whereas no visible MD lesions were found in the Md11p83-L group (Table 3). In this study, there was a high prevalence of tumors in heart (58%) and spleen (37%) induced by oncogenic Md11. However, the MDT bird infected with attenuated Md11 had tumorous liver and spleen tissue. Conversely, no liver tumors were found in MD"' birds infected with oncogenic Md11. IV. DISCUSSION Attenuation of oncogenic MDV offers a unique opportunity to examine mechanisms involved in herpesvirus-induced neoplasia in a natural host. Major changes in attenuated MDV genome structure, coincident with loss of oncogenicity, are well documented (Bradley et al., 1989; Chen and Velicer, 1991; Maotani et al., 1986; Silva and Witter. 1985). Disruption of RNA transcription due to amplification of 132 bp repeat sequences within the IR, region of the MDV genome may be related to MDV attenuation (Bradley et al, 1989). With the recent adaptation of pulsed field gel electrophoresis technology to the study of MDV (lsfort et al., 1990; 96 Wilson and Coussens, 1991), deletions in the BamHI-L fragment of the attenuated Md11 genome were identified (Wilson and Coussens, 1991). Subsequently, we were able to recover a purified population of an attenuated Md11 variant (Md11p83-L) from a heterogeneous stock of attenuated Md11. Data in this report indicates that the Md11p83-L genome contains only one of the two changes thus far identified in attenuated Md11. DNA physical mapping studies indicate that the deletion in the BamHI-L fragment of attenuated Md11 occurred early during serial passage in culture, since the oncogenic virus used in these studies, Md11 passage 15, contained a minor population of the deleted BamHl-L fragment, whereas earlier passages (Md11p10) did not contain detectable sub-populations of the deleted BamHI-L fragment (Wilson and Coussens, 1991). Md11p83-L was utilized in comparative studies to determine if deletions in the BamHl-L fragment of attenuated Md11 affected biological properties of the virus. in vitro, deletions in the BamHl-L fragment of attenuated Md11 genomes contributed to alteration of growth characteristics of the virus. Since plaque formation and DNA replication ofMd11p83-L was intermediate between oncogenic and attenuated Md11, deletion in the BamHl-L region was only partially responsible for increased plaque formation and DNA replication of attenuated Md11. Our results further suggest that amplification of the 132 bp repeat sequences, within IR,, also contributed to enhanced viral replication in vitro. DNA/PFU ratios for oncogenic Md11 were 8- and 5-fold higher than DNA/PFU ratios of attenuated Md11 and Md11p83-L, respectively. This difference may suggest that oncogenic Md11 was less able to spread to neighboring cells, leading to accumulation of non- 97 viable virus in vitra. Small size of oncogenic Md11 plaques as compared to attenuated Md11 plaques supports this nation. In addition, similarity of plaque size in Md11p83-L and attenuated Md11 infected cells suggests that the 200 bp deletion in the BamHl-L region may not contribute to initial virus spreading efficiency in vitro. However, preliminary results from our laboratory indicate that during extended in vitro cultivation (5 to 9 days), the efficiency of secondary plaque formation of oncogenic Md11 and Md11p83-L is higher than that of attenuated Md11. Studies designed to differentiate between vims spread to neighboring cells and virus replication within a single cell would help clarify these observations. Analysis of RNA transcripts indicated that temporal expression of genes encoded or partially encoded within the BamHI-L fragment, specifically the 0.8 kb IE transcript, were directly affected by deletion in the BamHI-L fragment of attenuated Md11 DNA. In addition, temporal class switching of the 0.8 kb IE transcript suggest that this deletion may reside within the promoter of an IE gene. Deletion within the BamHl-L fragment did not directly contribute to reduction in the expression of the 1.7 kb transcript of attenuated Md11. The 2.8 kb IE transcript was also affected by attenuation of Md11. However, data in this report indicates that the 200 bp BamHI-L deletion was not a major factor for alteration of the 2.8 kb IE transcription pattern. Lack of detection of the 2.8 kb IE transcript in PAA treated, attenuated Md11 or Md11p83-L infected cells suggests that attenuation of Md11 may affect RNA stability of the 2.8 kb IE transcript. Alternately, expression of the 2.8 kb IE transcript may have been subject to repression by a cellular or viral-encoded regulatory protein in attenuated Md11 and Md11p83-L 98 infected cells. Previous reports indicate that transcripts, similar in size to the 0.8 kb IE transcript affected by attenuation, are expressed in producer as well as non- producer MDV transformed cell lines (Schat et al, 1989). In addition, this transcript was tmncated by approximately 300 bp in a non-producer MDV transformed cell line. Since non-producer MDV transformed cells arise after prolonged in vitro cultivation, our results, taken together with previous results (Schat et al, 1989) suggest that the BamHI-L region is affected by prolonged in vitro cultivation of MDV. The observation that alteration of IE transcripts, DNA replication, and growth properties in vitra are related to mutation in the BamHl-L fragment of attenuated Md11 genomes and the ca-linear location of MDV genes with VZV and HSV (Buckmaster et al, 1988; Binns and Ross, 1989; Chen and Velicer, 1991b; Coussens et al., 1990; Ross et al., 1989; Brunovskis and Velicer, 1992) leads to speculation that a protein functionally similar to HSV-l ICPO is encoded within the BamHI-L region affected by attenuation. This is substantiated by previous results which indicate that mutation in the HSV-l ICPO gene affects lytic growth, expression of early and late proteins, and particle/PFU ratios of HSV-l in vitra (Cal and Schaffer, 1992). In vivo studies indicated that the deletion in the BamHI-L fragment of attenuated Md11 DNA did not contribute directly to MDV-induced neoplasms. However, MDV-induced oncogenicity is most likely a complex multi-step process involving the viruses ability to specifically infect lymphocytes, induce cellular and 99 viral transactivator proteins, and maintain lymphocytes in a transformed state. Expression of BamHl-L immediate early transcripts in MDV transformed cell lines (Schat et al, 1989; Sugaya et al, 1990), similar in size to the 0.8 kb transcript affected by Md11 attenuation, suggests that this region may be involved in latent infection of MDV transformed cells. In viva studies, utilizing Md11p83-L, to determine if the deletion in the BamHI-L fragment of attenuated Md11 genomes affects latent MDV infection of lymphocytes or early cytolytic infection are in progress. Comparison of the BamHl-L fragment and transcripts encoded within the BamHl-L fragment of other oncogenic MDV strains to their attenuated counterparts would be of interest to determine if deletions occurring in this region is a general phenomenon of attenuation or specific to v.v. MDV attenuation. 100 ACKNOWLEDGMENTS We thank Jessica Marcus for excellent technical assistance. We also thank S. J. Triezenberg for careful review of this manuscript and gratefully acknowledge Telmo Oleas and John Gill for helpful discussions and assistance with statistical analysis. This work was supported by the Michigan Agricultural Experiment Station, the Research Excellence Fund, State of Michigan, and by Grant Nos. 88- 37266-3983, 90-34116-5329, and 92-37204-802 awarded to P. M. Coussens under the Competitive Research Grants Program administered by the US. Department of Agriculture. M. Wilson was supported in part by a Michigan State University Biotechnology Research Fellowship. 101 Figure 1. Identification of an attenuated MDV strain Md11 variant. Attenuated Md11 cell-free genomes were electroporated into CEF cell suspensions as described (Wilson and Coussens, 1991). Resultant viral plaques were expanded, lysed in agarose plugs and cell-free MDV genomes were isolated on low melting point agarose CHEF gels as described (Wilson and Coussens, 1991). Cell-free MDV genomes from electroporated cells were digested with BamHI, electrophoresed on 0.8% agarose gels, transferred to nylon membranes, and hybridized under high stringency conditions to cell-free or- 32P labeled MDV genomic probes. BamHl digests of attenuated Md11 parent genomes (first lane) and cell-free MDV genomes recovered after electroporation (CHEF: 6 and 7). Dot on right denotes altered BamHl-L fragment in the CHEF 7 lane. Viral bands were assigned letter designations according to published maps of the MDV genome (Fukuchi et al., 1984). Sizes of radiolabeled Hinolll digested Lambda DNA are located to the right in kilobase pairs. 102 CHEF 6 7 PARENT A\ -< B, 23.13 C? E/ F/ ‘942 e/ . H’ h «6 56 I- ‘ can. J- 4.36 K.— o O L— " ‘ M... «2.32 N; ' «2.03 p- 2: S" T..— Figure 1. 103 Figure 2. Identification of the altered DNA fragment in the attenuated Md11 variant as BamHl-L. Cell-free MDV genomes of oncogenic (LP), attenuated (HP) and the attenuated variant (V) were isolated, digested with BamHI, and subjected to Southern transfer analysis as described in the Figure 1 legend. Left panel: autoradiographs of membrane hybridized to a radiolabeled MDV BamHl-L DNA fragment. Right panel: membrane depicted in the left panel was stripped and hybridized to a radiolabeled MDV BamHl-H DNA fragment. 104 LPV HP LPVHP Figure 2. 