.4. t .. x. o. u 1.... a ‘ u... 23.3....me $38.... km...“ fawn O ....... z .3 .. . “Mm...“ film. , iii...“ .20. u x . $394M. a»: ta... .3. . u. «£wa V’Inl a»... whet... 3. u Ari... 1...... 12.15.... . cu... .Hwnu€s._utv .a...,.......,....:: P 1.". ‘ .. . 5.- 33.311)... 3. u. .6 A. .t.{l’(5 0.0.533... Ii); 1 Id"... 9.. Hull“... 1| am. It...“ 2. rfiir )\ i a. .§lllla!(tll‘l.1||’.:np:l Intfiyaillttlvr :9” J.- 0‘. "”3? "we-mama». o m t wasmutvmcuwuv :3 V! ‘am’fi'w . .. to: ‘ it? . .. , . , . . _ .. aw, a ... ... fir... _ . , , . ._. . . . . . . . , . m I. . . .. . . .. . . . .. .. . . . . , . . ka......\..:§. . I... 27 . LIBRARY 1"- '3 Michigan State University“ This is to certify that the dissertation entitled CHARACTERIZING CHICKEN STEM CELL ANTIGEN 2, A PUTATIVE MAREK’S DISEASE GENE presented by WEIFENG MAO has been accepted towards fulfillment of the requirements for the Ph.D degree in GENETICS at... m 0,. Majorprfissor’s Signature DEC. [7’ 3007 Date MSU is an Affirmative Action/Equal Opportunity Employer PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K:/Proj/Aoc&Pres/CIRC/Dateoue.indd . .___.._ CHARACTERIZING CHICKEN STEM CELL ANTIGEN 2, A PUTATIVE MAREK’S DISEASE RESISTANT GENE By WEIFENG MAO A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILSOPHY Genetics 2009 ABSTRACT CHARACTERIZING CHICKEN STEM CELL ANTIGEN 2, A PUTATIVE MAREK’S DISEASE RESISTANT GENE By WEIF ENG MAO Marek’s disease virus (MDV) is an alpha-herpesvirus that is the causative agent of Marek’s disease (MD), a lymphomatous disease of chicken. MD is a serious chronic disease of concern to the poultry industry worldwide, although it has been controlled through the use of vaccination since the 1970’s. MD vaccines prevent the formation of lymphomas but are not sterilizing and do not prevent MDV infection or replication in the bird. Consequently, the observed increase in MDV virulence over the years may be an undesirable response to the widespread use of MD vaccines. With increases in MDV virulence expected in the future and the most effective MD vaccine (Rispens) already in use, additional approaches such as genetic resistance to control MD are needed to augment vaccinal control of MD. The overall purpose of this project is to characterize stem cell antigen 2 (SCA2), the product of the putative MD resistant gene, and its role in MDV biology I To analyze the biological properties of chicken SCA2, SCA2 protein was expressed and purified in E. coli, and a polyclonal antibody was developed. Utilizing this anti-SCA2 antibody, SCA2 was identified to be a 13 kDa cell surface protein anchored by a GPI moiety as is the case for most other Ly6 family members. In vivo studies showed unique SCA2 expression pattern in bursal cortical medullary epithelial cells (CMEC), which are surrounded by MHC class II presenting cells. Expression profiles of bursal cells induced by contact with SCA2-expressing cells in vitro demonstrated down- regulation of numerous genes that are involved in the B cell receptor and immune response signaling pathways. These data suggest that SCA2 plays a role in regulating chicken B lymphocytes. The viral USIO protein was previously demonstrated to interact with SCA2 in an E. coli two hybrid screen followed by confirmation using an in vitro binding assay. To analyze the fimctional interaction of SCA2 and USlO for MDV growth properties, two recombinant MDVs were developed in which viral US10 gene was fused to or replaced with an enhanced green fluorescent protein (EGFP) coding region. Over-expressing SCA2 impairs both MDV plaque size and the percent of fibroblasts infected in vitro but this effect is dependent on the presence of U810. We conclude that MDV USIO is both sufficient and required for this growth impairment via association with SCA2. A missense point mutation in MDV UL41 gene was found out when cloning the viral genome into bacterial artificial chromosome (BAC). Monitoring the frequency of each SNP by pyrosequencing during virus passages determined the ratio of each viral genome in a single monolayer. This point mutation in UL41 gene enhanced the viral fitness in the competitive growth assay in vitro, but abolished the virion host shutoff (vhs) activity of UL41. This result suggests that the enhanced fitness in vitro for MDV with inactive vhs was due to reduced degradation of viral transcripts. Dedicated to: Yunli Liu Eric Mao ACKNOWLEDGEMNETS I would like to thank my mentor, Dr. Hans Cheng, for introducing me to a fascinating and challenging project, his guidance through my graduate program and my academic career planning. Without his help and support, the writing of this thesis will not have been done. I would also like to express my gratitude to my committee members, Dr. Jerry Dodgson, Dr. Michele Fluck, Dr. Suzanne Thiem and Dr. Cathy Earnst for their additional guidance and suggestions on how to improve myself as a scientist. I would also like to thank Dr. Henry Hunt for his guidance and help for my research. I would like to thank my lab-mate and friend, Laurie Molitor, for her friendship and invaluable assistance. I would also like to thank Dr. Melinda Frame, Dr. Lucy Lee, Dr. Robert Silver, Dr. Huanmin Zhang, Dr. Masahiro Niikura, Dr. Kyle Maclea, Dr. Tajoong Kim, Lonnie Milam, Noah Koller for their help and useful suggestions. I would like to thank my parents, Shudui Lin, and Hanchong Mao, for their love and support. I would like to thank my wife, Yunli Liu. She is always along with me during this long journey. My son, Eric Mao, joined us three years ago. They gave me innumerable loves and wishes. Finally, I would like to thank Genetics program, and the USDA, ARS, avian disease and oncology laboratory. Without their support, completion of my research and degree would not have been possible. TABLE OF CONTENTS LIST OF TABLES ................................................................................. ix LIST OF FIGURES .................................................................................. x LIST OF ABBREVIATIONS .................................................................. xii Chapterl. Introduction and literature review ................................................................... 1 Introduction ........................................................................................ 2 History of Marek’s disease (MD) .............................................................. 3 Marek’s disease vaccines ......................................................................... 4 Marek’s disease virus (MDV) ................................................................... 5 Pathogenesis of MDV infection ............................................................. 5 The genomic structure MDV ............................................................... 7 MDV virion host shutoff gene (vhs; UL41) ........................................ 8 MDV ICP4 gene ....................................................................... 10 MDV USIO gene ....................................................................... 10 Bacterial artificial chromosome-cloned MDV genomes ........................... 11 The morphogenesis of MDV ............................................................ 11 Immune responses to MDV infection ...................................................... 12 Innate immune responses ................................................................. 12 Adaptive immune responses .............................................................. 13 Genetic resistance to MD ....................................................................... 14 Stem cell antigen 2 (SCA2; Ly-6E/TSA1) .................................................. 17 Lymphocyte complex antigen 6 (Ly-6) family .......................................... 17 Chicken Stem cell antigen 2 ............................................................... 19 Avian lymphoid tissues ......................................................................... 22 Avian bursa of Fabr1c1u322 Avian thymu323 Chapter 2. Cloning of and functional characterization of chicken stem cell antigen 2. . . ..............37 Abstract ...................................................................................... 38 Introduction .................................................................................... 39 Materials and methods .................................................................... 41 Cells and animals ......................................................................... 41 Development of anti-SCA2 rabbit antisera ........................................... 41 Generation of chicken SCA2 eukaryotic expression vector and stable transfection cell line ..................................................................... 42 Flow cytometry analysis .................................................................. 43 Cleavage of GPI—linked cell surface SCA2 molecules ............................... 43 Western blots and immunoprecipitation ............................................. 44 Immunohistochemistry .................................................................... 44 Affymetrix array processing and data analysis ......................................... 45 vi Real-Time quantitative reverse transcription PCR ................................... 47 Results .......................................................................................... 49 Chicken SCA2 is a 13 kDa cell surface protein anchored by a GPI moiety ........ 49 Tissue distribution of chicken SCA2 ..................................................... 52 Gene expression profiling of bursal cells in contact with SCA2-expressing cells .................................................................... 58 Discussion ................................................................................... 64 Chapter 3. Chicken stem cell antigen 2 modulates Marek's disease virus growth properties in vitro via USIO ...................................................................... 73 Abstract .......................................................................................... 73 Introduction ...................................................................................... 76 Materials and methods ....................................................................... 79 Cells ........................................................................................ 79 Recombinant BAC clones ................................................................. 80 Generation of viruses ...................................................................... 82 One step growth curve assay and plaque area determination ........................ 82 Immunofluorescence and confocal image analysis ................................. 83 Western blot analysis ..................................................................... 84 Flow cytometry analysis .................................................................. 84 Results ............................................................................................ 86 Construction of USIOEGFP and AUSlOEGFP MDV-BAC clones ............... 86 Generation and growth properties of the mutant viruses in CEF .................... 89 Cellular localization of USlO-EGFP and SCA2 proteins ............................. 92 Growth properties of US lOEGFP and AUS lOEGFP mutant viruses on Vector-DFl and SCA2Flag-DF1 cells ............................................... 96 Discussion ...................................................................................... 100 Chapter 4. Quantitative evaluation of viral fitness due to a single nucleotide polymorphism in the Marek’s disease virus UL41 gene via an in vitro competition assay ........................................................................ 108 Abstract ....................................................................................... 109 Introduction .................................................................................... 1 10 Materials and methods ..................................................................... 113 Culture method for MDV and DNA purification .................................. 113 Mini-7t mutagenesis to generate pMDV-UL41*Arg and pMDV-UL41*Cys-rev .................................................................. 113 Growth rates of individual viruses ................................................... 116 Standard curve .......................................................................... 116 Competitive assay ....................................................................... 116 UL41 activity assay in vitro ............................................................. 115 Results ........................................................................................ 117 Generation of pMDV-UL41*Arg and pMDV-UL41*Cys—rev by mini-7t mutagenesis ................................................................ 119 vii Growth rates of individual viruses in vitro ............................................ 120 Standard curve for pyrosequencing .................................................... 120 Increasing virus fitness of MDV—UL41*Cys in virus populations in vitro ...................................................................... 121 Wild type MDV UL41 (MDV-UL41*Arg) has vhs activity and this enzyme activity is abolished in MDV-UL41*Cys .......................... 125 Discussion ...................................................................................... 127 Chapter 5. Future work ........................................................................................ 134 viii LIST OF TABLES Table 2-1 Primers used for quantitative real-time PCR validation of microarray data .......................................................................................................... 48 Table 2-2 Genes under-express in SCA2 induced bursal cells ............................... 59 Table 2-3 Genes over-express in SCA2 induced bursal cells ............................. 61 Table 2-4 Real-time quantitative reverse transcription PCR analysis of differential expression for bursal cells incubated with SCA2-expressing or control DFl cells ......................................................................................................... 63 LIST OF FIGURES Figure 1—1 Protein sequences of chicken SCA2 with homologous mice Ly-6 family members. ............................................................................................. 21 Figure 2-1 Western blots using anti-Flag mAb or anti-SCA2 antisera ..................... 50 Figure 2-2 Flow cytometry analysis of SCA2Flag-DF1 and vector-DFl with anti-Flag mAb and anti-SCA2 antisera ..................................................................... 51 Figure 2-3 Immunohistochemical analysis of chicken SCA2 expression in chicken tissues ................................................................................................ 53 Figure 2-4 Immunoprecipitation and western blot analysis of SCA2 expression in chicken tissues ................................................................................................ 55 Figure 2-5 Immunohistochemical analysis of SCA2 expression in bursal follicles ........ 56 Figure 3-1 Schematic diagram of recombinant MDV-BAC clones ........................... 87 Figure 3-2 BamHI digestion pattern of purified USlOEGFP-BAC and AUSIOEGFP-BAC clones ................................................................................................ 88 Figure 3-3 Western blot analysis of USlOEGFP and AUSIOEGFP viruses .......................................................................................... 90 Figure 3-4 One step growth curve assay of er5-B40, USlOEGFP, AUSIOEGFP viruses in CEF cells ................................................................................ 91 Figure 3-5 The confocal image of sub-cellular localization of U810 in CEF infected with USlOEGFP virus through green fluorescence ................................................ 93 Figure 3-6 Confocal immunofluoresence images of SCA2 in SCA2Flag-DF1 cells ...... 94 Figure 3-7 Confocal immunofluoresence images of USlO-EGFP and SCA2 in SCA2Flag- DF1 cells infected with USlOEGFP virus ....................................................... 95 Figure 3-8 Plaques of USlOEGFP and AUSIOEGFP viruses on Vector-D’Fl and SCA2Flag-DF1 cells .............................................................................. 97 Figure 3-9 Relative plaque areas of US lOEGFP and AUSlOEGFP viruses on Vector-DFl and SCA2FIag-DF1 cells ................................................................................................... 98 Figure 3-10 The percentage of infected cells within Vector-DFl and SCA2Flag-DF1 cells with USlOEGFP and AUSIOEGFP viruses ................................................... 99 Figure 4-1 Partial UL41 DNA and amino acid sequences for the parental Mdll strain, the original MDV-BAC clone (MDV-UL41*Cys), and the two recombinant viruses (MDV- UL41*Arg and MDV-UL41*Cys-rev) ......................................................... 1 15 Figure 4-2 Growth rates of individual viruses in vitro ....................................... 120 Figure 4-3 Standard curve of pyrosequencing for UL41 alleles.............................121 Figure 4-4 The percentage of MDVs genomes in the competition assay ................... 124 Figure 4-5 The [3- galactosidase assay for MDV UL41*Arg, UL41*Cys activity and the controls .............................................................................................. 126 Figure 5-1 SCA2 is up-regulated on the surface of BrdU-induced RPl (MDV transformed) T cell line ........................................................................................... 139 Figure 5-2 Partial ICP4 DNA and amino acid sequences for the parental Mdll strain, the recombinant virus (MDV-ICP4*Ser), the original MDV-BAC clone (MDV-ICP4*Leu), and the recombinant viruse (MDV-ICP4*Leu-rev) ........................................... 140 Figure 5-3 The percetange of MDVs genomes containing SNPs in ICP4 alleles in a competition assay .................................................................................. 142 xi ADOL BAC BCR BLNK CEF CMEC CTL EGFP EID ERC FBS GAPDH GPI HC HSC HSV-1 HSV-2 LIST OF ABBREVIATIONS xii Avian Disease and Oncology Laboratory Bacterial artificial chromosome B cell receptor B cell linker Chicken embryo fibroblasts Cortico-medullary epithelial cells Cytotoxic T lymphocyte Enhanced green fluorescent protein Embryonic incubation day Epithelial reticular cells Fetal bovine serum Glyceraldehyde 3-phosphate dehydrogenase Glycosylphosphatidyl-inositol Hematopoietic cells Hematopoeitic stem cells Herpes simplex virus 1 Herpes simplex virus 2 Immediate-early genes Interferon Interleukin 2 Latency-associated transcripts LM Ly-6 MD MDV MHC ORFs pAb PCR PFU PI-PLC PLCG2 qRT-PCR QTL RMA SAM SCA2 SNP SYK T2ECs TCR UL xiii Liebovitz's L-15 and McCoy 5A Lymphocyte complex antigen 6 Monoclonal antibody Marek's disease Marek's disease virus Major histocompatibility complex Open reading frames Polyclonal antibody Polymerase chain reaction Plaque-forming units Phosphatidylinositol-specific phospholipase C Gamma 2-phospholipase C Quantitative reverse transcription PCR Quantitative trait loci Robust multichip average Significance analysis of rnicroarrays Stem cell antigen 2 Single nucleotide polymorphism Spleen tyrosine kinase Transforming growth factor-induced erythrocytic cells T cell receptor Unique long US lOEGFP AUS lOEGFP US vhs vv+MDV vaDV VZV xiv er5-B40-US lOEGFP erS-B40-AUS 1 OEGFP Unique short Virion host shutoff Very virulent plus MDV Very virulent MDV Varicella—zoster virus Chapter 1 Introduction and literature review Introduction Marek’s disease virus (MDV) is a highly oncogenic alpha-herpesvirus that induces Marek’s disease (MD), a lymphomatous disease of chicken. MD is considered the most serious persistent infectious disease of concern to the poultry industry worldwide, although it has been primarily controlled through vaccination and changes in animal husbandry since the 19703. Even with vaccines, MD still has a significant economic impact to the poultry industry as there are sporadic outbreaks, which require the continuous vaccination and monitoring of birds. Consequently, there is a need for alternative methods, such as enhanced genetic resistance to MD, to augment vaccinal control of MD. This chapter reviews the pathogenesis of MDV infection, the genomic structure of MDV, and genetic resistance to MD, while also discussing the prevention and control of the disease. MDV UL41, ICP4, USIO genes, as well as chicken SCA2, a putative MD resistant gene, are discussed. The possibility that SCA2 plays role in MDV spread through regulating MDV maturation and cell-cell adhesion is hypothesized. The overall purpose of this thesis was to determine and characterize the biological function of chicken SCA2, and through its interaction with MDV U810, investigate the role of SCA2 in MDV infection. By identifying the properties of SCA2 and its putative functions, we hoped to obtain a better understanding of the host protein and its involvement in the MDV life cycle. The impact of single nucleotide polymorphisms (SNPs) in MDV UL41 and ICP4 genes were also elucidated in this thesis. 1. History of Marek’s disease (MD) The first report of what is later called Marek’s disease (MD) occurred in 1907 by Dr. Joseph Marek, an eminent veterinary clinician and pathologist at the Royal Hungarian Veterinary School in Budapest (Marek, 1907). In the report, Dr. Marek described a disease in four adult cockerels that were affected by paralysis of the legs and wings, and the pathological changes that occurred in the central and peripheral nervous system. The disease is characterized by a mononuclear infiltration of the peripheral nerves, gonads, various viscera, muscles, skin, and partial or complete paralysis as a result of the accumulation of tumor cells in peripheral nerves. In the 19405 and 19503, diagnosis relied on pathological examination and MD was confused with other leukotic diseases under the term “avian leukosis complex” because of the difficulty of distinguishing between the visceral lymphoma of MD and lymphoid leukosis. Subsequently, on the basis of the studies in the field and transmission experiments, it was concluded that this avian lymphomatous disease would more appropriate to MD than leukosis. Biggs and Payne (1963, 1967) described that the transmission of this disease with blood or lymphoma cells from diseased birds and the transmission was dependent on whole cells as the destruction of cells resulted in loss of infectivity. The behavior of MD in chickens suggests that it was an infectious disease, and the causative agent was likely to be a cell-associated virus. In 1967, a cell-associated herpesvirus was isolated from blood and kidney tumor cells of infected chickens (Churchill and Biggs, 1967), and then in infected duck embryo fibroblasts (Nazerian et a1. 1968; Solomon et al. 1968). Because the herpesvirus was associated with cases of MD, the ability of the herpesvirus-infected cells to produce cytopathic effects in cell culture was quantitatively associated with lesions of MD, and the virus was highly cell-associated similar to the infectious agent of MD, this cell- associated herpesvirus was concluded to be the aetiological agent of MD and came to be referred to as the Marek’s disease virus (MDV). 