HUMAN T LYMPHOTROPIC VIRUS TYPE 1 (HTLV-1) ACCESSORY

PROTEIN P30(II) MODULATES CELLULAR AND VIRAL GENE EXPRESSION

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the Graduate

School of The Ohio State University

by

BINDHU MICHAEL, B.V.Sc. & A.H., M. Sc., M.S.

* * * * *

The Ohio State University

2004

Dissertation Committee: Approved by Professor Michael Lairmore, advisor

Professor Patrick Green

Professor Michael Oglesbee ______Adviser Professor Stefan Niewiesk Veterinary Biosciences Graduate Program ABSTRACT

Human T-lymphotropic virus type-1 (HTLV-1) is a deltaretrovirus that causes

adult T-cell leukemia/lymphoma, and is implicated in a variety of lymphocyte-mediated disorders. In addition to structural and enzymatic proteins, the HTLV-1 provirus encodes

various regulatory and accessory genes in four pX open reading frames. Our laboratory

has elucidated the functional characteristics of the HTLV-1 accessory proteins, and it has

become clear that these proteins are, in fact, essential for the virus life cycle and may

determine disease outcome associated with HTLV-1 infection. We have identified the

functional properties of the pX ORF-II encoded p30II, but the role of this viral protein in

virus replication or pathogenesis remain incompletely defined.

Studies from our laboratory have demonstrated that pX ORF-II mutations

diminish the ability of the virus to maintain high viral loads in vivo in our rabbit model of

HTLV-1 infection. The ORF-II encoded p30II, a nuclear-localizing protein that interacts

with the KIX domain of CREB-binding protein (CBP)/p300, disrupts CREB-Tax-

CBP/p300 complexes bound to viral Tax-Responsive Element (TRE) repeats and

differentially modulates CREB and TRE-mediated . Herein, we have further

characterized the role of p30II in regulation of cellular gene expression, using a stable

p30II expression system employing lentiviral vectors to test cellular gene

ii

expression with Affymetrix U133A arrays, representing ~33,000 human genes. Our data reveals alterations of interrelated pathways of cell proliferation, T-cell signaling,

apoptosis and cell cycle in p30II expressing Jurkat T-lymphocytes. Moreover, we have

verified our findings by reporter assays and RT-PCR in Jurkat and primary CD4+ T-

lymphocytes. We are the first to test the overall effect of an HTLV-1 accessory protein,

on cellular gene expression and demonstrate that p30II activates many key transcription

factors involved in T-cell signaling/activation, such as nuclear factor of activated T-cells,

nuclear factor-kappa B and activator protein-1. Collectively, these data indicate that this

complex retrovirus, associated with lymphoproliferative diseases, rely upon accessory

gene products to modify cellular environment to enhance clonal expansion of the virus

genome and thus maintain proviral loads in vivo.

Next, we further characterized the role of p30II in regulation of viral gene

expression, by identifying motifs critical in binding p300 and regulating TRE-mediated

transcription in the absence or presence of provirus. Analysis of the amino acid domain

of p30II (100-179) critical for repressing LTR mediated transcription, was used to identify

a lysine residue at amino acid 106 (K3) of p30II, that is critical for repressing TRE- mediated transcription. Additionally, we have found that p300 reverses the p30II- dependent repression of TRE-mediated transcription, in the absence or presence of the provirus, in a dose-responsive manner. Our data confirms the role of p30II as a regulator

iii

of viral gene transcription, in association with p300. However, in contrast to wildtype

p300, p300 histone acetyl transferase (HAT) mutants only partially rescue p30II-mediated

LTR repression. In addition, we also show that p30II is acetylated and that deacetylation enhances p30II-mediated HTLV-1 LTR repression. Inhibition of deacetylation decreased

p30II-mediated LTR repression. We are the first to demonstrate that the HAT activity of

p300 is crucial in modulating viral gene expression from HTLV-1 LTR by p30II. Overall, our data indicates that HTLV-1, a complex retrovirus associated with lymphoproliferative disorders, uses accessory genes to modulate viral gene expression in conjunction with host cell proteins in part, through the acetylation of p30II. Collectively, our data suggests

that HTLV-1 enhances clonal expansion of the virus genome by regulating viral and

cellular gene expression, avoids immune elimination by the host and thereby maintains proviral loads in vivo.

iv

Dedicated to my wonderful parents, siblings and husband who gave me the freedom to be

who I want to be and were there for me always, no matter what, and

Vavlo who reminds me of the little joys in everyday life

v

ACKNOWLEDGMENTS

First and foremost, I would like to express my sincere gratitude and appreciation to my thesis advisor, Dr. Michael Lairmore for his valuable guidance, motivation, support and leadership. As stated in Dr. Lairmore’s recent University Distinguished Scholar

Award biography, he is truly a mentor who invests a great deal of himself in his students and believes it is through their success that he is best reflected. With his exceptional

mentoring skills, he always keeps his students’ best interest in his mind and trains us to

be independent investigators. Moreover, he was gracious enough to support me through my Pathology Training. In spite of his busy schedule, he is committed to meeting with every student frequently and providing constant support and guidance. I am fortunate to have the opportunity to learn from Dr. Lairmore, who successfully handles every aspect of his teaching, research, pathology services and administrative responsibilities.

I would like to thank my committee members Dr. Patrick Green, Dr. Michael

Oglesbee and Dr. Stefan Niewiesk for their helpful ideas and suggestions. I am also thankful to Dr. Larry Mathes and Dr. Kathleen Boris-Lawrie for their helpful discussions

in the Retrovirus Group lab meetings. I would also like to acknowledge Dr. Louis

Mansky who was a member of my committee during my candidacy examination and Dr.

Weisbrode who has been a great teacher during my Pathology residency training.

vi

I was fortunate to be a part of Dr. Lairmore’s research group and would like to

thank all my colleagues, both past and present. I extend my sincere appreciation to Antara

Datta, Hajime Hiraragi and Wei Ding for their help me with some of the experiments discussed in this dissertation, valuable scientific discussions and for their friendship. I also thank Chris Premandandan, Seungjae Kim, Rashade Haynes, Andy Montgomery,

Lee Silverman, Andrew Phipps, John Nisbet, Jessica Alcorn and Laurie Millward for their helpful ideas, productive scientific discussions and great times shared together. I appreciate the effort made by Cynthia Sagel, Jennifer Huck, Stacey Dietrich, Susie Vogel and Allison Blanton in the maintenance of the laboratory, equipment and reagents. I must also thank members of the Boris-Lawrie, Green and Mathes laboratories for their discussions, reagents, and friendship.

I am previlaged to have had many mentors who taught me that the sky is the limit and to not be afraid to pursue my dreams and ambitions: right from Ashan who taught me the alphabets, Sr. Rosia and other teachers who invested a great deal of themselves for me, Dr. Francis Xavier and Dr. Nanu E who supported me immensely while pursuing my ambitions. I have also been lucky to have some great friends like Sangeetha and Sajan who believed in me. Special thanks to Desi friends, David Joseph, Tanya, Stan, Jordan,

Rachelmol, Thomas and George who made my stay in Columbus enjoyable.

vii

I am blessed with a wonderful and loving family. My parents Michael and

Kuttiyamma have always been a source of motivation, inspiration and support throughout

my life. In a small town where women’s higher education particularly in a foreign

country, was not supported much, what my parents have done needs special appreciation.

Moreover, they have always been there even when I did not listen to them and was faced

with difficult circumstances in life. My sister Rani and brother Bibin have always

believed in me and have been very supportive, sincere and loving. I also thank my

cousins Baiju, Asha, Jose and Bitto who have been loving and supportive to my family. I

could not have completed my PhD and training, without the unconditional love and support of my husband, friend and colleague Amrithraj, who have always been very

encouraging and understanding. Last but not least, special thanks to Vavlo, who reminds

me of the little joys and fun of everyday life.

viii

VITA

Birthday...... Born – 12th August, 1974

1989-91 ...... Pre-Degree- Biology, Chemistry, Physics Mahathma Gandhi University, Kottayam, Kerala, India

1991-97 ...... BVSc and AH (DVM equivalent) Kerala Agricultural University, India

1997-1998 ...... Research Associate Department of Veterinary Public Health Kerala Agricultural University, India

1998-99…………………………………….. MSc (Wild animal Health) University of London, UK

1999-2001………………………………….. Graduate Teaching Assistant Department of Microbiology The Ohio State University

2001-present………………………………...Graduate Research Associate, Department of Veterinary Biosciences The Ohio State University

PUBLICATIONS

Research Publications

1. Ding, W., Kim, S., Nair, A., Michael, B., Boris-lawrie, K., Tripp, A., Feuer G and Lairmore, M. Human T-Lymphotropic Virus Type-1 p12I Enhances Interleukin-2 Production during T Cell Activation. Journal of Virology. 77: 11027-11039, 2003

2. Michael, B., Smith, J. N., Swift, S., Heffron, F. and Ahmer, B. M. SdiA of Salmonella enterica is a LuxR Homolog that Detects Mixed Microbial Communities. Journal of Bacteriology.183: 5733-42, 2001.

ix

3. Francis Xavier, Gigi K. Joseph and Bindhu Michael. A Coprological Survey of Parasites of Seven Mammal Groups at Silent Valley National Park, Kerala, Zoo’s Print. 15, 2000

4. Amrithraj, M., Bindhu Michael and Madhavan Pillai, K. Artyfechinostomum malayanum (Leiper, 1911) Mendheim, 1943 from a Small Indian Civet (Viverricula indica). Zoo's Print. 14: 6, 1999

5. Bindhu Michael, Gigi. K. Joseph and Francis Xavier. Comparison of Morphometric Indices of Large and Small Flying Squirrels. Zoo’s Print. 13: 46-47, 1998.

6. Vijayan, R., Devada, K., Balakrishnan, V.S., Bindhu Michael and Aleyas, N. M. Ehrlichiosis in Dog - a Case Report. Journal of Veterinary and Animal Sciences. 28: 101-103, 1997.

7. Bindhu Michael, M. Amrithraj and K. Madhavan Pillai. A Note on Isospora Infection in a Southern Red Whiskered Bulbul (Pycnonotus jocosus fascicaudatus). Zoo’s Print. 12: 5, 1997.

8. Bindhu Michael, Francis Xavier and Leo Joseph. Anomalous Animal Behaviour Preceding Earthquake. Zoo’s Print. 12: 30-31, 1997.

9. Bindhu Michael, Amrithraj M., Sajan George and P. J. Rajkamal. Case Study of a Specialised Buffalo Farmer in the Foot-hills of Western Ghats. Dairy Guide. 18: 13- 14, 1997.