105 Figure 3. Comparison of plaque morphology of oncogenic (1) and attenuated (2) Md11 to Md11p83-L (3). Photomicrographs at max magnification were taken 7 days after infection of chick (A) or duck (B) embryo fibroblasts. Infections were performed as described in Materials and Methods. 106 Figure 3. 107 Figure 4. Panel A: comparison of plaque formation of oncogenic and attenuated Md11 to Md11p83-L in CEF. Growth curves forindividual strains were determined as described in Materials and Methods. Panel B: Viral DNA replication of oncogenic and attenuated Md11, and Md11p83-L in CEF. Isolation of total infected cell DNA, slot bolt assays, and replication curves were performed as described in Materials and Methods. On all graphs, each data point represents the average of three separate determinations. mmDZJU 7: mEDOI .e 959.... mm mv vm O .-l--. -- .r II 8.0 .II./# . CI 0 a \ p .mfio «I. /.z m \ Goo; N M d \\ m .3, o N V . .00— a 33 £2 {E m mm_.;-.-m.z_mm301 me me 4m 0 .l I - -I...Il ii? 000 .mmd Iolo .w OGQ: / . M d m. .2; a N uv eon.— v—QFFUE 62m Nb mdiI «.9. manssu z. mmDOI we 3N blmma__32 mmDZDU z. m>92 E. _ m; m. _ E 117 Table 3. Incidence of Marek’s Disease with Md11 strains. In vivo studies were performed as described in Materials and Methods. % Marek’s Disease was determined from number of birds which displayed gross lymphoma formation. 118 Table 3. Incidence of Marek's Disease with Md11 strains. STRAIN NO. BIRDS AGE AT % MAREK’S AND TERMINATION DISEASE PASSAGE (WEEKS) Md11p14 Md11p83 Md11p83-L Chapter 4. MOLECULAR ANALYSIS OF THE GLYCOPROTEIN C-NEGATIVE PHENOTYPE OF ATTENUATED MAREK'S DISEASE VIRUS. Melinda R. Wilson, James T. Pulaski, Ronald A. Southwick, Virginia L. Tieber', Yu Hong, and Paul M.Coussens. Molecular Virology Laboratory Department of Animal Science Michigan State University East Lansing, Michigan 48824 Submitted to Virology 2/17/93. ' Present Address: Department of Biochemistry, Michigan State University. 119 ABSTRACT Serial passage of oncogenic serotype 1 Marek’s Disease vian (MDV) in cultured avian cells results in attenuation of viral oncogenicity and pathogenicity. Coincident with attenuation, expression of MDV glycoprotein C (90) is significantly reduced. Regulation of MDV gC may have important implications for virus-host interactions, and have important consequences in the process of viral attenuation. To investigate the mechanism by which MDV gC expression is reduced during attenuation of the very virulent serotype 1 MDV strain Md11, protein levels, gene structure, steady state RNA levels, and transcription rates of MDV 90 from oncogenic and attenuated isolates of Md11 were determined. Comparison of these data with similar studies on MDV proteins whose expression is not altered during attenuation indicate that reduced expression of MDV 90 is directly related to reduction in transcription rate of the MDV gC gene in attenuated Md11. Reduced transcription rates and lack of any gross structural alterations in the attenuated MDV 90 gene suggests that MDV regulatory protein(s), which interact with the MDV gC promoter are altered during attenuation of MDV strain Md11. These regulatory proteins may be encoded within highly polymorphic regions of the attenuated MDV strain Md11 genome. Further study of MDV antigens and transcription factors at the molecular level may provide important clues for understanding virus-host interactions and immune response to MDV and other herpesvirus infections. 120 I. INTRODUCTION Marek’s Disease virus (MDV), an oncogenic avian herpesvirus, produces malignant lymphomas in infected hosts (Calnek and Witter, 1991). MDV is classified into three serotypes: oncogenic serotype 1, non-oncogenic serotype 2, and non-pathogenic serotype 3 (Payne, 1982). Serotypes 2 and 3 have been used as vaccines against serotype 1 MDV (Payne, 1982; Witter, 1983). There are at least six antigenically active viral proteins found in all three serotypes of MDV (lkuta et al., 1983; Silva and Lee, 1984). Two of these, MDV A (90) and B (98) antigens are homologous to herpes simplex virus-1 (HSV-1) glycoproteins C and B, respectively (Binns and Ross, 1989; Buckmaster et al., 1988; Chen and Velicer, 1991b; Coussens et al., 1990; Ross et al., 1989). In HSV-1, 90 is non-essential for viral replication in vitra whereas 98, highly conserved among alphaherpesviruses, is required for virus attachment to host cells (Roizman, 1991). In all MDV serotypes tested, expression of 90 is greatly reduced after 30-50 serial passages in vitra (Bulow and Biggs, 1975; Churchill et al., 1969; lkuta et al., 1983a; lkuta et al., 1983b; Nazerian, 1980; Purchase at al., 1971; Silva and Lee, 1984). Coincident with reduced gC expression, MDV serotypes 2 and 3 lose their vaccine efficacy and serotype 1 becomes attenuated with respect to oncogenicity (Calnek and Witter, 1991; Churchill et al., 1969; Kato and Hirai, 1985; Silva and Witter, 1985). MDV gB expression does not change significantly after serial in vitra passage (Silva and Lee, 1984). MDV gC, an abundant viral glycoprotein antigen, is found in sera of infected birds, infected cell culture fluids, and to a lesser extent associated with the plasma 121 an. M h .‘ 122 membrane of infected cells (lsfort et al., 1986). Evidence from pulse-chase experiments and tunicamycin treated cells suggests that MDV 9C is synthesized on the endoplasmic reticulum and processed through the Golgi via two pathways. Secretion of N-Iinked glycosylated MDV gC occurs through a major pathway while an unglycosylated form is secreted through a minor pathway (lsfort et al., 1986). Approximately 95% of serotype 1 MDV 90 is secreted while the remaining 5% is associated with plasma membranes of infected cells (lsfort et al., 1986). In comparison, MDV 98 is found primarily in infected cell lysates and the glycoprotein processing pathway is similar to that of other herpesvirus gB homologues (Chen and Velicer, 1991b; Ross et al., 1989; Velicer et al., 1978). Despite an apparent correlation between MDV gC expression and MDV oncogenicity, isolation of naturally occurring non-oncogenic (serotype 2) MDV which express gC and clonal isolates of oncogenic (serotype 1) strains which do not express gC (Bulow and Biggs, 1975a; Bulow and Biggs, 1975b; Glaubiger et al, 1983; Purchase at al., 1971), suggests that 90 does not directly affect the oncogenic potential of MDV. However, in viva data indicates that attenuation of MDV reduces the efficiency of lymphocyte infection (Schat et al., 1985). Therefore, MDV gC may play a role in lymphocyte targeting. Regulation of MDV 90 may have important implications for virus-host interactions, and have important consequences in the process of viral attenuation. Although the gC-negative phenotype of attenuated MDV has been documented for more than 20 years, the molecular basis for this alteration has remained obscure. The inability of attenuated MDV to produce 90 may be due to 123 mutations or deletions of the MDV 90 gene or promoter. Alternatively, decreased 90 expression may result from alterations of MDV encoded or induced trans-acting factors which interact with the 9C promoter. To distinguish between these possibilities we investigated the mechanism by which MDV gC expression is reduced during attenuation of the very virulent (v.v.) serotype 1 MDV strain Md11. Protein levels, gene structure, steady state RNA levels, and transcription rates of MDV gC from oncogenic and attenuated isolates of Md11 were determined. Comparison of these data with similar studies on MDV proteins whose expression is not altered during attenuation indicated that reduced expression of MDV 9C is directly related to reduction in transcription rate of the MDV 96 gene in attenuated Md11. Reduced transcription of the MDV 90 gene in attenuated Md11 are not due to major structural alterations of the 90 promoter or coding sequence. Reduced transcription rates and lack of any gross sthctural alterations in the attenuated MDV gC gene suggests that MDV regulatory protein(s), that interact with the MDV gC promoter are altered during attenuation of MDV strain Md11. These regulatory proteins may be encoded within highly polymorphic regions of attenuated MDV such as the BamHI-D, -H, and -L fragments located in repeat regions of the MDV strain Md11 genome. II. MATERIALS AND METHODS Cells and virus. Preparation, maintenance and infection of duck embryo fibroblasts (DEF) have been described (Glaubiger et al., 1983). MDV strain Md11 has also been 124 described (Silva and Witter, 1985; Tieber et al., 1990; Witter et al., 1983). Oncogenic MDV strain Md11 at passage level 9 (Md11p9) and attenuated Md11 at passage 80 (Md11p80) were kindly provided by the USDA. Avian Diseases and Oncology Laboratory, East Lansing, Michigan. Md11p9 to p12 and Md11p80 to p83 are designated low passage (LP) and high passage (HP), respectively. All infections were initiated with 7.0 x 10‘ PFU per 1x106 DEF in 100 mm tissue culture dish. Because high passage MDV yields higher virus titers in culture than low passage MDV, end point titers were determined at the time of harvest for evaluation of protein, RNA, and transcription rates. Plasmids and Radiolabeled DNA Probes. MDV BamHl plasmids, derived from serotype 1 MDV strain GA (Fukuchie et al., 1984), were a generous gift from M. Nonoyama (Showa University Research Institute for Biomedicine in Florida, St. Petersberg, FL). BamHl-l3 contains coding sequence of the MDV 93 gene (Chen and Velicer, 1991b; Ross et al., 1989). BamHI-K2 contains coding sequence for the MDV TK gene (Ross, . personal communication; Scott et al., 1989). p19MDA2.35 contains the promoter and complete coding sequence of the MDV gC gene (Coussens and Velicer, 1988). lzlE is a cDNA clone containing partial coding sequence of an immediate early gene located within genomic repeat regions (Hang, manuscript in preparation). pA/X6.1 was derived from the MDV BamHl-A clone and contains the MDV ICP4 gene (Coussens, manuscript in preparation). pAl contains the chicken B-actin gene (Cleveland at al., 1980). pBR322 was obtained commercially 125 (Bethesda Research Laboratories Life Technologies, Inc., Gaithersburg, MD). Full length cell-free MDV genomes were isolated as described (Wilson and Coussens, 1991) with the following modifications: 0.8% gels were electrophoresed for 16 hours at 180 volts, 50-90 sec. ramp. Viral bands were excised and electroeluted in an IBI electroelution apparatus, as per manufacturer's specifications (IBI, New Haven, CT). 