2. Marek’s disease vaccines MDV and its close relatives are grouped by serotype. Serotype 1 viruses include MDV (gallid herpesvirus type 2, GaHV-2) and are the only ones that are oncogenic. Serotype 2 MDVs (GaHV-3, MDV-2) are naturally occurring strains that are non- oncogenic. Serotype 3 is turkey herpes virus 1 (INT; MDV-3), a close relative (Osterrieder et al., 2006). Serotype 1 is further categorized into very virulent plus (vv+MDV), very virulent (vaDV), virulent (vMDV), rrrild (mMDV) and attenuated strains based on their ability to induce disease in unvaccinated and vaccinated flocks (Witter, 1997). MDV can lead to tremendous economic losses to the poultry industry. After the introduction and widespread use of vaccines in the early 19705, the disease was first controlled. The first widely used MDV vaccine was the antigenically similar HVT. HVT has been widely used in the United States, and appears to offer good protection (Witter et al., 2000). The current and most effect MD vaccine is CV1988 (also referred to as Rispens), a cell culture-attenuated serotype 1 MDV strain that offers effective protection against challenge with the very virulent pathotypes of MDV. The principle underlying successful immunization remains unknown. One speculation is “substrate deprivation” where the targeted B and T cells infected by the live attenuated vaccine are no longer susceptible to wild type virus infection or transformation. The major concern of MD vaccines is that they are not sterilizing, i.e., the vaccines block tumor formation but do not prevent viral replication or spread in birds. MDV continues to replicate and potentially evolve in MD vaccinated birds, consequently, the co—existence of the vaccines and wild type viruses in vaccinated flocks may speed up the virus recombination and viral evolution. The evolution of MDV to greater virulence (Witter, 1997; Nair, 2005) in the near future is one of the major driving forces to develop new alternative approaches to augment vaccinal control of MD. For example, recombinant MDV vaccines such as a Meg-deleted MDV serotype 1 strain, has shown improved vaccine protection (Lee et al., 2008). Apart from its economic impact, MDV is of interest to biomedical researchers in that it is the only naturally-occurring alpha-herpesvirus that induces tumors in its host and is controlled by vaccination. These characteristics of MDV make it a good model for the study of the mechanisms of viral induced T cell lymphoma. 3. Marek’s disease virus 3.1. Pathogenesis of MDV infection MDV is a highly cell-associated herpesvirus, and horizontal transmission occurs through enveloped virion particles produced at the superficial layers of the feather follicle epithelium. Poultry dusts containing these epithelial cell-enveloped MDV are infectious to both chickens and cultured cells (Biggs, 2000). These results provide a reasonable explanation for why the cell-associated virus can be highly contagious. Lyrnphomagenesis results from MDV infection in susceptible birds. Lymphomas may develop in most of the visceral organs: spleen, liver, kidney, heart, pancreas, gonads, and also in other organs including muscle, peripheral nerves, eye, and skin. Lymphomas in visceral organs may be the most proliferative but the lesions in other tissues may also be obvious. The involvement of lymphoma in various organs is influenced to some extent by both the virus strain and the genotype of the host. The pathogenesis of MDV in susceptible chickens has been divided into four phases: (1) early cytolytic, (2) latent, (3) late cytolytic and (4) transforming. The initial infection of chicken occurs via the respiratory tract following the inhalation of the cell- associated MDV from the contaminated environment. In the respiratory tract, macrophages or dendritic cells are infected and presumed to carry the virus to lymphoid organs such as the bursa of Fabricius (bursa), thymus, and spleen (Calnek et al., 2000). A cytolytic infection in these lymphoid organs occurs between 3 and 6 days postinfection. During the Iytic phase, the virus is able to infect B cells and activated CD4+ T cells, and rarely CD4-CD8- T cells or CD8+ T cells (Calnek et al., 2000). After 7 to 8 days, the infection in lymphoid organs switches from Iytic to latent infection mostly in CD4+ T cells. No extracellular virus can be detected in the blood in any infected bird, so it is concluded that virus spread is cell-to-cell. Interestingly, despite significant efforts, the putative receptor for MDV attachment or entry has not been resolved. The latent phase is defined as the presence of the viral genome without production of infectious virions. In contrast to other members of the alpha-herpesvirus subfamily, MDV goes latent in activated T cells but not in sensory nerve ganglia (Calnek et al., 2000). During latency, latency-associated transcripts (LAT s) are among the few active viral transcripts. Latently infected cells can become transformed cells leading to lymphoma formation; lymphomatous tumors caused by MDV can appear as early as 12-14 days postinfection. Various potentially important genes and latency-related transcripts that could contribute to oncogenicity of MDV have been reported. The most studied one to date is Meq. Meq is a gene that encodes a protein with 339 amino acids and is completely contained in the MDV EcoRI Q fragment, hence the name Meq. Overexpressed Meq is associated with lymphocytic tumors and confer these cells with anti-apoptotic activities. Anti-sense transcripts of Meq inhibit the growth of MDV-transformed T cells (Xie et al., 1996). Thus, the evidence strongly supports Meq as a significant viral oncogene for transformation. Further studies have shown Meq has transactivating activity, and binds to Jun/Fos cellular oncogenes to activate oncogenic pathways (Jones et al., 1992; Liu et al., 1999; Shamblin 2004; Levy et al., 2005; Brown et al., 2009). In addition to the viral oncogene, certain host cell genes, such as the tumor suppressor gene, p53, and proto- oncogene, Bel-2 could also play roles in the pathogenesis of MD (Venugopal and Payne, 1995). 3.2. The genomic structure of MDV Like other herpesviruses, MDV contains a linear, double-stranded DNA genome surrounded by an icosahedral protein capsid core, which in turn is surrounded by tegument proteins and a lipid-containing envelop to enclose the mature virion (Silva et al., 2000). The 180 kb MDV genome consists of unique long (UL) and unique short (US) regions, each flanked by inverted repeats: terminal repeat long (TRL), internal repeat long (IRL), internal repeat short (IRS), and terminal repeat short (TRS). The genomic structure resulted in MDV being classified as an alpha-herpesvirus (Roizman et al., 1992) although it was originally thought to have been a gamma-herpesvims due to its biological characteristics, especially the ability to produce lymphoid tumors. MDV genomic structure is very similar that of herpes simplex virus 1 (HSV-1; the prototype alpha- herpesvirus) and human herpesvirus 3 (varicella-zoster virus, VZV) (Osterrieder et al., 2006). As a herpesvirus, MDV genes are classified as immediate-early genes (IE), early genes and late genes based on the timing of expression of their homologs during the lytic cycle. The IE proteins migrate to the nucleus to transactivate other MDV IE genes. The IE genes in turn regulate the expression of early and late genes. Some of the early genes are enzymes required for viral DNA replication, while the late genes include most of the structural genes such as capsid, envelope, and tegument proteins (Lupiani et al., 2000). 3.2.1. MDV virion host shutoff gene (vhs; UL41) One viral defense mechanism common to alpha-herpesviruses is to degrade the host mRNA and inhibit processing of new synthesized host mRNA. This degradation is an important feature of the early stage of infection, and this activity is contained in the virion host shutoff (vhs or UL41) gene of HSV-1 (Read and Frenkel, 1983; Schek and Bachenheimer, 1985; Kwong et al., 1988). It was later demonstrated that UL41 is an endoribonuclease that degrades mRNA and does not discriminate between cellular or viral origin (Jones et al., 1995; Pak et al., 1995; Taddeo and Roizman, 2006). However, because the viral mRNA is much more abundant, the effect is to downregulate host gene expression. HSV UL41 is a late gene and the gene product is packaged into the virion as a part of the tegument. Immediately after ingress of mature virions into cells, UL41 is released from the virion and degrades host mRNA in the cytoplasm. As a result, there is down regulation of MHC class I and class II proteins, and the degradation of Jakl and Stat2 to block IFN signaling, suggesting that UL41 modulates the host’s immune defense to viral infection. UL41 also contributes to the ability of the virus to evade the host adaptive immune response due to reduced cytotoxic T-lymphocyte recognition (Suzutani et al., 2000; Kopper-Lalic et al., 2001; Murphy et al., 2003). In addition to evasion of the host immune response, HSV UL41 also facilitates the transition from IE to E (Kwong and Frenkel, 1987; Oroska and Read, 1987). While UL41 represents a significant virulence factor in virus latency and pathogenesis, deletion of UL41 in HSV-l did not alter one-step growth curves in vitro but impaired growth in corneas, trigeminal ganglia, and brains of mice. Viral DNA within trigeminal ganglia infected with the UL41-deleted virus was reduced by 30—fold compared to wild type HSV-1 (Strelow and Leib, 1995, Strelow et al., 1997). In herpes simplex virus 2 (HSV-2), U1A1 also plays an important role in vivo by interfering with the IFN-alphafbeta-mediated antiviral response and is a determinant in HSV-2 pathogenesis (Smith et al., 2002, Murphy et al., 2003). The central region of MDV UL41 is highly conserved among all MDV serotypes as well as exhibiting high protein sequence identity with herpes simplex (29%) and varcella-zoster (33%) virus homologs (Gimeno and Silva, 2008). A recombinant MDV lacking UL41 replicated in cell culture as well as the parental MDV, but resulted in a longer early cytolytic infection in the lymphoid organs. The deletion did not affect either levels of latency or the ability to reactivate from latency and did not decrease the replication of the virus in lymphoid organs or feather follicle epithelium. Thus, MDV UL41 may facilitate the normal cascade of immediate early, early, and late gene expression but the role of MDV UL41 in MDV pathogenesis is not nearly as great compared to HSV vhs, although the MDV UL41 is functionally equivalent to the HSV-1 vhs. 3.2.2. MDV ICP4 gene MDV ICP4 is a homologue of HSV ICP4 and lies in the inverted repeat that flanks the unique short region of the genome (Anderson et al., 1992). HSV ICP4 protein is a transcriptional regulator required for the induced transcription of most of the remaining HSV genes and is essential for viral replication (Compel and DeLuca, 2003; Su et al., 2006). The exact role of ICP4 in MDV replication has not been completely determined although it is speculated to play a role similar to HSV ICP4, is associated with maintenance of transformation, and is essential for MDV growth in chicken embryo fibroblasts (CEF) (Endoh, 1996; Chattoo et al., 2006). A number of studies demonstrated that LATs of MDV include a 10 kb RNA that is antisense to ICP4, and the suppression of ICP4 by LATs plays an important role in MDV latency as well as maintenance of transformation (Cantello et al., 1994; Xie et al., 1996; Cantello et al., 1997; Li et al., 1998). 3.2.3. MDV US10 gene HSV-1 USlO is a phosphorylated tegument protein that copurifies with the nuclear matrix (Yamada et al., 1997). Homologues of US10 protein are also encoded by other alpha-herpesviruses including HSV-2, VZV, and MDV (Brown and Harland, 1987; Davison and Scott, 1986; Sakaguchi et al., 1992). MDV USlO is not essential for virus replication in CEF but a ~4.8 kb deletion in the MDV US region that includes U510 and adjacent MDV reading open frames (USI, U52, SORF 1, SORFZ, SORF3) decreased early cytolytic infection, mortality, and tumor incidence in viva (Parcell et al., 1994; Parcell et al., 1995). The role of US10 in the MDV life cycle remains uncertain. Liu et a1. (2003) demonstrated that MDV US10 interacts with chicken SCA2 protein through a E. coli two-hybrid screen followed by an in vitro binding assay, suggesting that US10 may be involved in MD genetic resistance via SCA2 (Liu et al., 2003; Niikura et al., 2004). 3.2.4. Bacterial artificial chromosome-cloned MDV genomes The development of recombinant MDV DNA using infectious bacterial artificial chromosome (BAC)-cloned MDV genomes has contributed greatly to the ability of researchers to study viral genes and gene function including viral morphogenesis. Through recombinant MDVs, viral gB, gE, g], and UL49-VP22 genes have been demonstrated to be indispensible for MDV growth in vitro (Schumacher et al., 2000; Schumacher et al., 2001; Dorange et al., 2002). The viral genes Meq, vLIP, gC, pp38, US3, vIL-8, vTRm are involved in pathogenesis and immune evasion in vivo (reviewed in Osterrieder et al., 2006). 3.2.5. The morphogenesis of MDV MDV morphogenesis is proposed to be like a typical alpha-herpesvirus. The assembly process begins in the nucleus, where the viral genome is packaged into capsids. Then, nucleocapsids which follow with the envelope at the nuclear membrane exit from the nucleus to the cytOplasm, bind to tegument proteins, and are enveloped a second time 11 at the trans-Golgi network or endosomal membranes. The final egress step is assumed through exocytosis of vesicles (reviewed in Johnson and Huber, 2002; Mettenleiter 2002; Campadelli-Fiume and Roizman, 2006; Sugimotoqet al., 2008). Using a recombinant MDV that expresses a structural protein (VP22, which is encoded by UL49)-EGFP fusion protein, an ultrastructural study of MDV morphogenesis showed that, compared to other alpha-herpesviruses, MDV seems deficient in a few crucial steps of viral morphogenesis, i.e., release from the nucleus, secondary envelopment and the final egress (Denesvre et al., 2007). The number of enveloped MDV virions in the cytoplasm is very low, which may partially explain why infectious MDV cannot be recovered from cell lysates by freeze- thawing or sonication. However, due to its strict cell association, MDV morphogenesis is still not well understood. 3.3. Immune responses to MDV infection The innate and acquired immune responses of the chicken play important roles in different phases of MDV infection. After the virions are engulfed by macrophages, they are delivered to replicate in bursal follicles, in which MDV mostly replicates in B lymphocytes. Later, activated T lymphocytes are infected through intimate contact between B and T cells (Schat and Marlowski-Grimsrud, 2000). A complex set of innate and acquired immune responses are consequences of the maintenance of viral latent infection and tumor formation. 3.3.1. Innate immune responses Innate immune responses of MDV-infected chickens include activation of macr0phages, natural killer (NK) cells, cytokines, and other soluble mediators. The role of macrophages in MDV pathogenesis is significant. Phagocytosis of MDV by macrophages is important to the transportation of MDV from lung, the initial entry point, to other lympoid tissues. However, MDV does not replicate in these macrophages and, in fact, activation of macrophage inhibits viral DNA replication in chicken spleen cells (Lee et al., 1979). In vivo studies also have shown an increasing number of activated macrophages reduces virus replication and delays the onset of MD tumors (Gupta et al., 1989). These results demonstrate that macrophages play an important role in inhibiting MDV replication. MDV Iytic infection, as well as the establishment of latency in lymphocytes, causes the differential expression of cytokines and other soluble mediators. The up- regulation of IFN-gamma is one of the first immune responses after MDV infection (Schat and King, 2000). Some expression of IE, early, and late MDV genes is down- regulated by IFN-alpha (Volpini et al., 1996). Buscaglia and Calnek demonstrated that two soluble factors maintained MDV latency in spleen cell culture, but these factors have not been characterized (Buscaglia and Calnek, 1988). The increased NK cells activities induced by MDV suggests its role in protection against MDV infection (Lessard et al., 1996). 3.3.2. Adaptive immune responses Adaptive immunity to MDV consists of the development of viral antigen-specific antibodies and cytotoxic T lymphocyte (CTL) responses. During infection, virus- 13 neutralizing antibodies are generated to protect against infection. Antibodies to several MDV glycoproteins have been demonstrated but only antibodies to MDV gB are believed to play an important role in MDV protection based on its ability to neutralize MDV (lkuta et a., 1984; Nazerian et al., 1992). The non-neutralizing antibodies also play a role in adaptive immunity. Maternal antibodies passed from vaccinated hens or transferred by injection can reduce the severity of morbidity, mortality, and tumor formation of MD (Schat and Markowski-Grimsrud, 2000). The CTL response is of particular importance in immunity to MDV due to its highly cell-associated nature. Effector cells were characterized as CD4'CD8‘TCR0LBI+ (Omar and Schat, 1997). MDV glycoproteins have been implicated as determinants of CTL responses. MDV gB was shown to induce a significant CTL response (Omar and Schat, 1996). The down-regulation of MHC class I on the surface of infected cells by MDV is involved in the immune evasion of MDV in Iytic infection (Hunt et al., 2001). This escape from MHC class I may restrict cytotoxicity and facilitate Iytic MDV spread and the maintenance of latency. The hypothesis of MHC class I down-regulation and antigen processing blocked by MDV proteins has been pr0posed, but the mechanism of CTL evasion through MDV proteins is still inclusive. 4. Genetic resistance to MD Chickens resistant to MD are those that fail to deve10p characteristic symptoms upon exposure to MDV. Genetics resistance to MD is controlled by multiple genes or QTL (Schat and Xing, 2000; Vallejo et al., 1998; Yonash et al., 1999). The United States Department of Agriculture (USDA), Avian Disease and Oncology Laboratory (ADOL), l4 has generated inbred chicken lines selected for resistance and susceptibility to MD since the 19403. Three of the original lines are presently maintained. These include line 6 which is MD resistant, and lines 7 and 15 which are MD susceptible. Less than 1% of line 6 chickens compared to 76-80% of line 7 chickens were shown to develop MD after infection with GA or JM strains of MDV (Bacon et al., 2000). The best understood mechanism for the involvement of genetic resistance to MD involves the MHC or, as it is known in the chicken, the B complex. The MHC contains three tightly linked regions known as 8-17 (class 1), 8—0 (class II), and B—L (class IV), which control cell surface antigens. The B-G locus is expressed in erythrocytes, which enables convenient typing of blood groups. By measuring the allelic frequency of specific blood groups, it has been observed that certain B alleles are associated with resistance or susceptibility. Chickens with the B*2l allele have been found to be more resistant than those with other B haplotypes (Bacon and Witter, 1994). Other studies have allowed for the relative ranking of the other B alleles: moderate resistance, B*2, B*6, B*14; susceptibility, B*l, B*3, B*S, B*13, B*15, B*19, B*27 (Bacon, 1987; Bacon and Witter, 1992) The MHC also influences vaccinal immunity as some haplotypes develop better protection with vaccines of one serotype than of a different serotype (Bacon and Witter, 1992; 1994). ADOL lines 6 and 7 chickens share the same MHC, B haplotype, yet differ with respect to resistance to MD (Lee et al., 1981; Hunt and Fulton, 1998). Thus, non-MHC factors also play a major role in genetic resistance to MD. Using genetic markers spaced throughout the chicken genome, 14 QTL (7 significant and 7 suggestive) were discovered that explain one or more MD-associated traits (Vallejo et al., 1998; Yonash et al., 1999). 15 Also, a few non-MHC genes have been correlated with resistance to MD. The BUl allele (antigens on bursal lymphocytes) of line 6 was associated with resistance to MD in progeny of a (6x7)x7 backcross population (Tregaskes et al., 1996; Bacon et al., 2000). Chicken SCA2 was significantly associated with length of survival and the indicence of proventricular tumors following MDV challenge in a commercial layer resource population (Liu et al., 2003). The chicken growth hormone allele is linked with MD QTL and polymorphism in the growth hormone gene is associated with the number of tissues with tumors in commercial White Leghorn chickens with the MHC B*2/B*15 genotype (Liu et al., 2001). DNA microarrays have been employed to identify candidate genes for QTL and pathways involved in MD resistance. Host genes that were reproducibly induced by infection of CEF with MDV included macrophage inflammatory protein (MIP), interferon response factor 1 (IRFl), interferon-inducible protein, SCA2, MHC class I, and MHC class 11 (Morgan et al., 2001). To sufficiently resolve the QTL containing the genes conferring genetic resistance to MD, the expression of genes in peripheral blood lymphocytes from uninfected and MDV-infected line 6 and line 7 chickens were analyzed through microarray. The microarray results were refined by genetic mapping of the differentially-expressed genes. Microarray results showed that up to 25% of the genes detectable show differential expression between the resistant and susceptible lines. Integration of microarrays with genetic mapping data was achieved with a sample of 15 genes in which the SCA2, growth hormone, and MHC class I were involved (Liu et al., 2001). DNA microarrays provided a complementary approach for identifying positional candidate MD resistant genes. Aside from the the DNA and RNA variation associated l6 with MD resistance, host and MDV protein interactions associated with MD resistance were examined through a bacterial two-hybrid system (Niikura et al., 2004). One example is chicken growth hormone which interacted with MDV SORF2 and combined with linkage and differential expression, was defined as a MD resistant gene (Liu et al., 2000). Another putative MD resistant gene is SCA2 which is associated with the genetic resistance QTL mapping, differentially transcribed between Line 6 and Line 7 following MDV infection, and interacts with MDV US10 (Liu etal., 2003). In addition to defining host genes associated with MD resistance, other alternatives to enhancing MD resistance include the transgenics approach such as RNA interference to MDV. A successful report of RNA interference that inhibits MDV in vitro and in vivo has been described (Mo et al., 2008). 5. Stem cell antigen 2 (SCA2; Ly-6E/TSA1) Chicken SCA2 has been proposed as a MD resistance gene due to the following criteria: (1) association with MD resistance in a commercial resource population, (2) transcriptional differences between MD resistant and susceptible experimental chicken lines following challenge with MDV, and (3) a direct protein interaction with MDV USlO as determined by a two-hybrid screen followed by an in vitro binding assay. These potential effects in MD encouraged us to investigate the properties of chicken SCA2 and its role in MDV infection. 