10. Bindhu Michael, Gigi. K. Joseph and Francis Xavier. Morphometry and Feeding Habit of Small Travancore Flying Squirrel. Zoo’s Print. 11: 4-5, 1996.

Review Publications

1. Bindhu Michael, Amrithraj M. Nair and Michael D. Lairmore. Role of Accessory Proteins of HTLV-1 in Viral Replication, T Cell Activation, and Cellular Gene Expression. Frontiers in Bioscience. 9: 2556-2576, 2004.

2. M. Lairmore, Michael, B., Silverman. L and Nair. A. Human T-Lymphotropic Virus-Associated Neurological Disorders: Pathogenesis, Immunology and Clinical Management. Accepted as Chapter in textbook “Inflammatory disorders of the nervous system”

x

Previous Thesis / Dissertations

Detection of Heterologous Bacterial Pheromones by SdiA, a LuxR Homolog in Salmonella enterica serovar typhimurium: Master’s Dissertation, The Ohio State University, Columbus, Ohio, 2001 Identification of Type C Retrovirus and Simian T-Lymphotropic Virus from a Brown Headed Spider Monkey (Ateles fusciceps robustus). Master’s Dissertation, The Royal Veterinary College, University of London, London, UK, 1999

FIELDS OF STUDY

Major Field: Veterinary Biosciences.

xi

TABLE OF CONTENTS

Page Abstract...... ii

Dedication ...... v

Acknowledgments...... vi

Vita...... ix

List of Tables ...... xv

List of Figures...... xvi

Chapters:

1. Literature Review: Molecular mechanisms of HTLV-1 infection, LTR mediated gene expression and role of accessory proteins in HTLV-1 pathogenesis.

1.1 First Recognition of Human T-lymphotropic virus (HTLV-1) as Etiologic Agent of Human Disease ...... 1 1.2 Geographical Distribution of HTLV-1 ...... 2 1.3 Transmission of HTLV-1...... 3 1.4 HTLV-1 Associated Diseases...... 4

1.4.1 Adult T cell Leukemia/Lymphoma...... 4 1.4.2 Tropical spastic paraparesis/HTLV-1-associated myelopathy ..... 8 1.4.3 Other diseases associated with HTLV-1-infection ...... 10

1.5 Genomic Organization and Replication of HTLV...... 12 1.6 Structural Proteins of HTLV-1 ...... 16 1.7 Regulatory Proteins of HTLV-1: Tax and Rex...... 18

1.7.1 Tax and Transactivation of the HTLV-1 LTR ...... 18 1.7.2 Tax and Regulation of Cellular Gene Expression...... 19 1.7.3 HTLV-1 Rex: Post-transcriptional Regulator...... 21

1.8 HTLV-I accessory proteins...... 24 xii

1.8.1 pX ORF I p12I...... 26

1.8.1.1 Biochemical Characteristics of p12I: Features of a Signaling Molecule ...... 26 1.8.1.2 Subcellular Localization of p12I and Interactions with Cellular Proteins...... 27 1.8.1.3 Role of p12I in Regulation of Viral Infectivity In vitro and In vivo...... 30 1.8.1.4 Role of p12I in Calcium-Mediated T Cell Activation..... 31 1.8.1.5 Putative Role of p12I Variants in HAM/TSP...... 36

1.8.2 pX ORF II p13II ...... 37

1.8.2.1 Biochemical Characteristics of p13II: Mitochondrial Targeting...... 37 1.8.2.2 p13II Alteration of Mitochondrial Morphology and Inner Membrane Potential...... 37 1.8.2.3 p13II Cellular Protein Interactions ...... 39

1.8.3 pX ORF II p30II ...... 40

1.8.3.1 Biochemical Characteristics of p30II: Features of a Transcription modulator...... 41 1.8.3.2 Role of p30II in Modulating Viral and Cellular Gene Expression...... 42 1.8.3.3 HTLV-1 Mediated T Cell Activation /Transformation and the Role of p30II ...... 45

1.8.4 References...... 48

2. Human T Lymphotropic Virus Type-1 P30II Influences Cellular Gene Expression and Enhances Transcription Activation of T Lymphocytes

2.1 Introduction...... 87 2.2 Materials and Methods...... 90 2.3 Results...... 96 2.4 Discussion...... 101 2.5 References...... 107

3. Motifs of Human T Lymphotropic Virus Type 1 P30II critical for its Interaction with p300 and Regulation of LTR Transcriptional Activity. xiii

3.1 Introduction...... 121 3.2 Materials and Methods...... 125 3.3 Results...... 129 3.4 Discussion...... 137 3.5 References...... 144

4. Histone Acetyltransferase (HAT) Activity of p300 Modulates Human T Lymphotropic Virus Type 1 P30II-mediated Repression of LTR Transcriptional Activity.

4.1 Introduction...... 158 4.2 Materials and Methods...... 162 4.3 Results...... 165 4.4 Discussion...... 171 4.5 References...... 178

5. Synopsis and future studies

5.1 Role of p30II in T cell signaling / activation and transformation. 191 5.2 Identify the mechanism by which p30II activates NF-κB, NFAT and AP-1 mediated transcription ...... 192 5.3 Determine if p30II has an effect on IL-2 production and lymphocyte proliferation...... 194 5.4 Does p30II affect cell cycle progression in T lymphocytes?...... 195 5.5 Does p30II affect cell-to-cell adhesion between T lymphocytes? 196 5.6 Determine the significance of critical functional domains and specific amino acids within p30II, in vivo ...... 197 5.7 Does HTLV-1 p30II form complexes at the HTLV-1 , to modulate LTR mediated transcription? ...... 198 5.8 Identify the functional domains and specific amino acids within p30II, which are critical in nuclear retention of tax/rex mRNA and protein-protein interactions...... 199 5.9 Temporal expression pattern and interaction between HTLV-1 regulatory and accessory proteins during various stages of the HTLV-1 infection ...... 200 5.10 References...... 202

Bibliography ...... 207

xiv

LIST OF TABLES

Table Page

2.1 Genes modulated by HTLV-1 p30II ...... 113

3.1 Lysine residues in p30II are conserved among various HTLV isolates...... 149

xv

LIST OF FIGURES

Figure Page

1.1 Diagram of HTLV-1 genome and alternatively spliced mRNA and protein species from pX ORFs...... 85

1.2 Diagram of p12I, p30II, and p13II with predicted functional motifs…...... 86

2.1 Schematic illustration of lentiviral vectors expressing both p30IIHA and GFP (sample vector) as bicistronic messages and GFP alone (control vector) from elongation factor 1 alpha promoter...... 117

2.2 Triplicate p30II samples express GFP and p30II while triplicate controls express only GFP ...... 118

2.3 Semiquantitative RT-PCR of CHP, JUN and NFATc in controls and p30II expressing Jurkat T lymphocytes and primary CD4+ T lymphocyte samples...... 119

2.4 p30II activates NFAT, AP-1 and NF-κB transcriptional activity in Jurkat T lymphocytes ...... 120

3.1 p30II localizes to the nucleus and colocalizes with p300 in the nucleus of transiently transfected HeLa-Tat cells ...... 150

3.2 Amino acids 1-132 of HTLV-1 p30II are critical for its physical interaction with p300 ...... 151

3.3 Subcellular localization characteristics of various p30II serial deletion mutants p30II in transiently transfected HeLa-Tat cells...... 152

3.4 HTLV-1 p30II differentially modulates LTR-mediated transcription in the absence or the presence of the provirus...... 153

3.5 p300 expression reverses the p30II-dependent repression of LTR driven transcription, in the absence or the presence of the provirus...... 154

xvi

3.6 HTLV-1 p30II and Tax appear to compete with each other in modulating the LTR mediated transcription...... 155

3.7 Amino acids 100-179 of p30II is critical for its repression of LTR driven transcription, in the absence or the presence of the provirus ...... 156

3.8 Lysine 106 (K3) of p30II is critical for its repression of LTR driven transcription in the absence or the presence of the provirus...... 157

4.1 p300 HAT mutants only partially rescue p30II-dependent repression of LTR driven transcription ...... 183

4.2 Addition of HDAC-1 enhances p30II-dependent repression of HTLV-1 LTR driven transcription ...... 184

4.3 Addition of 100 nM Trichostatin A (TSA) causes a decrease in the enhancement of p30II-dependent repression of HTLV-1 LTR driven transcription by HDAC-1...... 185

4.4 Addition of Trichostatin A causes a decrease in the enhancement of p30II- dependent repression of HTLV-1 LTR driven transcription by HDAC-1 ...... 186

4.5 Expression of P/CAF does not rescue p30II-dependent repression of HTLV-1 LTR driven transcription...... 187

4.6 HTLV-1 p30II appears to be acetylated by the transcriptional co-activator p300 at lysine 106 (K3)...... 188

4.7 Schematic model diagram of the potential mechanism by which HTLV-1 p30II modulates LTR-driven transcription...... 189

xvii

CHAPTER 1

LITERATURE REVIEW: MOLECULAR MECHANISMS OF HTLV-1

INFECTION, LTR-MEDIATED GENE EXPRESSION AND ROLE OF

ACCESSORY PROTEINS IN HTLV-1 PATHOGENESIS

1.1 First Recognition of Human T-Lymphotropic Virus Type 1 (HTLV-1) as

Etiologic Agent of Human Disease

Takatsuki et al 334reported a unique and aggressive CD4+ T cell malignancy called

adult T cell leukemia (ATL) in a small geographically clustered group of patients with

lymphoid neoplasms in Japan In 1977 245,334,343. Human T-lymphotropic virus type 1 was

first detected and isolated as a type C retrovirus particle from two T cell lymphoblastoid cell lines, HUT 102 and CTCL-3 and fresh peripheral blood lymphocytes of a patient in the United States with cutaneous T cell lymphoma (mycosis fungoides) by Poisez, Gallo and their colleagues in 1980 293. This patient was diagnosed with cutaneous T cell

lymphoma; however he had widespread systemic involvement typical of ATL and is now

thought to have been possibly misdiagnosed and was in fact suffering from ATL. In

1981, Hinuma et al 123demonstrated the presence of type C retroviral particles termed 1 adult T cell leukemia virus (ATLV) from a cell line termed MT-1, which was established

from a patient diagnosed with ATL 123,378. About the same time, Miyoshi et al

234,236,237established another T cell line, MT-2, from an ATL patient which was later

found to have type C retrovirus particles 377. Further sero-epidemiologic, immunologic,

genetic and molecular studies showed that the viruses in these two cell lines were the

same and helped to establish the close linkage between ATL and HTLV-1. Thus HTLV-

1 was established as the first retrovirus associated with a human disease122,377,378.