5’gC and 3’gC gene fragments were isolated from p19MDA2.35. The 5'gC fragment spans nucleotides # 300 to 735. The 3’gC fragment spans nucleotides It 736 to 2350. or-“P labeled DNA probes were prepared using a random prime DNA labeling kit (Boehinger Mannheim, Indianapolis, IN). All radiolabeled DNA probes were prepared from gel purified inserts of plasmids described above. Immunoprecipitations. Cells infected with either LP or HP Md11 and uninfected control cells were labeled for 16 hours with 50 uCi/ml of [”Sjmethionine (EXPRE3533‘S protein labeling mix, 1232.7 Ci/mmol, NEN, Boston, MA) 72 hours post infection. Collection of cells and immunoprecipitations were performed as described (lsfort et al., 1986). Immune chicken sera (ICS) and rabbit antisera against MDV gC (RorA) were generous gifts from L. F. Velicer (Michigan State University, East Lansing, MI). Both sera have been described previously and have been used extensively to characterize MDV antigens (Glaubiger et al., 1983; lsfort et al., 1986; Long et al., 1985; Silva and Lee, 1884). CPM/ul of 3"’S-Iabeled Iysate or media were used to normalize sample volumes between uninfected, high and low 126 passage MDV strain Md11. Immunoprecipitations were analyzed using 10% denaturing polyacrylamide gels (Laemmli, 1970), electrophoresed in tris-glycine buffer, fixed and enhanced using Amplify (N EN, Boston, MA) according to manufacturer's specifications. Dried gels were exposed to X-ray film (Kodak, Rochester, NY) for 24 hrs to 6 days at - 70°C. “C-Iabeled protein standards were used as molecular size markers (Bethesda Research Labs, Bethesda, MD). DNA extractions and restriction fragment length polymorphism (RFLP) analysis. LP and HP Md11 infections of DEF were allowed to progress until cell monolayers appeared fully infected (4-10 days). Total infected cell DNA was isolated as described (Sambrook et al., 1989). The amount of viral DNA in total infected cell DNA was determined by slot-blot assays as described (T ieber et al., 1990). Similar amounts of LP and HP viral DNAs were digested with excess restriction enzymes for 16 hrs according to manufacture’s specifications, electrophoresed through agarose gels and transferred to nylon membranes in 0.4 N NaOH as described (Budowle and Baechtel, 1990). Radiolabeled probes were hybridized to nylon filters at 65°C as described (Budowle and Baechtel, 1990). Filters were exposed to X-ray film for 30 min to 48 hrs at room temperature. Radiolabeled lambda DNA digested with Hind III or BsiEll was used as size markers. 127 RNA extractions and Northern blots. Total RNA from LP and HP Md11 infected and uninfected cells was isolated 96 hours post-infection using the proteinase K/SDS method (Sambrook et al., 1989) followed by DNAse l digestion. RNA samples were treated with DNAse I and 20 ug of total RNA was loaded per well, electrophoresed through formaldehyde gels and transferred to supported nitrocellulose (Schleicher and Scull, Keen, NH) as described previously (Sambrook et al., 1989). Radiolabeled probes were hybridized to filters using high stringency conditions (Sambrook et al., 1989). RNA band sizes were determined by comparison to RNA standards (Bethesda Research Labs, Bethesda MD) run on the same gel. RNA blots were stripped of probe DNA and hybridized with chicken B-actin gene specific probes. Nuclear run-off transcription assays. Nuclei were collected 96 hours post-infection from uninfected and LP or HP Md11 infected DEF. Labeled RNA transcript probes were prepared and hybridized to nitrocellulose (Schleicher and Schull, Keene, NH) strips as described (Pulaski et al., 1992; Stewart of al., 1987). Each nitrocellulose strip contained triplicate spots (5 pg each) of linearized plasmids which contained the genes for MDV 90, 93, TK, 5’90, 390, chicken B-actin, pBR322, and cell-free MDV strain Md11 genomes. Specific plasmids are defined in the Plasmid and Radiolabeled Probe section. 1 28 Data analysis. Intensity of bands on autoradiographs were analyzed on a FB910 Densitometer using Zeineh 1-D Videophoresis ll software (FisherBiotech). Graphs were generated using Sigma Plot version 3.1 (Jandel Scientific, Corte Madera, CA). Where appropriate, statistical analysis was performed with ANOVA using the general linear models procedure of SAS (1985). III. RESULTS Comparison of protein levels from oncogenic and attenuated MDV strain Md11. A decrease in MDV gC protein levels after serial in vitro passage is well documented for many MDV strains and serotypes (lkuta et al., 1983a; lkuta et al., 1983b; Silva and Lee, 1984). To determine if reduction in 90 is also observed in v.v. MDV strain Md1 1, immunoprecipitations were performed using media and cell Iysates from DEF infected with oncogenic (LP) or attenuated (HP) MDV strain Md11. Immune chicken sera (ICS) contains polyclonal antibodies reactive to gB (gpt 00/60/49), p79 (Silva and Lee, 1984), and 9C (gp57-65). The four most predominant proteins visible in immunoprecipitations with ICS from infected cell lysates are gB (gp100/60/49) and p79 (Chen and Velicer, 1991b; Silva and Lee, 1984). Analysis of optical densities from autoradiographs indicated that similar amounts of 9B and p79 were detected in LP and HP Md11 infected cell lysates, suggesting that relative amounts of these proteins are not significantly affected by 129 attenuation of Md11 (Figure 1A: A 4-fold higher end point titer for HP Md11 resulted in higher protein amounts in the H lane.) MDV gC was not detected in ICS immunoprecipitates of media from HP Md11 infected cells (Figure 1A). In contrast, significant amounts of MDV 90 protein were detected in immunoprecipitates of media from LP Md11 infected cells (Figure 1A). To determine if an accumulation of intracellular 90 occurs in HP Md11, rabbit antisera against purified MDV gC (RorA) was used to specifically immunoprecipitate 90 from both media and cell Iysates. The rabbit antisera prepared against gC will also recognize DEF proteins (gC was purified from DEF infected cells for rabbit challenges). Thus, DEF specific bands are visible in all lanes (Figure 13). Levels of MDV 90, associated with infected cells, were 9-fold lower in HP Md11 infected cells than in LP Md11 infected cells (Figure 1B). In addition, comparison of RorA immunoprecipitations of media from HP and LP infected cells indicated a 20-fold reduction in QC protein levels of HP relative to LP infected cell media. These results are consistent with data obtained using ICS, and data for other MDV strains examined previously (Chen and Velicer, 1991 b; Glaubiger et al., 1983; lsfort et al, 1986; Silva and Lee, 1984). RFLP analysis of LP and HP MDV strain Md11. Although both ICS and RM antisera contain polyclanal antibodies reactive to many continuous and discontinuous epitopes on the 90 molecule, gross structural alterations of the MDV gC gene in HP Md11 may cause dismption of major antigenic epitopes resulting in an inability to react with ICS or RorA. To rv-“—‘~1' 130 determine if Md11 gC gene structure was altered during attenuation, restriction fragment length polymorphism (RFLP) analysis were performed. Comparison of restriction fragments from LP and HP Md11 DNA by enzymes which cut outside flanking regions of the Md11 90 gene locus (EcoRl, Bgll, Smal, Hind III, and Xhol) indicate no alteration or polymorphism of 9C flanking regions in HP Md11 (fig. 2B). Evaluation of restriction sites within the Md11 gC gene locus were based upon frequency of cutting, within the GA strain gC gene (Coussens and Velicer, 1988), of a variety of enzymes (Figure 2A). Comparison of BamHl, Smal and Hindlll restriction fragments of LP MDV strain GA and LP Md11 DNA were identical (data not shown), suggesting that overall structure of LP Md11 and LP GA DNA are very similar. A total of 27 restriction sites spanning the 90 gene were assayed using this strategy. This type of analysis can detect changes at the level of 10-50 nucleotides. No polymorphisms were detected within the 9C gene locus of HP Md11 relative to LP Md11 (Figure 23). Data indicated that no gross structural alterations outside of, or within, the Md11 gC gene locus occurred during attenuation of Md11. These data are consistent with previous reports that 90 protein structure was unchanged after attenuation of MDV (Van Zaane et al., 1982). Additional blots were probed with full length MDV genomic probes. For each enzyme used (same as above) multiple changes in banding patterns of HP