5.1. Lymphocyte complex antigen 6 (Ly-6) family 17 The majority of research on Ly-6 family members has been conducted using mice. The mouse Ly-6 locus encodes a family of 10-12 kDa proteins are characterized by their remarkable conservation of cysteine residues and, with few exceptions, are linked to the cell surface via a g1ycosylphosphatidyl-inositol (GPI) anchor. Ly-6 proteins have a wide range of functions but, in general, are thought to be involved in cell signaling and cell adhesion (reviewed in Bamezai, 2004). Ly-6 proteins are broadly expressed in tissues of hematopoietic stem cells (HSC) origin and, consequently, are often used as differential markers for developmental stages in these cells (reviewed in Rock et al., 1986; Shevach et al., 1989; Gumley et al., 1995). Ly—6A.2 (also known as stem cell antigen 1 and T cell- activating protein) is expressed on the murine stem cells of bone marrow and T cell lineages, and contributes to regulating HSC renewal and development of progenitor cells (Uchida et al., 1994; Ito et al., 2003). Ly-6C is expressed on a majority of bone marrow cells, some subsets of immature T cells and B cells, and was shown to promote differentiation of B cells into antibody-generating plasma cells (reviewed in Gumley et al., 1995; Schlueter et al., 1997; McHeyzer—Williams and McHeyzer-Williams, 2004). Ly-6E (also known as stem cell antigen 2, SCA2; thymic shared antigen 1, TSAl) is expressed on the majority of immature double negative thymocytes, B cells, and also expressed in lymphoid epithelial cells (Godfrey et al., 1992; MacNei et al., 1993; Classon and Coverdale, 1996; Antica et al., 1997; Zammit et al., 2002). Ly-6G is a granulocyte- differentiation antigen and serves as a marker for identifying mature granulocytes (Fleming et al., 1993). Ly—6M is expressed on hematopoietic cells (HC), especially monocytes and myeloid precursors (Patterson et al., 2000). In the chicken genome, there are 3 annotated genes showing similar molecular size and conserved cysteine residues with mice Ly-6 genes, but it is not certain whether these chicken genes belong to the same Ly-6 family. 5.2. Chicken stem cell antigen 2 Chicken SCA2 transcript was first identified through a subtraction cDNA library of v-Rel (a NF-kappaB family member) — transformed avian bone marrow cells regulated by a conditional v-Rel estrogen receptor (Petrenko et al., 1997). Bresson-Mazet et al. reported that SCA2 was involved in the self-renewal of chicken erythroid progenitor (Bresson-Mazet et al., 2008). But the protein properties of chicken SCA2 were not defined due to the lack of antibodies to SCA2. Chicken SCA2 has the 10 cysteine residues conserved among Ly—6 members (Fig. I 1-1), however, it is not known if this chicken homolog has the same biological characteristics of mice SCA2 (Ly-6E) or other Ly-6 members. While the biological role of chicken SCA2 protein is not defined, its homologues in the mouse Ly-6 family provide clues for investigation. Mature mice SCA2 is a cell surface protein of 82 amino acids anchored on the cell membrane by a C-terminal GPI moiety; highly expressed in immature T cells, bone marrow cells, and peripheral B cells, with low levels in peripheral mature T cells. SCA2 is implicated in regulating intracellular signaling events via TCR and cell-cell adhesion as a receptor on the lymphocyte membrane. The ability of anti- SCA2 mAb to block thymopoiesis in fetal thymic organ culture indicates that SCA2 functions in T cell development (Randle et al., 1993). Fleming and Malek (1994) demonstrated that anti-SCA2 mAb inhibited anti-CD3-induced interleukin 2 (IL-2) production and this inhibition is independent on the GPI anchor (Fleming and Malek, 1994. This finding was further supported by Kosugi et al. (1994). Saitoh et a1. (1995) demonstrated that SCA2 plays an important role in the TCR signaling pathway in that the tyrosine phosphrylation of CD3§-chains following TCR stimulation was reduced by anti- SCA2 mAb. Anti-TCR/CD3-induced T cell apoptosis was blocked by anti-SCA2 mAb, which confirmed the function of SCA2 through the TCR signaling pathway (Noda et al., 1996), and the physical interaction between SCA2 and CD3 may clarify their functional interaction (Kosugi et al., 1998). Aside from lymphocytes, SCA2 is also expressed in other tissues, including the embryonic adrenal gland and liver and adult kidney and liver and in various epithelial cells (MacNeil et al., 1993; Classon and Coverdale, 1996). SCA2 deficient (SCA2-/-) mice exhibited abnormal development, impaired function of the embryonic adrenal gland and midgestational lethality at embryonic day 15 due to the cardiac abnormalities, but the lymphoid development is not impaired, suggesting SCA2 is not an obligate requirement for normal differentiation of lymphocytes (Zammit et al., 2002). One common expression pattern of SCA2 shared within different species is that SCA2 transcripts are overexpressed in malignant cells which suggest it may play role in tumorigenesis (Eshel et al., 1995; Petrenko et al., 1997; He and Chang, 2004; Darniola et al., 2004). 20 5.8 57mm rmméfifiznmmqmnel--mimeéoméiomifiémo?mxSmx>moman>ozHm5momom-xx_.§_~omrmnt-mn>l-m<3xo_uz:~m3mmmmormz 23.8 8% romlomx7-->m_.:x> ......... orofizxmzxmmxmzmzofiawrxmzlmrqéxxermz ninxm: mn>~ Smnmnmc>mmzz>m5naHz..mH>>>m<~ 2:: soaoamocm 38o Era 335. 303603. .38 8:328 Seam: 839.3 30 38.on 3. Ba. 2] 6. Avian lymphoid tissues Avian lymphoid tissues include the bursa of Fabricius, thymus, spleen, and other tissues including bone marrow and blood system. The avian bursa of Fabricius is the site for hematopoiesis, B lymphocyte differentiation and proliferation, and lies between the cloaca and the sacrum. The avian thymus is where T lymphocytes differentiate and proliferate, and lies parallel to the vagus nerve and internal jugular veins on each side of the neck where there are 7-8 separate lobes. The spleen is not a primary site for lymphocyte antigen-independent differentiation and proliferation but has an important role in embryonic lymphopoiesis (Olah et al., 2008). Here we only discuss the structure of bursa and thymus as these are the organs where SCA2 may potentially function. 6.1. Avian bursa of Fabricius The bursa is unique to birds and has a critical role in development of antibody responses. The bursa is an epithelial and lymphoid organ, and contains 15-20 longitudinal folds in the lumen. The bursal folds are filled with hundreds of bursal follicles. Each bursal follicle consists of a medulla and cortex with the supporting epithelium cells between the medulla and the cortex. The cellular composition of the medulla consists of epithelial cells and blood-bome HC including dendritic cells, lymphoid cells, and macrophages. The cortex is separated from the medulla by the cortico—medullary epithelial cells (CMEC). Once the cortical and medullary regions have been separated, bursal B cell proliferation continues in the cortex but the rate of cell division in the medulla is reduced more than 10 fold (Reynolds, 1987). This result suggests that 22 medullary and cortical regions have functional as well as structural differences (Ratcliffe, 2008). The function of the CMEC in bursa follicles is not understood but it is suggested to have a kind of “nursing “or regulatory function for B cell development. The murine Notch 1 plays important roles in T and B lymphocyte development (reviewed in Schmitt and Zuniga-Pflucker, 2006). The avian Notch ligand, Serrate2, is expressed exclusively in the CMEC and is suggested to regulate B cell differentiation by association with Notch displayed on the B cells (Morimura, et al., 2001). Ram et al., (1991) suggested that the CMEC represent a macrophage subpopulation. The CMEC express neither B cell antigen nor dendritic cell antigen and are resistance to infectious bursal disease virus (Ramm et aL,199l) 6.2. Avian thymus The avian thymus lies parallel to the vagus nerve and internal jugular veins. The avian thymus begins to develop as an epithelial outgrowth of the pharyngeal pouches at the age of 5 days of embryonic development, and the thymic epithelium begins to attract blood-bome HC at embryonic incubation days (BID) 6. 12 and 18 (Coltey et al., 1989). At these stages, some molecules expressed on the surface of the bone marrow T cell progenitor have been identified, including MHC class II, c-kit, and aIIB3integrin (Ody et al., 2000). During embroyonic development, the thymic mass gradually increases with colonization of HSC and results in the appearance of the medulla. Blood-bome HC invade the epithelial analage of the thymus, and immature and proliferating T lymphocytes are present in the subcapsular zone from where they migrate to the medulla during T cell maturation. The cortex and medulla are established by EID 12 and stromal epithelial cells form a network containing lymphoid cells (Coltey et al., 1987). In the thymic cortex, the epithelial reticular cells (ERC) are densely packed with thymocytes and may have a nursing function (Rieker et al., 1995). The development of the T cell lineage was traced using mAb recognizing different chains of the T cell receptor (TCR). At EID 12, a subpopulation of thymocytes begins to express the ySTCR-CD3 complex (TCR"'1) on the thymoctye surface (Sowder et al., 1988). The expression patterns of TCR+2 and TCR+3 are similar to mammalian aBTCR cells where the dual expression of the CD4 and CD8 molecules are on the surface of both of TCR+2 and TCR+3 cortical thymocytes (Coltey et al., 1989). These CD4+CD8+ double-positive cells mature to become either single positive CD4+ or CD8+ T cells (Davidson and Boyd, 1992). The histological structure in the thymic medulla is an epithelial cell aggregation but it is not certain if they are actual epithelial cells or stages of differentiation of lympoid cells, and their functional significance is not known, but an endocrine function has been suggested (Isler, 1976; Audhya et al., 1986; Watanabe et al., 2005). The cortico-medullary border contains a dendritic cell barrier expressing the MHC class II and a large number of peroxidase-positive cells, which may play a role in the negative selection of T cells (Guillemot et al., 1984). 24 References l. 10. ll. 12. 13. Anderson, A. S., A. Francesooni, and R. W. Morgan. 1992. Complete nucleotide sequence of the Marek's disease virus ICP4 gene. Virology 189:657-67. Antica, M., L. Wu, and R. Scollay. 1997. Stem cell antigen 2 expression in adult and developing mice. Immunol Lett 55:47-51. Audhya, T., D. Kroon, G. Heavner, G. Viamontes, and G. Goldstein. 1986. Tripeptide structure of bursin, a selective B-cell-differentiating hormone of the bursa of fabricius. Science 231:997-9. Bacon, L. D. 1987. Influence of the major histocompatibility complex on disease resistance and productivity. Poult Sci 66:802-11. Bacon, L. D., Hunt, H.D., and Cheng, H.H. 2000. Genetic resistance to Marek's Disease, p. 121-142. In K. Hirai (ed.), Mareks' Disease. Springer, Berlin. Bacon, L. D., and R. L. Witter. 1992. Influence of turkey herpesvirus vaccination on the B-haplotype effect on Marek's disease resistance in 15.8- congenic chickens. Avian Dis 36:378-85. Bacon, L. D., and R. L. Witter. 1994. B haplotype influence on the relative efficacy of Marek's disease vaccines in commercial chickens. Poult Sci 73:481-7. Bamezai, A. 2004. Mouse Ly—6 proteins and their extended family: markers of cell differentiation and regulators of cell signaling. Arch Immunol Ther Exp (Warsz) 52:255-66. Biggs, P. M. 2000. The History and Biology of Marek's Disease Virus, p. 1-24. In K. Hirai (ed.), Marek's Disease. Springer, Berlin. Bresson-Mazet, C., O. Gandrillon, and S. Gonin-Giraud. 2008. Stem cell antigen 2: a new gene involved in the self-renewal of erythroid progenitors. Cell Prolif 41:726-38. Brown, A. C., L. P. Smith, L. Kgosana, S. J. Baigent, V. Nair, and M. J. Allday. 2009. Homodimerization of the Meq viral oncoprotein is necessary for the inductidn of T cell lymphoma by Marek's disease virus (MDV). J Virol. Brown, S. M., and J. Harland. 1987. Three mutants of herpes simplex virus type 2: one lacking the genes U810, U811 and USl2 and two in which Rs has been extended by 6 kb to 0.91 map units with loss of Us sequences between 0.94 and the Us/TRs junction. J Gen Virol 68 (Pt 1): 1-18. Burnside, J ., E. Bernberg, A. Anderson, C. Lu, B. C. Meyers, P. J. Green, N. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. Jain, G. Isaacs, and R. W. Morgan. 2006. Marek's disease virus encodes MicroRNAs that map to meq and the latency-associated transcript. J Virol 80:8778-86. Buscaglia, C., and B. W. Calnek. 1988. Maintenance of Marek's disease herpesvirus latency in vitro by a factor found in conditioned medium. J Gen Virol 69 ( Pt 11):2809-18. Calnek, B. W. 2000. Pathogenesis of Marek's disease, p. 25-55. In K. Hirai (ed.), Marek's Disease. Springer, Berlin. Campadelli-Fiume, G., and B. Roizman. 2006. The egress of herpesviruses from cells: the unanswered questions. J Vir0180:6716-7; author replies 6717-9. Cantello, J. L., A. S. Anderson, and R. W. Morgan. 1994. Identification of latency-associated transcripts that map antisense to the ICP4 homolog gene of Marek's disease virus. J Virol 68:6280-90. Cantello, J. L., M. S. Parcells, A. S. Anderson, and R. W. Morgan. 1997. Marek's disease virus latency-associated transcripts belong to a family of spliced RNAs that are antisense to the ICP4 homolog gene. J Virol 71:1353-61. Chattoo, J. P., M. P. Stevens, and V. Nair. 2006. Rapid identification of non- essential genes for in vitro replication of Marek's disease virus by random transposon mutagenesis. J Virol Methods 135:288-91. Chen, M., W. S. Payne, H. Hunt, H. Zhang, S. L. Holmen, and J. B. Dodgson. 2008. Inhibition of Marek's disease virus replication by retroviral vector-based RNA interference. Virology 377:265-72. Churchill, A. E., and P. M. Biggs. 1967. Agent of Marek's disease in tissue culture. Nature 215:528-30. Classon, B. J., and R. L. Boyd. 1998. Thymic-shared antigen-1 (TSA-l). A lymphostromal cell membrane Ly—6 superfamily molecule with a putative role in cellular adhesion. Dev Immunol 6: 149-56. Classon, B. J., and L. Coverdale. 1994. Mouse stem cell antigen Sea-2 is a member of the Ly-6 family of cell surface proteins. Proc Natl Acad Sci U S A 91:5296-300. Classon, B. J ., and L. Coverdale. 1996. Genomic organization and expression of mouse thymic shared antigen-1 (TSA-l): evidence for a processed pseudogene. Immunogenetics 44:222-6. Coltey, M., R. P. Bucy, C. H. Chen, J. Cihak, U. Losch, D. Char, N. M. Le 26 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Douarin, and M. D. Cooper. 1989. Analysis of the first two waves of thymus homing stem cells and their T cell progeny in chick-quail chimeras. J Exp Med 170:543-57. Coltey, M., F. V. Jotereau, and N. M. Le Douarin. 1987. Evidence for a cyclic renewal of lymphocyte precursor cells in the embryonic chick thymus. Cell Differ 22:71-82. Compel, P., and N. A. DeLuca. 2003. Temperature-dependent conformational changes in herpes simplex virus ICP4 that affect transcription activation. J Virol 77 :3257-68. Damiola, F., C. Keime, S. Gonin-Giraud, S. Dazy, and O. Gandrillon. 2004. Global transcription analysis of immature avian erythrocytic progenitors: from self-renewal to differentiation. Oncogene 23:7628-43. Davidson, N. J., and R. L. Boyd. 1992. Delineation of chicken thymocytes by CD3-TCR complex, CD4 and CD8 antigen expression reveals phylogenically conserved and novel thymocyte subsets. Int Immunol4:1175-82. Davison, A. J., and J. E. Scott. 1986. The complete DNA sequence of varicella- zoster virus. J Gen Virol 67 ( Pt 9): 1759-816. Denesvre, C., C. Blondeau, M. Lemesle, Y. Le Vern, D. Vautherot, P. Roingeard, and J. F. Vautherot. 2007. Morphogenesis of a highly replicative EGFPVP22 recombinant Marek‘s disease virus in cell culture. J Virol 81:12348- 59. Dorange, F., B. K. Tischer, J. F. Vautherot, and N. Osterrieder. 2002. Characterization of Marek's disease virus serotype 1 (MDV-1) deletion mutants that lack UL46 to U149 genes: MDV-l UL49, encoding VP22, is indispensable for virus growth. J Virol 76:1959-70. Endoh, D. 1996. Enhancement of gene expression by Marek's disease virus homologue of the herpes simplex virus-1 ICP4. Jpn J Vet Res 44:136-7. Eshel, R., M. Firon, B. Z. Katz, O. Sagi-Assif, H. Aviram, and I. P. Witz. 1995. Microenvironmental factors regulate Ly-6 A/E expression on PyV- transforrned BALB/c 3T3 cells. Immunol Lett 44:209-12. Feng, P., D. N. Everly, Jr., and G. S. Read. 2001. mRNA decay during herpesvirus infections: interaction between a putative viral nuclease and a cellular translation factor. J Virol 75: 10272-80. Feng, P., D. N. Everly, Jr., and G. S. Read. 2005. mRNA decay during herpes simplex virus (HSV) infections: protein-protein interactions involving the HSV 27 37. 38. 39. 40. 41. 42. 43. 45. 46. 47. virion host shutoff protein and translation factors eIF4H and eIF4A. J Virol 79:9651-64. Fleming, T. J., M. L. Fleming, and T. R. Malek. 1993. Selective expression of Ly-6G on myeloid lineage cells in mouse bone marrow. RB6-8C5 mAb to granulocyte-differentiation antigen (Gr—1) detects members of the Ly-6 family. J Immunol 151:2399-408. Fleming, T. J., and T. R. Malek. 1994. Multiple g1ycosylphosphatidylinositol- anchored Ly-6 molecules and transmembrane Ly-6E mediate inhibition of IL-2 production. J Immunol 153:1955-62. Gimeno, 1., and R. F. Silva. 2008. Deletion of the Marek’s disease virus UL41 gene (vhs) has no measurable effect on latency or pathogenesis. Virus Genes 36:499-507. Godfrey, D. I., M. Masciantonio, C. L. Tucek, M. A. Malin, R. L. Boyd, and P. Hugo. 1992. Thymic shared antigen-l. A novel thymocyte marker discriminating immature from mature thymocyte subsets. J Immunol 148:2006-11. Guillemot, F. P., P. D. Oliver, B. M. Peault, and N. M. Le Douarin. I984. Cells expressing Ia antigens in the avian thymus. J Exp Med 160:1803-19. Gumley, T. P., I. F. McKenzie, and M. S. Sandrin. 1995. Tissue expression, structure and function of the murine Ly-6 family of molecules. Immunol Cell Biol 73:277-96. Gupta, M. K., H. V. Chauhan, G. J. Jha, and K. K. Singh. 1989. The role of the reticuloendothelial system in the immunopathology of Marek's disease. Vet Microbiol 20:223-34. He, J., and L. J. Chang. 2004. Functional characterization of hepatoma-specific stem cell antigen-2. Mol Carcinog 40:90-103. Hunt, H. D., and J. E. Fulton. 1998. Analysis of polymorphisms in the major expressed class I locus (B-FIV) of the chicken. Immunogenetics 47 :456-67. Hunt, H. D., B. Lupiani, M. M. Miller, I. Gimeno, L. F. Lee, and M. S. Parcells. 2001. Marek's disease virus down-regulates surface expression of MHC (B Complex) Class 1 (BF) glyc0proteins during active but not latent infection of chicken cells. Virology 282:198-205. Ikuta, K., S. Ueda, S. Kato, and K. Hirai. 1984. Identification with monoclonal antibodies of glycoproteins of Marek's disease virus and herpesvirus of turkeys related to virus neutralization. J Virol 49:1014-7. 28 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. Isler, H. 1976. Fine structure of chicken thymic epithelial vesicles. J Cell Sci 20:135-47. Ito, C. Y., C. Y. Li, A. Bernstein, J. E. Dick, and W. L. Stanford. 2003. Hematopoietic stem cell and progenitor defects in Sca—l/Ly-6A-null mice. Blood 101:517-23. Johnson, D. C., and M. T. Huber. 2002. Directed egress of animal viruses promotes cell-to-cell spread. J Virol 76:1-8. Jones, D., L. Lee, J. L. Liu, H. J. Kung, and J. K. Tillotson. 1992. Marek disease virus encodes a basic-leucine zipper gene resembling the fos/jun oncogenes that is highly expressed in lymphoblastoid tumors. Proc Natl Acad Sci U S A 89:4042-6. Jones, F. E., C. A. Smibert, and J. R. Smiley. 1995. Mutational analysis of the herpes simplex virus virion host shutoff protein: evidence that vhs functions in the absence of other viral proteins. J Virol 69:4863-71. Koppers-Lalic, D., F. A. Rijsewijk, S. B. Verschuren, J. A. van Gaans-Van den Brink, A. Neisig, M. E. Ressing, J. Neefjes, and E. J. Wiertz. 2001. The UL41-encoded virion host shutoff (vhs) protein and vhs-independent mechanisms are responsible for down-regulation of MHC class I molecules by bovine herpesvirus 1. J Gen Virol 82:2071-81. Kosugi, A., S. Saitoh, S. Narumiya, K. Miyake, and T. Hamaoka. 1994. Activation-induced expression of thymic shared antigen-1 on T lymphocytes and its inhibitory role for TCR-mediated H.-2 production. Int Immunol 6:1967-76. Kosugi, A., S. Saitoh, S. Noda, K. Miyake, Y. Yamashita, M. Kimoto, M. Ogata, and T. Hamaoka. 1998. Physical and functional association between thymic shared antigen-l/stem cell antigen-2 and the T cell receptor complex. J Biol Chem 273: 12301-6. Kwong, A. D., and N. Frenkel. 1987. Herpes simplex virus-infected cells contain a function(s) that destabilizes both host and viral mRNAs. Proc Natl Acad Sci U S A 84: 1926-30. Kwong, A. D., J. A. Kruper, and N. Frenkel. 1988. Herpes simplex virus virion host shutoff function. J Virol 62:912-21. Lee, L. F. 1979. Macrophage restriction of Marek's disease virus replication and lymphoma cell proliferation. J Immunol 123: 1088-91. Lee, L. F., B. Lupiani, R. F. Silva, H. J. Kung, and S. M. Reddy. 2008. Recombinant Marek's disease virus (MDV) lacking the Meq oncogene confers 29 60. 61. 62. 63. 65. 66. 67. 68. 69. protection against challenge with a very virulent plus strain of MDV. Vaccine 26:1887-92. Lee, L. F., P. C. Powell, M. Rennie, L. J. Ross, and L. N. Payne. 1981. Nature of genetic resistance to Marek's disease in chickens. J Natl Cancer Inst 66:789-96. Lessard, M., D. L. Hutchings, J. L. Spencer, H. S. Lillehoj, and J. S. Gavora. 1996. Influence of Marek's disease virus strain AC-l on cellular immunity in birds carrying endogenous viral genes. Avian Dis 40:645-53. Levy, A. M., O. Gilad, L. Xia, Y. Izumiya, J. Choi, A. Tsalenko, Z. Yakhini, R. Witter, L. Lee, C. J. Cardona, and H. J. Kung. 2005. Marek's disease virus Meq transforms chicken cells via the v-Jun transcriptional cascade: a converging transforming pathway for avian oncoviruses. Proc Natl Acad Sci U S A 102:14831-6. Li, D., G. O'Sullivan, L. Greenall, G. Smith, C. Jiang, and N. Ross. 1998. Further characterization of the latency-associated transcription unit of Marek's disease virus. Arch Virol 143:295-311. Liu, H. C., H. H. Cheng, V. Tirunagaru, L. Sofer, and J. Burnside. 2001. A strategy to identify positional candidate genes conferring Marek's disease resistance by integrating DNA microarrays and genetic mapping. Anim Genet 32:351-9. Liu, H. C., H. J. Kung, J. E. Fulton, R. W. Morgan, and H. H. Cheng. 2001. Growth hormone interacts with the Marek's disease virus SORF2 protein and is associated with disease resistance in chicken. Proc Natl Acad Sci U S A 98:9203- 8. ' Liu, H. C., M. Niikura, J. E. Fulton, and H. H. Cheng. 2003. Identification of chicken lymphocyte antigen 6 complex, locus E (LY6E, alias SCA2) as a putative Marek's disease resistance gene via a virus-host protein interaction screen. Cytogenet Genome Res 102:304-8. Liu, J. L., Y. Ye, Z. Qian, Y. Qian, D. J. Templeton, L. F. Lee, and H. J. Kung. 1999. Functional interactions between herpesvirus oncoprotein MEQ and cell cycle regulator CDK2. J Virol 73:4208-19. Lupiani, B., Lee, LE, and Reddy, S.M. 2000. Protein-coding content of the sequence of Marek's disease virus serotype-1., p. 159-190. In K. Hirai (ed.), Marek's disease. Springer, Berlin. MacNeil, I., J. Kennedy, D. I. Godfrey, N. A. Jenkins, M. Masciantonio, C. Mineo, D. J. Gilbert, N. G. Copeland, R. L. Boyd, and A. Zlotnik. 1993. Isolation of a cDNA encoding thymic shared antigen-1. A new member of the 30 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. Ly6 family with a possible role in T cell development. J Immunol 151:6913-23. Marek, J. 1907. Mutiple nervenentzundung(Polyneuritis) bei Huhnem. Dtsch. Tierarztl. Wschr. 15:417-421. McHeyzer-Williams, L. J., and M. G. McHeyzer-Williams. 2004. Developmentally distinct Th cells control plasma cell production in vivo. Immunity 20:231-42. Mettenleiter, T. C. 2002. Herpesvirus assembly and egress. J Virol 76:1537-47. Morgan, R. W., L. Sofer, A. S. Anderson, E. L. Bernberg, J. Cui, and J. Burnside. 2001. Induction of host gene expression following infection of chicken embryo fibroblasts with oncogenic Marek's disease virus. J Virol 75:533-9. Morimura, T., S. Miyatani, D. Kitamura, and R. Goitsuka. 2001. Notch signaling suppresses IgH gene expression in chicken B cells: implication in spatially restricted expression of Serrate2/Notch1 in the bursa of Fabricius. J Immunol 166:3277-83. Murphy. J. A., R. J. Duerst, T. J. Smith, and L. A. Morrison. 2003. Herpes simplex virus type 2 virion host shutoff protein regulates alpha/beta interferon but not adaptive immune responses during primary infection in vivo. J Virol 77:9337- 45. Nair, V. 2005. Evolution of Marek's disease -- a paradigm for incessant race between the pathogen and the host. Vet J 170:175-83. Nazerian, K., L. F. Lee, N. Yanagida, and R. Ogawa. 