1.2 Geographical Distribution of HTLV-1

Although it is estimated that HTLV-1 infects 15 to 25 five million people worldwide, it is not uniform in its distribution103, it is in fact found in geographical

clusters208. HTLV-1 infection is endemic in southern Japan 122, the Caribbean basin 29,

Central and South America 229,298, among some Melanesian Islands in the Pacific basin 13 and in regions of central Africa 308. It is also a serious public health problem among at-

risk groups in Europe and North America and therefore a target of intervention policies

including blood donor screening57,80,84,250,252,253,310. HTLV-1 infection is also prevalent in

regions such as southern India, northern Iran and northern Australia103. The seroprevalence of HTLV-1 varies widely within endemic regions, between 0.1% and 30

% of the population, depending on age and gender103.

Recent studies indicate the presence of HTLV-1 sequences in a 1,500-year-old

Andean mummy, indicating the ancient nature of HTLV-1 infections200. Primate T

lymphotropic viruses are thought to have originated in Eastern Africa and spread 2 worldwide346. It is hypothesized that an early African lineage of the simian T

lymphotropic virus (STLV) evolved into STLV-1 in Asia and subsequently infected

humans prior to their migration to Melanesia/Australia, resulting in the origin of

"Melanesian" clad or subtype HTLV-1c, one of the 4 major widely-recognized clads of

HTLV-1 127,171,204,322. The Asian STLV-1 strains possibly re-entered Africa at a much

later stage, and multiple interspecies transmissions resulted in the establishment of the

second clad of HTLV-1, the African clad, which can be further subdivided into five sub-

clads (HTLV-1a, -b, -d, -e and –f) 347. The third clad, referred to as the "Cosmopolitan"

clad is considered to have originated from HTLV-1a, which spread throughout the world

due to the increased human host mobility, particularly during the slave trade 102,204.

Recent reports indicate the presence of subtype D in Central Africa, mainly in

Pygmies212. In addition, two other African HTLV-1 variants from Zaire (Efe1) and Gabon

(Lib2) have been discovered and may be yet considered as the sole human prototypes of

new subtypes (E and F) 344.

1.3 Transmission of HTLV-1

HTLV-1 is a highly cell-associated virus and free HTLV-1 particles have low

infectivity. Cell-cell contact through virus-infected T cells is required for efficient

transmission 373. The most efficient route of HTLV-1 transmission is by exposure to

infected blood or whole cell blood products, but not plasma (e.g., transfusion or through

145 the use of contaminated needles), with a seroconversion rate of approximately 50 %

and an estimated median time for seroconversion of 51 days 221. In fact, sharing of

3 needles among intravenous drug users constitutes one of the most common modes of blood-to-blood transmission among endemic population in the United States 153,154.

Transmission by infected blood products has also been a major problem in endemic areas, particularly in Japan 221,268. Nonetheless, most cases of HTLV-1 infections are due to

transmission from mother to child through prolonged breastfeeding for more than 6

months (through infected milk-borne lymphocytes) or by sexual contact later in life

120,121. The probability of mother-to-infant transmission is estimated to be 18–30% and

the risk factors involved are maternal factors, including higher HTLV-1 antibody titer,

prolonged ruptured membranes during delivery, and low socioeconomic status 358. The sexual route of transmission is less efficient than transmission involving blood, but has the potential to introduce infection into previously unexposed groups. Transmission from male to female through sexual contact is about four times as frequent as female to male transmission, at a rate of 4.9 per 100 person-years among females married to an infected male compared with 1.2 among males married to an infected female323. Transmission

risk to females is believed to be greater if the partner has high antibody titers or antibody

to Tax proteins while the risk of female-to-male transmission is associated with penile sores/ulcers and diagnosis of syphilis among males 251.

1.4 HTLV-1-Associated Diseases

1.4.1 Adult T cell Leukemia/Lymphoma

Typically, 1-5 % of the HTLV-1 infected individuals develop ATL after a latent

period of 20-30 years 50,371,372. ATL is an aggressive T cell malignancy with a leukemic

phase characterized by circulating, activated CD4+/CD25+ T cells343. Infection early in 4 life is associated with the development of ATL and the estimated lifetime risk is about 5

% in individuals infected before the age of 20 years 50,359. The incidence rate is 2–4 per

100,000 person-years and males have higher risk than females 50,170. The average age of

onset is 40 years in Jamaica, Trinidad, and Brazil and 60 years in Japan 50,295.

The clinical features of ATL are similar to non-Hodgkin’s lymphoma: malaise, fever, lymphoadenopathy, hepatosplenomegaly, jaundice, drowsiness, weight loss, and opportunistic infections. Typically widespread or localized skin involvement (in 40% of cases, there are large nodules, plaques, ulcers, and papular rash on the limbs, trunk, or face) and lytic bone lesions are noticed in these patients. Immunosuppression with bacterial and opportunistic infections such as Pneumocystis carinii pneumonia, fungal

infections, and strongyloidiasis has been reported in these patients, contributing to poor

prognosis 357. Unique clinical chemistry characteristics of this condition include

hypercalcemia, high serum concentrations of lactate dehydrogenase and elevated serum

levels of soluble interleukin-2 (IL-2) receptor α chain. Circulating ATL lymphocytes are

morphologically unique abnormal T cells with characteristic convoluted multi-lobulated

nuclei, which are defined as “flower cells” 317. Typical criteria to establish a diagnosis of

ATL in patients include seropositivity for HTLV-1, marked leukocytosis, characteristic

morphologic appearance of ATL cells, histology or cytology of malignant cells with a T

cell immunophenotype, usually CD4+/CD3+, hypercalcemia and increased circulating

levels of the IL-2 receptor α-chain (IL-2Rα/CD25) and elevated serum lactate

dehydrogenase (LDL) levels317. Diagnostic tools for asymptomatic carriers include an immunoassay to detect HTLV-1 antibodies and polymerase chain reaction (PCR)-based

5 amplification of lymphocyte genomic DNA to detect provirus 353. Based on restriction fragment length polymorphism (RFLP) analysis, the proviral integration into host cells is believed to be polyclonal, but the RFLP analysis of leukemic cells isolated from ATL patients show a monoclonal proviral integration pattern 353,378, suggesting that a sub- population of infected cells are clonally selected to proliferate while rest of the infected lymphocytes either die or get eliminated.

Based on clinical and laboratory criteria, ATL is classified into four subtypes: acute, lymphoma, chronic or smoldering 151,317,370. Distribution of these subtypes vary depending on the geographic location 317, 115 and the median survival for ATL patients is

4-6 months depending on the subtype of ATL220. While the acute subtype has the highest incidence and is characterized by increased ATL cells, skin lesions, systemic lymphoadenopathy, hepatosplenomegaly and poor prognosis, particularly with high level of serum LDH, calcium and bilirubin 369,369,371, the chronic subtype ATL is characterized by increased white blood cells counts, cough and skin lesions. Smouldering subtype

ATL has the best prognosis among the four subtypes and is reported to have few ATL cells in the peripheral blood over a long time period.

The development of ATL has been the focus of many investigations, but the exact mechanism is not completely understood. After infecting CD4+CD25+ T cells, the

HTLV-1 provirus is randomly integrated into the host genome, where it persists for years.

Although HTLV-1 replicates via reverse transcription in vivo, HTLV-1 replication is thought to occur via mitosis rather than via reverse transcription242. In ATL, tumor cell

6 clones were identified against a general background of oligo to polyclonal expansion of

infected but non-transformed cells40. Typically there is a progression from polyclonal to

oligoclonal and then to monoclonal proliferation in vivo, which is achieved while the

cells become IL-2 independent 85,125,378. Based on the long clinical latency and the low

percentage of individuals who develop ATL, T cell transformation is believed to occur

after a series of cellular alterations and/or mutations. Transformation of infected

lymphocytes is believed to be initiated through induction of cellular genes and alterations in critical cellular activation and death pathways by the viral transactivator Tax 97.

Although the mechanism of cellular transformation by Tax is not fully understood, it is thought to involve dysregulation of cell cycle and suppression of DNA repair pathways, leading to the gradual accumulation of mutations over time and development of ATL.

However, transformed cells often contain latent or defective proviruses 87,88,175, indicating

the possibility that viral protein expression may not be essential to maintain the

transformed phenotype.

Intriguingly, despite a strong immune response mounted against the HTLV-1, the

virus is able to persist in the host. Several possible mechanisms have been suggested to

explain this observation: Often the genome is partly deleted, resulting in defective virus

which may provide a mechanism for escape from immune surveillance 18-20. Recent

reports 14 suggest that HTLV-1 may not be latent and the virus might be far more dynamic than previously thought. Based on in vitro studies on the half-life of Tax-

expressing cells, it is hypothesized that infected cells start expressing viral proteins as

early as within one day after infection and might have a turn over rate of ~ 109 to 1011

7 infected cells per day. Cytotoxic T lymphocytes (CTLs) are thought to be highly efficient in significantly reducing the viral load in vitro and possibly in vivo, resulting in equilibrium between the rate of production and the rate of clearance of cells expressing viral proteins. Based on this proposed mechanism, cell-free virus, viral mRNA and viral protein may be often below detection limits and infectious transmission is limited, explaining the relative lack of sequence diversity 14.

1.4.2 Tropical Spastic Paraparesis / HTLV-1-Associated Myelopathy (TSP/HAM)

In the mid-1980's HTLV-1 began to be associated through epidemiologic studies as a contributing factor in chronic neurodegenerative disorders. After detecting anti-

HTLV antibodies in the serum and cerebrospinal fluid (CSF) of infected patients, HTLV-

1 was associated with the development of a progressive neurological disease called

tropical spastic paraparesis (TSP) in Martinique (French West Indies) in 1985. In 1986,

Osame et al reported similar clinical findings in HTLV-1-positive individuals in Japan

and they named the syndrome HTLV-1-associated myelopathy (HAM) 101,278.

Subsequently, in 1988, both syndromes were understood to be identical in clinical presentation and the World Health Organization (WHO) declared HAM and TSP as the same disease, now known collectively as HAM/TSP.

HAM/TSP is characterized as a progressive chronic myelopathy, with preferential damage of the thoracic spinal cord 2,95,149,167,253. In affected patients the myelopathy

presents as paraparesis with spasticity in the lower extremities, hypereflexia, muscle

8 weakness, sphincter disorders including dysfunction of the urinary bladder and intestines.

Less frequently it may precede, or give rise to, a cerebellar syndrome with ataxia and intention tremor 39. Some HAM/TSP patients exhibit a predominance of sympathetic

nervous system dysfunction 6. Magnetic resonance imaging (MRI) of HAM/TSP patients

reveals the presence of multiple white matter lesions in both the spinal cord and the brain

involving perivascular demyelination and axonal degeneration 93,104,162,163.