viral DNA were observed relative to LP viral DNA. All restriction fragments which were altered in HP Md11 DNA map to repeat regions which flank the unique long 131 region of MDV DNA (data not shown). These results were consistent with previous analyses of genome structure changes in HP MDV (Silva and Witter, 1985; Wilson and Coussens, 1991). Steady State 90 RNA levels In LP and HP Md11. To evaluate expression of Md11 gC at the RNA steady state level and determine if steady state RNA levels of Md11 gC were altered in HP Md11 infected cells, northern blots of total RNA isolated from LP and HP Md11 infections were probed with a full length gC gene fragment. Strong hybridization to a 1.8 kb transcript was clearly evident in RNA from LP Md11 infected cells (Figure 3). Weak hybridization to a larger transcript, which is proposed to be a read through message of 9C (Coussens and Velicer, 1988), was also visible. Analysis of optical densities from autoradiographs indicated a 20-fold reduction in steady state 90 RNA in HP infected cells relative to LP Md11 RNA (Figure 3, Table 1). After prolonged exposure of filters to X-ray film, no size differences between LP and HP Md11 90 RNA were observed (Figure 3). To determine if reduction in steady RNA is a general phenomenon of attenuation or specific to gC RNA, additional northern blots of RNA isolated from LP and HP Md11 infected cells were probed with 98, ICP4, and IZIE gene fragments. Analysis of optical densities from autoradiographs indicated a 2-fold reduction of 9B RNA in HP Md11 infections. Steady-state RNA levels of the immediate early genes, ICP4 and lzlE, were also slightly lower in HP Md11 relative to LP Md11 (Table 1). 132 Nuclear run off transcription assays. To test whether the rate of transcription across the 9C gene is reduced in HP Md11 infected cells, nuclear run off transcription assays were performed using nuclei isolated from uninfected, and Md11 LP or HP infected DEF. For each gene tested, differences in transcription rates between LP and HP Md11 genes were statistically significant. Transcription rate of the 9C gene in HP Md11 infected cell nuclei showed the most significant decrease, 6-fold lower than that in LP Md11 infected cell nuclei (f-test, p=0.0001) (Figure 4). In addition, comparison of 5’ and 3’ 9C transcription rates in LP and HP Md11 infected cell nuclei indicated a significant decrease in 5’ (p=0.0001) as well as 3' (p=0.0001) transcriptional efficiencies in HP Md11 infected cell nuclei. 9B and TK transcription rates were 2.4- (p=0.002) and 2.5- (p=0.04) fold lower, respectively, in HP Md11 than in LP Md11 infected cell nuclei. Analysis of total genome transcription revealed a 1 .3- fold decrease (p=0.02) in HP Md11, relative to LP Md11 infected cell nuclei. To determine if premature termination of transcription was occurring in HP Md11 infected cell nuclei, transcription efficiencies of the first 485 bp (5’90) and downstream 916 bp of 90 RNA were tested separately. Theoretically, if premature termination of transcription was absent, % transcription efficiency [%TNS=(O.D.IO.D. 5'gC +3'gC )x100)] would be 35% for 5’gC and 65% for 3’gC. If premature termination of transcription was present, %TNS of 5’ 90 would be greater than 35% and %TNS for 3’gC would be less than 65%. In LP Md11 infected cell nuclei, %TNS for 5'gC and 3’gC was 31% and 69% respectively. In HP Md11 infected cell nuclei, %TNS for 5'gC and 3'gC was 35% and 65% 133 respectively. These data indicated that premature termination of transcription of the 90 gene did not occur in HP Md11 infected cell nuclei. IV. DISCUSSION Attenuation of oncogenic MDV by serial passage in culture offers a unique opportunity to examine mechanisms involved in herpesvirus induced neoplasia and viral attenuation. Changes in MDV genome structure, coincident with attenuation, are well documented. In particular, expansion of a 132 bp repeat sequence within the BamHl-H and -D fragments and a 200 bp deletion in the BamHl-L fragment (all located within repeat regions flanking the unique long) may be important for viral attenuation (Bradley et al., 1989; Chen and Velicer, 1991a; Maotani et al., 1986; Wilson and Coussens, 1991). Additional changes within the MDV genome, coincident with attenuation have not been documented. It has been assumed that reduced expression of MDV 90 during attenuation results from changes in genome structure within or near the MDV gC gene locus. Results presented in this report indicate that, in attenuated MDV strain Md11, no gross structural alterations in genome structure occur within or near the MDV gC gene. In addition, RFLP analysis of DNA from MDV strains JM (serotype 1) and SB-I (serotype 2) indicate no stmctural alteration in the MDV gC gene locus of attenuated isolates (Wilson and Coussens, unpublished data). Thus, it seems unlikely that loss of MDV gC expression in attenuated MDV is due to gross structural alterations within or near the MDV gene locus. Minor changes in nucleotide sequences, however, cannot be completely 134 ruled out on the basis of RFLP analysis presented in this report. Given the similarity of MDV gC gene structure in HP and LP Md11, it seems likely that MDV gC protein expression in HP Md11 was reduced by another mechanism. Possible mechanisms include: gC promoter mutations, reduced protein or RNA stability, altered translational efficiency or mRNA transcription. Analysis of RNA from cells infected with HP and LP MDV strain Md11 indicated a decrease in steady state levels of the 1.8 kb 90 transcript in HP Md11 infected cells, suggesting that reduced MDV gC expression results from a mechanism affecting RNA stability or transcription efficiency. To distinguish between mRNA stability and transcription efficiency, nuclear run off assays were performed. Nuclear run off assays using various M DV genes suggests that HP Md11 may exhibit an overall reduced efficiency of gene transcription, relative to LP Md11. Although 98 protein levels in HP Md11 infected cells were similar to 98 protein levels of LP Md11 infected cells, a 2.5-fold decrease in gB transcription rates was observed in HP Md11 infected cell nuclei. A similar decrease in steady state 98 RNA in HP Md11 infected cells suggests that decreased 98 transcription in HP Md11 infected cell nuclei was responsible for decrease in steady state gB RNA in HP Md11 infected cells. Discrepancies between gB protein levels and 9B transcription rates was due to higher virus titers in HP Md11 infected cells used for immunoprecipitation studies, thus higher viral protein levels resulted from these infections, and protein level differences were not readily apparent. A 6-fold decrease in MDV gC transcription rate suggests that transcription of the MDV gC gene is adversely affected by serial passage of MDV 135 strain Md11. Our nuclear run off assays utilized a full length MDV gC probe as well as 3' and 5' regions of the MDV gC gene. Thus, even prematurely thncated gC mRNA species would have been detected. A 6-fold decrease of the HP Md11 gC transcription rate did not account for the 20-fold decrease observed for 90 protein and 9C steady state RNA levels in HP Md11 infected cells. This discrepancy suggests that decreased RNA stability may also contribute to decreased gC expression in HP Md11 infected cells. Reduced gC transcription rates observed in HP Md11 infected cell nuclei then suggest that reduced 90 expression was related to alteration of 9C RNA stability as well as alteration of 90 gene transcription rates. These alterations would be consistent with minor mutations within the 90 promoter or coding sequence as well as perturbation of an essential MDV encoded regulatory factor which interacts with the 90 gene promoter. The overall reduction in gene expression in HP Md11 infected cells, relative to LP Md11 infected cells further supports the hypothesis that a critical regulatory element is perturbed in HP Md11. The model that alteration of MDV encoded transcription factors affect reduced expression of 90 in HP Md11 is consistent with previous reports that transactivation of an exogenous RSV promoter-CAT construct by HP MDV is 3-fold lower relative to LP Md11 (T ieber et al., 1990). Preliminary results from our laboratory indicate that HP Md11 is unable to transactivate expression of the CAT gene linked to 3 LP MDV 90 promoter. Although RFLP analysis of the 9C gene locus suggests a lack of gross DNA structural alterations, there may be single nucleotide changes in the HP Md11 gC 136 gene locus which affect promoter activity causing a reduction in QC transcription rates. Alternately, well characterized mutations within distinct regions of the HP Md11 genome (Bradley et al., 1989; Chen and Velicer, 1991a; Hirai et al., 1981; Maotani at al., 1986; Silva and Witter, 1985; Wilson and Coussens, 1991) may adversely affect factors important to MDV 90 gene expression. Single nucleotide changes in the HP Md11 gC coding region are unlikely since 1) low levels of unaltered 90 protein, recognized by RorA and ICS, are present in HP Md11. Van Zanne et al. (1982) has established that the decrease in 90 protein levels is a quantitative as opposed to a qualitative difference, 2) low levels of full length gC RNA are present, and truncated RNA were not detected, in HP Md11 infected cells. Transcriptional efficiency of 5’gC and 3'gC in attenuated Md11 are the same as theoretical values and 3) RFLP analysis strongly suggests that the 90 gene locus is unaltered