1992. Protection against Marek's disease by a fowlpox virus recombinant expressing the glycoprotein B of Marek's disease virus. J Virol 66: 1409-13. Nazerian, K., J. J. Solomon, R. L. Witter, and B. R. Burmester. 1968. Studies on the etiology of Marek's disease. H. Finding of a herpesvirus in cell culture. Proc Soc Exp Biol Med 127:177-82. Niikura, M., J. Dodgson, and H. Cheng. 2006. Direct evidence of host genome acquisition by the alphaherpesvirus Marek's disease virus. Arch Virol 151:537-49. Niikura, M., H. C. Liu, J. B. Dodgson, and H. H. Cheng. 2004. A comprehensive screen for chicken proteins that interact with proteins unique to virulent strains of Marek's disease virus. Poult Sci 83:1117-23. Noda, S., A. Kosugi, S. Saitoh, S. Narumiya, and T. Hamaoka. 1996. Protection from anti-TCR/CD3-induced apoptosis in immature thymocytes by a signal through thymic shared antigen-l/stem cell antigen-2. J Exp Med 183:2355- 3| 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 60. Ody, C., S. Alais, C. Corbel, K. M. McNagny, T. F. Davison, O. Vainio, B. A. Imhof, and D. Dunon. 2000. Surface molecules involved in avian T-cell progenitor migration and differentiation. Dev Immunol 7 :267-77. Olah, I. a. V., L. 2008. Structure of the avian lymphoid system, p. 13-50. In F. Davison (ed.), Avian Immunology. Elsevier, London. Omar, A. R., and K. A. Schat. 1996. Syngeneic Marek's disease virus (MDV)- specific cell-mediated immune responses against immediate early, late, and unique MDV proteins. Virology 222:87-99. Omar, A. R., and K. A. Schat. 1997. Characterization of Marek's disease herpesvirus-specific cytotoxic T lymphocytes in chickens inoculated with a non- oncogenic vaccine strain of MDV. Immunology 90:579-85. Oroskar, A. A., and G. S. Read. 1987. A mutant of herpes simplex virus type 1 exhibits increased stability of immediate-early (alpha) mRNAs. J Virol 61:604-6. Osterrieder, N., J. P. Kamil, D. Schumacher, B. K. Tischer, and S. Trapp. 2006. Marek's disease virus: from rniasma to model. Nat Rev Microbiol 4:283-94. Pak, A. S., D. N. Everly, K. Knight, and G. S. Read. 1995. The virion host shutoff protein of herpes simplex virus inhibits reporter gene expression in the absence of other viral gene products. Virology 211:491-506. Parcells, M. S., A. S. Anderson, J. L. Cantello, and R. W. Morgan. 1994. Characterization of Marek's disease virus insertion and deletion mutants that lack USl (ICP22 homolog), U810, and/or US2 and neighboring short-component open reading frames. J Virol 68:8239-53. Parcells, M. S., A. S. Anderson, and T. W. Morgan. 1995. Retention of oncogenicity by a Marek's disease virus mutant lacking six unique short region genes. J Virol 69:7888-98. Patterson, J. M., M. H. Johnson, D. B. Zimonjic, and T. A. Graubert. 2000. Characterization of Ly-6M, a novel member of the Ly-6 family of hematopoietic proteins. Blood 95:3125-32. Petrenko, 0., I. Ischenko, and P. J. Enrietto. 1997. Characterization of changes in gene expression associated with malignant transformation by the NF-kappaB family member, v-Rel. Oncogene 15:1671-80. Ramm, H. C., T. J. Wilson, R. L. Boyd, H. A. Ward, K. Mitrangas, and K. J. Fahey. 1991. The effect of infectious bursal disease virus on B lymphocytes and bursal stromal components in specific pathogen-free (SPF) White Leghorn 32 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. chickens. Dev Comp Immunol 15:369-81. Ramm, H. C., T. J. Wilson, R. L. Boyd, H. A. Ward, K. Mitrangas, and K. J. Fahey. 1991. The effect of infectious bursal disease virus on B lymphocytes and bursal stromal components in specific pathogen-free (SPF) White Leghorn chickens. Dev Comp Immunol 15:369-81. Randle, E. S., G. A. Waanders, M. Masciantonio, D. I. Godfrey, and R. L. Boyd. 1993. A lymphostromal molecule, thymic shared Ag-l, regulates early thymocyte development in fetal thymus organ culture. J Immunol 151:6027-35. Ratcliffe, M. 2008. B cells, the bursa of Fabricius and the generation of antibody repertoires., p. 67-90. In F. Davison (ed.), Avian Immunology. Elsevier, London. Read, G. S., and N. Frenkel. 1983. Herpes simplex virus mutants defective in the virion-associated shutoff of host polypeptide synthesis and exhibiting abnormal synthesis of alpha (immediate early) viral polypeptides. J Virol 46:498- 512. Reynaud, C. A., V. Anquez, H. Grimal, and J. C. Weill. 1987. A hyperconversion mechanism generates the chicken light chain preimmune repertoire. Cell 48:379-88. Rieker, T., J. Penninger, N. Romani, and G. Wick. 1995. Chicken thymic nurse cells: an overview. Dev Comp Immunol 19:281-9. Rock, K. L., H. Reiser, A. Bamezai, J. McGrew, and B. Benacerraf. 1989. The LY-6 locus: a multigene family encoding phosphatidylinositol-anchored membrane proteins concerned with T-cell activation. Immunol Rev 111:195-224. Roizmann, B., R. C. Dosrosiers, B. Fleckenstein, C. Lopez, A. C. Minson, and M. J. Studdert. 1992. The family Herpesviridae: an update. The Herpesvirus Study Group of the International Committee on Taxonomy of Viruses. Arch Virol 123:425-49. Sakaguchi, M., T. Urakawa, Y. Hirayama, N. Miki, M. Yamamoto, and K. Hirai. 1992. Sequence determination and genetic content of an 8.9-kb restriction fragment in the short unique region and the internal inverted repeat of Marek's disease virus type 1 DNA. Virus Genes 6:365-78. Schat, K. A., and Z. Xing. 2000. Specific and nonspecific immune responses to Marek's disease virus. Dev Comp Immunol 24:201-21. Schat, K. A. a. M.-G., C. J. 2000. Immune Responses to Marek's Disease Virus Infection. In K. Hirai (ed.), Marek‘s Disease. Springer, Berlin. 33 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. Schek, N., and S. L. Bachenheimer. 1985. Degradation of cellular mRNAs induced by a virion-associated factor during herpes simplex virus infection of Vero cells. J Virol 55:601-10. Schlueter, A. J., T. R. Malek, C. N. Hostetler, P. A. Smith, P. deVries, and T. J. Waldschmidt. 1997. Distribution of Ly-6C on lymphocyte subsets: 1. Influence of allotype on T lymphocyte expression. J Immunol 158:4211-22. Schmitt, T. M., and J. C. Zuniga-Pflucker. 2006. T-cell deveIOpment, doing it in a dish. Immunol Rev 209:95-102. Schumacher, D., B. K. Tischer, W. Fuchs, and N. Osterrieder. 2000. Reconstitution of Marek's disease virus serotype 1 (MDV-1) from DNA cloned as a bacterial artificial chromosome and characterization of a glycoprotein B- negative MDV-l mutant. J Virol 74:11088-98. Schumacher, D., B. K. Tischer, S. M. Reddy, and N. Osterrieder. 2001. Glycoproteins E and I of Marek‘s disease virus serotype 1 are essential for virus growth in cultured cells. J Virol 75: 1 1307-18. Shamblin, C. E., N. Greene, V. Arumugaswami, R. L. Dienglewicz, and M. S. Parcells. 2004. Comparative analysis of Marek's disease virus (MDV) glyc0protein-, Iytic antigen pp38- and transformation antigen Meq-encoding genes: association of meq mutations with MDVs of high virulence. Vet Microbiol 102: 147-67. Shevach, E. M., and P. E. Korty. 1989. Ly-6: a multigene family in search of a function. Immunol Today 10:195-200. Silva, R. F. 2000. The Genome Structure of Marek‘s Disease Virus, p. 143-158. In K. Hirai (ed.), Marek's Disease. Springer, Berlin. Smith, T. J., L. A. Morrison, and D. A. Leib. 2002. Pathogenesis of herpes simplex virus type 2 virion host shutoff (vhs) mutants. J Virol 76:2054—61. Solomon, J. J., R. L. Witter, K. Nazerian, and B. R. Burmester. 1968. Studies on the etiology of Marek's disease. I. Propagation of the agent in cell culture. Proc Soc Exp Biol Med 127:173-7. Sowder, J. T., C. L. Chen, L. L. Ager, M. M. Chan, and M. D. Cooper. 1988. A large subpopulation of avian T cells express a homologue of the mammalian T gamma/delta receptor. J Exp Med 167:315-22. Strelow, L., T. Smith, and D. Leib. 1997. The virion host shutoff function of herpes simplex virus type 1 plays a role in corneal invasion and functions independently of the cell cycle. Virology 231:28-34. 34 117. 118. 119. 1 20. 121. 122. 123. 124. 125. 126. 127. Strelow, L. 1., and D. A. Leib. 1995. Role of the virion host shutoff (vhs) of herpes simplex virus type 1 in latency and pathogenesis. J Virol 69:6779-86. Su, Y. H., X. Zhang, X. Wang, N. W. Fraser, and T. M. Block. 2006. Evidence that the immediate-early gene product ICP4 is necessary for the genome of the herpes simplex virus type 1 ICP4 deletion mutant strain d120 to circularize in infected cells. J Virol 80:11589-97. Sugimoto, K., M. Uema, H. Sagara, M. Tanaka, T. Sata, Y. Hashimoto, and Y. Kawaguchi. 2008. Simultaneous tracking of capsid, tegument, and envelope protein localization in living cells infected with triply fluorescent herpes simplex virus 1. J Vir0182:5198-211. Suzutani, T., M. Nagamine, T. Shibaki, M. Ogasawara, I. Yoshida, T. Daikoku, Y. Nishiyama, and M. Azuma. 2000. The role of the U14] gene of herpes simplex virus type 1 in evasion of non-specific host defence mechanisms during primary infection. J Gen Vir0181:1763-71. Taddeo, B., and B. Roizman. 2006. The virion host shutoff protein (UIAI) of herpes simplex virus 1 is an endoribonuclease with a substrate specificity similar to that of RNase A. J Virol 80:9341-5. Taddeo, B., M. T. Sciortino, W. Zhang, and B. Roizman. 2007 . Interaction of herpes simplex virus RNase with VP16 and VP22 is required for the accumulation of the protein but not for accumulation of mRNA. Proc Natl Acad Sci U S A 104:12163-8. Tregaskes, C. A., N. Bumstead, T. F. Davison, and J. R. Young. 1996. Chicken B-cell marker chB6 (Bu-1) is a highly glycosylated protein of unique structure. Immunogenetics 44:212-7. Uchida, N., H. L. Aguila, W. H. Fleming, L. Jerabek, and I. L. Weissman. 1994. Rapid and sustained hematopoietic recovery in lethally irradiated mice transplanted with purified Thy-1.110 Lin-Sca-1+ hematopoietic stem cells. Blood 83:3758-79. Vallejo, R. L., L. D. Bacon, H. C. Liu, R. L. Witter, M. A. Groenen, J. Hillel, and H. H. Cheng. 1998. Genetic mapping of quantitative trait loci affecting susceptibility to Marek's disease virus induced tumors in F2 intercross chickens. Genetics 148:349-60. Venugopal, K., and L. N. Payne. 1995. Molecular pathogenesis of Marek's disease-recent developments. Avian Pathol 24:597-609. Volpini, L. M., B. W. Calnek, B. Sneath, M. J. Sekellick, and P. 1. Marcus. 35 128. 129. l 30. 131. 132. 133. l 34. 135. 1996. Interferon modulation of Marek's disease virus genome expression in chicken cell lines. Avian Dis 40:78-87. Watanabe, N., Y. H. Wang, H. K. Lee, T. Ito, W. Cao, and Y. J. Liu. 2005. Hassall's corpuscles instruct dendritic cells to induce CD4+CD25+ regulatory T cells in human thymus. Nature 436:1181-5. Witter, R. L. 1997. Increased virulence of Marek’s disease virus field isolates. Avian Dis 41:149-63. Witter, R. L. and Kreager, K. S. 2004. Serotype 1 viruses modified by backpassage or insertional mutagenesis: approaching the threshold of vaccine efficacy in Marek’s disease. Avian Dis 48: 149-163. Xie, Q., A. S. Anderson, and R. W. Morgan. 1996. Marek's disease virus (MDV) ICP4, pp38, and meq genes are involved in the maintenance of transformation of MDCC-MSBI MDV-transformed lymphoblastoid cells. J Virol 70:1125-31. Yamada, H., T. Daikoku, Y. Yamashita, Y. M. Jiang, T. Tsurumi, and Y. Nishiyama. 1997. The product of the US10 gene of herpes simplex virus type 1 is a capsid/tegument-associated phosphoprotein which copurifies with the nuclear matrix. J Gen Virol78 (Pt 11):2923-31. Yao, Y., Y. Zhao, H. Xu, L. P. Smith, C. H. Lawrie, M. Watson, and V. Nair. 2008. MicroRNA profile of Marek's disease virus-transformed T—cell line MSB-l: predominance of virus-encoded microRNAs. J Vir0182:4007-15. Yonash, N., L. D. Bacon, R. L. Witter, and H. H. Cheng. 1999. High resolution mapping and identification of new quantitative trait loci (QTL) affecting ‘ susceptibility to Marek's disease. Anim Genet 30: 126-35. Zammit, D. J., S. P. Berzins, J. W. Gill, E. S. Randle—Barrett, L. Barnett, F. lKoentgen, G. W. Lambert, R. P. Harvey, R. L. Boyd, and B. J. Classon. 2002. Essential role for the lymphostromal plasma membrane Ly-6 superfamily molecule thymic shared antigen 1 in development of the embryonic adrenal gland. Mol Cell Biol 22:946-52. 36 Chapter 2 Cloning of and functional characterization of chicken stem cell antigen 2 37 Abstract: Stem cell antigen 2 (SCA2) is a Ly-6 family member whose function is largely unknown. To characterize biological properties and tissue distribution of chicken SCA2, SCA2 was expressed in E. coli, purified, and a polyclonal antibody developed. Utilizing the polyclonal antibody, SCA2 was detected to be a 13 kDa cell surface protein anchored by a glycosyl-phosphatidylinositol (GPI) moiety. SCA2 is expressed in connective tissues of thymus and bursa based on immunohistochemistry, immunoprecipitation, and western blots. In bursal follicles, SCA2 is specifically expressed on the cortical-medullary epithelial cells (CMEC) surrounded by MHC class H presenting cells. Expression profiles of bursal cells induced by contact with SCA2-expressing cells show down-regulation of numerous genes including CD793, B cell linker (BLNK), spleen tyrosine kinase (SYK), and gamma 2-phospholipase C (PLCGZ) that are involved in the BCR and immune response signaling pathways. These results suggest chicken SCA2 plays a role in regulating B lymphocytes. 38 Introduction Stem cell antigen 2 (SCA2), also known as lymphocyte antigen 6 complex, locus E (Ly-6E) and thymic shared antigen-1 (TSA-l), is a member of the lymphostromal cell membrane Ly-6 superfamily. Ly-6 family members are low molecular weight (10-12 kDa) proteins with 8-10 conserved cysteine residues, anchored to the cell surface via a glycosyl-phosphatidylinositol (GPI) moiety. In mouse, SCA2 is expressed on thymic medullary epithelial cells and immature T lymphocytes, B lymphocytes, and activated T cells (Godfrey et al., 1992; Classon et al., 1994; Antica and Scollay, 1997). Aside from lymphoid organs, SCA2 is also expressed in other mouse tissues including embryonic adrenal gland and liver, and adult kidney and liver (MacNeil et al., 1993; Classon and Coverdale, 1996; Zammit et al., 2002). SCA2 has been suggested to function in cell-cell adhesion (Classon and Boyd, 1998), signaling transduction via T cell receptor, and T cell development (Randle et al., 1993; Fleming and Malek, 1994; Saitoh et al., 1995; Noda et al., 1996). Some studies showed mouse SCA2 expression was highly associated with tumorigenicity (Eshel et al., 1995; He and Chang, 2004) although the role of SCA2 in tumor progression is not well defined. The SCA2 homologue in chicken was initially identified as a transcript highly expressed in v-Rel-transformed avian bone marrow cells (Petrenko et al., 1997). In self- renewing avian erythrocytic progenitor cells called T2ECs (transforming growth factor (TGF)-0t, TGF-B induced erythrocytic cells), SCA2 mRNA is up-regulated T2ECs in comparison to differentiated T2ECs, and inhibition of SCA2 expression through RNA interference substantially reduced cell proliferation and increased differentiation of T2ECs (Gandrillon et al., 1999; Damiola et al., 2004; Bresson-Mazet et al., 2008). These 39 results imply that SCA2 supports the self-renewal of chicken lymphoid cells, and the up- regulation of SCA2 may promote the transformation of lymphoid cells. We previously reported that the allele of the chicken SCA2 can influence resistance to Marek’s disease (MD), a viral-induced T cell lymphoma, based on linkage to MD resistance and related traits, induced transcription in Marek’s disease virus (MDV)-infected chicken embryo fibroblasts (CEF), transcriptional differences between MD resistant and susceptible experimental chicken lines following challenge with MDV, and direct protein interaction with MDV USlO (Liu et al., 2003). Although studies (Gandrillon et al., 1999; Damiola et al., 2004; Bresson-Mazet et al., 2008) have shown that chicken SCA2 may participate in cell proliferation, differentiation, or tumorigenesis, the biochemical and biological properties of SCA2 are not understood. These studies are further hindered by the lack of a specific antibody to SCA2. In this report, we describe the purification of chicken SCA2 protein in E. coli and development of a polyclonal antibody to SCA2. Utilizing this antibody, the biological properties of chicken SCA2 were investigated. Chicken SCA2 is a 13 kDa protein that is anchored on the cell surface by a GPI moiety. In vivo studies show chicken SCA2 is expressed in liver cells, and cells in connective tissues of thymus and bursa. Within bursa follicles, the MHC class II presenting cells are adjacent to the cortical-medullary epithelial cells (CMEC) expressing SCA2. Bursa] cells incubated with a SCA2- expressing cell line down-regulate many genes involved in BCR and immune response signaling pathways. These results suggest a role of SCA2 as a microenvironment factor regulating B cell deve10pment, and. it may act as a functional regulator in lymphomatous disease. 40 Materials and methods Cells and animals DFl, an immortalized cell line from chicken embryonic fibroblasts (Himly et al., 1998), was used for transfection of the SCA2Flag-pSELECT-zeo-mcs vector (Invivogen, San Diego, CA, USA). Chicken lines 0 (lacks endogenous avian leucosis viruses), 63 (inbred and MD resistant), and 72 (inbred and MD susceptible) were maintained at the Avian Disease and Oncology Laboratory (ADOL), USDA. White rabbits were purchased from Harlan Laboratories, Inc. (Indianapolis, IN, USA). Development of anti-SCA2 rabbit antisera Primers (SCA2F: 5’ GC GAA TCC CAC TCA TCT GCT TIT CGT GCT CG 3’; SCA2R: 5’ CG CTC GAG TGG GCC GCC TAA TAG CTG GCT TTA ACG C 3’; restriction sites are underlined) based on GenBank accession no. NM_204775 were designed to amplify chicken SCA2 without the signal peptide and GPI anchor (bases 115 to 375). Using a chicken spleen cDNA library as template. the amplicon was cloned into the pET-28c prokaryotic expression vector between the EcoRI and XhoI sites (Novagen, Madison, WI, USA), forwarded by DNA sequencing to verify the construct. The recombinant vector containing both N- and C-terminal His tags was transformed into E. coli strain BL21 (DE3) (Novagen). The expression and purification of the SCA2-His recombinant protein followed the pET system protocol (Novagen). The expressed SCA2 was examined through western blotting using anti-His tag monoclonal antibody (mAb; Invitrogen, Carlsbad, CA, USA) and purified through His . Bind kits (Novagen). Purified chicken SCA2 was electrophoresed through a 10% SDS-PAGE gel, and the 13 kDa band 41 removed and mixed with Irnject Alum (Pierce, Rockford, IL, USA). The antigen-alum mixture was administered into multiple intramuscular, intradermal, or subcutaneous sites of two rabbits. Four weeks after the primary immunization, booster immunizations were given and the animals were bled 7 days later. The anti-SCA2 sera were obtained after the fifth boost. Generation of chicken SCA2 eukaryotic expression vector and stable transfection cell line Primers (SCA2SalIF: 5’ ATCG GTC GAC ATG AAG GCG TIT CTG TTC GC 3’; SCA2NCOIR: 5’ ATCG CCA TGG TCA CTC ACG AGC CCT GAG G 3’; restriction sites are underlined) were designed to amplify the entire coding portion of chicken SCA2 (bases 55 to 435). The amplicon was inserted between the Sall and NcoI sites of the pSELECT—zeo-mcs vector (Invivogen) to generate SCA2-pSELECT-zeo—mcs. A Flag tag was inserted immediately downstream of the signal peptide of SCA2 by PCR using SCA2-pSELECT-zeo—mcs as the template with the primers (SCA2SalIF; Sca2FlagNsiIR: 5’ GTT GGA GG ATG CAT CCG AGC ACG AAA AGC AGA TGA GCT TAT CGT CGT CAT CCT TGT AAT CCG TGT GGG CTC TCT CCA CG 3’; restriction sites are underlined, Flag tag sequences are in italic). The amplicon was inserted between the Sall and Nsil sites of SCA2-pSELECT-zeo-mcs to construct SCA2F1ag-pSELECT, which was confirmed by DNA sequencing. The SCA2F1ag- pSELECT-zeo—mcs and empty pSELECT—zeo—mcs vector were independently transfected into the DFl cell line using the Basic Nucleofector Kit (Amaxa Inc., Gaithersburg, MD, USA). The transfected cells were incubated in medium containing Liebovitz’s L-15 and McCoy 5A media (1:1) supplemented with 5% fetal bovine serum (FBS), penicillin, streptomycin, and zeocin (300 ng/ml) until single colonies appeared. The expression of SCA2 was tested by western blots using anti-Flag mAb and anti-SCA2 antisera. The single colony with the best SCA2 expression was maintained and called the SCA2F1ag- DFl cell line. The single colony with the empty vector was maintained and named the Vector-DFl cell line. Flow cytometry analysis Cell surface expression of SCA2 was measured as previously reported (Hunt et al., 2001). SCA2F1ag-DF1 cells and DFl cells were trypsinized and harvested (lOOOxg, 5 min.). Thymocytes were collected from 2 to 4-week old ADOL Line 72 and Line 63 chicken thymus, followed by centrifugation (lOOOxg, 5 min.). 105 cells were washed with cold flow cytometry buffer (phosphate buffered saline containing 1% FBS and 0.1% sodium azide). Anti-Flag M2 mAb (1:100; Sigma-Aldrich Inc, Milwaukee, WI, USA) and anti-SCA2 antisera (1:100) were diluted in flow cytometry buffer and reacted with cells for 20 min. at 4 0C, followed by washing and incubation with goat anti-mouse and goat anti-rabbit IgG conjugated with FITC (1:100; Zymed, South San Francisco, CA, USA). The cells were analyzed using a Becton-Dickinson FACScaliber flow cytometer and CellQuest Pro software (BD Bioscience, San Jose, CA, USA). Cleavage of GPI-linked cell surface SCA2 molecules 105 SCA2F1ag-DF1 and DFl cells were incubated with 0.2-0.5 unit of phosphatidylinositol-specific phospholipase C (PI-PLC; Sigma-Aldrich Inc.) in 200 pl 43 flow cytometry buffer at 37 0C for 30 min. Cells were analyzed by flow cytometry analysis as described in section 2.4. Western blots and immunoprecipitation Cells and tissues were lysed in Cellytic M Mammalian Cell lysis reagent (Sigma- Aldrich Inc., Milwaukee, WI, USA), supplemented with 1% P8340 protease inhibitor cocktails (Sigma-Aldrich Inc.), followed by sonication and centrifugation. Cell lysates were mixed with loading and reducing buffer, heated to 70°C for 10 min. and resolved using 10% SDS-PAGE gels (Invitrogen). The samples were blotted on Immobilon-PSQ PVDF transfer membranes (Millipore, Billerica, MA, USA). After blocking in PBS containing 5% non-fat milk for 1 hour at room temperature, membranes were probed with anti-Flag M2 mAb (1:100) or anti-SCA2 antisera (1:100) at 4 0C overnight. Membranes were washed three times in PBS before incubation with goat anti-mouse or goat anti- rabbit IgG conjugated with horseradish peroxidase (Zymed) at room temperature for 1 hour. Proteins were detected using ECL Western blotting substrate (Pierce) after washing three times. Immunoprecipitation using anti-SCA2 antisera followed the instructions of the Immunoprecipitation Kit (Roche, Indianapolis, IN, USA). Immunohistochemistry Multiple tissues were collected from 4-week old ADOL Line 72 birds. The tissues were soaked into compound optimum cutting temperature embedding medium (Tissue- Tek, Torrance, CA, USA) and stored at -80 0C. Samples of flash frozen chicken bursa were sectioned at 5 pm on a Cryotome FSE (Thermo Fisher, Pittsburgh, PA, USA) at - 44 20 0C. Slides were placed in -70 DC for storage until staining. Reagent incubation was as follows: anti-SCA2 antisera (1:1000) in normal antibody diluent (NAD) (Scytek Laboratories, Logan, UT, USA) for 60 min. at room temperature; biotinylated goat anti- rabbit IgG (h+l) (Vector Laboratories, Burlingame, CA, USA) in NAD for 30 min.; Alkaline Phosphatase Complex Reagent (Kirkeguard & Perry Laboratories, Gaithersburg, MD, USA) for 60 min.; and developed using Vector Red substrate I (Vector Laboratories, Burlingame, CA, USA) for 8 min. Double staining was performed on the frozen bursa follicles. Tissue sections were incubated with anti-SCA2 antisera (121000) in NAD for 60 min. at room temperature; biotinylated goat anti-rabbit IgG(h+l) in NAD for 30 min.; Vectastain Elite ABC Reagent (Vector Laboratories) for 30 min.; and developed using DAB (Vector Laboratories) for 15 min.; followed with the second antibody anti-MCH II mAb TAPl (1:200; Developmental Studies Hybridoma Bank. Iowa City, IA, USA) in NAD for 60 min. at room temperature; biotinylated horse anti-mouse IgG(h+l) (Vector Laboratories) in NAD for 40 min.; Alkaline Phosphatase Complex Reagent for 60 min.; and developed using Vector Red substrate I for 8 min. Affymetrix array processing and data analysis Bursa of 3-week old line 0 chickens were collected and manually teased to harvest the bursal cells. 