Within the spinal cord, regions mainly affected are the so-called ‘watershed’

zones, suggesting a nonrandom distribution 132. Lesions include degeneration of lateral

corticospinal tract and spinocerebellar or spinothalamic tract of the lateral column as well

as perivascular and parenchymal cellular infiltration, consisting of mainly T cells (both

CD4+ and CD8+ T cells), macrophages, astrocytes, and glial cells 276. Typically, cellular

destruction and inflammation is a feature of HAM/TSP. CD4+ T cells, CD8+ T cells,

and activated macrophages are present within white matter lesions of HAM/ TSP patients133. Peripheral blood and the CSF of HAM/TSP patients have high levels of

proinflammatory cytokines, such as IFN-γ, TNF-α, IL-1 and IL-6, suggesting the

importance of these proinflammatory cytokines in HAM/TSP15,150,179,227. Additionally,

CSF of these patients contains activated lymphocytes, suggesting that activated

lymphocytes and macrophages also play a role in the development of HAM/TSP133.

The lifetime risk of developing HAM/TSP among carriers is estimated to be

0.23%, based on multiple risk factors 148. The progression to HAM/TSP is influenced by

multiple risk factors such as host genetic factors254, host immune response, high HTLV-1

9 proviral load 19,20,166,264,374, route of infection (exposure via blood transfusion) 277 and perhaps by specific viral characteristics such as variations in HTLV-1 tax19,65,82,106,212,306.

HAM/TSP patients develop a vigorous expansion of CD8+ T cells 307 and their

CSF contain high levels of anti-HTLV-1 antibodies 100 and HTLV-1 Tax specific

CTLs176. These CTLs are implicated in the cellular destruction and inflammation within

the central nervous system (CNS) 176, probably by directly killing of HTLV-1 infected

cells expressing Tax, CNS cells expressing a cross-reactive cellular determinant and

neighboring uninfected neuronal cells by apoptosis mediated via the toxic inflammatory

cytokines 21. Although the pathogenesis of HAM/TSP is not completely resolved,

immune mechanism is widely accepted due to the increased cellular and humoral immune

responses in HAM/TSP patients. Autoimmune mechanism has also been proposed due to

the presence of a unique T cell receptor CDR3 motif in infiltrating lymphocytes in the

spinal cord lesions of HAM/TSP patients, as in brain lesions of MS and experimental

autoimmune encephalomyelitis patients116. Moreover, antibodies to hnRNP-A1 may

cross-react with HTLV-1-Tax, during the immune response associated with HAM/TSP,

leading to a recent theory that molecular mimicry between HTLV-1 and autoantigens in

CNS play a role in the pathogenesis of HAM/TSP 195,196.

1.4.3 Other Diseases Associated with HTLV-1-Infection

HTLV-1 infection has been linked to many other immune-mediated diseases,

although some of these remain controversial. A high incidence of idiopathic uvetis with a

high rate of anti-HTLV-1 antibodies in HTLV-1 endemic areas of Kyushu, Japan led to

the speculation that HTLV-1 might be the etiological agent 165,239. Proviral DNA was 10 detected in cells from the patients’ aqueous humor. Moreover, peripheral blood

leukocytes of these patients had higher proviral loads than HAM/TSP patients 269.

Interestingly, proviral loads showed significant positive correlations with the severity of disease among uveitis patients with a previous history of Grave's disease 269.

Immunological and virological studies further confirmed these findings 270,270,305,305.

Moreover, all patients with HTLV-1-associated uveitis had hyperthyroidism preceding the onset of uveitis, implying a potential role for hyperthyroidism in the onset of HTLV-

1-associated uveitis 238,280.

In 1990, HTLV-1 was associated with infective dermatitis, characterized by

watery nasal discharge with crusting of the anterior nares and severe exudative dermatitis

in scalp, neck, ears, axillae and groin183. Later, these infective dermatitis patients were

confirmed to be seropositive for HTLV-1, however the pathogenesis is not yet

understood 182,183,185. Interestingly, the presence of infective dermatitis in children may

predispose them to subsequent development of ATL or TSP/ HAM 184.

HTLV-1 has also been associated with cutaneous T cell lymphoma including

mycosis fungoides and Sezary syndrome, supported by the detection of HTLV-1 DNA in a minority of cutaneous T cell lymphoma patients155,213. Although cutaneous T cell lymphoma is clinically and histologically similar to ATL, its association with HTLV-1 is debated, due to the failure to detect HTLV-1 proviral DNA in patients199,362,363, HTLV-1

seronegative nature in most of these patients114,213,281 and uncertainty on whether HTLV-

1 specific sequences in these patients are intact, full-length proviral genomes or non-

functional deleted proviruses199,362,363.

11 HTLV-1 has been linked to Sjögren’s syndrome based on epidemiological evidence in southwest Japan, a high endemic region of HTLV-1 infection. Sjögren's syndrome is a chronic inflammatory disorder characterized by lacrimal and salivary gland insufficiency, due to high HTLV-1 337. These patients also had high serum titers of

HTLV-1 antibodies 223, but low viral load and only tax sequences were detected in their labial salivary glands222 222,325. The pathogenesis and the causative role of HTLV-1 in

Sjögren's syndrome remains a topic of debate.

HTLV-1 has also been linked to HTLV-1 associated arthropathy 263. However, whether HTLV-1 associated arthropathy represents a separate clinical entity remains controversial. HTLV-1 infection is also associated with a variety of additional autoimmune disorders such as polymyositis 194, T lymphocyte alveolitis324, lymphadenitis

267, chronic respiratory disease159, acute myeloid leukemia 367 and conditions such as systemic sarcoidosis as well as Strongyloides stercoralis infection 368. Since the evidence linking these diseases to HTLV-1 infection are not as strong as those for ATL and

HAM/TSP, the etiological role of HTLV-1 in these diseases remains controversial.

Nonetheless, the association of HTLV-1 to an array of clinical disorders indicates the huge public-health impact of this human pathogen.

1.5 Genetic Organization and Replication of HTLV-1

Based on its morphologic appearance, HTLV-1 is a C-type retroviral particle, which is typically assembled directly at the cell membrane and does not mature until after budding from the infected cell 51. It is a complex retrovirus and a member of the 12 deltaretrovirus group, which includes bovine leukemia virus (BLV) and STLV. Unlike

simple retroviruses, complex retroviruses carry additional genes to encode for several

regulatory and accessory proteins. Mature HTLV-1 particles are spherical and have a

diameter about 110 to 140 nm with an outer envelope composed of both viral and cellular

proteins and lipids 51. The core of the virion is composed of the ribonucleoprotein

complex consisting of two copies of an identical ~9 kb RNA genome that contains the m7Gppp capping group, the primer tRNAPro of host cell origin, coding regions and 3’

poly(A) sequences51 with structural and enzymatic proteins, including nucleocapsid,

capsid, matrix, reverse transcriptase, integrase and protease.

The HTLV-1 life cycle starts when viral particles attach to the cell surface

through an interaction between envelope and HTLV-1 receptors, (which will be discussed

in later section) which are not yet fully characterized. After the outer membrane of the

viral particles fuse with cell membrane, the contents of virions enter the cytoplasm and

the viral genomic RNA is reverse transcribed to generate double-stranded DNA (dsDNA)

using the virally encoded reverse transcriptase (RNA-dependent/ DNA-dependent

polymerase). Viral dsDNA traffics to the nucleus and randomly integrates into the host

genome using the viral encoded integrase (the genome is now referred to as the

provirus)108. Host cellular machinery drives transcription of the proviral DNA using host cell RNA polymerase II to produce viral mRNA. The unspliced full-length (~9 kb) viral genomic RNA is used to generate progeny virus particles (virions) and to generate Gag

(internal group specific antigen), Pol (polymerase, reverse transcriptase and integrase

enzymes needed for viral replication and maturation) and Pro (protease enzyme required

13 for maturation of structural and enzymatic proteins) proteins while the singly spliced

subgenomic ~4.3 kb mRNA is used to generate Env protein (envelope glycoprotein) and doubly-spliced (two introns removed) RNA (~ 2.1 kb) is used to translate the regulatory proteins Tax (open reading frame (ORF) IV) and Rex (ORF III). Immature virus particles are assembled at the plasma membrane and become mature only after budding

52. Additionally, by alternatively splicing, several accessory proteins are also produced

from pX ORF I and II, which will be discussed further in the following sections 24.

HTLV-1 proviral genome is 9032 nucleotide long, 312 has 5’ and 3’ long terminal repeats (LTR). The LTR is 755 nucleotides long and contain cis-acting regulatory sequences necessary for viral gene expression and replication, including sequences essential in transcription start and termination, splicing, mRNA polyadenylation, strand transfer during reverse transcription and RNA transport 24,47,173. LTR is divided into three

regions, termed U3, R and U5 110. The U3 region contains the viral transcriptional

and promoter elements 316, including three imperfect 21-base-pair repeats,

known as Tax-responsive elements-1 (TRE-1). The TRE-1 in 5’ LTR of HTLV-1

proviral genome at positions –251 to –231 (promoter distal repeat, pd), -203 to –183

(promoter central repeat, pc) and –103 to –83 (promoter proximal repeat, pp) relative to

the transcription start site 107,339,355,356, and is necessary for

transactivation34,107,164,339,355,356. The TRE-1 regions contain a conserved TGACG

sequence, with consensus to the TGACG cyclic adenosine monophosphate (cAMP)

response element (CRE), present in the promoters of several cellular genes. Thus, TRE-1

elements are in fact defective/ imperfect CREs, which bind multiple transcription factors

14 such as cyclic AMP response element binding protein (CREB)326,382, cAMP response

element modulator (CREM) 326, activating transcription factors (ATFs) 111 and other

cellular proteins that facilitate transcription including TRE binding proteins (TREB-1 335,

TREB-5 379, TREB-7 379, c-Jun 135, c-Fos 91, activator protein-2 (AP-2) 244,266, HTLV enhancer binding protein (HEB)-1p67 and HEB-1p94 205. Between the pc and pp repeats

is another second enhancer element, namely, Tax-responsive element 2 (TRE-2), also

critical in viral transcription 224,336. In addition, many cellular proteins such as Ets family proteins (Ets-1, Ets-2, Elf-1 and TIF-1) and c-Myb bind to TRE-232,33,49,53,340. Based on sequence analysis, NF-κB, AP-2 and PU.1 are also thought to potentially interact with

TRE-2 69. Additionally, inducible cAMP early repressor protein (ICER) binds TRE-1 and/or TRE-2 and decreases transcription from the LTR. Newbound et al257 reported that

Tax-mediated transcription can be repressed by ICER protein in peripheral blood

mononuclear cells, with possible significance in decreasing viral gene expression, leading

to immune evasion and subsequent viral persistence. Tax mediated transactivation

through TRE is discussed further in the following sections. In addition, many cellular

proteins such as Sp-1 266 and nuclear factor-1 (NF-1)266 bind the LTR at regions outside

TRE and enhance transcription. Moreover, LTR also contains the Rex-responsive

element (RxRE). RxRE secondary structure directly interacts with Rex and is essential

for proper polyadenylation of HTLV-1 transcripts and Rex-mediated cytoplasmic

expression of incompletely-spliced RNAs. R and U5 provide the 1-119 leader sequence

for HTLV-1 transcripts 3,4.