in HP Md11. Ideally, a direct comparison of nucleotide sequence of the LP and HP Md11 90 gene would mle out any contribution of nucleotide base changes in reduction of 90 expression in HP Md11. However, due to the tightly cell-associated nature of MDV and presence of heterogeneous vims populations in HP, attenuated MDV (T anaka, et al., 1980), DNA clones containing the HP 90 gene locus may not represent the attenuated phenotype. [A few plaque isolates of HP MDV strain JM express high levels of gC (Zalinskis and Coussens, unpublished abservations)]. Our suggestion that MDV encoded transcription factors are altered during in vitro attenuation suggests several hypothesis. Alteration of MDV encoded transcription factors may be responsible for loss of oncogenicity as well as final. ' 137 contribute to reduced 90 expression in attenuated serotype 1 MDV. In addition, non-oncogenic serotypes 2 and 3 MDV may produce a slightly different transcription factor which is non-oncogenic but functions to enhance gC transcription. Lastly, the putative transcription factor responsible for efficient expression of 90 may be encoded within viral repeat regions known to be altered during serial in vitro passage of MDV. Elucidation of the precise mechanism responsible for decreased transcription of the 90 gene in attenuated Md11 must await development of efficient methods to purify clonal isolates of MDV, efficient systems to produce site- directed MDV mutants, and characterization of M DV encoded transcription factors. Isolation and characterization of MDV encoded transcription factors and cellular factors induced by MDV will have a significant impact on our understanding of MDV attenuation and possibly on MDV-induced neoplasia. I r. i 138 ACKNOWLEDGMENTS We thank Jessica Marcus for excellent technical assistance. We also thank S. J. Triezenberg for careful review of this manuscript and gratefully acknowledge Telmo Oleas and John Gill for helpful discussions and assistance with statistical analysis. This work was supported by the Michigan Agricultural Experiment Station, the Research Excellence Fund, State of Michigan, and by Grant Nos. 88- 37266-3983, 90-34116-5329, and 92-37204-802 awarded to P. M. Coussens under the Competitive Research Grants Program administered by the US. Department of Agriculture. M. Wilson was supported in part by a Michigan State University Biotechnology Research Fellowship. 139 Figure 1. Immunoprecipitation mass-labeled viral proteins from cell lysates and media of DEF infected with oncogenic or attenuated MDV strain Md11. Immunoprecipitation, denaturing acrylamide gel electrophoresis and flourography were performed as described in Materials and Methods. Antisera used were immune chicken sera (ICS) (Panel A) and rabbit anti-MDVgC (RorA) (Panel B). “C-labeled protein standards were Myosin-Heavy Chain (200 kDa), Phosphorylase B (97.4 kDa), Bovine Serum Albumin (68 kDa), Ovalbumin (43 kDa), and Carbonic Anhydrase (29 kDa). Location of standards are indicated to the right of each panel. D: uninfected DEF, L: low passage (oncogenic) MDV strain Md11 infected DEF, and H: high passage (attenuated) MDV strain Md11 infected DEF. Bands corresponding to previously characterized MDV proteins are indicated by dots to the left of each panel. 140 ONV QOV 3v 89 I I L I r... 4 OZ...— ._ o— ooar 3’» AV. 9% <8: .m— 37 f! . ‘ BIG .— 3.5m."— auv a?! oovnn I a. QC! 2.. NOV ”GO—hm ocuv _I ._ OZ: 4 n: GI. 92%. mo. .< 141 Figure 2. RFLP analysis of the MDV strain Md11 gene locus and flanking regions. Panel A: Location of restriction sites within the MDV gC gene. Bold horizontal line illustrates the MDV gC gene, sizes are in base pairs. Restriction sites are numbered from the BamHI site. Panel B. Autoradiographs from Southern transfer analysis of 0.8% (left) and 1.2% (right) agarose gels. Total DNA from DEF infected with oncogenic low passage (L) and attenuated, high passage (H) MDV strain Md11 was isolated, digested with enzymes indicated, electrophoresed, transferred to nylon membranes, and hybridized to a radiolabeled full length MDV gC gene fragment as described in Materials and Methods. Enzymes are indicated above each low and high passage pair. Fragment sizes (kbp) of radiolabeled BstEll or Hind Ill digested lambda DNA standards are indicated by arrows to the right of each autoradiograph. Ins. line are sler )EF IDV ;ed. IDV tled aled the 142 o 250 500 750 1000 1250 1500 1750- 2000 CAP ATG BamHl 5‘3“ I r IF’ I I l I I l I 310 H II | | pa 399 818 1536 Hae III I I I 109 1031 1095 Hinfl I I I 347 812 854 1243 1616 1654 1925 Fokl I I 262 434 564 1086 1393 1745 Hhal | l 76 455 457 790 1492 Clal I 1543 TOTAL II I IIII I III III | I IIIII I | B. \ \ \ \\ \ \ \\ \\‘ \ \ \ \ \ \ \ o \ 0 .(P o e o o o e 1- 0 4° 9° 9“ x» o‘ 9‘ +“ 9° 3° 8 4° 9“ 'L H' T117 r1. Hi Ii. 117 IL 11' ll. HI .- ‘1 I’ Q - .1 11 ‘. I. .’ 423.13 49.42 -. 46.56 44.36 C. -4232 . -.<2.03 I Figure 2. ‘ IL HI IL HI IL HI IL HI IL HI IL "1 I. s ' . * . '. I . . -1 ‘ _ u , _ ‘ , v -- up - 143.68 42.32 41.93 0. t 41. 7 , .13. " - up- - - - - g ”40.70 a... 3“}: .' a? :- fl 0' Q t“ , n .2 1" 40.22 _50 KB: SIZE INDEPENDENT >50 KB: SIZE DEPENDENT MIGRATION MIGRATION Pr .0... 0.... f g Q Q o O O O O ' Q . . . REPTATION O . Q Q ‘ . . o g g P O o o o O o o o O o o CONVENTIAL CHEF - = = = = = = = C: - _ - - - _ - _ Figure 2. 169 IMMUNOPRECIPITATION OF PROTEINS FROM MDV INFECTED CELLS p5S-methionine 16 hrs ME IA / MDV INFECTED CELLS ® \ CE "I LL LYSATE ff} a: SPECIFIC ANTISERA . AQ/ I y '1' 7* STAPHA ........ ......... 22's.". SPIN, RINSE, @ HEAT @ F—l I——1 = Ax°k .... j“. Figure 3. 170 NUCLEAR RUN OFF TRANSCRIPTION ASSAY NRO MDV INFECTED CELLS /- %- , dATP, dCTP, dGTP, radiolabeled dUTP ISOLATE RNA W ff 6) SLOT BLOT m o HYBRIDIZATION AUTORADIOGRAPHY AND DENSITOMETRY Figure 4. LIST OF REFERENCES [Lt-r. .r “d LIST OF REFERENCES Aihara,S., C. M. Rao, Yu-Yx, T. Lee, K. Watanabe, T. Komiya, H. Sumiyoshi, H. Hashimoto, and A. Nomoto. (1991). Identification of mutations that occurred on the genome of Japanese encephalitis virus during the attenuation process. Virus genes. 5:95-109. Anderson, A. S., A. Francesconi, and R. W. Morgan. (1992). Complete nucleotide sequence of the Marek’s disease virus ICP4 gene. Virology 189:657-667. Andino, R., G. E. Reickhof, D. Trono, and D. Baltimore. (1990). Substitution in the protease (3PM) gene of poliovirus can suppress a mutation in the 5’ noncoding region. J. Virology 64:607-612. Bacon, L. D., R. L. Witter, and A. M. Fadly. (1989). Augmentation of retrovirus induced lymphoid Ieukosis by Marek's Disease herpesvirus in white leghom chickens. J. Virol. 63:504-512. Binns, M. M. and N. L. J. Ross. (1989). Nucleotide sequence of the Marek's disease virus (MDV) B-1 B A-antigen and the identification of the MDV A antigen as the herpes simplex vims-1 glycoprotein C homologue. Virus Research. 12:371- 382. Bishop, J. M. (1987). Trends in oncogenes. In Oncogenes and growth factors, (R. A. Bradshaw and S. Prentis, eds.), Elsevier Science Publ. Co., New York. Blair, E. D. and B. W. Snowden. (1991). Comparative analysis of the parameters that regulate expression from promoters for two late HSV-1 gene products. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida. Bostock, C. J. (1988). Parameters of field inversion gel electrophoresis for the analysis of pox virus genomes. Nucl. Acids Res. 16:4239-4252. Bradley, G., G. Lancz, A. Tanaka, and M. Nonoyama. (1989). Loss of Marek’s disease virus tumorigenicity is associated with truncation of RNAS transcribed within BamHl- H. J. Virology 63:4129-4135. 171 172 Bradley, G., M. Hayashi, G. Lancz, A. Tanaka, and M. Nonoyama. (1989). Structure of the Marek's Disease virus BamHI-H gene family: Genes of putative importance for tumor induction. J. Viral. 63:2534-2542. Buckmaster, A. E., S. D. Scott, M. J. Sanderson, M. E. G. Boursnell, N. L. J. Ross, and M. M. Binns. (1988). Gene Sequence and mapping data from Marek's disease virus and herpesvirus of turkeys: Implications for herpesvirus classification. J. gen. Virol. 69: 2033-2042. Budowle, B. and F. S. Baechtel. (1990). Modifications to improve the effectiveness of restriction fragment length polymorphism typing. Appl. Theor. Electra. 1 :181-187. Bulow, V. van, and P. M. Biggs. (1975). Precipitating antigens associated with Marek's disease virus and a herpesvirus of turkeys. Avian Pathol. 4:147-162. Bulow, V. van, and P. M. Biggs. (1975). Differentiation between strains of Marek’s disease virus and turkey herpesvirus by immunofluorescence assays. Avian Pathol. 4:133-146. Burkitt,D. (1962). A children’s cancer dependent upon climatic factors. Nature (London) 194:232. Cai, W. and P. A. Schaffer. (1992). Herpes simplex virus type 1 ICPO regulates expression of immediate-early, early, and late genes in productively infected cells. J. Viral. 66:2904-2915. Calnek, B. W., K. A. Schat, L. J. N. Ross,W. R. Shek, and C. L. H. Chen. (1984). Further characterization of Marek’s Disease virus infected lymphocytes. I. In vivo infection. lntl. J. Cancer 33:389-398. Calnek, B. W., and R. L. Witter. (1991). Diseases of Poultry: Marek's disease, pp. 342-385. (B. W. Calnek, H. J. Barnes, C. W. Beard, W. M. Reid, and H. W. Yoder, Jr., eds.) Iowa State University Press, Ames, Iowa. Camacho, A. and P. G. Spear. (1978). Transformation of hamster embryo fibroblasts by a specific fragment of the herpes simplex virus genome. Cell 15:993. Cann, A. J., G. Stanway, P. J. Hughes, P. D. Minor, D. M. A. Evans, G. C. Schild, and J. W. Almond. (1984). Reversion to neurovirulence of live-attenuated Sabin type 3 oral poliovirus vaccine. Nucleic Acids Res. 12:7787-7792. Card, J. P., M. E. Whealy, A. K. Robbins, and L. W. Enquist. (1992). Pseudorabies virus envelope glycoprotein gl influences both neurotropism and virulence during infection of the rat visual system. J. Virology 66:3032-3041. 