2x106 bursal cells were spread on the monolayers of SCA2F1ag- DFl cells or control Vector-DFI cells in triplicate. After growth in the mixture of Liebovitz’s L-15 and McCoy 5A media (1:1) supplemented with 4% FBS, penicillin, and streptomycin for overnight at 37 °C in 5% CO2, the bursal cells were recovered by 45 washes from the adhesive DFl cells followed by gentle centrifugation at 1000xg for 10 min. Total RNA was isolated using the absolutely RNA microprep kit (Stratagene, La Jolla, CA, USA). The quantity and quality of total RNA were analyzed on an Agilent 2100 Bioanalyzer using the RNA 6000 Nano Chip Kit according to the manufacturer’s instructions (Agilent Technologies, Santa Clara, CA, USA). The cDNA synthesis, labeling, hybridization and scanning of Affymetrix GeneChip Chicken Array (Affymetrix, Inc., Santa Clara, CA, USA) were carried out at the Michigan State University Research Support Technology Facility (RSTF). The .CEL files (GEO accession no.GSE18506) were provided and normalized at the probe level using the robust multichip average (RMA) method (Irizarry et al., 2003) implemented into the Affymetrix Expression Console software; RMA is a robust linear model analyzing data for background adjustment, quantile normalization, and log2-transformation of the perfect match probe intensities for each probe set. Statistical comparison of differences between the normalized mRNA expression in SCA2-induced and control bursal cells was performed through significance analysis of microarrays (SAM) software (Tusher, et al., 2001). The significant differential expressions between SCA2-induced and control samples were filtered based on the following criteria: 1) false discovery rate<5% and 2) fold change 22. Gene ontology annotations associated to Affymetrix DNA probe sets were analyzed through Ingenuity Pathways Analysis software (Ingenuity Systems Inc, Redwood, CA, USA). Biological processes were identified by inputting differentially expressed gene sets into the Metacore database (GeneGo Inc., St. Joseph, MI, USA) and the biological processes were ranked according to their P value. 46 Real-Time quantitative reverse transcription PCR Selected chicken transcripts were measured using real time quantitative reverse transcription PCR (qRT-PCR). Briefly, 1 pg of total RNA was used for cDNA synthesis with the Improm-II Reverse Transcription System according to the manufacturer’s protocol (Promega, Madison, WI, USA). The primers were designed using Primer Express Software v3.0 (Applied Biosystems, Foster City, CA, USA). Primers listed in Table 2-1 were synthesized by Operon Biotechnologies, Inc (Huntsville, Al, USA). 2.5 uM of each forward and reverse primer were used in the presence of 25 ul SYBR GREEN PCR master mix (Applied Biosystems, Foster City, CA, USA) in a final volume of 50 pl. The real-time PCR program was as follows: 95 °C for 10 min., 40 cycles at 95 °C for 15 sec followed by 60 °C for 1 min., in an ABI 7500 Sequence Detection System (Applied Biosystems). The expression of each gene was normalized to the expression of glyceraldehyde 3-phosphate dehydrogenase (GAPDH) through relative quantification using the 2'AACT method (Livak and Schmittgen, 2001). 47 Table 2-1. Primers used for quantitative real-time PCR validation of microarray data Gene Primer Sequence (5’ to 3’) GAPDH Forward TGCCATCACAGCCACACAGAAG Reverse ACT I ' I 'CCCCACAGCCT'I‘AGCAG CD798 Forward AGGAGTI’CCACGTGCTGGAT Reverse GAAGCTGATGCGGTCA'I'I'I'GT S YK Forward CCAAAAACCTCAGCTGGAGAA Reverse TCTGACTCTI‘CCCGAGAGATCCT BTK Forward CCCAGAGCTCATCACGTACCA Reverse GACACGGGATACTTCAGTCTGGAT BLNK Forward GCC'ITGCCCAGACCAAAGA Reverse CCCTCGGCTI‘CAGTGGTAATI‘ PLCGZ Forward GAA'I'I'I'GATCTGCGTCTCACTGA Reverse CCTCGACTCAAGTTGCTATAGTACCA TH YI Forward CAGCGTCTCCGAGAACGTCTA Reverse GAGGCACACCAGGTTCTTGTG 48 Results Chicken SCA2 is a 13 kDa cell surface protein anchored by a GPI moiety The eukaryotic expression vector SCA2F1ag-pSELECT-zeo-mcs includes the entire SCA2 cDNA sequence and a Flag tag immediately downstream of the SCA2 signal peptide. In SCA2F1ag-DP] stably transfected cells, SCA2F1ag ran with a mobility of about 13 kDa on denatured SDS-PAGE gels as measured by western blot analysis utilizing both anti-Flag mAb and anti-SCA2 polyclonal antisera (Fig. 2-1). SCA2 encodes 126 amino acids and the mature SCA2 protein is predicted to possess 85 amino acids. The 13 kDa molecular size suggests potential posttranslational modifications of SCA2 in DFl cells. In SCA2F1ag-DF1 cells, SCA2 is displayed on the cell surface as determined by flow cytometry using both anti-Flag mAb and anti-SCA2 antisera (Fig. 2-2). When the viable SCA2F1ag-DF1 cells were incubated with PI-PLC, a phospholipase that specifically cleaves the GPI anchor, the SCA2 molecule is cleaved from the surface of SCA2F1ag-DP] cells in a dose dependent manner. The surface expression of control DFl cells were not affected by the enzyme. These results confirm the surface expression of SCA2 with a direct attachment of the GPI anchor to the C terminus of SCA2. 49 IdG'JODQA [d0'10139A U) o 8 > > N N 3 3 vi a? a a 75 kDa 75 kDa 50”” 50 kDa 37 kDa 37kDa 23:3: 2.... 20 kDa 131$: 15 kDa 10 kDa Figure 2-1. Western blots using anti-Flag mAb or anti-SCA2 antisera. DFl cells transfected with SCA2F1ag or empty vector probed with a) anti-Flag mAb or b) anti- SCA2 antisera. The size of the protein standards are indicated. The molecular size of SCA2 is about 13 kDa. 50 a) b) Figure 2—2. Flow cytometry analysis of SCA2F1ag-DP] and vector-DFI with anti-Flag mAb and anti-SCA2 antisera. a) SCA2F1ag-DF1 (green line) and Vector-DFI (black line) stained with anti-Flag mAb; SCA2F1ag-DP] treated with 0.2 units and 0.5 units of PI- PLC stained with anti-Flag mAb (pink and blue dotted lines, respectively). b) SCA2F1ag- DFl stained with anti-SCA2 antisera (green line) and prebleed antisera (black line); and SCA2F1ag-DP] treated with 0.2 units and 0.5 units of PI-PLC and stained with anti- SCA2 antisera (pink and blue dotted lines, respectively), SCA2F1ag-DF1 treated with 0.5 units of PI-PLC and stained with prebleed antisera (purple dotted line). 51 Tissue distribution of chicken SCA2 Frozen sections of selected organs were stained with the anti-SCA2 polyclonal antisera or prebleed antisera as a negative control. As shown (Fig. 2-3), SCA2 is strongly expressed in liver cells and connective cells in thymus and bursa. In the liver, hepatocytes were stained uniformly for SCA2 and the endothelial cells also stained strongly. In the thymus, SCA2 is expressed on connective epithelial cells. In bursa, SCA2 is specifically expressed on CMEC and medullary reticular epithelial cells. In the spleen, weak staining was observed in red pulp areas. Weak staining was also observed in gonad, heart and lung tissues. The molecular weight of SCA2 expression in these organs was determined by western blot. SCA2 ran with a mobility of 13 kDa, consistent with SCA2F1ag expressed in SCA2F1ag-DF1 cells (Fig. 2-4). The results show abundant SCA2 expression in thymus, bursa, liver and spleen and relatively sparse expression in gonads, heart and lung. In bursal follicles, SCA2 is specifically expressed in CMEC and reticular epithelial cells in the medulla (Fig. 2-5a). In the immunohistochemistry image of bursa follicles stained with both anti-SCA2 antisera and anti-MHC class H mAb, the majority of MHC class II presenting cells are located adjacent to the CMEC on which SCA2 is strongly and specifically stained (Fig. 2-5b, c). Under the fluorescent microscope, the areas positive for MHC class II and the areas positive for SCA2 do not overlap. The immunohistochemistry results show CMEC and MHC class II presenting cells are two distinct types of cells, and the MHC class II presenting cells are probably attracted to the CMEC on both sides of the medullar cortex. 52 Figure 2-3. Immunohistochemical analysis of chicken SCA2 expression in chicken tissues. Tissues from 4-week old line 72 birds were stained with anti- SCA2 antisera. Magnification of the bursal follicle is 400x. All the other magnifications are 100x. 53 Figure 2-3 -l I" l— C) :1: u) a s 2 s a 2 5 5 3 0° 2 n: :1 8 (’3 g a ‘1 = 6 75kDa 50kDa 37kDa 25kDa 20kDa 15kDa lOkDa Figure 2-4. Immunoprecipitation and western blot analysis of SCA2 expression in chicken tissues. Cell lysates fi'om chicken tissues were immunoprecipitated with anti- SCA2 antisera and detected with anti-SCA2 antisera through western blots. The size of the protein standards are indicated. 55 Figure 2-5. Immunohistochemical analysis of SCA2 expression in bursal follicles. a) Bursa] follicles were doubly stained with anti-SCA2 antisera (gray/black) and anti—MHC class H mAb (red). An identical section viewed under b) bright and c) fluorescent lighting; with the fluorescent microscope, only MHC class II is illuminated in red. 56 a) b) C) Figure 2—5 Gene expression profiling of bursal cells in contact with SCA2-expressing cells Bursa] cells incubated with SCA2-expressing (SCA2F1ag-DF1) or negative control (Vector-DFI) cells were profiled using Affymetrix chicken genome arrays and analyzed. 95 genes are significantly under-expressed (Table 2-2) and 68 genes are over- expressed (Table 2-3) in the SCA2-induced bursal cells in comparison to the control. In contrast to the over-expressed genes, many of the genes with decreased expression were related to B cell proliferation, development, and immune response. The results of biological process analysis through the MetaCore database showed the set of down- regulated genes is induced in a statistically significant manner in comforrnent of several pathways including the BCR signaling pathway (P _<_ 1.6 x 107) and immune response signaling in B lymphocytes (P g 3.2 x 10'“). To validate the microarray results, selected under-expressed (CD798, SYK, BLNK BTK, PLCGZ) and over-expressed (THY-I) genes were analyzed by qRT-PCR. Using the chicken GAPDH gene for normalization, all the queried genes showed an expression pattern similar to the results from the microarray (Table 2-4). 58 Table 2-2. Genes under-express in SCA2 induced bursal cells Probe ID Gga.5273.1.Sl_at GgaAffx.9065.1.Sl_at GgaAffx.2l699.1.Sl_at Gga.7334.1.Sl_at member 1 GgaAffx.7881.l.Sl_s_at Gga.19387. l.Sl_at G gaAffx.20426. 1 .S l_at Gga.3738.1.Sl_at GgaAffx.] 1885.1 .Sl_s_at Gga.10069.l.Sl_at translocated to, 3 Gga.5128.l.Sl_at GgaAffx.] 1849.1.Sl__s_at Gga.998.2.Sl_a_at Gga.14078.1.Sl_s__at GgaAffx. 151.1.Sl_s_at Gga.17250.1.Sl_at Gga.9816.l.Sl_s_at GgaAffx. 12918.1.Sl_s_at GgaAffx.21308.1.Sl__s_at Gga.19843.].S1_s_at Gga.3573.2.Sl_a_at Gga.6002. l .S l_a_at Gga.6032.1.Sl_at Gga.445.1.Sl_at Gga.8031.2.Sl_a_at Gga.13533.1.S1_at Gga.4564.2.Sl_a_at Gga.588. 1 .S l_at GgaAffx.]3238.1.Sl_s_at Gga.9138.l.Sl_at Gga.5289.1.Sl_at Gga.1092.1.Sl_s_at Gga.10880.1.Sl_at GgaAffx.] 1705. l .Sl_at GgaAffx.] 1387.1.Sl__s__,at Gga.11016.1.Sl_at Gga.13336.2.Sl_x_at Gga.16655.l.Sl_at Gga.14571.1.Sl_s_at Gga.11539.l.Sl_at Gga.3366. 1 .Sl_at Gga.4326. 1 .S l_at G gaAffx.26566. l .Sl_at Gga.1235.2.Sl_at subfamily, beta member 2 Gga.9912.1.S1_at GgaAffx.] l727.l.Sl__at GgaAffx. 13207.1 .Sl_s_at Gene Svmbol ADRB K2 AICDA AN KRD44 APBB NP 2.86 ARHGAP l 5 ATXN7 BCLl 1A BLNK BTK CBFA2T3 2.25 CCL20 CD36 CD3D CD72 CD7 93 CDKN l B CLIC2 CYTH4 DCBLDl DENNDSB DKK3 DNMT3A DOCK8 EG667604 FAM 1 8A FAM6SB FMO6 GFRAI GLCCIl GMIP GPD] GPR174 GRAP HAGHL HCLSl HLA-DMA HLA-DRB 1 HMHAI HPSS IFI3O IL20RA JAK3 IP6K2 KCNAB2 2.70 KLHL6 LCPl LIMD2 Gene Name adrenergic, beta, receptor kinase 2 activation-induced cytidine deaminase ankyrin repeat domain 44 amyloid beta (A4) precursor protein-binding. family B, Rho GTPase activating protein 15 ataxin 7 B-cell CLL/lymphoma 11A (zinc finger protein) B-cell linker Bruton agammaglobulinemia tyrosine kinase core-binding factor, runt domain, alpha subunit 2; chemokine (C-C motif) ligand 20 CD36 molecule (thrombospondin receptor) CD3d molecule, delta (CD3-TCR complex) CD72 molecule CD79b molecule, immunoglobulin-associated beta cyclin-dependent kinase inhibitor 1B (p27, Kip1) chloride intracellular channel 2 cytohesin 4 discoidin, CUB and LCCL domain containing 1 DENN/MADD domain containing SB dickkopf homolog 3 (Xenopus laevis) DNA (cytosine-5-)-methyltransferase 3 alpha dedicator of cytokinesis 8 predicted gene, EG667604 family with sequence similarity 18, member A family with sequence similarity 65, member B flavin containing monooxygenase 6 pseudogene GDNF family receptor alpha 1 glucocorticoid induced transcript 1 GEM interacting protein glycerol-3-phosphate dehydrogenase l (soluble) G protein-coupled receptor 174 GRB2-related adaptor protein hydroxyacylglutathione hydrolase-like hematopoietic cell-specific Lyn substrate 1 major histocompatibility complex, class 11, DM alpha major histocompatibility complex, class 11, DR beta ‘1 histocompatibility (minor) HA-l Hermansky-Pudlak syndrome 5 interferon, gamma-inducible protein 30 interleukin 20 receptor, alpha Janus kinase 3 (a protein tyrosine kinase, leukocyte) inositol hexakisphosphate kinase 2 potassium voltage-gated channel, shaker-related kelch-like 6 (Drosophila) lymphocyte cytosolic protein 1 (L-plastin) LIM domain containing 2 59 Continued Table 2-2. Genes under-express in SCA2 induced bursal cells GgaAffx.5229.l.S1_at Gga.449l.l.Sl_s_at Gga.4354.1.S1_at GgaAffx.6232.l.Sl_s__at Gga.5323.1.Sl_at GgaAffx.]2551.1.Sl__at Gga.8736. 1.S l_at domain containing 1 Gga.8880.1.Sl__at Gga.16267.1.Sl_at GgaAffx.] 1413.1.S1_s_at G gaAffx. l 1420.1.Sl_s_at Gga.17430.l.Sl_at Gga.1329l.l.Sl_at Gga.12076.1.Sl_at Gga.11516.1.Sl__at G gaAffx.9707. 1 .S 1_s_at domain containing 1 G ga.49 16. 1.52_s_at inhibitor) Gga.71.1.Sl_at LRRNI LSPI M6PR MDM l MEF2C MGC33894 MICALI 2.28 MME MR1 NCOA7 NECAP2 NUP210 OIT3 P2RX5 P2RY6 PCMTDI 2. 10 PDCD4 2.40 PDGFB sarcoma viral oncogene homolog) Gga.10030.1.Sl_s_at Gga.7 148.2.A1_at Gga.19667.1.Sl_at exchange factor 1 Gga.8329.1.Sl__at Gga.9379. 1 .S 1_s_at GgaAffx. 12720. 1 .S l_at GgaAffx.] l475.1.Sl_s_at Gga.9386. l .Sl_at Gga.15728. l.Sl_at Gga.352.1.Sl_at Gga.41 l3.l.Sl_at GgaAffx.235 l .28 1_s_at G ga.408. l .S l_at GgaAffx.]2264.1.Sl_s_at Gga.7956.1.Sl_at Gga.6286. l .S l_at Gga.8508. 1.82_s_at GgaAffx.] 1711.1.S1_s_at Gga.6976. 1 .S l_at GgaAffx.4780.1.Sl_at GgaAffx.943. l .S 1_s_at Gga.1148.1.S2_at GgaAffx. 12873. LS l_at GgaAffx. 12646.1.S1_at Gga.11781.1.S1_at Gga.16833.1.Sl_at GgaAffx. 12557. LS 1_s_at GgaAffx. 12475.1.Sl_at Gga.15780.1.S1_at Gga.17034.1.S1_s_at PIK3R5 PLCG2 PREX] 3.33 PRKCH PTPN6 RABGAP l L RASSFZ RBP7 RCSDI RGSZO RHOF RORA RPE65 SASH3 SH3BGRL SIAH2 SLA SNAP91 3.29 SPINKS STAT4 SPNS3 ST6GAL1 STRBP SYK TM6SF] TP53INPI TPST2 TRAFS WDR66 YPEL2 leucine rich repeat neuronal l lymphocyte-specific protein 1 mannose-6-phosphate receptor (cation dependent) Mdml nuclear protein homolog (mouse) myocyte enhancer factor 2C transcript expressed during hematopoiesis 2 microtubule associated monoxygenase, calponin and LIM membrane metallo-endopeptidase major histocompatibility complex, class I-related nuclear receptor coactivator 7 NECAP endocytosis associated 2 nucleoporin 210kDa oncoprotein induced transcript 3 purinergic receptor P2X, ligand-gated ion channel, 5 pyrimidinergic receptor P2Y, G-protein coupled, 6 protein-L-isoaspartate (D-aspartate) O-methyltransferase programmed cell death 4 (neoplastic transformation platelet-derived growth factor beta polypeptide (simian 2.27 ph03phoinositide-3-kinase, regulatory subunit 5 phospholipase C, gamma 2 (phosphatidylinositol-specific) phosphatidylinositol-3,4,5-trisphosphate-dependent Rac protein kinase C, eta protein tyrosine phosphatase, non-receptor type 6 RAB GTPase activating protein 1-like Ras association (RalGDS/AF-6) domain family member 2 retinol binding protein 7, cellular RCSD domain containing 1 regulator of G-protein signaling 20 ras homolog gene family, member F (in filopodia) RAR-related orphan receptor A retinal pigment epithelium-specific protein 65kDa SAM and SH3 domain containing 3 SH3 domain binding glutamic acid-rich protein like seven in absentia homolog 2 (Drosophila) Src-like-adaptor synaptosomal-associated protein, 91kDa homolog (mouse) serine peptidase inhibitor, Kaza] type 5 signal transducer and activator of transcription 4 spinster homolog 3 (Drosophila) ST6 beta-galactosamide alpha-2,6-sialyltranferase l spennatid perinuclear RNA binding protein spleen tyrosine kinase transmembrane 6 superfamily member 1 tumor protein p53 inducible nuclear protein 1 tyrosylprotein sulfotransferase 2 TNF receptor-associated factor 5 WD repeat domain 66 yippee-like 2 (Drosophila) 60 Table 2-3. Genes over-express in SCA2 induced bursa] cells Probe ID GgaAffx.6058.].S1_at Gga.10984.1.S1__at Gga.1232l.1.S1_at Gene Symbol AADAT AIDA ALG] 1 mannosyltransferase homolog (yeast) Gga.2301.l.Sl_at Gga.] 1793.1.S1_at Gga.8363.1.S1_at Gga.2840.1.Sl_at polypeptide 1 Gga.12194.1.S1_at GgaAffx.218l7.l.Sl_s_at Gga.2592.1.Sl_at Gga.4257. l .S l_at Gga.2584. 1 .Sl_s_at Gga.4211.l.Sl_at Gga.184. 1 .S l_a_at Gga.16050.1.Sl_s_at Gga.9493.1.Sl__at GgaAffx. l 1017.5.Sl_s_at Gga.17040.2.Sl_a_at Gga.8858.1.Sl_s_at Gga.2663. l .S l_at GgaAffx. 1557.1.S1_at GgaAffx.4054.1.Sl_at Gga.13785.1.S1_at Gga.3066.1.Sl__at Gga.4724. 1 .S2_at Gga.892.l.Sl_at loop-helix protein GgaAffx.20617. l.Sl_at Gga.14123.1.Sl_at Gga.16230.2.Sl_a_at Gga.7898. 1 .S l__at Gga.15876.].S1_at Gga.13l62.l.Sl_at Gga.2550.1.Sl_at Gga.6218.l.Sl_at Gga.10617.].S1_a_at Gga.3332.1.S1_at Gga.18504.1.S1_at Gga.5300.1.S1_at Gga.3398.2.Sl_a_at Gga.4726.l.Sl_at Gga.9989.l.Sl_at Gga.4856. 1 .S 1_s_at Gga.11468.1.Sl__a_at Gga.1546. l .S2_a_at Gga.5640. l .S l_at Gga.9612.1.Sl_at Gga.5921.1.Sl_at Gga.8500. l .S l_at Gga.10018.l.S1_at APCDDI ATG l 0 AXlN2 B4GALT1 2.59 CBYI CDH13 COL5A] COL6A2 CPNE2 CRTAP CSPGS DNAJ C3 DPYSL3 DYNCZHI FGF2 FKBP7 FKBP9 GPCI HECT D2 HNRPLL HS lBP3 HSP9OB 1 ID] 5.1 l IFI81 ISM 1 KCT D3 LCLATI LRIG3 MANEA MAT l A MFAPS MSRB3 MXRA8 PARVA PDE6D PITX2 PLODl PLXDC2 PMP22 PPIC PRRXI PRS823 RDH l4 REEPS RG32 RIN 3 Gene Name aminoadipate arninotransferase axin interactor, dorsalization associated asparagine-linked glycosylation 11, alpha-1,2- 2.02 adenomatosis polyposis coli down-regulated 1 ATGIO autophagy related 10 homolog (S. cerevisiae) axin 2 UDP-GalzbetaGlcNAc beta 1,4- galactosyltransferase, chibby homolog 1 (Dr030phila) cadherin 13, H-cadherin (heart) collagen, type V, alpha I collagen, type V], alpha 2 c0pine II cartilage associated protein chondroitin sulfate proteoglycan 5 (neuroglycan C) DnaJ (Hsp40) homolog, subfamily C, member 3 dihydropyrimidinase-like 3 dynein, cytoplasmic 2, heavy chain 1 fibroblast growth factor 2 (basic) FK506 binding protein 7 FK506 binding protein 9, 63 kDa glypican l HECT domain containing 2 heterogeneous nuclear ribonucleoprotein L-like HCLSI binding protein 3 heat shock protein 90kDa beta (Grp94), member 1 inhibitor of DNA binding 1, dominant negative helix- intraflagellar transport 81 homolog (Chlamydomonas) isthrnin 1 homolog (zebrafish) potassium channel tetramerisation domain containing 3 lysocardiolipin acyltransferase l leucine-rich repeats and immunoglobulin-like domains 3 mannosidase. endo-alpha methionine adenosyltransferase 1, alpha microfibrillar associated protein 5 methionine sulfoxide reductase B3 matrix-remodelling associated 8 parvin, alpha phosphodiesterase 6D, cGMP-specific, rod, delta paired-like homeodomain 2 procollagen-lysine l, 2-oxoglutarate 5—dioxygenase 1 plexin domain containing 2 peripheral myelin protein 22 peptidylprolyl isomerase C (cyclophilin C) paired related homeobox 1 protease, serine, 23 retinol dehydrogenase l4 (all-trans/9-cis/ l l -cis) receptor accessory protein 5 regulator of G-protein signaling 2, 24kDa Ras and Rab interactor 3 61 Continued Table 2-3. Genes over-express in SCA2 induced bursal cells Gga.4516.2.Sl_at Gga.7890. l .S l_at Gga.13449.l.S1_s_at GgaAffx.8046. l .S 1_s_at Gga.19099.2.Sl_a_at Gga.17868. l.Sl_s_at Gga.16722.1.Sl_at Gga.485. l .Sl_a_at Gga.9244.2.Sl_a_at GgaAffx.4539. l .S l_at Gga.4153. 1.32_a_at GgaAffx. 13062.1.S1_at Gga.19252.1.Sl_s_at Gga.10985.1.Sl_at GgaAffx.6948. l .S l_at G ga.4975. l .S l_a_at GgaAffx. 12952.1.Sl_at Gga.3250. 1 .S l_at Gga.116ll.l.Sl_at SEPT] l SFRP4 SGCE SLC4 1 A2 SPARC SPATA6 SRGAP l TEAD4 TEKT4 TEKTS THYI TMEM129 TMEM184C TMEM45A TMTC2 TPM3 USP28 VSNLI WIF] septin ll secreted frizzled-related protein 4 sarcoglycan, epsilon solute carrier family 41, member 2 secreted protein, acidic. cysteine-rich (osteonectin) sperrnatogenesis associated 6 SLIT-ROBO Rho GTPase activating protein 1 TBA domain family member 4 tektin 4 tektin 5 Thy-l cell surface antigen transmembrane protein 129 transmembrane protein 184C transmembrane protein 45A transmembrane and tetratricopeptide repeat containing 2 tropomyosin 3 ubiquitin Specific peptidase 28 visinin-like 1 WNT inhibitory factor 1 Table 2-4. Real-time quantitative reverse transcription PCR analysis of differential expression for bursal cells incubated with SCA2-expressing or control DFl cells Gene SCA21 Controll Fold Change CD798 6.561082 3.21 $0.39 -10.2 BLNK 6.38 $0.63 4.82 10.36 -2.9 BTK 11.94 10.89 7.86 $0.36 -l6.9 PLCGZ 10.67 10.74 7.62 10.41 -8.2 S YK 10.79 10.76 7.48 $0.39 -9.9 THY-1 4.95 3:080 8.33 10.53 10.4 lCt mean and standard deviation for three replicates. 