15 1.6 Structural Proteins of HTLV-1

HTLV-1 uses ribosomal frame shifting and alternative splicing mechanisms to

produce an array of viral proteins from the same coding region 51. Ribosomal frame

shifting is the predominant mechanism employed to generate the structural and enzymatic

proteins and precursor proteins. First, the gag gene encoding the viral structural proteins is translated from an unspliced, full-length mRNA as a single precursor polyprotein p55(Gag) 275, which is myristoylated at its N-terminus, targeted to inner plasma

membrane 271 and later gets proteolytically cleaved by viral protease into three final

products of 19 kDa (matrix), 24 kDa (capsid) and 15 kDa (nucleocapsid) proteins 119,275.

The p19 matrix protein remains myristoylated and is localized to the inner surface of the viral envelope 256. HTLV-1 Gag protein alone has been found to be adequate for the

assembly and release of virus particles 188. The p19 matrix protein (MA) contains eleven

basic amino acids, critical in the release of HTLV-1 virus particles and cell-to cell

transmission 189.

The pol gene message is translated into the Gag/Pol precursor from the same

genomic mRNA as the Gag precursor, but in a different reading frame through a

ribosomal frame shift 255. This precursor is available in lower levels than the Gag

precursor and is subsequently cleaved by the protease to produce reverse transcriptase

(RT), integrase (IN) and the remaining structural Gag proteins 256. Pol also has RNase H

activity that is required during the process of reverse transcription 52.

16 Protease (PR) is encoded by a reading frame extending between 3' part of gag and

5' part of pol and is made by another frame shift256. Due to the absence of a methionine

codon at the 5’ end of the pro ORF, an mRNA processing event or translational frame

shifting is thought to occur prior to the synthesis of protease 256. In addition, protease is

synthesized in an immature form that subsequently undergoes self-cleavage to generate

the active form 110.

The viral envelope glycoprotein (Env) is synthesized from a ~4.3 kb singly-

spliced mRNA transcript as a gp61 envelope precursor in the endoplasmic reticulum

(ER), gets heavily glycosylated and cleaved into the mature surface protein (SU) gp46

and a transmembrane (TM) gp21 protein 289. SU and TM proteins form a heterodimer necessary for receptor binding and the fusion events involving the virus and host cell109.

The SU protein contains two functionally well-defined regions71 while the TM protein contains a leucine-zipper like region required for fusion169,304 and the region between C

terminus of TM ectodomain to the end of the cytoplasmic domain critical for viral

transmission, likely for post-fusion steps in viral infection72,139. In addition, TM protein

is recognized by neutralizing antibodies75. Although Env has been extensively

investigated, the identity of HTLV-1 receptor that facilitates its spread has remained

elusive until recently. Transferrin receptor, an inducible receptor which plays a crucial

role in the regulation of iron uptake and cell growth, is overexpressed in ATL cells285 and

a neutralizing monoclonal antibody directed against transferrin receptor induced

apoptosis of ATL cells in patients243. More importantly, Manel et al recently identified

GLUT-1, a ubiquitous glucose transport protein, as a receptor for HTLV-1215. It is not

17 clear how interactions of the HTLV-1 envelope with GLUT-1 affect cell-to-cell spread of

the virus, however if GLUT-1 is essential for the spread of both cell-free and cell-

associated HTLV-1, further characterization of this receptor may be valuable in

understanding of the complexities of HTLV-1 pathogenesis and in designing strategies to

prevent HTLV-1 transmission279.

1.7 Regulatory Proteins of HTLV-1: Tax and Rex

As a complex retrovirus, HTLV-1 contains regulatory genes in pX region, located

between env and 3’ LTR. By means of alternative splicing, utilizing sequences within the

5’ LTR, pol and pX region and internal initiation codons, virus produces regulatory and

accessory proteins from four open reading frames in the pX region. ORF IV and ORF III

encode regulatory proteins Tax and Rex respectively 168,313 while ORF I and ORF II encode four accessory proteins, p12I, p27I, p30II and p13II 24,47,94,173,41(Fig. 1.1).

1.7.1 Tax and Transactivation of the HTLV-1 LTR

Tax, (Transcriptional Activator of pX region) is a 40 kd nuclear protein, produced

from a doubly spliced ~2.1 kb mRNA, and is known as the viral transactivator, which

regulates viral gene expression from the HTLV-1 LTR 37,83,92,320,321. As discussed in a

previous section, HTLV-1 gene expression is dependent on three 21-bp imperfectly

repeated elements, termed TRE-1 34,107,164,339,355,356. Among the 3 repeats within TRE-1, at least two have been demonstrated to be necessary for Tax transactivation of LTR. Tax

18 does not directly bind to TRE-1 or TRE-2 sequences, but by protein-protein interactions,

Tax increases the DNA-binding activity of CREB/ATF proteins to the TRE-1 and several

basic leucine zipper (bZIP) proteins to the TRE-2 282. Interaction of Tax with CREB and

serum response factor (SRF) enhances the stability of the transcription enhancers - DNA complexes. However, Tax transactivation of the LTR requires direct binding of Tax to

GC-rich sequences flanking the TRE-1, at a region within the CREB binding protein

(CBP/p300) binding domain 160,161,193,207. Once cells are stimulated, protein kinase A

(PKA) phosphorylate CREB, which then binds CBP/p300 45,105,180. Tax recruits

CBP/p300 to the DNA bound CREB, allowing CBP/p300 to further associate with DNA

and facilitate transcription. By binding both CREB and CBP, Tax acts as a bridge

between these two proteins without the need for CREB phosphorylation, enabling the

virus to maintain constitutive activation (without CREB phosphorylation) through TRE-1 in HTLV-1 infected cells 25,333. However, CBP/p300 recruitment by Tax is not sufficient for transactivation 118 and requires CBP/p300 associated factor (P/CAF), which associates with the C-terminal activation domain of Tax 117,137.

1.7.2 Tax and Regulation of Cellular Gene Expression

Tax regulates the gene expression of numerous cellular genes as well, mostly by induction of the transcription factors NF-κB and SRF, independent of CREB activation.

An increasing list of cellular genes activated by Tax include cytokines such as IL-2 226,

IL-3 361, IL-4202, IL-6 249, IL-8 240, IL-2Rα 226, IL-1 241, GM-CSF 233, TNFα 58 and TNFβ

203, transcription factors such as c-myc 79, c-fos 89, c-sis 297, erg-1 90, c-rel 198, as well as

19 261,342 genes involve in apoptosis Bcl-xL and DNA repair (proliferating cell nuclear antigen or PCNA) 300. Tax has also been shown to repress expression of β-polymerase

(an enzyme involved in DNA repair), lck,and bax 35,136,192.

ATL cells have characteristic IL-2Rα overexpression 338, which is believed to be

induced by Tax-mediated NF-κB activation 128. Tax directly binds multiple NF-κB

family proteins such as p50, p65, c-Rel 327,329 and lyt10 248. Since NF-κB proteins and

Tax both bind CBP/p300 99,283,383, it is possible that Tax links NF-κB and CBP/p300 to

enhance the stability of the DNA-protein complex 25. In addition, Tax localizes in the

cytoplasm and associates with IκBα, the inhibitor of NF-κB, destabilizes IκBα, thereby

allowing NF-κB to translocate to the nucleus210,328. Moreover, Tax interacts with and

induces the activation of cytokine-inducible IκB kinase (IKK), which contains IKKα,

IKKβ and IKKγ (NEMO) subunit, resulting in dephosphorylation and degradation of

IκBα and subsequent NF-κB activation 201. Additionally, Tax activates cellular gene promoters controlled by the serum response element (SRE), such as c-fos. Similar to activation of CREB, Tax induces dimerization of the serum response factor (SRF) protein p67SRE in a complex with the SRE 10,90.

Tax has a significant effect on cell cycle regulation, by means of trans-repression

of multiple cellular genes including p18INK4c, and Bax 35. By sequestering

CBP/p300, Tax downregulates the promoter activity of p18INK4c, an inhibitor of CDK4

that can arrest the cells at G1 phase of the cell cycle 181,333. Although similar mechanism

has been implicated in Tax-mediated repression of p53 12,345, it is not consistent with the

20 elevated p53 levels detected in HTLV-1 transformed 96,299,375 and Tax immortalized cell

lines 5. This led to the hypothesis that Tax does not impair the expression of p53, instead

impairs its function, supported by the detection of hyperphosphorylated and inactive p53

in HTLV-1 infected cells 247,292. In addition, Mulloy et al 247found CREB binding

domain of Tax to cause Tax-mediated p53 repression while Pise-Masison et al 291 suggested an NF-κB dependent pathway for p53 inactivation by Tax. In addition, Tax associates with other cell cycle regulators, such as p16INK4a 330 and p15INK4b 144,331, negatively influences their inhibitive functions on cyclin dependent kinase 4/6 (CDK4/6), leading to Rb phosphorylation and induction of G1/S phase transition. Tax also binds to

HsMAD1, a component of mitotic checkpoint that prevents anaphase until chromosomal alignment is completed and may abrogate the function of HsMAD1 to facilitate cell division 138. Furthermore, Tax expression leads to accumulation of DNA damage

expression 232,286, by inhibition of topoisomerase I 332 and DNA polymerase β 136.

Overall, Tax suppresses DNA repair, enhances DNA replication and cell-cycle progression, thereby increase cellular mutation frequency and contribute to HTLV-1- mediated cellular transformation 97.

1.7.3 HTLV-1 Rex: Post-transcriptional Regulator

Rex, a nucleolar localizing 27KDa protein encoded by pX ORF III, is responsible for viral post-transcriptional regulation. Rex facilitates transport of unspliced RNA

(gag/pol) and singly spliced RNA (env) from nucleus to cytoplasm and inhibits the

splicing, transport of doubly spliced RNAs that encode the regulatory proteins in the pX region, which include Rex itself. Overall, Rex is necessary for viral replication 129,130,168,

21 by regulating the balance between messages encoding for viral structural proteins and

messages encoding regulatory / accessory proteins. Rex functions by binding to the Rex response element (RxRE), a highly stable RNA stem-loop structure in the U3/R region of the HTLV-1 3’LTR 17,22,26,27,156,314, critical in polyadenylation of viral RNAs 3,73. RxRE

is present in doubly spliced, singly spliced and unspliced messages, and thus other cis-

acting sequences possibly determine the selectivity of Rex on various messages. In fact, cis-acting repressive sequence (CRS) in the U5 region of the viral LTR negatively regulate Rex 28. Additionally, Rex may inhibit mRNA splicing by altering early

splicosome assembly and stabilizing the mRNA of cellular gene, IL-2Rα chain by acting

in trans on its coding sequence 147.