173 Carter, J. K. and R. F. Silva. (1990). Cell culture amplification of a defective Marek’s Disease virus. Virus Genes 4:225-237. Cebrian, J., C. Kaschka-Dierich, N. Berthelot, and P. Scheldrick. (1982). Inverted repeat nucleotide sequences in the genome of Marek’s Disease Virus and the herpesvirus of the turkey. Proc. Natl. Acad. Sci. USA 79:555-558. Chen, X. and L. F. Velicer. (1991). Multiple bi-directional initiations and terminations of transcription in Marek's disease virus long repeat regions. J. Virol. 65 :2445-2451 . Chen, X. P. J. Sondermeijer and L. F. Velicer. (1992). Identification of a unique Marek’s disease virus which encodes a 38 kilodalton phosphoprotein and is expressed in both lytically infected cells and latently infected Iymphoblastoid tumor cells. J. Viral. 66:85-94. Chen, X. an L. F. Velicer. (1993). The Mark’s disease virus B antigen complex (gp 100, gp 60, gp 49) is the homolog of the herpes simplex virus glycoprotein B (98). (in press). Christodoulou, C., F. Colbere-Garapin, M. Macadam, L. F. Taffs, S. Marsden, P. Minor, and F. Horaud. (1990). Mapping of mutations associated with neurovirulence in monkeys infected with Sabin 1 poliovirus revertants selected at high temperature. J. Viral. 64:4922-4929. Chu, G., D. Vollrath, and R. W. Davis. (1986). Separation of large DNA molecules by contour-clamped homogeneous electric fields. Science 234:1582-1585. Churchill, A. E. and P. M. Biggs. (1967). Agent of Marek's disease in tissue culture. Nature (London) 215:528-530. Churchill, A. E., R. C. Chubb, and C. W. Baxendale. (1969). The attenuation with loss of oncogenicity of the herpes type virus of Marek’s disease (strain HPRS-16) on passage in cell culture. J. gen. Virology. 4:557-564. Cleveland, D. W., M. A. Lopata, R. J. MacDonald, N. J. Cowan, W. J. Rutter, and M. W. Kirschner. (1980). Number and evolutionary conservation of alpha- and beta-tubulin and cytoplasmic beta- and gamma- actin genes using specific cloned cDNA probes. Cell 20:95-105. Collingham, K. E., T. M. Pollock, and M. O. Roebuck. (1978). Paralytic poliomyelitis in England and Wales. Lancet i:976-977. Coussens, P. M., V. L. Tieber, C. S. Mehigh, and M. Marcus. (1991). Identification of a novel transcription factor, ACF, in cultured avian fibroblast cells that interacts 174 with a Marek's disease virus late gene promoter. Virology 185:80-89. Coussens, P. M. and L. F. Velicer. (1988). Structure and complete nucleotide sequence of the Marek’s disease herpesvirus gp 57-65 gene. J. Viral. 62:2373- 2379. Coussens, P. M., M. R. Wilson, H. Camp, H. Roehl, R. J. lsfort, and L. F. Velicer. (1990). Characterization of the gene encoding herpesvirus of turkeys gp 57-65: Comparison to Marek’s disease virus gp 57-65 and herpes simplex vims glycoprotein C. Virus Genes 32291-307. Cui, 2., L. F. Lee, J-L Liu, and H-J Kung. (1991). Structural analysis and transcription mapping of the Marek's disease virus gene encoding pp38, an antigen associated with transformed cells. J. Viral. 65:6509-6515. Dallo, S. and M. Estbam. (1987). Isolation and characterization of attenuated mutants of vaccinia virus. Virology 159:408-422. Elmubarak, A. K., J. M. Sharma, R. L. Witter and V. L. Sanger. (1981). Failure of herpesvirus of turkeys to protect turkeys against Marek's Disease. Am. J. Vet. Res. 42:2122-2128. Epstein, M. A., B. G. Achong, and Y. M. Barr. (1964). Virus particles in cultured lymphoblasts from Burkitt’s lymphoma. Lancet i:702. Evans, D. M. A., G. Dunn, P. D. Minor, G. C. Schild, A. J. Carin, G. Stanway, J. W. Almond, K. Currey, and J. V. Maizel. Jr. (1985). Increased neurovirulence associated with a single nucleotide change in a non-coding region of the Sabin type 3 poliovaccine genome. Nature (London) 314:548-550. Everett, R. D., C. M. Preston, and N. D. Snow. (1991). Functional and genetic analysis of the role of VMW110 in herpesvirus replication. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida. Flint, S. J. (1982). Transformation by adenoviruses. In RNA tumor viruses 2"" Edition, Cold Spring Harbor laboratory, New York. Fraenkel-Conrat, H., P. C. Kimball, J. A. Levy. (1988). Virology. 2"" Edition, Prentice-Hall, Inc., New Jersey. Fukuchi, K. M. Sudo, Y. -S. Lee, A. Tanaka, and M. Nonoyama. (1984). Stmcture of the Marek’s disease virus DNA: Detailed restriction enzyme map. J. Viral. 51:102-109. Fynan, E., T. M. Block, J. DuHadaway, W. Olson, and D. L. Ewart. (1992). 175 Persistence of Marek’s Disease virus in a subpopulation of B cells that is transformed by avian leukosis virus, but not normal bursal cells. J. Viral. 66:5860- 5866. Glaubiger, C., K. Nazerian, and L. F. Velicer. (1983). Marek's disease herpesvirus: IV. Molecular characterization of Marek’s disease herpesvirus A antigen. J. Viral. 45:1228-1234. Grodzicker, T. and N. Hopkins. (1981). Origins of contemporary DNA tumor virus research. In DNA tumor viruses 2"6 Edition, Cold Spring Harbor laboratory, New York. Harris, A. and D. R. Bentley. (1988). Separation of episomal Epstein-Barr virus from Burkitt’s lymphoma host cell DNA in pulsed field gels. Nucl. Acids Res. 16: 4172. Hayward, S. D. and J. M. Hardwick. (1991). Epstein-barr virus transactivators and their role in reactivation. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida Henderson, L. M., R. L. Levings, A. J. Davis, and D. R. Sturtz. (1991). Recombination of pseudorabies virus vaccine strains in swine. Am. J. Vet. Res. 52:820-825. Henle, G. and W. Henle. (1966). Studies on cell lines derived from Burkitt’s lymphoma. Trans. N. Y. Acad. Sci. 29:71-79. Hirai, K., K. lkuta, K. Maotani, and S. Kata. (1984). Evaluation of DNA homology of Marek's disease virus, herpesvirus of turkeys and Epstein-Barr virus under varied stringent hybridization conditions. J. Biochem. 95:1215-1218. Hirai, K., K. lkuta, and S. Kata. (1981). Structural changes of the DNA of Marek’s disease virus during serial passage in cultured cells. Virology 115:385-389. Hirai, K., K. lkuta, and S. Kata. (1979). Comparative studies on Marek’s Disease virus and herpesvirus of turkey DNAs. J. Gen. Virol. 45:119-131. Hirai, K., M. Yamada, Y. Arao, S. Kato, and S. Nii. (1990). Replicating Marek's disease virus (MDV) serotype 2 DNA with inserted MDV serotype 1 DNA sequences in a Marek’s disease lymphoblastoid cell line MSBt-41C. Arch. Virol. 114:153-165. Hirt, B. (1967). Selective extraction of polyoma DNA from infected mouse cell cultures. J. Mol. Biol. 26:365-369. 176 Hovi, T., A. Pitkaranta, A. Macadam, P. Minor, and J. Almond. (1991). Covariance of lowered capacity to induce interferon in human Ieucocytes and temperature- sensitivity of type 3 poliovirus. J. Interferon Res. 11:105-110. lgarashi, T., M. Takahashi, J. Donovan, J. Jessip, M. Smith, K. Hirai, A. Tanaka, and M. Nonoyama. (1987). Restriction enzyme map of herpesvirus of turkeys and its collinear relationship with Marek’s disease virus DNA. Virology 157:351- 358. lkuta, K. S., S. Ueda, S. Kato, and K. Hirai. (1983). Monoclonal antibodies reactive with the surface and secreted glycoproteins of Marek's disease virus and herpes virus of turkeys. J. gen. Virol. 64: 2597-2610. lkuta, K. S., S. Ueda, S. Kato, and K. Hirai. (1983). Most virus-specific polypeptides in cells productively infected with Marek’s disease virus or herpesvirus of turkeys possess cross-reactive determinants. J. gen Virol. 64:961- 965. lsfort, R. J., R. A. Stringer, H. -J. Kung, L. F. Velicer. (1986). Synthesis, processing and secretion of the Marek's disease herpesvirus A antigen glycoprotein. J. Viral. 57:464-474. lsfort, R., D. Jones, R. Kost, R. Witter, and HJ Kung. (1992). Retrovims insertion into herpesvirus in vitro and in viva. P. N. A. 8. USA 89:991-995. lsfort, R. J., D. Robinson, H.-J. and Kung. (1990). Purification of genomic sized herpesvirus DNA using pulsed-field electrophoresis. Journ. Virol. Methods 27:311-318. lwata, A., S. Ueda, A. lshihama, and K. Hirai. (1992). Sequence determination of cDNA clones of transcripts from the tumor-associated region of the Marek's disease virus genome. Virology 187:805-808. Jones, D., L. Lee, J-L Liu, H. J. Kung, and J. K. Tillotson. (1992). Marek’s disease virus encodes a basic-leucine zipper gene resembling the fos/jun oncogenes that is highly expressed in Iymphoblastoid tumors. P. N. A. S. USA 89:4042-4046. Kato, S. and K. Hirai. (1985). Marek’s disease vims. In "Advances in Virus Research" vol. 30, pp. 225-277 (K. Maramorosch, F. A. Murphy, and A. J. Shatkin, eds. ), Academic Press. Katz, J. B., L. M. Henderson, and G. A. Erickson. (1990). Recombination in viva of pseudorabies vaccine strains to produce new virus strains. Vaccine 8:286-288. Kawai, S. and M. Niahizawa. (1984). New procedures for DNA transfection with polycation and dimethylsulfoxide. Mol. Cell. Biol. 4:1172-1174. 177 Kuhn, R. J., D. E. Griffin, H. Zhang, H. G. Nesters, and J. H. Strauss. (1992). Attenuation of sinibis virus neurovirulence by using defined mutations in non- translated regions of the genome RNA. J. Viral. 