63 Discussion Although chicken SCA2 gene expression was associated with tumorigenicity and self-renewal of avian erythroid progenitors, the natural biochemical and biological pr0perties of SCA2 protein are unknown. We here describe the development of a polyclonal antibody to chicken SCA2 and its application to the analysis of the biological properties and tissue distribution of SCA2. SCA2 is a 13 kDa small molecular weight, cell surface protein anchored by a GPI moiety, which is similar to homologs in other species. The expression pattern of SCA2 in chicken is relatively wide, and it is abundantly expressed in some tissues. In thymus, strong staining using anti-SCA2 antisera was on the connective cells. In bursa, SCA2 was specifically expressed on the CMEC. SCA2 also shows strongly expression in hepatocytes and endothelial cells in liver. Anti-SCA2 antisera stained weakly in the spleen, lung, heart and gonad. The expression pattern of SCA2 is similar in chicken and mouse in which SCA2 was abundantly expressed in thymic epithelial cells in thymus and in other tissues including the embryonic adrenal gland and liver and adult kidney and liver. Although the function of mouse SCA2 is not well defined, the literature suggests its participation in cell-cell adhesion, signal transduction, and T cell differentiation. The deletion of SCA2 in mouse resulted in abnormal adrenal gland and midgestational lethality due to cardiac abnormalities (Zammit et al., 2002). Chicken SCA2 has been suggested to have an association with tumorigenesis based on the up-regulation of the SCA2 transcript in transformed cell lines, and the role of SCA2 in self-renewal and proliferation of avian erythroid progenitors (Petrenko et al., 1997; Gandrillon et al., 1999; Damiola et al., 2004; Bresson-Mazet et al., 2008). Based on our anti-SCA2 antisera, 64 chicken SCA2 is specifically expressed on medullary reticular epithelial cells and CMEC that are surrounded by MHC class II presenting cells in bursal follicles. Therefore, SCA2 may function in maintaining or mediating the proliferation or differentiation of B lymphocytes located at the cortex-medullar border. To provide insights into the biological function induced by SCA2, we performed a microarray experiment on bursal cells in contact with SCA2. After sorting for differential expression at Z2-fold changes defined by SAM, the expression of 95 genes were reduced in SCA2-induced bursal cells in comparison to control bursal cells, and many of them are related to signaling, proliferation, immunity response and control of transcription of B cells. Although 68 genes were up-regulated in SCA2-induced bursal cells, the majority of these genes are not specific to B cells. Real time PCR assays with a limited number of genes indicate the array results are valid. The development of B lymphocytes from hematopoietic stem cells is regulated by Specific signaling mechanisms and the BCR pathway serves a central role during this progression (Reviewed in Kurosaki, 1998; Kurosaki 1999; Gauld et al., 2002; Jumaa et al., 2005). Of the down-regulated genes profiled and validated in the SCA2-induced bursal cells, CD79B, SYK, BLNK BTK, and PLCGZ are components of the BCR signaling pathway and play critical roles in B cell development. CD79B is a subunit of the BCR complex, and phosphorylation of CD79B by protein tyrosine kinase results in the recruitment of SYK, another tyrosine kinase, that facilitates down-stream signaling cascades (Sanchez et al., 1993; Pao et al., 1998; Kurosaki et al., 1995). BTK is a nonreceptor tyrosine kinase, and BTK-deficient mice show an X-linked immunodeficiency phenotype with immature peripheral B cells (Kemer et al., 1995; 65 Khan et al., 1995; Hendriks et al., 1996). Phosphorylated BLNK provides docking sites for various signaling proteins including BTK and PLCG2, and pre-B cells deficient in BLNK show an increased proliferation rate resulting in a pre-B cell leukemia, suggesting that BLNK is a tumor suppressor for limiting the proliferation of pre-B cells (Pappu et al., 1999; Minegishi et al., 1999; Jumaa et al., 2003). Therefore, the ability of chicken SCA2 in down-regulating critical components for normal B cell deve10pment and proliferation, such as the putative tumor suppressor BLNK, indicate SCA2 may be functionally associated with transformation by delivering proliferation signals. The role of CMEC, where SCA2 is specifically localized, is still not clear although these cells are suggested to provide a “nursing” function for regulating the development of bursa cells. Notch family transmembrane receptor proteins function in mediating cell development and differentiation (reviewed by Robey, 1999), and have been known to be expressed in medullary B cells located close to CMEC, and the Notch ligand Serrate2 is expressed exclusively in the CMEC (Morimura et al., 2001). Although the ligands for SCA2 are not known, the tissue localization of SCA2 is closely related to Serrate2 and may also function similarly as a B cell developmental signal. K-l antigen has also been identified to be expressed in CMEC and reticular epithelial cells in bursa (Kaspers et al., 1993; Olah et al., 2002). K-l antigen is a cell surface antigen consisting of a heterodimer with polypeptide chains of molecular weights of 135 kDa and 61-68 kDa, and it is expressed on chicken thrombocytes, macrophages, and a virus transformed phagocytic cell line, but not on T and B cells. Although both K-l and SCA2 are located on CMEC, no apparent homology is observed between SCA2 and K-l antigen. 66 Mice Ly—6 family members are broadly used as a differential marker for hemotopoeitic stem cells. In this report, we did not thoroughly investigate SCA2 as a differential marker of chicken hemotopoeitic stem cells. We did not obtain the detectable level of cell surface expression of SCA2 on thymocyte cells from 1-4 week old birds. Therefore, either the expression of SCA2 is restricted to certain stages of lymphoid cells, or it is mainly expressed in epithelial cells. In conclusion, chicken SCA2 has been identified as a cell surface antigen anchored by GPI moiety, and the tissue distribution of SCA2 was determined. SCA2 is Specifically expressed on CMEC in bursal follicles, which are surrounded by MHC class II expressing cells. The down-regulation of genes in the BCR signaling pathway induced by contact with SCA2-expressing cells suggests SCA2 may function in mediating B cell proliferation and differentiation via interaction with receptors on the B cell surface. Further analysis of the role of SCA2 in B cell proliferation is warranted and may elucidate its function in B cell deve10pment. Acknowledgments We thank L. Molitor and N. Koller for technical support, and AS. Porter, K.A. Joseph, and RA. Rosebury for their assistance with immunohistochemistry, which was carried out at the Investigative Histophathology Laboratory, Department of Physiology/Division of Human Pathology, Michigan State University. We thank J. Landgraf and A. Thelen for their assistance with the Affymetrix array experiment, which was carried out at Research Technology Support Facility, Michigan State University. We 67 also thank Dr. Heidari for technical assistance with real-time PCR. This work was supported in part by the USDA NRICGP grant #2003-05414. 68 References 1. 10. 11. 12. Antica, M., L. Wu, and R. Scollay. 1997. Stem cell antigen 2 expression in adult and developing mice. Immunol Lett 55:47-51. Bresson-Mazet, C., O. Gandrillon, and S. Gonin-Giraud. 2008. Stem cell antigen 2: a new gene involved in the self-renewal of erythroid progenitors. Cell Prolif 41:726-38. Cheng, A. M., B. Rowley, W. Pao, A. Hayday, J. B. Bolen, and T. Pawson. 1995. Syk tyrosine kinase required for mouse viability and B-cell development. Nature 378:303-6. Classon, B. J., and R. L. Boyd. 1998. Thymic-shared antigen-1 (TSA-l). A lymphostromal cell membrane Ly-6 superfamily molecule with a putative role in cellular adhesion. Dev Immunol 6: 149-56. Classon, B. J., and L. Coverdale. 1994. Mouse stem cell antigen Sea-2 is a member of the Ly-6 family of cell surface proteins. Proc Natl Acad Sci U S A 91:5296-300. Classon, B. J ., and L. Coverdale. 1996. Genomic organization and expression of mouse thymic shared antigen-1 (TSA-l): evidence for a processed pseudogene. Immunogenetics 44:222-6. Damiola, F., C. Keime, S. Gonin-Giraud, S. Dazy, and O. Gandrillon. 2004. Global transcription analysis of immature avian erythrocytic progenitors: from self-renewal to differentiation. Oncogene 23:7628-43. Eshel, R., M. Firon, B. Z. Katz, 0. Sagi-Assif, H. Aviram, and I. P. Witz. 1995. Microenvironmental factors regulate Ly-6 A/E expression on PyV- transformed BALB/c 3T3 cells. Immunol Lett 44:209-12. Fleming, T. J., and T. R. Malek. 1994. Multiple glycosylphosphatidylinositol- anchored Ly—6 molecules and transmembrane Ly-6E mediate inhibition of IL—2 production. J Immunol 153:1955-62. Gandrillon, O., U. Schmidt, H. Beug, and J. Samarut. 1999. TGF-beta cooperates with TGF-alpha to induce the self-renewal of normal erythrocytic progenitors: evidence for an autocrine mechanism. Embo J 18:2764—81. Gauld, S. B., J. M. Dal Porto, and J. C. Cambier. 2002. B cell antigen receptor signaling: roles in cell development and disease. Science 296: 1641-2. Godfrey, D. 1., M. Masciantonio, C. L. Tucek, M. A. Malin, R. L. Boyd, and P. Hugo. 1992. Thymic shared antigen-1. A novel thymocyte marker 69 l3. 14. 15. 16. 17. 18. 19. 20. 21. 23. 24. discriminating immature from mature thymocyte subsets. J Immunol 148:2006-11. He, J., and L. J. Chang. 2004. Functional characterization of hepatoma-specific stem cell antigen-2. Mo] Carcinog 40:90-103. Hendriks, R. W., M. F. de Bruijn, A. Maas, G. M. Dingjan, A. Karis, and F. Grosveld. 1996. Inactivation of Btk by insertion of lacZ reveals defects in B cell development only past the pre-B cell stage. Embo J 15:4862—72. Hunt, H. D., B. Lupiani, M. M. Miller, I. Gimeno, L. F. Lee, and M. S. Parcells. 2001. Marek's disease virus down-regulates surface expression of MHC (B Complex) Class I (BF) glycoproteins during active but not latent infection of chicken cells. Virology 282:198-205. Irizarry, R. A., B. Hobbs, F. Collin, Y. D. Beazer-Barclay, K. J. Antonellis, U. Scherf, and T. P. Speed. 2003. Exploration, normalization, and summaries of high density oligonucleotide array probe level data. Biostatistics 4:249—64. Jumaa, H., L. Bossaller, K. Portugal, B. Storch, M. Lotz, A. Flemming, M. Schrappe, V. Postila, P. Riikonen, J. Pelkonen, C. M. Niemeyer, and M. Reth. 2003. Deficiency of the adaptor SLP-65 in pre-B-cell acute lymphoblastic leukaemia. Nature 423:452-6. Jumaa, H., R. W. Hendriks, and M. Reth. 2005. B cell signaling and tumorigenesis. Annu Rev Immunol 23:415-45. Kaspers, B., H. S. Lillehoj, and E. P. Lillehoj. 1993. Chicken macrophages and thrombocytes share a common cell surface antigen defined by a monoclonal. antibody. Vet Immunol Immunopathol 36:333-46. Kerner, J. D., M. W. Appleby, R. N. Mohr, S. Chien, D. J. Rawlings, C. R. Maliszewski, O. N. Witte, and R. M. Perlmutter. 1995. Impaired expansion of mouse B cell progenitors lacking Btk. Immunity 3:301-12. Khan, W. N., F. W. Alt, R. M. Gerstein, B. A. Malynn, I. Larsson, G. Rathbun, L. Davidson, S. Muller, A. B. Kantor, L. A. Herzenberg, and et al. 1995. Defective B cell development and function in Btk-deficient mice. Immunity 3:283-99. Kurosaki, T. 1998. Molecular dissection of B cell antigen receptor signaling (review). Int J Mol Med 1:515-27. Kurosaki, T. 1999. Genetic analysis of B cell antigen receptor Signaling. Annu Rev Immunol 17:555-92. Kurosaki, T., S. A. Johnson, L. Pao, K. Sada, H. Yamamura, and J. C. 70 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. Cambier. 1995. Role of the Syk autophosphorylation site and SH2 domains in B cell antigen receptor signaling. J Exp Med 182:1815-23. Liu, H. C., M. Niikura, J. E. Fulton, and H. H. Cheng. 2003. Identification of chicken lymphocyte antigen 6 complex, locus E (LY6E, alias SCA2) as a putative Marek's disease resistance gene via a virus-host protein interaction screen. Cytogenet Genome Res 102:304-8. Livak, K. J., and T. D. Schmittgen. 2001. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25:402-8. MacNeil, 1., J. Kennedy, D. I. Godfrey, N. A. Jenkins, M. Masciantonio, C. Mineo, D. J. Gilbert, N. G. Copeland, R. L. Boyd, and A. Zlotnik. 1993. Isolation of a cDNA encoding thymic shared antigen-1. A new member of the Ly6 family with a possible role in T cell development. J Immunol 151:6913-23. Minegishi, Y., J. Rohrer, E. Coustan-Smith, H. M. Lederman, R. Pappu, D. Campana, A. C. Chan, and M. E. Conley. 1999. An essential role for BLN K in human B cell development. Science 286:1954-7. Morgan, R. W., L. Sofer, A. S. Anderson, E. L. Bernberg, J. Cui, and J. Burnside. 2001. Induction of host gene expression following infection of chicken embryo fibroblasts with oncogenic Marek's disease virus. J Virol 75:533-9. Morimura, T., S. Miyatani, D. Kitamura, and R. Goitsuka. 2001. Notch signaling suppresses IgH gene expression in chicken B cells: implication in spatially restricted expression of Serrate2/Notch] in the bursa of Fabricius. J Immunol 166:3277-83. Noda, S., A. Kosugi, S. Saitoh, S. Narumiya, and T. Hamaoka. 1996. Protection from anti-TCR/CD3-induced apoptosis in immature thymocytes by a signal through thymic shared antigen-I/stem cell antigen-2. J Exp Med 183:2355- 60. Olah, 1., K. H. Gumati, N. Nagy, A. Magyar, B. Kaspers, and H. Lillehoj. 2002. Diverse expression of the K-1 antigen by cortico-medullary and reticular epithelial cells of the bursa of Fabricius in chicken and guinea fowl. Dev Comp Immun0126:481-8. Pao, L. 1., S. J. Famiglietti, and J. C. Cambier. 1998. Asymmetrical phosphorylation and function of immunoreceptor tyrosine-based activation motif tyrosines in B cell antigen receptor signal transduction. J Immunol 160:3305-14. Pappu, R., A. M. Cheng, B. Li, Q. Gong, C. Chiu, N. Griffin, M. White, B. P. Sleckman, and A. C. Chan. 1999. Requirement for B cell linker protein (BLNK) 71 35. 36. 37. 38. 39. 40. 41. 42. 43. in B cell development. Science 286:1949-54. Petrenko, O., I. Ischenko, and P. J. Enrietto. 1997. Characterization of changes in gene expression associated with malignant transformation by the NF—kappaB family member, v-Rel. Oncogene 15:1671-80. Randle, E. S., G. A. Waanders, M. Masciantonio, D. I. Godfrey, and R. L. Boyd. 1993. A lymphostromal molecule, thymic shared Ag-l, regulates early thymocyte development in fetal thymus organ culture. J Immunol 151:6027-35. Ratcliffe, M. J. 2006. Antibodies, immunoglobulin genes and the bursa of Fabricius in chicken B cell development. Dev Comp Immunol 30:101-18. Robey, E. 1999. Regulation of T cell fate by Notch. Annu Rev Immunol 17 :283- 95. Saitoh, S., A. Kosugi, S. Noda, N. Yamamoto, M. Ogata, Y. Minami, K. Miyake, and T. Hamaoka. 1995. Modulation of TCR-mediated signaling pathway by thymic shared antigen-l (TSA-l)/stem cell antigen-2 (Sea-2). J Immunol 155:5574-81. Sanchez, M., Z. Misulovin, A. L. Burkhardt, S. Mahajan, T. Costa, R. Franke, J. B. Bolen, and M. Nussenzweig. 1993. Signal transduction by immunoglobulin is mediated through Ig alpha and Ig beta. J Exp Med 178:1049— 55. Turner, M., P. J. Mee, P. S. Costello, 0. Williams, A. A. Price, L. P. Duddy, M. T. Furlong, R. L. Geahlen, and V. L. Tybulewicz. 1995. Perinatal lethality and blocked B-cell development in mice lacking the tyrosine kinase Syk. Nature 378:298-302. Tusher, V. G., R. Tibshirani, and G. Chu. 2001. Significance analysis of microarrays applied to the ionizing radiation response. Proc Natl Acad Sci U S A 98:5116-21. Zammit, D. J., S. P. Berzins, J. W. Gill, E. S. Randle-Barrett, L. Barnett, F. Koentgen, G. W. Lambert, R. P. Harvey, R. L. Boyd, and B. J. Classon. 2002. Essential role for the lymphostromal plasma membrane Ly-6 superfamily molecule thymic shared antigen 1 in development of the embryonic adrenal gland. Mol Cell Biol 22:946-52. Chapter 3 Chicken stem cell antigen 2 modulates Marek’s disease virus grthh properties in vitro via U810 73 Abstract: Marek’s disease virus (MDV) is a highly cell-associated oncogenic alphaherpesvirus and the causative agent for Marek’s disease (MD), a lymphoproliferative disease in chickens. Understanding the molecular mechanisms for pathogenesis is of both academic and commercial interest. Chicken stem cell antigen 2 (SCA2) is a putative MD resistance gene based on (1) genetic linkage of SCA2 to MD resistance and related traits, (2) differential transcription of SCA2 between MD resistant and susceptible chickens following MDV challenge, and (3) a confirmed and direct protein interaction of SCA2 with MDV U810 in vitro. The functional relevance of the interaction between SCA2 and U810 proteins has not been elucidated. Two recombinant MDVS were developed: er5-B40-U810EGFP (referred to as USlOEGFP) expresses a U810-enhanced green fluorescent protein (EGFP) fusion protein, and er5-B40- AUSIOEGFP (referred to as AUSIOEGFP) replaces US 10 with EGFP. Neither modification impacted viral replication rate in vitro. These two mutant viruses were employed to examine the sub-cellular localization of U810 and SCA2, and the “potential influence of their interaction on viral growth properties. As judged by EGFP expression, U810 is found exclusively in the cytoplasm of DF1 cells, an immortalized chicken embryo fibroblast cell line. SCA2, which is normally found on the cell surface, co- localizes with U810 in DFI cells stably expressing a SCA2-Flag fusion protein. The expression of SCA2 in DFl cells results in MDV plaques that are significantly smaller at 6 days post-infection but only when U810 is present. The ability of SCA2 to impact MDV growth properties is in agreement with the percentage of USlOEGFP or AUSIOEGFP-infected DFl cells as measured using flow cytometry. Over-expression of SCA2 impairs MDV Spread in vitro but this influence is dependent on the presence of 74 U810. This is the first report showing a direct influence of SCA2 on MDV, and it provides insights on how this chicken protein may impact MDV replication and MD pathogenesis, and helps to confirm its role as an MD resistance gene. 75 Introduction Marek’s disease (MD) is a lymphoproliferative disease of chickens characterized by T cell lymphomas and nerve enlargements. The causative agent is the highly oncogenic alphaherpesvirus Marek’s disease virus (MDV). The 180 Kb MDV genome has a typical alphaherpesvirus structure with a linear double-stranded DNA genome consisting of unique long (UL) and unique short (US) regions, each flanked by inverted repeats (Tulman et al., 2000; Silva et al., 2000). The viral genome is enclosed by viral nucleocapsids surrounded by a tegument protein layer and an outer viral glycoprotein envelope (Kato and Hirai, 1985). Viral products and host elements participate within different structures of the cell for virion packaging and egress. We have reported nine specific chicken-MDV protein-protein interactions via a bacterial two-hybrid screen following by in vitro binding assays for confirmation (Niikura et al., 2004). MD is considered the most serious persistent infectious disease of concern to the poultry industry worldwide, although it has been controlled through vaccination since 19703. Even with vaccines, MD still has a significant economic impact to the poultry industry as there are sporadic outbreaks, which require the continuous vaccination and monitoring of birds. Consequently, there is a need for alternative methods, such as enhanced resistance to MD, to augment vaccinal control of MD. Chicken stem cell antigen 2 [SCA2, also known as lymphocyte antigen 6 complex, locus E (LY6E) and thymic shared antigen 1 (TSA1)], which directly interacts with MDV U810, was further investigated and shown to be a putative MD resistance gene based on its genetic association with MD incidence and related traits, and differential transcription following MDV challenge between MD resistant and susceptible chicken lines (Liu et al., 2006). To 76 address the putative function(s) of SCA2 in MDV biology through association with viral U810 product, validating their co-localization in chicken cells and determining the relevance of this interaction is important. MDV U810 is located in the US region of the viral genome and encodes a 213 amino acid protein. U810 homologues are encoded by other herpesviruses including herpes simplex virus type 1 (HSV-1), varicella—zoster virus (VZV), and equine herpesvirus type 1 (EHV-l) (McGeoch et al., 1985; Brown and Harland, 1987; Davison and Scott, 1986; Telford et al., 1992; Sakaguchi et al., 1992). Based on properties of HSV-1, MDV U810 in virions is predicted to be a capsid/tegument phosphoprotein (Yamada et al., 1997). MDV U810 in lymphoblastoid cell lines derived from MD tumors show U810 is localized in the cytoplasm near patches of glycoprotein B (gB) on the cell surface, suggesting U810 is associated with virion assembly (Parcells et al., 1999). The deletion of the MDV U8 region containing U810 and adjacent MDV open reading frames (ORFs) indicates that U810 is not essential for MDV growth in vitro, however, this deletion decreases early cytolytic infection, mortality, and tumor incidence in vivo (Parcells et a1, 1994; Parcells et a1, 1995). Chicken SCA2 transcription is induced by MDV infection in cultured chicken embryo fibroblast (CEF) cells (Morgan et al., 2001). However, the function of chicken SCA2, as in other Species, is not well defined. Chicken SCA2 is associated with v-Rel— transformed avian bone marrow cells and plays a role in self-renewal of avian erythroid progenitors (Petrenko et al., 1997; Bresson-Mazet et al., 2008). The majority of work characterizing SCA2 has been done in mice. Mouse SCA2 is a member of the lymphostromal cell membrane Ly6 superfamily, and it localizes on the cell surface via a 77 glycosyl-phosphatidylinositol (GPI)-anchor (Classon et al., 1994). Mouse SCA2 functions in regulating intracellular signaling events via T cell receptor and cell-cell adhesion as a receptor on the lymphocyte membrane (Randle et al., 1993; Fleming and Malek, 1994; Saitoh et al., 1995; Noda et al., 1996). Here we report the development of recombinant MDVS that either lack or have U810 fused with EGFP. Utilizing chicken cell lines stably expressing SCA2, we find that SCA can co—localize with U810 in the cytoplasm. Furthermore, SCA2 alters the size and percentage of MDV infected cells but this effect is dependent on the MDV genome containing US10. 