Rex contains both an N-terminal located NLS responsible for nuclear localization

and RxRE binding 265,318, as well as a nuclear export sequence (NES) located in the

middle, with similarity to human immunodeficiency virus (HIV) Rev, influenza virus

NS2, herpes simplex virus type 1 (HSV-1) ICP27 protein 61 and cellular protein IκBα 140.

The molecular mechanism of Rex-mediated RNA export is not completely elucidated, but

current understanding involves the CRM1/exportin1 pathway 31,112, which is also

employed by HIV Rev to facilitate viral RNA transport 61,294. Interestingly, Rex can

functionally substitute for HIV Rev 131,301. Rex and Rev directly bind both the viral

mRNA and hCRM1 in the presence of RanGTP via their RNA binding domains and NES,

respectively, acting as adaptor proteins that bridge the target mRNA and the export

receptor. Interestingly, multimerization of Rex/Rev along the viral mRNA is necessary

for RNA export67,209,354, possibly because multimerization allows a number of Rex/Rev

22 proteins to shield RNA, preventing the attachment of factors involved in splicing and nuclear retention of RNA and thus increasing the number of CRM1 molecules that associate with the complex. The association of multiple CRM1 molecules may be important in overcoming the factors that retain RNA in the nucleus and allow RNA to pass efficiently through NPCs. Human CRM1 (hCRM1) functions in the Rex-mediated mRNA export of HTLV-1 as an export receptor and as an inducing factor for Rex multimerization on its cognate RNA113. Recently, Ye et al376 reported that Rex is critical for efficient infection of cells and persistence in vivo, but its function to modulate viral gene expression and virion production is not required for in vitro immortalization by

HTLV-1.

Interestingly, a variety of HTLV-1-infected cell lines including MT-1, MT-2, TL-

Su, H582, HUT-102 contain mRNA and protein for a truncated form of Rex (p21Rex)

41,94,272, which can be generated from both the doubly-spliced message encoding Tax and

Rex, as well as the message singly spliced from nt 119 directly to the splice acceptor at nt

6875 and only encoding p21Rex. The second type of message is highly expressed in primary cells isolated from asyptomatic HTLV-1 carriers, ATL and HAM/TSP patients

94,274. This protein contains only the NES sequence and not the NLS sequence and when coexpressed with the full-length Rex, it repressed the shuttle function of the full-length protein and increased the nuclear accumulation of full-length protein 177. HTLV-2 and bovine leukemia virus (BLV) have p21Rex homolog mRNAs 273, however the exact function of p21Rex in viral pathogenesis remains to be elucidated.

23 1.8 HTLV-1 Accessory Proteins

The HTLV-1 pX genome region includes ORF I and II that produce alternatively

spliced forms of mRNA, which encode four accessory proteins, p12I, p27I, p13II, and p30II (Fig. 1.1)24,46,172,173. pX ORF I mRNA is produced by alternative splicing events

that combine the second exon of Tax with additional downstream sequences and encodes

p27I (152 amino acids long) and p12I (99 amino acid long). p12I can be translated from a

singly spliced message produced by direct splicing of nucleotide 119 to the splice

acceptor at position 6383 or by initiation at an internal methionine codon in the p27I ORF

I47. p12I is thought to be preferentially expressed from the p27I mRNA since transfection

of expression plasmids containing HA1-tagged versions of either the full-length p27I

cDNA or the cDNA for the singly spliced p12I produced only the smaller p12I protein173.

However, using in vitro transcription-translation systems, Ciminale et al 47 produced p27I

from the doubly spliced mRNA. Therefore, removal of the internal p12I AUG start codon

could yield detectable levels of p27I. Interestingly, Pique et al 290 demonstrated

production of cytotoxic T cells in HTLV-1-infected subjects that were reactive against

peptides representing all putative pX accessory proteins, including p27I.

The accessory proteins encoded by pX ORF II of HTLV-1 are produced from two

alternatively spliced mRNAs. The larger protein, p30II, is encoded by a doubly spliced

message including the first and second exon of Tax spliced to the splice acceptor site at position 6478 24. p13II, the smaller protein contains the C-terminal 87 amino acids of

p30II and is produced from a singly spliced message by splicing of the first Tax exon

24 directly to the splice acceptor at position 6875 or translated from an internal methionine

codon within p30II24,173. HTLV-1 accessory proteins were originally thought to be

dispensable for viral replication 74. However, recent investigations performed by our laboratory and others have shed light into the role of the HTLV-1 accessory proteins in viral infectivity, maintenance of high viral loads, host cell activation, and regulation of gene transcription 7,8,23,48,55,64,70,76,141,157,186,341,380,381.There is evidence that pX ORFs I and

II mRNAs and proteins are expressed in infected cell lines and freshly isolated cells from

asymptomatic carriers, ATL and HAM/TSP patients 173,41. In addition, HTLV-1 infected

patients and asymptomatic carriers have antibodies 43,70 and cytotoxic T cells290 against recombinant proteins or peptides of the pX ORF I and II proteins. However, to date, there are no conclusive reports on the temporal expression of Tax, Rex and the four accessory proteins messages in terms of their specific quantities or relative levels.

Chronically infected cell lines were found to have pX-tax/rex mRNA at 500-fold to 2500- fold higher levels than pX tax-ORF II mRNA and 1000-fold higher levels than pX-rex-

ORF I mRNA296.

Interestingly, analogous gene regions encoding the accessory proteins, especially

the pX ORF I-encoded p12I, are highly conserved in the closely related virus HTLV-2

and the nonhuman primate counterpart, simian T cell lymphotropic virus type 1 (STLV-

1)9. Further illustration of the conserved nature of these gene regions comes from studies

of another member of the deltaretroviruses, bovine leukemia virus (BLV). BLV, like

HTLV-1, contains an X region between the env sequences and the 3’ long terminal

repeat. Two proteins are expressed from this region of BLV: the protein R3, which

25 shares a common NLS with the Rex protein of HTLV-1, and G4, an arginine rich protein

that may exist as two isoforms following protease processing 152,191. Similar to HTLV-1,

deletion of homologous sequences from BLV infectious molecular clones encoding these

accessory proteins, R4 and G3, results in decreased viral loads in the experimental sheep

model 152,190,360. Collectively these studies illustrate that these retroviruses, which are all

associated with lymphoproliferative disorders, during the course of their evolution have

retained conserved gene regions that apparently serve analogous functional roles.

1.8.1 pX ORF I p12I

1.8.1.1 Biochemical Characteristics of p12I: Features of a Signaling Molecule

HTLV-1 p12I is a highly hydrophobic protein, which contains a significant

percentage of leucine (32%) and proline (17%) residues173, a minimal number of soluble

regions and two putative transmembrane domains (aa 12 to 30 and aa 48 to 67)77,341, which overlap with two predicted leucine zipper motifs that form alpha-helices. These distinct structural features (Fig. 1.2) could contribute to membrane localization and dimerization341. Additionally, p12I contains four predicted SH3-binding motifs (Fig.

1.2), which are typically proline rich cellular signaling proteins with a minimal core of

PXXP 38. Interestingly, aa 8 to 11 and 70 to 74, encoding the first and third PXXP motifs

of p12I are highly conserved among viral strains, suggesting a possible role for these domains in the function of p12I. In addition, there is a conserved sequence (PSLP(I/L)T)

extending from aa 48 to 99 in p12I, highly homologous to the PXIXIT calcineurin-binding

motif of nuclear factor of activated T cells (NFAT) and found to be critical for the

26 interaction between p12I and calcineurin157. In addition, p12I contains a dileucine motif

(DXXXLL) at amino acid positions 26 to 31, with no functional role ascertained yet (Fig.

1.2). However, dileucine motifs have been described in viral proteins such as HIV Nef and are commonly involved in directing protein trafficking through endosomal compartments by mediating association of the protein with adapter protein 1 (AP-1) to

AP-360. Sequence analysis of p12I suggests possible post-translational modifications of

the protein due to the presence of potential phosphorylation sites, glycosylation sites and

a ubiquitynation motif surrounding the lysine at position 88, which might significantly

enhance the half-life of the protein341are also present in p12I. However, p12I was

determined to be not a glycoprotein or a phosphoprotein 77.

1.8.1.2 Subcellular Localization of p12I and Interactions with Cellular Proteins

HTLV-1 p12I was originally reported to localize in cellular endomembranes,

which was suggested by a spider-like staining of cells expressing the viral protein 172.

Subsequent immunofluorescent confocal microscopy, electron microscopy and subcellular fractionation studies confirmed that p12I accumulates in the ER and cis-Golgi

apparatus. In addition, p12I directly binds calreticulin and calnexin, two resident ER-

proteins which regulate calcium storage and calcium-mediated cell signaling, suggesting a role in the establishment of HTLV-1 infection by activation of host cells 77.

HTLV-1 p12I shares sequence homology with the bovine papilloma virus (BPV)

E5 protein and Epstein-Barr virus (EBV) LMP-1 protein 86. The region of highest

27 homology starts after the first transmembrane domain and extends into the second. Like

E5, p12I also binds to the 16 kD subunit of the vacuolar H+-ATPase (16K). Although this

association appears to be required for the E5-mediated transformation of epithelial

cells,86,174 additional studies are necessary to determine its functional significance.

Interestingly, Nef, a key accessory protein of simian immunodeficiency virus (SIV) and

HIV, binds the catalytic subunit NBP-1 of the ATPase206. NBP-1 association of Nef

mediated by the Nef C-terminal flexible loop is critical for Nef-dependent internalization

of CD4 and viral infectivity 214.

Analogous to E5, p12I also associates with immature form of IL-2 receptor β and

γ chain when transiently co-expressed, resulting in reduced surface expression of the

receptor chains246. The p12I-binding site on the IL-2R chain overlaps with the binding

site for Janus kinases, JAK 1 and 3 and the adapter protein Shc. p12I does not influence

JAK3 kinase activity directly, however it is considered to be accountable for the modest

increase in STAT5 DNA binding activity262. As a consequence, p12I-expressing cells

displayed a decreased requirement for IL-2 to induce proliferation during suboptimal

stimulation with anti-CD3 and anti-CD28 antibodies 262. However, peripheral blood

derived lymphocyte cell lines immortalized by a HTLV-1 proviral clone ablated for pX

ORF I expression, ACH.p12I, have intact IL-2 receptor signaling pathways 54. A possible

explanation for these conflicting observations is that p12I may modestly activate IL-2 receptor pathways during the early stages of HTLV-1 infection before immortalization.