66:7121-7127. Laemmli, U. K. (1970). Cleavage of stmctural proteins during the assembly of the head of bacteriophage T4. Nature (London) 227: 680-685. Lee, L. F., X. Liu, and R. L. Witter. (1983). Monoclonal antibodies with specificity for three different serotypes of Marek’s disease virus in chickens. J. Immunol. 130:1003-1006. Lee, M. S., J. M. Roos, L. C. McGuigan, K. A. Smith, N. Corrnier, L. K. Cohen, B. E. Roberts, and L. G. Payne. (1992). Molecular attenuation of vaccinia virus: mutant generation and animal characterization. J. Viral. 66:2617-2630. Lee, L. F., E. D. Kieff, S. L. Bachenheimer, B. Roizman, P. G. Spear, B. R. Burmester, and K. Nazerian. (1971). Size and composition of Marek's disease virus deoxyribonucleic acid. J. Virol. 7:289-294. Lobigs, M., R. Usha, A. Nestorowicz, I. D. Marshall, R. C. Weir, and L. Dalgarno. (1990). Host cell selection of Murray Valley encephalitis virus variants altered at an RGD sequence in the envelope protein and in mouse virulence. Virology 1 76:587-595. Long, P. A., J. L. Clark, and L. F. Velicer. (1975). Marek’s disease herpesvirus. II. Purification and further characterization of Marek’s disease herpesvirus A antigen. J. Viral. 15:1192-1201. Macadam, A. J., G. Ferguson, J. Burlison, D. Stone, R. Skuce, J. W. Almond, and P. D. Minor. (1992). Correlation of RNA secondary structure and attenuation of Sabin vaccine strains of poliovirus in tissue culture. Virology 189:415-422. Maotani, K., A. Kanamori, K. lkuta, S. Ueda, S. Kato, and K. Hirai. (1986). Amplification of tandem direct repeats within inverted repeats of Marek's disease virus DNA during serial in viva passage. J. Viral. 58:657-660. Maray, T., M. Malkinson, and Y. Becker. (1988). RNA transcripts of Marek's disease virus (MDV) serotype-1 in infected and transformed cells. Virus genes 2:49-68. Martin, G. S. (1970). Rouse sarcoma virus: a function required forthe maintenance of the transformed state. Nature (London) 227:1021-1023. McCall, K. A., B. W. Calnek, W. V. Harris, K. A. Schat, and L. F. Lee. (1987). Expression of a tumor-associated surface antigen on normal versus Marek’s 178 disease virus-transformed lymphocytes. J. Natl. Cancer Inst. 79:991-1000. Melnick, J. L. (1982). Towards the eradication of poliomyelitis. in Medical Virology, pp 261 -299 (L. Maza and E. Peterson, eds. ) Elsevier Science Publishing Co., Inc. Mettenleiter, T. C., C. Schreurs, F. Zuckermann, and T. Ben-Porat. (1987a). Role of pseudorabies virus glucoprotein gl in virus release from infected cells. J. Virol. 61 :2764-2769. Mettenleiter, T. C., L. Zsak, A. S. Kaplan, T. Ben-Porat, and B. Lomniczi. (1987b). Role of a structural glycoprotein of pseudorabies in vims virulence. J. Virol. 61 :4030-4032. Mikiko, K., M. Hayashi, T. Fumichi, M. Nonoyama, E. Isogai, and S. Namioka. (1991). The inhibitory effects of oligonucleotides, complimentary to Marek's disease virus mRNA transcribed from the BamHI-H region, on the proliferation of transformed Iymphoblastoid cells, MDCC-MSBI. J. Gen. Virol. 72:1105-1111. Morgan, R. W., J. Gelb Jr., C. S. Schreurs, and P. J. A. Sondermeijer. (1992). A live recombinant HVT vaccine expressing the newcastle disease virus fusion proteins protects chickens from both newcastle disease and Marek's Diseases. Proc. XIX World’s Poultry Congress, vol. 1, Pansen and Looijen, Wageningen, The Netherlands. Moriuchi, H., M. Moriuchi, H. A. Smith, S. E. Straus, and J. I. Cohen. (1992). Varicella-zoster virus open reading frame 61 protein is functionally homologous to herpes simplex virus type 1 ICPO. J. Viral. 66:7303-7308. Moss, B. (1990). Poxviridae and their replication. In Fundamental Virology. 2"‘1 Edition, (B. Fields and D. M. Knipe et al., eds.) Raven Press, Ltd, New York. Mass, B., E. Winters, and N. Cooper. (1981). Instability and reiteration of DNA sequences within the vaccinia virus genome. P. N. A. S. USA 78:1614-1618. Moss, E. G., R. E. O’Neill, and V. R. Racaniello. (1989). Mapping of attenuating sequences of an avirulent poliovirus type 2 strain. J. Viral. 63:1884-90. Moss, E. G. and V. R. Racaniello (1991). Host range determinants located on the interior of the poliovirus capsid. EMBO J. 10:1067-1074. Nazerian, K. (1970). Attenuation of Marek’s disease virus and study of its properties in two different cell cultures. J. Natl. Cancer Inst. 44:1257-1267. Nazerian, K. (1980). Viral Oncology: Marek's disease: a herpesvirus induced malignant lymphoma of the chicken. p. 665-682. (G. Kein, ed.), Raven Press. New York. 179 Nazerian, K., and R. F. Silva. (1988). Properties of producer and non-producer clones of a Marek's disease turkey Iymphoblastoid cell line. Avian diseases 32:486-493. Nazerian K. and R. L. Witter. (1970). Cell-free transmission and in viva replication of Marek’s disease virus (MDV). J. Virol. 5:388-397. Nazerian K. and B. R. Burmester. (1968). Electron microscopy of a Herpesvims associated with the agent of Marek’s disease in cell culture. Cancer research 28:2454-2462. Nazerian, K., L. F. Lee, N. Yanagida, and R. Ogawa. (1992). Protection of Marek's Disease by a fowlpox virus recombinant expressing the glycoprotein B of Marek’s disease virus. J. Viral. 66:1409-1413. Neumann, E., M. Schaefer-Ridder, Y. Wang, and P. H. Hofschneider. (1982). Gene transfer into mouse lyoma cells by electroporation in high electric fields. EMBO J. 1:841-845. Nitayaphan, S., J. A. Grant, G. J. Chang, and D. W. Trent. (1990). Nucleotide sequence of the virulent SA-14 strain of Japanese encephalitis virus and its attenuated vaccine derivative, SA-14-14-2. Virology 177:541-542. Ohashi, K. and K. A. Schat. (1992). cDNA clones derived from the Marek’s disease virus tumor cell line MDCC-CU41. Proc. XIX World’s Poultry Congress, vol. 1, Pansen and Looijen, Wageningen, The Netherlands. Pattison, M. (1985). Control of Marek’s disease by the poultry industry: Practical considerations. In Marek’s Disease, scientific methods and basis of control. (L. N. Payne, ed), Martinus Niijhoff Publishing, Boston, Mass. Payne, L. N. (1985). Historical review. In Marek’s Disease, scientific methods and basis of control. (L. N. Payne, ed.), Martinus Niijhoff Publishing, Inc., Boston, Mass. Payne, L. N. (1982). Biology of Marek’s disease virus and the herpesvirus of turkeys. In "The Herpesviruses" (B. Roizman, ed.), Vol. 1 pp. 347-431. Plenum, NY. Payne, L. N. (1985). Pathology. In Marek’s Disease, scientific methods and basis of control. (L. N. Payne, ed), Martinus Niijhoff Publishing, Inc., Boston, Mass. Peng, F., G. Bradley, A. Tanaka, G. Lancz, and M. Nonoyama. (1992). Isolation and characterization of cDNAs from the BamHI-H gene family RNAS associated with tumorigenicity of Marek’s disease virus. J. Virol. 66:7389-7396. 180 Powell, P. C., K. Howes, A. M. Lawn, B. M. Mustill, L. N. Payne, M. Rennie, and M. A. Thompson. (1984) Marek's disease in turkeys: the induction of lesions and the establishment of lymphoid cell lines. Avian Pathol. 13:201-214. Powell, P. C., L. N. Payne, J. A. Frazer, and M. Rennie. (1974). T-lymphoblastoid cell lines from Marek’s disease Lymphomas. Nature (London) 251 :79-80. Pratt, W. D., R. W. Morgan, and K. A. Schat. (1992). Characterization of reticuloendotheliosis vims-transformed avian T Iymphoblastoid cell lines infected with Marek's disease vims. J. Virol. 66:7239-7244. Pritchard, A. E., M. A. Calenoff, S. Simpson, K. Jensen, and H. L. Lipton. (1992). A single base deletion in the 5’ non-coding region of Theiler’s virus attenuated neurovirulence. J. Virol. 66:1951-1958. Pruthi, A. K., R. K. Gupta, and J. R. Sadana. (1989). Studies on the pathology of Marek’s Disease following challenge in chicks vaccinated with three different vaccines. J. Comp. Pathol. 101:295-305. Pulaski, J. T., V. L. Tieber, and P. M. Coussens. (1992). Marek’s disease virus- mediated enhancement of avian leukosis virus gene expression and virus production. Virology 186:1 13-121. Purchase, H. C., B. R. Burmester, and C. H. Cunningham. (1971). Responses of cell cultures from various avian species to Marek's disease virus and herpesvirus of turkeys. Am. J. Vet. Res. 32:1821-1823. Purchase, H. G. (1984). Clinical disease and its economic impact. In Marek’s Disease, scientific methods and basis of control. (L. N. Payne, ed.,) Martinus Niijhoff Publishing, Boston, Mass. Putnak, J. R. and B. A. Phillips. (1981). Picornaviral structure and assembly. Microbiol. Rev. 45:287-315. Pyles, R. B., N. M. Sawtell, and R. L. Thompson. (1992). Herpes simplex virus type 1 dUTPase mutants are attenuated for neurovirulence, neuroinvasiveness, and reactivation from latency. J. Virol. 66:6706- 6713. Rhozinski, J., L. E. Girton, and J. L. Van Etten. (1989). Chlorella viruses contain linear nonpermutated double-stranded genomes with covalently closed hairpin ends. Virology 168:363-369. Raab-Traub, N. and K. Gilligan. (1991). Comparison of EBV transcription in lymphoid and epithelial cells. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc, Florida. 181 Reychan, W. T., P. Ling, P. R. Kinchington, and J. Hay. (1991). The correlation between varicella-zoster virus transcription and the sequence of the viral genome. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida. Robbins, A. K., J. P. Ryan, M. E. Whealy, and L. W. Enquist. (1989). The gene encoding the glll envelope protein of Pseudorabies virus vaccine strain BaItha contains a mutation affecting protein localization. J. Virol. 63:250-258. Robertson, L. M., A. R. MacLean, and S. M. Brown. (1992). Peripheral replication and latency reactivation kinetics of the non-neurovirulent herpes simplex virus type 1 variant 1716. J. Gen. Virol. 