78 Materials and methods Cells MDV was propagated on secondary CEF from 10—day old line 0 chicken embryos and cultivated on a mixture of Liebovitz’s L-15 and McCoy 5A (LM) media (1:1) supplemented with 4% fetal bovine serum (FBS), penicillin, and streptomycin. DF-l cells were cultured in LM media with 4% FBS, penicillin and streptomycin. The Vector—DF] cell line is the DF-l stably transfected with the pSELECT-zeo-mcs vector (Invivogen, CA, USA). The SCA2F1ag-DP] cell line is identical to Vector-DFl except the vector was engineered to constitutively express a SCA2F1ag recombinant protein with the Flag tag inserted immediately downstream of the signal peptide of SCA2. In brief, the entire coding portion of chicken SCA2 (bases 55 to 435 based on GenBank accession no. NM_204775) was amplified from a chicken spleen cDNA library by PCR using primers (SCA28allF: 5’ atcg gt£ga_c_ atg aag gcg ttt ctg ttc gc 3’; SCA2NcoIR: 5’ atcg ga_tgg tca ctc acg agc cct gag g 3’; restriction sites are underlined). The amplicon was inserted between the Sall and Ncol sites of the pSELECT-zeo-mcs vector to generate SCA2- pSELECT-zeo-mcs, and DNA sequenced to confirm that the construct was correct. A Flag tag was inserted immediately downstream of the signal peptide of SCA2 by PCR using SCA2-pSELECT—zeo-mcs as the template with the primers (SCA2SalIF; Sca2FlagNsiIR: 5’ gtt gga gg a_tg_c_at ccg agc acg aaa agc aga tgag ctr atc gtc gtc atc ctr gta arc cgt gtg ggc tct ctc cac g 3’; restriction sites are underlined, Flag tag sequences are in italic). The amplicon was inserted between the Sall and N311 sites of SCA2-pSELECT- zeo-mcs to produce SCA2F1ag-pSELECT-zeo-mcs. The SCA2F1ag-p8ELECT—zeo—mcs and pSELECT-zeo-mcs vectors were each transfected into the DP] cell line using the 79 Basic Nucleofector Kit (Amaxa Inc., Gaithersburg, MD, USA). The transfected cells were incubated in medium containing LM media supplemented with 5% FBS, penicillin, streptomycin, and zeocin (300 ng/ml) until single colonies appeared. The expression of SCA2 was tested by western blots using anti-Flag mAb and anti-SCA2 antisera. The single colony with the best SCA2 expression was maintained as the SCA2F1ag-DP] cell line. Recombinant BAC clones To insert EGFP into MDV U810 gene, two step Red-mediated recombination techniques were applied (Tischer et al., 2006). All PCR amplifications used in BAC recombineering were conducted with Expand High Fidelity PCR system (Roche, Indianapolis, IN, USA). Briefly, the kanamycin gene (aphAI) and I-SceI cassette from pEP-kanS (kindly provide from Niklaus Osterrieder, Cornell University) were amplified with pEPEGFP.F (5’-TCC TGTACA AGT AAA GCG GCC GCG ACT CTA GAT CAT AAT CAG CCA T AC CAC ATT TAG GGA TAA CAG GGT AAT CGA TT-3’; italicized sequence for BerI restriction sites, underlined sequence for Red recombination, bold sequence to amplify I-Scel cassette from pEP-kanS) and pEPEGFP.R (5’-TCC TGTACA CTA GCC AGT GTT ACA ACC AAT TAA CC-3’; italicized sequence for BerI restriction Sites, bold sequence to amplify I-Scel cassette from pEP-kanS). The primer sets were synthesized by Integrated DNA Technologies (IDT, Coralville, IA, USA). The PCR amplicon and pEGFP-Nl (Clontech, E Palo Alto, CA, USA) were digested with BerI and the resulting fragment (~1053 bp) cloned into pEGFP-Nl to create the pEP-EGFP—Nl construct. The EGFP, aphA] and I-SceI cassette 8O was amplified from the pEP-EGFP—Nl construct with the primers, U810-EGFP-fus.F (5’- TAT CTG ACA AAT CTT CGG GAA TCG CCA ACA GGA GAC GGG GAA TCC TAC TIA ATG GTG AGC AAG GGC GAG GAG CTG T-3’; underlined sequences for homologous recombination, bold sequences to amplify EGFP, aphAI and I-Scel cassette) and U810-EGFP.R (5’-CGT TAG TAG CAG TIT TTC CTA AAA TCC TAT TAA TAA TTG TGC GAT TAG TTA AAA GCA AGT AAA ACC TCT ACA AAT-3’; underlined sequences for homologous recombination, bold sequences to amplify EGFP, aphAI and I-Scel cassette). The PCR amplicon with U810-EGFP-fus.F and U810- EGFP.R was used for Red recombination in SW105 E. coli as earlier described (Warming et a] 2005, Tischer et a], 2006) to insert between the 212th amino acid sequence and the downstream stop codon of U810 (155616-155617 nucleotide in Gallid herpes virus 2 genome, GenBank accession number: NC_002229) of er5-B4O BAC (Niikura et a1, 2006). To replace the MDV U810 gene with EGFP, the EGFP, aphAI and I-Scel cassette was amplified from pEP-EGFP-Nl construct with USlO-EGFP-rep.F (5’ TIT GAA TAC TGG AGA CGA GCG CCG TGT AAG ATI‘ AAA ACA TAT TGG AGA GGT ATG GTG AGC AAG GGC GAG GAG CTG T-3’; underlined sequences for homologous recombination, bold sequences to amplify EGFP, aphAI and I-SceI cassette) and U810- EGFP.R. The PCR amplicon with U81-EGFP-rep.F and U810-EGFP.R was used for Red recombination to replace the U810 from the start to stop codons (154978-155617 nucleotide in Gallid herpes virus 2 genome, GenBank accession number: NC_002229) of er5-B40 BAC (Tulman et al., 2000; Niikura et a1, 2006). After Red recombination using the above PCR fragment, kanamycin-resistant BAC clones were selected, then I- 81 SceI induced from pBAD-I-Scel plasmid by treatment with arabinose—containing media. The digestion of I—SceI recognition site in MDV-BAC and subsequent second Red recombination at duplicated sequences were conducted in SW105 cells. After the second Red recombination, the positive MDV-BAC clone was kanamycin sensitive, thus, screened via PCR amplification using U810fl.F (5’-GAGAGGAGA ACGTGT'ITGAATAC-3’) and EGFP.R (5’-CTGCTTCATGTGGTCG GGGTAG-3’) primers. The resulting MDV-BAC clone containing the U810-EGFP fusion protein was named USlOEGFP and the MDV-BAC clone with the replacement of U810 by EGFP was named AUSIOEGFP. Both BAC clones were purified, digested with BamHI, and their digestion patterns analyzed by gel electrophoresis. Generation of viruses To generate MDVS from all three BAC clones, purified plasmids (er5-B40 BAC, USlOEGFP and AUSIOEGFP) were individually transfected into CEF. The nucleofector II (Amaxa Biosystems, Gaithersburg, MD, USA) was used to transfect MDV-BAC into CEF. All CEF cultures were incubated at 37 C°, 5% CO2 incubator in LM media supplemented with 4% FBS, penicillin, and streptomycin. The PBS was reduced from 4 % to 1 % after virus inoculation until MDV plaques were observed. One step growth curve assay and plaque area determination Plaque growth assays were performed as previously reported (Mao et al., 2008). Briefly, 100 PFU of various viruses were inoculated onto CEF cells seeded on 35-mm dishes. On 1, 2, 3, 4, and 5 days after inoculation, the infected cells were trypsinized, serial dilutions inoculated onto fresh cells, and plaques of the different dilutions counted 7 days later in triplicate to determine growth characteristics of er5-B40-derived, US lOEGFP and AUSIOEGFP viruses. 500 PFU USlOEGFP or AUSIOEGFP were inoculated onto 5x105 Vector-DFI and SCA2F1ag-DF1 cells seeded on 35-mm dishes, and the cell cultures incubated with LM supplemented with 1% FBS, penicillin, and streptomycin. Virus plaques on Vector- DF] and SCA2F1ag-DF1 cells 7 days post-infection in three replicates were examined under a LEICA DM [RB/E inverted fluorescence microscope (Leica Wetzlar, Germany). 7-10 plaques were counted on each dish and 21-30 plaques for each virus were counted in total. Images were captured by a DEL-750CE digital system (Optronics, Goleta, CA, USA), and plaque sizes determined using the Image] software (rsb.info.nih.gov/ij/index.html). Statistical significance of plaque areas were performed using the Student’s T—test. Immunofluorescence and confocal image analysis SCA2F1ag-DP] and fresh CEF cells grown on glasscovers were inoculated with 1000 PFU of recombinant viruses, and incubated in LM medium with 4% FBS, penicillin, and streptomycin at 37°C. The CEF with recombinant viruses were fixed with 100% acetone, washed 3 times with PBS , and stained with Hoechest 33342 (Invitrogen, Carlsbad, CA, USA). SCA2F1ag-DP] cells with recombinant virus were incubated with anti-Flag M2 mAb (1:100) for 20 min at 4°C, then Rhodamine-conjugated anti-mouse goat IgG (1:400; Santa Cruz Biotechnology Inc, Santa Cruz, CA, USA) as second Abs 83 for 20 min at 4°C, and stained with Hoechest 33342 after 3 washes with PBS. The stained cells were analyzed with Olympus fluoview 1000 confocal microsc0pe (Olympus, Tokyo, Japan) and the images analyzed using FVlO-ASW software (Olympus, Tokyo, Japan). Western blot analysis CEF with recombinant viruses were lysed in Cellytic M Mammalian Cell lysis reagent (Sigma-Aldrich Inc., Milwaukee, WI, USA), supplemented with 1% P8340 protease inhibitor cocktails (Sigma-Aldrich Inc., Milwaukee, WI, USA) and followed with sonication and centrifugation. Cell lysates were mixed with loading and reducing buffer, heated at 70°C for 10 min and loaded on 10% SDS-PAGE gel (Invitrogen, Carlsbad, CA, USA). The samples were blotted on Immobilon-PSQ PVDF transfer membrane (Millipore, Billerica, MA, USA). After blocking in PBS containing 5% non- fat milk for 1 hour at room temperature, membranes were probed with anti-GFP mAb (Sigma-Aldrich Inc., Milwaukee, WI, USA) at 4°C overnight. Membranes were washed three times in PBS before incubation with a goat anti-mouse IgG conjugated with horseradish peroxidase (Zymed, South San Francisco, CA, USA) at room temperature for 1 hour. Proteins were detected using ECL Western blotting substrate (Pierce, Rockford, IL, USA) after washing for three times. Flow cytometry analysis Three replicates of SCA2F1ag-DE] and Vector-DP] cells infected with USlOEGFP or AUSlOEGFP viruses were trypsinized and harvested (1000xg, 5 min). After 7 days post-infection, 106 cells were washed with cold flow cytometry buffer 84 (phosphate buffered saline containing 1% FBS and 0.1% sodium azide). The cells were analyzed using a Becton-Dickinson FACScaliber flow cytometer and CellQuest Pro software (BD Bioscience, San Jose, CA, USA). 85 Results Construction of USlOEGFP and AUSlOEGFP MDV-BAC clones Insertion of EGFP to generate either a fusion protein with U810 on the C- tenninal end or to delete U810 were performed using the er5-B40-BAC clone, which generates virulent MDV, and BAC modification techniques (Fig. 3-1). To produce a recombinant MDV-BAC that expresses a U810-EGFP fusion protein, the EGFP- containing amplicon was inserted between the 212th amino acid and the stop codon of U810 resulting in er5-B40-U810EGFP (referred to as USlOEGFP). Likewise, the MDV-BAC clone lacking USIO was produced by replacing U810 with EGFP from the start to Stop codons to generate er5-B40-AU810EGFP (referred to as AUSIOEGFP). Both USlOEGFP and AUSlOEGFP BAC clones retain the natural USIO promoters for downstream expression of U810-EGFP and AUSlO-EGFP recombinant genes. Nucleotide analysis of the two BAC clones confirmed the genetic modifications. Analysis of the BAC clones compared to the parental er5-B4O also indicates that no unexpected modifications occurred in the generation of the recombinant MDV-BAC clones (Fig. 3-2); unfortunately, the U810-containing BamHI fragment is too large to detect a change in size after the EGFP insertion into er5-B40 (21.6, 21.5, and 20.9 kb for er5-B40, USlOEGFP, and AUSIOEGFP, respectively). 86 a) TRL - IRL 1R3 US TRS [:1 Llfi—ZZI J U510 EGFP b) TRL '_. IRL IRS US TRS 1:1 4 l 1--——-E::I [ :1 T J 22> r:l> fi- T‘) in the MDV UL41 gene that resulted in a missense mutation (AA801683:p.Arg377Cys), bacterial artificial chromosome (BAC)-derived MDVs that differed only in the UL41 SN P were evaluated using a head-to-head competition assay in vitro. Monitoring the frequency of each SNP by pyrosequencing during virus passage determined the ratio of each viral genome in a single monolayer, which is a very sensitive method to monitor viral fitness. MDV with the UL41*CyS allele showed enhanced fitness in vitro. To evaluate the mechanism of altered viral fitness caused by this SNP, the virion-associated host shutoff (vhs) activity of both UL41 alleles was determined. The UL41*Cy3 allele had no vhs activity, which suggests that enhanced fitness in vitro for MDV with inactive vhs was due to reduced degradation of viral transcripts. The in vitro competition assay should be applicable to other MDV genes and mutations. Introduction Marek’s disease virus (MDV; family Herpesviridae, genus Mardivirus, species Gallid herpesvirus 2) is an infectious pathogen and the cause of Marek’s disease, which is characterized by T cell lymphomas, neurological disorders, and death in susceptible chickens (Osterrieder et al. 2006). Periodic yet unpredictable disease outbreaks that vaccines cannot adequately protect make Marek’s disease a serious problem for the poultry industry worldwide. As the molecular basis of the increasing virulence is unknown, there is a need to understand the role of naturally-occurring mutations in MDV genes on virulence (Witter, 1996). In this report, a head-to-head competition assay is described that contributes to the understanding of the molecular basis for viral fitness variance, and should enable more efficacious vaccines and other methods that provide improved disease control. MDV has an alpha—herpesvirus structure that is similar to herpes simplex virus type-1 (HSV-1) with ~100 genes contained within its 175 kb DNA genome (Silva et al., 2001). Recently, MDV genomes have been cloned into bacterial artificial chromosome (BAC) vectors (Schumacher et al., 2000; Petherbridge et al., 2004; Niikura et al., 2006). Besides providing a platform for infectious molecular clones, MDV-BACS can be readily manipulated even at the nucleotide level, which enables the direct association of specific sequences to phenotypes. Ideally, recombinant MDVs would be engineered in sites where mutations may account for the observed variation in virulence (e.g., Spatz and Silva, 2007; Spatz et al., 2007). During the cloning and full genome sequencing of virulent MDV Mdll strain (.Niikura et al., 2006), a nonsynonymous single nucleotide polymorphism (SNP; 110 AY510475zg.108,206C>T) was found in the UL41 gene that resulted in Arg to Cys substitution in residue 377 (AA801683zp.Arg377Cys). UL41 encodes the virion- associated host shutoff (vhs) protein, which degrades both host and viral mRNA during the viral cytolytic infection and assists the virus in evading the host immune response (Read and Frenkel, 1983; Kwong and Frenkel, 1988; Read et al., 1993; Suzutani et al., 2000; Murphy et al., 2003; Barzilai et al., 2006). The UL41 point mutation identified in the MDV clone is interesting because it is located in a critical amino acid that is conserved among alpha-herpesvirus genomes (Everly and Read, 1999). Thus, questions arise as to whether this point mutation would affect viral fitness or activity of the MDV UL41 gene product, and is there any relationship between the MDV UL41 gene and virulence? The straightforward and preferred way of quantifying viral fitness is by a head-to- head competition experiment using Stable genetic markers. This is because the two different viruses must compete in the same environment, and assuming the viruses do not recombine, the genetic markers provide proxies for each genome. Competition experiments have applied in other viruses including vesicular stomatitis virus (Novella, 2003), human immunodeficiency virus type-l (Holland et al., 1991; Yuste et al., 2005), and tobacco etch virus (Carrasco et al., 2007). This method might be particularly useful for MDV as the traditional plaque assay counts infectious foci generated from a single infected cell rather than a single virion due to the strict cell—associated nature of this virus. Furthermore, many studies examining defined recombinant MDVs have shown that growth rate measurements are not very sensitive and, more importantly, are not good predictors of in vivo effects (Reddy et al., 2002; Silva et al., 2004; Prigge et al., 2004; ll] Kamil et al. 2005; Schumacher et al., 2005), thus, a more sensitive in vitro method is desired. For MDV, pyrosequencing was used to monitor the UL41 SNP. Pyrosequencing is a DNA sequencing technology based on a 4-enzyme real-time monitoring of DNA synthesis, and accurately measures the relative abundance of each SNP nucleotide (Ronaghi et al., 1998; Ronaghi, 2001). This technology has been applied to determine the influence of specific HIV SNPs on resistance to protease inhibitors among others (O’Meara et al., 2001; Lahser et al., 2003). In this study, viral fitness was quantified by evaluating the relative SNP frequency for MDV populations that differed only in UL41 alleles following repeated passages in cell culture. Also, the UL41 vhs activity for each allele was measured in an attempt to help explain the affect on viral fitness. 112 Materials and methods Culture method for MDV and DNA purification MDV was propagated on secondary CEF from line 0 chicken embryos cultivated on a mixture of Liebovitz’s L-15 and McCoy 5A media (1:1) supplemented with 4% FBS, penicillin, and streptomycin. DNA of MDV-infected cells was extracted in Tris buffer (0.1 M, pH 8.0) containing 1% SDS, 1 mM EDTA, 0.1 M NaCl and 100 rig/ml proteinase K, followed by phenol extraction and ethanol precipitation. Mini-l mutagenesis to generate pMDV-UL41*Arg and pMDV-UIAI"Cys-rev As shown in Fig. 4-1, recombinant BAC clones containing the entire MDV strain Mdll genome were generated that differ only in the genomic sequence that encodes UL41 amino acid residue 377 (GenBank no. AA801683) using the mini-it mutagenesis as described by Court et a1. (2003) and Costantino and Court (2003). Briefly, the oligonucleotides UL41mut (5’ GCTGTCTCATCACGTCCCTCTTGCATI‘ATAGGGAACCGTITCAAAATACTAAC CCTGGAAGTTCTGCGCT 3’) and UL41mutRev (5' GCTGTCTCATCACGTCCCTCTTGCATTATAGGGAAGCACTTCAAAATACTAAC CCTGGAAGTTCTGCGCTI‘ 3') were designed to introduce specific point mutations and synthesized (Operon, Huntsville, AL, USA). UL41mut was mixed with competent cells containing the Mdl 1ngoxP5-12 MDV-BAC clone (Niikura et al., 2006), which has the UL41*Cys allele (Cys at amino acid residue 377) and referred to as pMDV-UL41*Cys, and screened by PCR using primers 97556F (5’ CGATAGTGCTAGCTICATCTAC 3’) and 97628AtGR (5’ GAACTTCCAGGGTTAGTA'I'I'I'IGAAACGG 3’) to generate ll3 pMDV-UL41*Arg. Similarly, UL41mutRev was mixed with the competent cells containing pMDV-UL41*Arg, subjected to electroporation, and screened by PCR using primers 97556F and 97628UL41rev (5’ GAACTTCCAGGGTI’AGTATITTGAAGTGC 3)’ to generate pMDV-UL41*Cys-rev. pMDV-UL41*Cys and pMDV-UL41*Cys-rev encode the same UL41 amino acid sequence but differ for a silent SNP (AY510475:g.108,207A>G), which allows molecular tracking of both viral genomes in mixed populations. The new BAC clones were purified using a Large-Construct Kit (Qiagen, Valencia, CA, USA) and a ~500 bp region that encompasses the UL41 SNP was confirmed by DNA sequencing. To generate MDV, BAC DNA (50-100 ng) was added to a CEF cell suspension and electroporated using a cell electroporation system at a setting of 330 uF and 400 V (Life Technologies, Rockville, MD, USA). 114 Parental MDV—U L41 *Cys MDV-U L41 *Arg MDV-U L41 *Cys-rev Figure 4-1. Partial UL41 DNA and amino acid sequences for the parental Mdll strain, the original MDV-BAC clone (MDV-UL41*Cys), and the two recombinant viruses (MDV-UL41*Arg and MDV-UL41*Cys-rev). The position of the given sequences for the first and last bases are 108.209 and 108,204 based on the Mdll strain sequence (GenBank no. AY510475). The asterisk (*) indicates the SNP between the parental Mdll and MDV-UL41*Cys that resulted in an amino acid difference at amino acid residue 377 (GenBank no. AA801683). In addition to altering this position by the mini-7t mutagenesis procedure, the third base of the same codon (position 108,204) was mutated to increase the efficiency of recovering the correct recombinants. Plus, the third base of the Lys codon was changed (AY510475:g.108207A>G) in MDV-UL41*Cys-rev to allow tracking and allele quantification by pyrosequencing between MDV-UL41*Cys and MDV-UL41*Cys-rev. 115 AAA Lys AAA Lys AAA Ly s l AAG Lys CGC Ar g * TGC Cys l l CGG Arg 1 I TGC Cys Growth rates of individual viruses Growth characteristics of each individual cloned virus was determined. Briefly, 100 PFU were inoculated onto CEF cells seeded on 35-mm dishes. On 1, 2, 3, 4, and 5 days after inoculation, the infected cells were trypsinized, serial dilutions inoculated onto fresh cells, and plaques of the different dilutions were counted 7 days later in triplicate. Standard curve UL41-containing fragments were cloned from MDV-BAC clones containing the UL41*Arg and UL41*Cys alleles by PCR using primers 97628F (5’ TTCAACTGCTGTCTCATCAC 3’) and 97628R (5’ TCCCAGAGCGCAGAACTTCC 3’). Amplicons containing the UL41*Arg and UL41*Cys alleles were cloned individually into the TOPO TA Vector (Invitrogen, Carlsbad, CA, USA) and the entire segment verified by DNA sequencing. DNA pools containing the two vectors in 10% increments were generated. The artificial DNA pools were analyzed by pyrosequencing using primers 97628fBiotin (5’ [BioTEG] TTCAACTGCTGTCTCATCAC 3’) and 97628R for PCR amplification, primer 97628Rseq (5’ CCAGGGTTAGTATTTI‘GAA 3’) for sequencing, and the PSQ 96M software (Biotage, Uppsala, Sweden). Competitive assay Three experimental groups were established. Viruses with the UL41*Arg, UL41*Cys. UL41*Cys—rev alleles at amino acid residue 377 (see Fig. 4—1), referred to as MDV-UL41*Arg, MDV-UL41*Cys, and MDV-UIAI*Cys-rev, respectively, were generated by transfections into CEF individually from pMDV-UL41*Arg, pMDV- 116 UL41*Cys, and pMDV-UL41*Cys-rev, respectively. Viral stocks of all three MDV types were made and the titers determined. CEF containing 1,000 unit PFU of MDV- UL41*Arg and MDV-UL41*Cys each, MDV—UL41*Arg and MDV-UL41*Cys-rev each, or MDV-UL41*CyS and MDV-UL41*Cys-rev each were mixed in triplicate. MDV- UL41*Arg, MDV-UL41*Cys, or MDV-UL41*Cys-rev were passed individually as controls. Virus DNA was extracted on every passage for SNP frequency determination and each sample was measured once by pyrosequencing. UL41 activity assay in vitro Chicken DFl cell line (Himly etal., 1998) was used for UL41 activity assays, and were cultured using the same conditions described for CEF. UL41 alleles from pMDV- UL41*Arg and pMDV-UL41*Cys were amplified using the primers UL41HindIIIF (5’ATCGAAGCTICGCCACCATGGGAGTATATGGATGTATG 3’) and UL41EcoRIR (5’ ATCGGAATTCTCAGTCATI‘AGATCGTGTTG 3’). The PCR products were digested with HindIII and EcoRI, cloned into the HindIII and EcoRI sites of the pcDNA3.1/Zeo vector (Invitrogen, Carlsbad, CA, USA), and the entire region verified by DNA sequencing. DFl cells were co—transfected with 0.5 pmoles of the B- galactosidase expression plasmid pON249 (Geballe et al., 1986) along with 0.01 to 0.5 pmoles of the UL41*Arg or UL41*Cys expressing plasmid. Two HSV-1 UL41- containing constructs, pCMVvhs and pCMVvhs-l, which express functional and inactive forms of the UL41 protein, respectively, were used as positive and negative controls (Jones et al-, 1995); kindly provided by Jeffery Cohen (NIAID, NIH). The pcDNA3.1/Zeo vector was added to each co-transfection reaction to maintain a constant 117 amount of the CMV promoter and total DNA. All transfections were performed in duplicate. Two days after co-transfection, the DP] cells were lysed by alternate cycles of freeze-thaw. B—galactosidase activity was measured at an OD of 420 nm in conjunction with the B—Gal Assay Kit (Invitrogen, Carlsbad, CA, USA). 118 Results Generation of pMDV-UL41*Arg and pMDV-UL41*Cys-rev by mini-2. mutagenesis The mini-2t phage recombinant mutagenesis system was used to introduce a point mutation into the original MDV-BAC clone to alter a specific UL41 amino acid (residue 377). Furthermore, additional point mutations were utilized to increase the efficiency of recovering the correct recombinant (AY510475:g.108,204C>G), or for differentiating between two viruses with the same amino acid sequence (AY510475:g.108,207A>G). A recombinant BAC clone that generates MDV with a UL41*Arg allele was generated from pMDV-UL41*CyS, the original MDV-BAC clone. pMDV-UL41*Cys-rev was generated from pMDV-UL41*Arg, which contains the same amino acid but a different nucleotide sequence compared to pMDV-UL41*Cys (Fig. 4-1). This reverse mutant was generated as a negative control for the head-to-head competitive assay. All of the MDV-BAC clones were confirmed by restriction enzyme digestions using 8amHI or HindIII for the intact viral genome, and sequencing of the UL41 SNP and flanking regions. 119 Growth rates of individual viruses in vitro MDV-UL41*Cys, MDV-UL41*Arg, and MDV-UL41*Cys-rev were generated from the appropriate MDV-BAC clones. To determine potential effects of the UL41 polymorphisms, each virus was grown individual in vitro. As shown in Fig. 4-2, there was no significant difference between any of the three MDVs. ~~e—--- UL41*Arg + UL41*Cys Plaques —+- UL41*Cys- I'BV Day Figure 4-2. Growth rates of individual viruses in vitro. MDV-UL41*Cys, MDV- UL41*Arg, and MDV-UL41*Cys-rev were grown on CEF, harvested and titered on the days indicated. The experiment was performed in triplicate and the average is given. Standard curve for pyrosequencing A standard curve was generated to calibrate the pyrosequencing technology using DNA pools containing a known mixture of cloned UL41*Arg and UL41*Cys alleles in 10% increments. The standard curve showed that the pyrosequencing signals for the 120 SNPs accurately reflected the original proportion of the two recombinant vectors without a need for extrapolation (Fig. 4-3). l 90 60 ' / 50 ’ / 4o 30 20 .,--.__.___- ____ _ -__.- s _. _at ~-__ss _.-_ v a ._ _. 10 Percent of UL41*Arg as determined by pyrosequencing O 10 20 3O 4O 50 60 70 80 90 100 Percent of UL41*Arg in the DNA pool Figure 4-3. Standard curve of pyrosequencing for UL41 alleles. DNA fragments containing the UL41*Arg and UL41*Cys alleles were mixed at known amounts to generate DNA pools that differed in 10% increments (X axis). The percent of UL41*Arg was determined by pyrosequencing (Y axis). Increasing virus fitness of MDV-UL41*Cys in virus populations in vitro In the group containing MDV-UL41*Arg and MDV-UL41*Cys, the average percentage of MDV-UL41*Arg steadily decreased from 52% to 5% after 20 passages, while the average percentage of MDV-UL41*Cys increased from 48% to 95% (Fig. 4-4a). In the MDV-UL41*Arg and MDV-UL41*Cys-rev mixed group, the average percentage of MDV-UL41*Arg decreased from 50% to 5%, and the percentage of MDV-UL41*Cys- rev increased from 50% to 95% relatively over 20 passages (Fig. 4—4b). Note that control groups that contained MDV-UL41*Cys and MDV-UL41*Arg separately, the SNP frequency was constant at 100%. To exclude the possibility that these results were attributed to unexpected mutations during the mutagenesis approach itself, the virus growth in a mixed group containing the same amino acid (Cys) but different SNPs were evaluated. In contrast to other mixtures, following 20 passages, the proportion of MDV- UL41*Cys and MDV-UIAl*Cys-rev genomes remained at a similar proportion (Fig. 4- 4C). Figure 4-4. The percentage of MDVs genomes in the competition assay. A. The percent of MDV genomes with the UL41*Arg alleles among a mixture of CEF infected with viruses containing either UL41*Arg or UL41*Cys alleles as determined by pyrosequencing. EXP 1, EXP2 and EXP3 represent three independent replicates. MDV-UL41*Cys Control is the result where CEF were infected with MDV containing the UL41*Cys allele only. B. Results from the same experimental design as A except that MDVs differed in having either UL41*Arg or UL41*Cys-rev alleles. MDV-UL41*Arg Control. results are from infected CEF where MDV containing the UMI*Arg allele only were passaged. C. Results from the same experimental design as A and B except that MDVS differed in having either UL41*Cys or UL41*Cys-rev alleles. MDV- UL41*Cys-rev Control results are from infected CEF where MDV containing the UL41*CyS-rev allele only were passaged. 