Nevertheless, p12I does not appear to be necessary for the activation of the IL-2R-

associated JAK1 and JAK3, or their downstream effectors STAT3 and STAT5, after

28 immortalization. Collectively, these studies indicate that p12I may induce STAT activity

to confer a growth advantage to infected cells during the early stages of infection, before

immortalization. Future studies are necessary to elucidate the JAK3-independent

pathway p12I uses to induce STAT5 activation.

HTLV-1 p12I was reported to associate with immature forms of the major

histocompatibility complex class-I (MHC-I), interfere with the interaction of MHC-I with

β2-microglobulin, decrease the surface expression of MHC-I and direct its degradation in the proteasome 143,143, suggesting that p12I might help the virus escape immune surveillance. However, levels of MHC-I and II were similar between T-lymphocytes

immortalized with the wild type and p12I-mutant HTLV-1 molecular clones (ACH and

ACH.p12 respectively), indicating that p12I-mediated modulation of MHC-I surface

expression likely occurs only during the early stages of infection54. Intriguingly, the

accessory proteins p10I and p11V of HTLV-2 also associate with MHC-I, however these

do not bind to either 16K or IL-2Rβ or γ142. Additionally, HIV-1 Nef also binds to and

down regulates the cell surface expression of MHC-I and is believed to contribute to

immune evasion by HIV-1287. However, down regulation of MHC-I of virus-infected

cells does not explain the early loss of infectivity of a molecular clone of HTLV-1 that

lacks ORF I expression, as virus infection is blocked as early as 1 week post-inoculation, before a possible active immune response occurs55. Future studies of early virus

replication immediately after inoculation of virus infected cells in animal models might

provide evidence to whether p12I indeed down regulates MHC-I expression on infected peripheral blood mononuclear cells (PBMC) in vivo and actively contributes to viral spread or persistence. 29 In addition, p12I associates with two ER-resident calcium-binding proteins, calreticulin and calnexin77 and the calcium/calmodulin-dependent serine/threonine

phosphatase, calcineurin 157. Calreticulin, a highly conserved and ubiquitous protein,

serves as one of the major calcium-binding proteins in the ER, participates in calcium

signaling, and has been linked to activation of the NFAT230. Through

protein-protein interactions, calcineurin regulates transcriptional activation of NFAT by

triggering the dephosphorylation and subsequent nuclear translocation of NFAT, which

results in transactivation of NFAT-inducible cytokine genes including IL-2178,228,385. It would be advantageous for a virus to target such conserved proteins to dysregulate calcium signaling pathways to activate and modulate NFAT in infected T lymphocytes.

1.8.1.3 Role of p12I in Regulation of Viral Infectivity In vitro and In vivo

Earlier reports showed that deletion of pX ORF I from HTLV-1 infectious molecular clones did not affect the viral infectivity and primary lymphocyte transformation mediated by HTLV-1 infection in vitro74,302. In contrast, our research

group demonstrated that selective elimination of pX ORF I from the molecular clone

ACH resulted in dramatically reduced viral infectivity in vivo 55. Rabbits inoculated with

ACH.p12, harboring selective mutations that abolish the expression of pX ORF I mRNA,

failed to establish persistent infection as indicated by reduced anti-HTLV-1 antibody

responses, failure to demonstrate viral p19 antigen production in PBMC cultures, and

only transient detection of provirus by PCR55. A major difference between these in vitro

and in vivo studies is that standard in vitro co-culture techniques used to transmit virus to

30 naive PBMC utilize target cells stimulated by IL-2 and mitogen. However, in vivo the

majority of circulating and tissue-associated lymphocytes are non-dividing.

In support of these findings, using co-culture assays that would allow

transmission of the virus to resting primary lymphocytes, our group demonstrated that pX

ORF I mRNA is critical for viral infectivity in non-activated/quiescent PBMCs in vitro7.

More importantly, viral infectivity was restored upon addition of IL-2 and mitogen to the co-cultured PBMCs7. These data provided the first evidence that HTLV-1 p12I is

required for optimal viral infectivity in nondividing primary lymphocytes and suggested a

role of p12I in T-lymphocyte activation and in the early stage of viral infection.

Analogously, studies of HIV-1 Nef indicate that the accessory protein is dispensable for

transmission of the virus to activated target cells in vitro but is required for viral

infectivity in nondividing lymphocytes219,284,364.

1.8.1.4 Role of p12I in Calcium-Mediated T Cell Activation

Reports from our laboratory illustrated that p12I expression in Jurkat cells results

in ~ 20-fold activation of NFAT dependent gene expression, while AP-1 or NF-κB-

mediated transcription remained unchanged8. p12I specifically activates NFAT in

synergy with Ras/MAP kinase activation stimulated by the phorbol ester, PMA. By inhibition of proximal signaling molecules, such as phospholipase C- γ (PLC-γ) and LAT,

as well as distal signals, including calcineurin and NFAT, the function of p12I was

mapped to be between PLC-γ and calcineurin8. p12I mediated NFAT activation was dependent on cytosolic calcium since this function was abolished by BAPTA-AM

31 treatment. Importantly, p12I functionally substituted for thapsigargin, which specifically

depletes the ER calcium store by blocking the ER calcium ATPase. Therefore, HTLV-1

p12I was found to activate NFAT-mediated transcription in lymphoid cells in a calcium- dependent manner. Indeed, p12I expression increases the base-line cytoplasmic calcium concentration, possibly secondary to reduced ER stores calcium content and subsequently

76 higher extracellular calcium entry . Both the calcium channel on ER membranes, IP3R,

and the calcium channels on plasma membranes, calcium release activated calcium

channels (CRAC), contribute to the p12I mediated NFAT activation, strongly indicating

the modulation role of p12I on calcium homeostasis. Induction of calcium release from the ER by p12I to activate NFAT 76, would be an advantage for the virus during the early stages of HTLV-1 infection. Interestingly, this is similar to the function of a cellular

protein CAML (Ca2+-modulating cyclophilin ligand), which also contain two putative

transmembrane domains like p12I, colocalizes with calreticulin in the ER, induces

calcium release from the ER and leads to NFAT activation 351. In addition, p12I mediated increase in NFAT activity could cause complete activation of cellular stimuli that would normally induce only partial activation of T cells (e.g., through AP-1). These stimuli could be triggered by cytokines or chemokines released from infected neighboring cells

or by direct contact between viral envelope proteins and certain cell surface receptors on

newly targeted lymphocytes prior to viral entry 16. Although localization of p12I to the

ER appears to be essential for NFAT activation, 78, direct binding between calreticulin

and p12I does not correlate with the NFAT activation76 and future studies are necessary to

identify the biological significance of p12I - calreticulin interaction in HTLV-1 infection

and pathogenesis.

32 Expression of NFAT induces a highly permissive state to overcome the blockade

at reverse transcription and permitted HIV replication in primary CD4+ T cells, therefore it is possible that p12I causes T cells to be hypersensitive to T cell receptor and CD28

stimulation and thus highly permissive for subsequent viral infection. Interestingly, susceptibility of these cells to HIV infection could be restored by mitogen treatment,

likely due to the phytohemagglutinin-induced upregulation of NFAT activity. This is

similar to earlier reports from our laboratory that addition of mitogens can rescue the

infectivity of a p12I mutant viral clone in resting PBMC7, likely by overriding the

requirement for p12I-induced activation of NFAT.

Other retroviruses also encode proteins regulating Ca2+- related signals by

analogous or different mechanisms in T lymphocytes or other cell types. HIV accessory

protein, Nef, also activates NFAT in synergy with Ras/MAPK pathway in a calcium dependent fashion217-219 and is dispensable for viral infection of activated target cells in vitro, but required for viral infectivity in quiescent lymphocytes216,352,364,384. Importantly

p12I appears to enhance the production of a downstream gene of NFAT activation, IL-2 in Jurkat T cells and primary lymphocytes, which could be abrogated by calcium chelator

BAPTA-AM and calcineurin inhibitor, cyclosporin A, suggesting the effect is calcium pathway-dependent78. This increase in IL-2 could account for the decrease in

requirement for the cytokine in proliferation of human primary lymphocytes in the

presence of p12I262. Overall, p12I expression promotes T cell activation and likely

facilitates the viral replication and productive infection, which correlates with our

previous finding that p12I is necessary for viral infectivity in a rabbit model of

infection55. 33 Recently, p12I was found to contain two positive (aa 33-47, aa 87-99) and two

negative (aa 1-14, aa 70-86) regions containing individual PXXP motif, that regulate the

NFAT activation78. In a recent report from our laboratory, proline residues in these

motifs were mutated into alanine residues to test the role of PXXP motifs in p12I- mediated NFAT activation. Interestingly, the third SH3 binding domain (aa 70-73) was responsible for the negative effect of region aa 70-86 on NFAT activation 78,157. Besides, mutations in the first two PXXP motifs (aa 8-11 and 35-38) caused only minimal changes in NFAT activation, while mutations in the last PXXP motif (aa 90-93) in the second positive region (aa 86-99), enhanced the NFAT activation. In addition, p12I competes

with NFAT for calcineurin binding. More strikingly, alanine substitution mutations in

PSLP(I/L) caused more NFAT nuclear translocation and increased NFAT transcriptional activity (~2-fold) than wild type p12I 157. Interestingly, PSLP(I/L)T calcineurin binding

site is within the aa 70–86 negative region and third PXXP motif which was responsible

for the negative effect of region aa 70-86 on NFAT activation. Additionally, NFAT-

inhibitory function of the PSLP(I/L)T motif in p12I was verified to be not from the

inhibition of calcineurin phosphatase activity. PSLP(I/L)T is homologous to the

conserved functionally critical calcineurin-binding motif (PXIXIT) in the N-terminal

regulatory domain of NFAT, which binds both inactivated and activated calcineurin 11.

Many calcineurin-binding proteins such as the anti-apoptotic protein Bcl-2, calcineurin B homologous protein, a kinase anchoring protein AKAP79, and myocyte-enriched calcineurin-interacting protein 1 inhibit either calcineurin phosphatase activity or its substrate NFAT transcriptional activity59,197,350. However, PXIXIT motif mediated binding itself does not inhibit the catalytic activity of calcineurin, because NFAT

34 activation requires enzymatic activity of calcineurin as well as binding via this motif.