73:967-70. Roizman, B. (1992). The family Herpesviridae: an update. Arch. Virol. 123:425-449. Roizman, B. . (1991). Fundamental Virology 2nd Edition: Herpes simplex viruses and their replication. pp. 849-895. (B. N. Fields, D. M. Knipe, eds.), Raven Press, Ltd., New York. Roizman, B. (1990). Herpesviridae: A bn’ef introduction. In Fundamental Virology, 2"‘1 Edition, (B. Fields and D. M. Knipe et al., eds.) Raven Press, Ltd, New York. Roizman, B. and A. E. Sears (1990). Herpes simplex viruses and their replication. In Fundamental Virology, 2"“ Edition, (B. Fields and D. M. Knipe et al., eds.) Raven Press, Ltd, New York. Ross, L. J. N., M. Sanderson, S. D. Scott, M. M. Binns, T. Doel, and B. Milne. (1989). Nucleotide sequence and characterization of the Marek’s disease virus homologue of glycoprotein B of herpes simplex virus. J. gen. Virol. 70:1789-1804. Ross, L. J. N., Milne, B., and Schat, K. A. (1984). Restriction enzyme analysis of MDV DNA and homology between strains. In ”Intemational Symposium on Marek's Disease“ (B. W. Calnek and J. L. Spencer, eds), pp. 51-67. American Association of Avian Pathologists, Kennett Square, PA. Sabin, A. B. and L. R. Boulger. (1973). History of Sabin attenuated poliovirus oral live vaccine strains. J. Biolo. Standardization 1:115-118. Sabin, A. B. (1957). Present status of attenuated live virus poliomyelitis vaccine. Bull. NY Acad. Med. 33:17-39. Salk, J. E. (1953). Principels of immunization as applied to poliomyelitis and influenza. J. Public Health 43:1384-1398. Sambrook, J., E. F. Fritsch, and T. Maniatis. (1989). ”Molecular Cloning: A 182 Laboratory Manual. " Cold Spring Harbor Laboratory, Cold Spring Harbor, NY. Sandri-Goldin, R. M. (1991). Analysis of the regulation activities of the HSV-l a protein ICP27. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida. SAS User's Guide: Statistics, version 5. (1985). SAS lnst., Inc., Cary, N. C. Schat, K. A., C-L. H. Chen, B. W. Calnek, and D. Char. (1991). Transformation of T-Iymphocyte subsets by Marek's disease herpesvirus. J. Virol. 65:1408-1413. Schat, K. A. (1985). Characteristics of the virus. In Marek's Disease, Scientific Basis and Methods of Control (L. N. Payne, ed.) Martinus Nijhoff Publishing, Inc, Boston, Mass. Schat, K. A., A. Buckmaster, and L. J. N. Ross. (1989). Partial transcription map of Marek’s disease herpesvims in lytically infected cells and lymphoblastoid cell lines. lntl. J. Cancer 44:101-109. Schat, K. A., B. W. Calnek, J. Fabricant, and D. L. Grahm. (1985). Pathogenesis of infection with attenuated Marek's disease vims strains. Avian pathol. 14:127- 146. Schwartz, D. C. and C. R. Cantor. (1984). Separation of yeast chromosome-sized DNAs by pulsed field gradient gel electrophoresis. Cell 37:67-75. Scott, S. D., N. L. J. Ross, and M. Binns. (1989). Nucleotide and predicted amino acid sequence of the Marek's Disease virus and turkey herpesvirus thymidine kinase genes; comparison with thymidine kinase genes of other herpesviruses. J. gen. virol. 70:3055-3065. Sharma, J. M. (1989). In situ production of interferon in tissues of chickens exposed as embryos to turkey herpesvirus and Marek’s disease virus. Am. J. Vet. Res. 50:882-886. Shek, W. R., B. W. Calnek, K. A. Schat, and C. L. H. Chen. (1983). Characterization of Marek's Disease virus infected lymphocyteszdiscrimination between cylolytically and latently infected cells. J. Natl. Cancer Inst. 70:485-491 Shope, R. E. (1932). A Filterable virus causing tumor-like conditions in rabbits and its relationship to virus myxomatosum. J. Exp. Med. 56:803. Silva, R. F. and L. F. Lee. (1984). Monoclonal antibody-mediated immunoprecipitation of proteins from cells infected with Marek's disease virus or turkey herpesvirus. Virology 136:307-320. 183 Silva, R. F. and R. L. Witter. (1985). Genomic expansion of Marek’s disease vims DNA is associated with serial in vitro passage. J. Virol. 54:690-696. Smiley, J. R., B. Panning, and C. A. Smibert. (1991). Regulation of cellular genes by HSV products. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida. Spear, P. and B. Roizman. (1981). Herpes simplex viruses. In DNA tumor viruses 2"‘1 Edition, Cold Spring Harbor laboratory, New York. Stevens, J. G., E. K. Wagner, G. Devi-Rao, M. L. Cook, and L. Feldman. (1987). RNA complementary to herpesvirus a-gene mRNA is predominant in latently infected neurons. Science 235:1056-1059. Stewart, C. J., M. Ito, S. E. Conrad. (1987). Evidence for transcriptional and post- transcriptional control of the cellular thymidine kinase gene. Mol. Cell. Biol. 7:1 156- 1163. Sugaya, K., G. Bradley, M. Nonoyama, and A. Tanaka. (1990). Latent transcripts of Marek’s disease virus are clustered in the short and long repeat regions. J. Virology 64:5773-5782. Svitkin, Y. V., N. Cammack, P. D. Minor, and J. W. Almond. (1990). Translation deficiency of the Sabin type 3 poliovirus genome: association with an attenuating mutation C 472 -- U. Virology 175:103-109. Takehara, K., M. K. Min, J. K. Battles, K. Sugiyama, V. C. Emery, J. M. Dalrympal, and D. H. Bishop. (1989). Identification of mutations in the M RNA of a candidate vaccine strain of Rift Valley fever vims. Virology 169:452-457. Tanaka, A., Y-S. Lee, and M. Nonoyama. (1980). Heterogeneous populations of Virus DNA in serially passaged Marek's Disease virus preparations. Virology 103:510-513. Tieber, V. L., L. L. Zalinskis, R. F. Silva, A. Finkelstein, and P. M. Coussens. (1990). Transactivation of the Rous Sarcoma virus long terminal repeat promoter by Marek's disease vims. Virology 179:719-727. Tillotoson, J. K., H-J Kung, and L. F. Lee. (1988). Accumulation of viral transcripts coding for a DNA binding protein in Marek’s disease tumor cells. In Advances in Marek’s disease virus research (S. Kato, T. Horiuchi, T. Mikami, and K. Hirai, eds.), Japanese Assoc. on Marek’s disease, Osaka, Japan. Tooze, J. (1981). DNA tumor viruses. Cold Spring Harbor laboratory, New York. 184 Toyoda, H., M. Kohara, Y. Kataoka, T. Suganuma, T. Omata, N. Imura, and A. Nomoto. (1984). Complete nucleotide sequence of all three poliovirus serotype genomes. J. Mol. Biol. 174:561-585. Tyler, K. L. and B. Fields. (1990). Pathogenesis of viral infections. In Fundamental Virology, 2"‘1 edition, (B. Fields and D. M. Knipe et al., eds.) Raven Press, Ltd, New York. vanZaane, D., J. M. A. Brinkhof, A. L. J. Gielkens. (1982). Molecular-biological characterization of Marek’s disease virus. ll. Differentiation of various MDV and HVT strains. Virology 121:133-146. Velicer, L. F., D. R. Yager, and J. L. Clark. (1978). Marek's disease herpesviruses. llI. Purification and characterization of the Marek’s disease herpesvirus B antigen. J. Virol. 27:205-217. Volk, W. A. (1982). Essentials of Medical Microbiology. J. B. Lippincott Co., Phila., Pa. Wagner, E. K. (1991). Herpesvirus transcription; general aspects. In Herpesvirus transcription and its regulation. (E. K. Wagner, ed.) CRC Press, Inc., Florida. Weinstock, D., and K. A. Schat. (1987). Virus specific syngenic killing of reticuloendotheliosis vims transformed cell line target cells by spleen cells. In Avian Immunology, pp253-263, (W. T. Weber and D. L. Ewert, eds.), Alan R. Liss, New York. Weiss, R., N. Teich, H. Varmus, J. Coffin. (1984). Origins of contemporary RNA tumor virus research. In RNA tumor viruses 2"“ Edition, Cold Spn'ng Harbor laboratory, New York. Wen, L. T., A. Tanaka, and M. Nonoyama. (1988). Identification of Marek's disease virus nuclear antigen in latently infected Iymphoblastoid cell lines. J. Virol. 62:3764- 3771. Westrop, G. D., K. A. Wareham, D. M. Evans, G. Minor, D. I. Magrath, F. Taffs, S. Marsden, M. A. Skinner, G. C. Schild, and J. W. Almond. (1989). Genetic basis of attenuation of the Sabin type 3 oral poliovirus vaccine. J. Virol. 63:1338-1344. Wilson, M. R. and P. M. Coussens. (1991). Purification and characterization of infectious Marek’s disease virus genomes using pulsed field electrophoresis. Virology 185:673-680. Witter, R. L. (1992a). Recent developments in the prevention and control of Marek’s disease. Proc. XIX World’s Poultry Congress, vol. 1, Ponsen and Looijen, 185 Wageningen, The Netherlands. Witter, R. L. (1992b). Safety and comparative efficacy of the CV1988/Rispens vaccine strain. Proc. XIX World’s Poultry Congress, vol. 1, Ponsen and Looijen, Wageningen, The Netherlands. Witter R. L. and L. F. Lee. (1984). Polyvalent Marek's disease vaccines: safety, efficacy, and protective synergism in chickens with maternal antibodies. Avian Pathol. 13:75-92. Witter, R. L. (1983). Characteristics of Marek’s disease viruses isolated from vaccinated commercial flocks: association of viral pathotype with lymphoma frequency. Avian diseases. 27:113-132. Witter R. L., E. A. Stephens, J. M. Shanna, and K. Nazerian. (1975). Demonstration of a tumor associated-associated surface antigen in Marek’s Disease. J. Immunol. 115:177-183. Witter, R. L. (1985). Principels of vaccination. In Marek's Disease, Scientific Basis and Methods of Control (L. N. Payne ed.) Martinus Nijhoff, Inc Witter, R. L., J. M. Sharma, and A. M. Fadly. (1980). Pathogenicity of variant Marek’s disease virus isolates in vaccinated and unvaccinated chickens. Avian Diseases 24:210-232. Witter. R. L. (1982). Protection by attenuated and polyvalent vaccines against highly virulent strains of Marek’s disease virus. Avian Pathol. 11:49-62. zurHausen, (1981). Oncogenic herpesviruses. In DNA tumor viruses 2"cl Edition, Cold Spring Harbor laboratory, New York. ”IC IIIIIIIIIIIIIIIIIIIIIII