123 Figure 4-4 124 iii} itii iii} EXP 1 EXP2 EXP3 Control EXP l EXP2 EXP3 Control EXP l EXP2 EXP3 Control Wild type MDV UL41 (MDV-UL41*Arg) has vhs activity and this enzyme activity is abolished in MDV-UL41*Cys When DFl cells were co-transfected with plasmids expressing MDV UL41*Arg and B—galactosidase, B—galactosidase activity decreased as the concentration of the UL41- expressing plasmid increased (Fig. 4-5). On the other hand, the B-galactosidase activity did not change with varying MDV UL41*CyS expressing plasmid amounts when co- transfected with a B-galactosidase expression plasmid into the DP] cells (Fig. 4—5). The HSV-1 vhs plasmid, the positive control, showed vhs activity that was dependent on the amount of plasmid DNA while the HSV-1 vhs-l plasmid, the negative control, did not show any detectable vhs activity (Fig. 4-5). The background B-galactosidase activity of DFl cells only was less than 0.15 when measured at an OD of 420 nm. 2.5 t... '5 +UL41*Arg OD420 +UL41*Cys +HSV-1 vhs 0.5 \ —x—Hsv-1 vhs—1 0 T l I 1 0 0.1 0.2 0.3 0.4 0.5 0.6 pmol of UL41-expressing plasmid Figure 4—5. The B— galactosidase assay for MDV UL41*Arg, UL41*Cys activity and the controls. Plasmids expressing MDV UL41*Cys, MDV UL41*Arg, HSV-l UL41 (vhs, positive control), or a mutant form of HSV-1 UL41 (vhs-1, negative control) were co- transfected into chicken DFl cells along with the fi-galactosidase expression plasmid pON249. and after two days, the cell lysates were measured for B—galactosidase activity. The values on the X axis indicate the amount of the UL41 expressing plasmid that was transfected; in addition, 0.5 pmole of pON249 was added along with varying amounts of pcDNA3. l/Zeo vector to keep CMV promoter constant. This figure shows the result from one of two replicates, which were consistent. 126 Discussion Understanding gene function in MDV has been greatly facilitated by the ability to make defined recombinant viruses through MDV-BAC clones. However, it is often the case that recombinant MDVs lacking a gene of interest show little to no altered phenotype in vitro but significant changes in viva (e.g., Reddy et al., 2002; Silva et al., 2004; Prigge et al., 2004; Kamil et al. 2005; Schumacher et al., 2005). This suggests that the in vitro assays are not very sensitive in detecting relevant mutations, which may be compounded by the fact that there are different environmental conditions and selective pressures in vitro compared to those in viva. To provide an alternative and more efficient screening method for evaluating recombinant MDVs in vitro, a head-to-head competition assay was employed, which is very sensitive to detect differences in viral fitness, combined with pyrosequencing to quantify each viral genome. The missense mutation in MDV-UL41 that results in Cys at amino acid 377 was found to enhance the viral fitness in cell culture by the head-to-head competition assay but not with the traditional in vitro grth curves of individual viruses. In addition, the growth advantage was not shown in the mixed viral population with MDV-UL41*Cy3 and MDV-UL41*Cys-rev, viruses that have identical U141 amino acids. Thus, mixed MDV populations favor the growth of viruses with inactive forms of vhs in cell culture, and the pyrosequencing approach could quantify this difference in viral fitness. During 20 passages, the MDV-UL41*Arg was almost replaced by MDV- UL41*Cys in the viral population when the initial ratio of the two viruses was 1:1. This adaptability of MDV-UL41*Cys is gradual, consistent between replicates, and constant, which suggests the UL41*Cys allele provides a moderate relative fitness advantage at each passage. Also, a similar fitness advantage was observed in viral p0pulations where the ratio of MDV-UL41*Arg to MDV-UL41*CyS was 4:1 suggesting that viral fitness depends on the number of passages rather than the initial ratio. The separate MDV- UL41*Arg and MDV-UL41*Cys viruses did not Show any SNP frequency change after 20 passages. These controls provide evidence that MDV with different UL41 alleles are capable of growing in cell culture, there is no contamination among plates, and the SNPS are stable. As the fitness advantage was dependent on the UL41 amino acid composition, UL41 activity was evaluated. The Arg in the variant amino acid site was reported to be conserved in all alpha-herpesviruses, and the replacement of Arg to His in HSV-1 abolishes endoribonuclease activity of UL41 (Everly and Read, 1999). Using the same [3- galactosidase functional assay, the effect of the Arg to Cys substitution in MDV UL41 was determined. UL41*Cys lacks any detectable endoribonuclease activity. However, the wild type MDV UL41 containing Arg has endoribonuclease activity at levels similar to HSV-l UL41. Thus, in the competitive assay, the null MDV UL41 gene is beneficial to MDV replication when competing with MDV with the UL41 wild type allele. It has been reported that HSV with a non—functional UL41 showed impaired growth comparing with wild type HSV in mice though there is no obvious growth difference in cell culture (Smith et al., 2002). The HSV-1 UL41 gene had been demonstrated to play an important role in the virus life cycle as it is capable of degrading the host mRNA although this degradation, for the most part, does not discriminate among host and viral mRNAs (Kwong and Frenkel, 1987; Esclatine et al., 2004). One possible hypothesis for this contradictory result is that a functional MDV UL41 gene is advantageous for virus growth and spread in viva where its vhs may interfere with the host immune response, however in vitro, a null UL41 allele confers a slight viral replication advantage as the viral transcripts are not degraded. This slight difference might not be detected by classical plaque assay as a previous report demonstrated that the effect of vhs deletion on viral replication in tissue is minimal (Smith et al., 2002). If true, then these results support a role for MDV UL41 in interfering with the host immune system although further studies are required to determine the detail mechanism of UL41 gene in MDV biology. In summary, this study presents a quantitative approach for head-to-head viral competition assays using SNPs as genetic markers for MDV. This approach is capable of directly detecting the ratio of virus genomes represented by the SNP on genomes and providing very sensitive monitoring at every passage. The use of this approach should help evaluate whether other MDV genes and genetic variation have functional roles in vitro and potentially in vivo. Acknowledgements We thank Thomas Goodwill, Laurie Molitor, and Lonnie Milam for their excellent technical assistance with pyrosequencing and B-Gal assays, and Dr. Dodgson for useful suggestions. This work was supported in part by USDA NRICGP #2002-03407 toHHC. 129 References l. 1.0. Barzilai, A., Zivony-Elbom, I., Sarid, R., Noah, E., Frenkel, N. 2006. The herpes simplex virus type 1 vhs-UL41 gene secures viral replication by temporarily evading apoptotic cellular response to infections with elements of cellular mRN A degradation machinery. J Vir0180: 505-513. Carrasco, P., Daros, J.A., Agudelo-Romero, P., Elena, SF. 2007. A real-time RT-PCR assay for quantifying the fitness of tobacco etch virus in competition experiments. J Virol Meth 139: 181-188. Costantino, N., Court, D.L. 2003. Enhanced levels of 7s Red-mediated recombinants in mismatch repair mutants. Proc Natl Acad Sci USA 100: 15748- 15753. Court, DL, Swaminathan, 8., Yu, D., Wilson, H., Baker, T., Bubunenko, M., Sawitzke, J ., Sharan, SK. 2003. Mini-2.: a tractable system for chromosome and BAC engineering. Gene 315: 63-69. Esclatine, A., Taddeo, B., Roizman, B. 2004. The UL41 protein of herpes simplex virus mediates selective stabilization or degradation of cellular mRNAs. Proc Natl Acad Sci USA 101: 18165-18170. Everly, D.N., Read, GS. 1999. Site-directed mutagenesis of the virion host shutoff gene (UL41) of herpes simplex virus (HSV): analysis of functional differences between HSV type 1 (HSV-1) and HSV-2 alleles. J Virol 73: 9117- 9129. Geballe, A.P., Spaete, R.R., Mocarski, ES. 1986. A cis—acting element within the 5’ leader of a cytomegalovirus B transcript determines kinetic class. Cell 46: 865-872. Himly, M., Foster, D.N., Bottoli, I., Iacovoni, J .S., Vogt, P.K. 1998. The DF-l chicken fibroblast cell line: transformation induced by diverse oncogenes and cell death resulting from infection by avian leucosis viruses. Virology 248: 295—408. Holland, J.J., de la Torre, J.C., Clarke, D.K., Duarte, EA. 1991. Quantification of relative fitness and great adaptability of clonal populations of RNA viruses. J Virol 65: 2960-2967. Jones, F.E., Smibert, C.A., Smiley, J .R. 1995. Mutational analysis of the herpes simplex virus virion host shutoff protein: Evidence that vhs functions in the absence of other viral proteins. J Virol 69: 4863-4871. 130 11. 12. 13. 14. 15. l6. 17. 18. 19. 20. 21. 22. Kamil, J.P., Tischer, B.K., Trapp, 8., Nair, V.K., Osterrieder, N., Kung, HJ. 2005. vLIP, a viral lipase homologue, is a virulence factor of Marek's disease virus. J Virol 79: 6984-6996. Kwong, A.D., Frenkel, N. 1987. Herpes simplex virus-infected cells contain a function(s) that destabilizes both host and viral mRNAs. Proc Natl Acad Sci USA 84: 1926-1930. Kwong, A.D., Frenkel, N. 1988. Herpes simplex virus virion host shutoff function. J Virol 62: 912-921. Lahser, F.C., Wright-Minogue, J., Skelton, A., Malcolm, B.A. 2003. Quantitative estimation of viral fitness using pyrosequencing. BioTech 34: 26-28. Murphy, J .A., Duerst, R.J., Smith, T.J., Morrison, LA. 2003. Herpes Simplex virus type 2 virion host shutoff protein regulates alpha/beta interferon but not adaptive immune responses during primary infection in vivo. J Virol 77: 9337- 9345. Niikura, M., Dodgson, J, Cheng, H.H. 2006. Direct evidence of host genome acquisition by the alphaherpesvirus Marek’s disease virus. Arch Virol 151: 537- 549. Novella, 1.8. 2003. Contributions of vesicular stomatitis virus to the understanding of RNA virus evolution. Curr Opin Microbiol 6: 399-405. O’Meara, D., Wilbe, K., Leitner, T., Hejdeman, B., Albert, J., Lundeberg, J. 2001. Monitoring resistance to human innunodeficiency virus type 1 protease inhibitors by pyrosequencing. J Clin Microbiol 39: 464-473. Osterrieder, N., Kamil, J.P., Schmacher, D., Tischer, B.K., Trapp, S. 2006. Marek’s disease virus: from miasma to model. Nature Rev Micro 4: 283-294. Petherbridge, L., Brown, A.C., Baigent, S.J., Howes, K., Sacco, M.A., Osterrieder, N., Nair, V.K. 2004. Oncogenicity of virulent Marek's disease virus cloned as bacterial artificial chromosomes. J Virol 78: 13376-13380. Prigge, J .T., Majerciak, V., Hunt, H.D. Dienglewicz, R.L., Parcells, MS, 2004. Construction and characterization of Marek's disease viruses having green fluorescent protein expression tied directly or indirectly to phosphoprotein 38 expression. Avian Dis 48: 471-487. Read, G.S., Frenkel, N. 1983. Herpes simplex virus mutants defective in the virion-associated shutoff polypeptide synthesis and exhibiting abnormal synthesis of alpha (immediate early) viral polypeptides. J Virol 46: 498-512. 131 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Read, G.S., Karr, B.M., Knight, K. 1993. Isolation of a herpes simplex virus type 1 mutant with a deletion in the virion host shutoff gene and identification of multiple forms of the vhs (UL41) polypeptide. J Virol 67: 7149-7160. Reddy, S.M., Lupiani, B., Gimeno, I.M., Silva, R.F., Lee, L.F., Witter, R.L. 2002. Rescue of a pathogenic Marek's disease virus with overlapping cosmid DNAs: Use of a pp38 mutant to validate the technology for the study of gene function. Proc Natl Acad Sci USA 99: 7054-7059. Ronaghi, M. 2001. Pyrosequencing sheds light on DNA sequencing. Genome Res 11: 3-11. Ronaghi, M., Uhlen, M., Nyren, P. 1998. A sequencing method based on real- time pyrophosphate. Science 281: 363-365. Schumacher, D., Tischer, B.K., Fuchs, W., Osterrieder, N. 2000. Reconstitution of Marek's disease virus serotype 1 (MDV-1) from DNA cloned as a bacterial artificial chromosome and characterization of a glycoprotein B- negative MDV-l mutant. J Virol 74: 11088-11098. Schumacher, D., Tischer, B.K., Trapp, 8., Osterrieder, N. 2005. The protein encoded by the U83 orthologue of Marek's disease virus is required for dfficient de-envelopment of perinuclear virions and involved in actin stress fiber breakdown. J Virol 79: 3987-3997. Silva, R.F., Lee, L.F. Kutish, G.F. 2001. The genomic structure of Marek’s disease virus, in: Hirai, K. (Ed.), Marek’s Disease. Springer, Berlin, pp. 143-158. Silva, R.F., Reddy, S.M., Lupiani, B. 2004. Expansion of a unique region in the Marek's disease virus genome occurs concomitantly with attenuation but is not sufficient to cause attenuation. J Virol 78: 733-740. Smith, T.J., Morrison, L.A., Leib, D.A. 2002. Pathogenesis of herpes simplex virus type 2 virion host shutoff (vhs) mutants. J Virol 76: 2054-2061. Spatz, S.J., Peterbridge, L., Zhao, Y., Nair, V. 2007. Comparative full-length sequence analysis of oncogenic and vaccine (Rispens) strains of Marek’s disease virus. J Gen Virol 88: 1080-1096. Spatz, S.J., Silva, R.F. 2007. Polymorphisms in the repeat long regions of oncogenic and attenuated pathotypes of Marek’s disease virus 1. Virus Genes 35: 41-53. Suzutani, T., Nagamine, M., Shibaki, T., Ogasawara, M., Yoshida, 1., Daikoku, T., Nishiyama, Y., Azuma, M. 2000. The role of the UL41 gene of 35. 36. herpes simplex virus type 1 in evasion of non-specific host defense mechanisms during primary infection. J Gen Virol 81: 1763-1771. Witter, R.L. 1996. Increased virulence of Marek’s disease virus field isolates. Avian Dis 41: 149-163. Yuste, E., Borderia, A.V., Domingo, E., Lopez-Galindez, C. 2005. Few mutations in the 5’ leader region mediate fitness recovery of debilitated human immunodeficiency type lviruses. J Virol 79: 5421-5427. 133 Chapter 5 Future work 134 This thesis described the cloning and functional characterization of chicken SCA2, and its influence on MDV growth properties in vitro via association with viral U810 protein. To investigate the biological properties of chicken SCA2, chicken SCA2 was expressed and purified in E. coli, and a polyclonal antibody was developed. Utilizing this anti-SCA2 antibody, SCA2 was identified to be a 13 kDa cell surface protein anchored by a GP] moiety as is the case for most other Ly-6 members (Classon and Coverdale, 1994). In viva studies showed that SCA2 is expressed in a number of chicken tissues. Of particular interest was the unique SCA2 expression pattern in bursal CMEC that were often surrounded by MHC class II presenting cells. Expression profiles of bursal cells induced by contact with SCA2-expressing cells shows down-regulation of numerous genes including CD798, B cell linker (BLNK), spleen tyrosine kinase (S YK), and gamma 2-phospholipase C (PLCGZ) that are involved in the BCR and immune response signaling pathways. These results suggest chicken SCA2 plays a role in regulating B lymphocytes. Based on the identified functions and expression patterns of SCA2, we hypothesize three putative roles for SCA2 in MDV biology. First, SCA2 promotes MDV spread as SCA2 may function in cell-cell adhesion with direct cell-cell contact required for MDV spread. Second, SCA2 transduces signals for cell proliferation that could affect transformation of cells harboring latent MDV. Third, SCA2 promotes virion maturation or egress via interacting with viral proteins. Some reports have demonstrated that chicken SCA2 is associated with v-Rel—transformed avian bone marrow cells and plays a role in self-renewal of avian erythroid progenitors (Petrenko et al., 1997; Bresson-Mazet et al., 2008). We also found the cell surface SCA2 was up-regulated in a MDV transformed T 135 cell line (RPl) when induced for MDV replication (Fig. 5-1). An important question that one could address is whether up-regulation of SCA2 is specific for MDV or it is a response from host immune system. The preliminary results suggest that cell surface SCA2 was not induced by a reticuloendotheliosis virus in RPl-transformed T cell line. Consequently, the up-regulation of SCA2 could be induced by MDV gene products. SCA2 transcription was induced by MDV infection in CEF, a fibroblast cell line. Therefore, these results support the conclusion that MDV gene products, probably US 10, induce the SCA2 transcription and further SCA2 cell surface expression. Interestingly, SCA2 transcription was up-regulated in many malignant tumor cell lines (Bresson-Mazet, 2008). Consequently, the question arises as to whether virus-induced SCA2 promotes virus-induced lymphomogenesis. The expression profiling results from B lymphocytes upon contact with SCA2-expressing cells indicated that SCA2 is involved in BCR and immune response signal pathways. Therefore, future efforts should be directed to investigate the role of SCA2 in signal transduction in chicken lymphocytes. While we are the first to show a biological role for MDV U810, the role of MDV U810 is stil not well defined. The interaction of U810 and SCA2, and the putative function of SCA2 in lymphocytes proliferation suggest that U810 may regulate chicken lymphocyte via interacting with SCA2. Therefore, to elucidate the role of MDV U810 protein and the functional interaction of SCA2 and U810 for transformation or immune responses of lymphocytes, we could profile the expression of host genes of lymphocytes infected with MDV and AUSl’O-MDV to find out the immune response pathways in which U810 is involved. Parcell et al. (1994) reported that MDV lacking USI, U82 SORFI, SORF2, SORF3 and USIO showed decreased early cytolytic infection, mortality, 136 tumor incidence, and horizontal transmission, but the role of U810 alone in MDV pathogenicity was not identified. A AUSIO-MDV recombinant BAC has been generated, which we could use to characterize U810 with respect to MDV pathogenicity. The US lOEGFP-MDV we developed could also be employed for studies on MDV morphogenesis. An ideal assay for MDV morphogenesis analysis would be to label MDV membrane, tegument, and capsid proteins each with unique fluorescent tags. By utilizing this recombinant MDV, one could track the MDV dynamics of viral replication in living fibroblast cells. This would allow one to address areas such as: l) The timing of new virion maturation and egress to the neighboring cells. 2) Is an intact virion required for MDV spread between cell-cell? 3) Does the virion transmit between cells mainly through cell—cell fusion? 4) The efficiency of super-infection. Due to the highly cell-associated characters, we would use MDVs labeled with fluorescent proteins to synchronize MDV infection. Furthermore, MDVs with fluorescent proteins might also be employed for tracking and collecting infected-lymphocytes in viva. Finally, on an unrelated subject, during the cloning and full genome sequencing of virulent MDV Mdll strain (Niikura et al., 2006), another nonsynonymous single nucleotide polymorphism (SNP; AY510475:g.l60,3l9 A>G) was found in the ICP4 gene that resulted in a Leu to Ser substitution in residue 256 (AA801711.1:p.Leu2568er; Fig. 5-2). Monitoring the frequency of each SNP by pyrosequencing during virus passages determined the ratio of each viral genome in a single monolayer, which is a very sensitive method to monitor viral fitness. MDV with the ICP4*Ser allele showed enhanced fitness in vitro (Fig. 5-3). HSV ICP4 protein is a transcriptional regulator required for the induced transcription of most of the remaining HSV genes and essential for viral 137 replication (Compel and DeLuca, 2003; Su et al., 2006). The exact role of ICP4 in MDV replication has not been well determined although it is speculated to play a role similar to the ICP4 of other alpha-herpesviruses as a transcription regulator, and it is associated with maintenance of transformation and is essential for MDV growth in CEF (Endoh, 1996; Chattoo et al., 2006). Therefore, we would identify the effect of the SNP on the role of ICP4 as a transcriptional activator, and further on the virus evolution. 138 RP 1 00118 BrdU-induced cells with anti-SCA2 anti-SCA2 antisera inactivated cells with anti-SCA2 inactivated cells with prebleed antisera BrdU-induced cells with prebleed antisera Figure 5-1. SCA2 is up-regulated on the surface of BrdU-induced RPl (MDV transformed) T cell line. 139 Parental MDV-ICP4*Ser MDV-ICP4*Leu MDV-ICP4*Leu-rev Figure 5-2. Partial ICP4 DNA and amino acid sequences for the parental Mdll strain, the recombinant virus (MDV-ICP4*Ser), the original MDV-BAC clone (MDV-ICP4*Leu), and the recombinant viruse (MDV-ICP4*Leu-rev). 140 GC C A] a GCC A] a GCC Al 21 l GCG A] a TT G LC u TCG Se r I TTG Leu TTG Leu Figure 5-3. The percetange of MDVs genomes containing SNPs in ICP4 alleles in a competition assay. A. The percent of MDV genomes with the ICP4*Leu alleles among a mixture of CEF infected with viruses containing either ICP4*Leu or ICP4*Ser alleles as determined by pyrosequencing during 8 passages. Exp 1, Exp 2 and Exp 3 represent three independent replicates. B. Results from the same experimental design as A except that MDVs differed in having either ICP4*Ser or ICP4*Leu-rev alleles. C. Results from the same experimental design as A and B except that MDVs differed in having either ICP4*Leu or ICP4*Leu-rev alleles. I41 EXP l EXP2 EXP3 Hi ' Y 1 fi fi— 1? a? a Passages EXP] EXP2 EXP3 Hi Passages EXP 1 EXP2 EXP3 Hi Passages Figure 5-3 142 References 1. Bresson-Mazet, C., O. Gandrillon, and S. Gonin-Giraud. 2008. Stem cell antigen 2: a new gene involved in the self-renewal of erythroid progenitors. Cell Prolif 41:726-38. Chattoo, J. P., M. P. Stevens, and V. Nair. 2006. Rapid identification of non- essential genes for in vitro replication of Marek's disease virus by random transposon mutagenesis. J Virol Methods 135:288-91. Classon, B. J., and L. Coverdale. 1994. Mouse stem cell antigen Sca-2 is a member of the Ly-6 family of cell surface proteins. Proc Natl Acad Sci U S A 91:5296-300. Compel, P., and N. A. DeLuca. 2003. Temperature-dependent conformational changes in herpes simplex virus ICP4 that affect transcription activation. J Virol 77:3257-68. Endoh, D. 1996. Enhancement of gene expression by Marek's disease virus homologue of the herpes simplex virus-l ICP4. Jpn J Vet Res 44:136-7. Niikura, M., J. Dodgson, and H. Cheng. 2006. Direct evidence of host genome acquisition by the alphaherpesvims Marek's disease virus. Arch Virol 151:537-49. Parcells, M. 8., A. S. Anderson, J. L. Cantello, and R. W. Morgan. 1994. Characterization of Marek's disease virus insertion and deletion mutants that lack U81 (ICP22 homolog), U810, and/or U82 and neighboring short-component open reading frames. J Virol 68:8239-53. Petrenko, O., I. Ischenko, and P. J. Enrietto. 1997. Characterization of changes in gene expression associated with malignant transformation by the NF-kappaB family member, v-Rel. Oncogene 15:1671-80. Su, Y. H., X. Zhang, X. Wang, N. W. Fraser, and T. M. Block. 2006. Evidence that the immediate-early gene product ICP4 is necessary for the genome of the herpes simplex virus type 1 ICP4 deletion mutant strain d120 to circularize in infected cells. J Virol 80:11589-97. 143