Not surprisingly, p12I binding to calcineurin via a motif similar to PXIXIT did not inhibit calcineurin catalytic activity but instead influenced NFAT and calcineurin interaction by competing for binding with NFAT similar to artificial peptides representing this motif 11.

Due to the existence of a calcineurin-binding motif in p12I, it may have at least

two regulatory actions to modulate NFAT activation: (1) positive modulation by

increasing cytosolic calcium concentration from ER stores and (2) negative modulation

by calcineurin binding. It is unclear why p12I has two regulatory functions for NFAT

transcriptional activity. It is notable that Bcl-2 has these similar properties like p12I, however the functional relationship between calcium release from the ER and calcineurin binding of Bcl-2 is also still unresolved. Bcl-2 maintains calcium homeostasis and prevents apoptosis by localizing not only at the mitochondrial membrane but also in the

ER membranes 1,288. p12I may function like Bcl-2 at the ER membrane, acting like an ion

channel protein to increase ER calcium permeability. Like Bcl-2, p12I may affect

apoptosis in HTLV-1-infected T cells. Another ER membrane protein, CAML that

activates NFAT by increasing calcium flux binds with calcineurin indirectly, through its

association with cyclophilin351. Overall, p12I interacts with calcineurin, an important

regulator of NFAT signaling, via a highly conserved PSLP(I/L)T motif, to further T cell

activation, an important antecedent to effective viral infection, via a

calcium/calcineurin/NFAT pathway.

35 1.8.1.5 Putative Role of p12I Variants in HAM/TSP

Proteasome destabilization of viral proteins is thought to be an intracellular

defense mechanism against viral infection 68,309. Lysine is a known target for covalent binding of ubiquitin 56,68,309and the metabolic instability p12I is mediated in part by

ubiquitylation at a single lysine residue at position 88 and subsequent proteasomal degradation, as well as by destabilizing residues at its amino terminus341. Interestingly,

lysine carrying allele was found only in some TSP-HAM cases irrespective of

geographical locations, suggesting that a selective pressure over p12I might occur in the

host 85,341. It is hypothesized that the reduced stability of p12I in HAM/TSP patients due to this sequence variation may facilitate generation of a viral-specific CTL response, since degradation of p12I would alleviate the reduction of MHC-I molecules at the cell

surface21. Interestingly, in a study conducted at the same geographic location, lysine

residue at position 88 in p12I was not found in all HAM/TSP patients, but found in an

asymptomatic HTLV-1 carrier225. However, since all individuals in this study were born

in the same geographic region, it might simply represent a particular HTLV-1 carrier

population in which the selective pressure on the p12I sequence would not be occurring.

Overall, the significance of natural p12I alleles is unclear, and it is possible that the lysine

at position 88 of p12I might have a significant effect on the biological effects of the

protein in the host, including giving a possible selective advantage in individuals with a

certain MHC-I. Future studies including screening of HTLV-1-infected individuals in

other populations may elucidate this further.

36 1.8.2 pX ORF II p13II

1.8.2.1 Biochemical Characteristics of p13II: Mitochondrial Targeting

Less is known about the function of the smaller protein, p13II, encoded by the 87

carboxy-terminal amino acids of the 241-residue Tof or p30II protein. Initial studies

demonstrated p13II localization to the nucleus173, but more-recent reports show

mitochondrial localization of the protein48,62, specifically in the inner mitochondrial

membrane and cristae62. This localization is mediated by an atypical mitochondrial targeting sequence (MTS) in the N terminus of p13II between amino acids 22-31 48(Fig

1.2). In addition, p13II MTS is, at least in part, was able to override the potent NLS of

Rev66. While the p13II MTS is also present in p30II, p30II is not localized to the

mitochondria, suggesting that the signal has to be near the amino-terminus to direct mitochondrial targeting63. Analysis of the amino acid sequence of p13II reveals no

characteristic DNA-binding motifs and previous data show neither DNA binding nor

transcriptional activity303.

1.8.2.2 p13II Alteration of Mitochondrial Morphology and Inner Membrane Potential

Functionally, expression of p13II disrupts the mitochondrial inner membrane

potential and alters mitochondrial morphology and architecture, leading to apparent

mitochondrial swelling and fragmentation of their normal interconnected string-like network, suggesting a role for p13II in induction of apoptosis48. Additionally, it is

thought that p13II might possibly cause changes in mitochondrial permeability and/or

alter processes like calcium signaling that relies on the mitochondrial network and the

ER. Since p13II does not cause substantial cytochrome c release48 or show any 37 involvement of the permeability transition pore (PTP) in driven cation fluxes62, the biological significance of p13II mitochondrial localization and disruption of membrane

potentials remains unclear. D’Agostino et al64 consider that p13II might initially induce

selective permeability changes causing swelling and subsequent de-energization and

irreversible swelling, and suggested that p13II triggers apoptosis. However, thus far,

increased levels of apoptosis have never been demonstrated in p13II-expressing cells,

leaving open the possibility for other mitochondrion-based functions of this viral protein.

Such functions could simply include an increased respiratory activity of mitochondria,

which is often accompanied by swelling or Ca2+ signaling and Ca2+ dependent regulation

of the oxidative phosphorylation machinery. Thus, p13II may facilitate later stages of

HTLV-1 infection such as assembly and release.

D’Agostino et al 62 demonstrated that amino acids 9–41 of p13II which includes

the MTS, folds into an α helix in the context of a membrane-like environment and has

specific effects on the permeability of isolated mitochondria to small cations.

Furthermore, the presence of the four arginines in the MTS is essential for the increase in mitochondrial ion conductance and in situ effects on mitochondrial morphology but not

for mitochondrial targeting. However, the observation that the p13II MTS is not cleaved

upon import into the mitochondria and that it does not require positively charged residues

distinguishes it from most MTS, and suggests that the protein might utilize a

mitochondrial import pathway distinct from that described for most MTS, which involves

binding to a series of acidic receptors followed by cleavage of the signal.

38 1.8.2.3 p13II Cellular Protein Interactions

Hou et al.,126 reported that p13II associates with two novel cellular proteins

designated C44 and C254. C254 appears to be rabbit actin binding protein 280

(ABP280) present in the cytoskeleton of many different cell types and functions in the

modulation of cell shape and polarity and is essential for migration in melanocytes and

other cultured cells. More importantly, ABP-280 is important in the insertion of adhesion

molecules into the cell membrane. While the region of ABP-280 that interacts with p13II is also involved in interactions with integrin B1, tissue factor, and presenilin, other regions of ABP-280 binds to glycoprotein and the cytoplasmic domain of furin126. C44 shares homology with archeal adenylate kinases, the eukaryotic homologues of which localize to mitochondria and are involved in energy metabolism. Interestingly, the human homologue of C44 is expressed in the Jurkat T cell line and proliferating, but not resting, PBMC 126. These studies were performed using two molecular clones of HTLV-

1, K30p and K34p, which have amino acid differences in rex, p13II, and p30II.

Interestingly, only p13II K34, but not p13II K30 allow the interaction with C44. The

amino acid sequence of the p13II variant used by Ciminale et al 48 was most similar to

p13II K30, as it had the 25- amino-acid carboxyl tail and that it would not bind C44. This

might provide clues to the role of specific amino acids within p13II, which are critical in binding C44 and in causing apoptosis. Additionally, p13II binds to farnesyl

pyrophosphate synthetase (FPPS), an enzyme involved in the mevalonate/squalene

pathway, and in the synthesis of farnesyl pyrophosphate, a substrate required for

prenylation of Ras190. Interestingly, G4, a mitochondrial targeting accessory protein of

39 BLV also binds to FPPS. Future studies on the functional significance of the interaction between p13II and these cellular proteins will possibly elucidate the role of p13II in altering mitochondrial physiology and thus in HTLV-1 replication and pathogenesis.

Furthermore, Mahana et al., 211 reported an increase in Vav phosphorylation in

rabbit cells transfected with an HTLV-1 molecular clone that contains two mutations in

pX ORF II, resulting in expression of truncated p13II and p30II. Vav is a

hematopoietically restricted guanine nucleotide exchange factor for the Rac/Rho family

of GTPases and is necessary for T cell activation146. These findings suggest that p13II may play a role in controlling the activation state of Vav, which may relate to viral infectivity and leukemogenesis.

1.8.3 pX ORF II p30II

Earlier studies suggested that ORF II was dispensable for expression of HTLV-1

proteins Tax, Rex, or Env, viral replication and immortalization of primary lymphocytes

in vitro74,302. In addition, the isolation of a viral clone containing a premature stop codon in pX ORF II, from leukemic cells led to the conclusion that pX ORF II was not necessary for the outgrowth of leukemic clones in vivo 44. However, possible functional

role of pX ORF II during early infection was not ruled out by these initial studies. To

specifically test the functional role of pX ORF II in viral replication in vivo, our

laboratory group inoculated rabbits with lethally irradiated cell lines expressing the wild-

type molecular clone of HTLV-1 (ACH) and a clone containing selected mutations in pX

40 ORF II (ACH.p30/13)23. While all ACH-inoculated rabbits became infected as early as 2 weeks postinoculation, ACH.p30/13-inoculated animals failed to become infected or maintained low proviral copy numbers in their blood leukocytes. These animals also had weak and transient ex vivo p19 antigen production from their PBMC cultures and anti-

HTLV-1 antibody titers declined towards the end of the study. Most strikingly, using quantitative competitive PCR, our laboratory group demonstrated a dramatically reduced

(up to 100-fold) viral load in the ACH.p30/13-infected animals 23. Taken together, these

data suggested that pX ORF II is necessary for maintenance of high viral loads in vivo.

1.8.3.1 Biochemical Characteristics of p30II: Features of a Transcription modulator

Several lines of evidence indicate that p30II acts as a transcription factor.

Importantly, the protein localizes to the nucleus, specifically the nucleolus of cells

transiently transfected with a p30II expression vector172,381. p30II contains a highly

conserved bipartite NLS between amino acids 71 to 98, which can be functionally

substituted for the NLS of Rex66. In addition, p30II contains serine and threonine-rich

regions that share distant homology to the activation domain of cellular transcription factors, such as Oct-1/2, Pit-1, and POU-1 47 (Fig. 1.2). Taken together, these characteristics suggest that p30II has a role in transcription of viral and cellular gene expression.

41 1.8.3.2 Role of p30II in Modulating Viral and Cellular Gene Expression

Our laboratory group has reported that, when provided in limiting concentrations,

p30II expression stimulates HTLV-1 LTR-driven reporter gene activity, even in the presence of Tax, whereas higher concentrations represses LTR and CRE-driven reporter gene activity381. These activities are analogous to herpes simplex virus type 1 (HSV-1)

regulation of its immediate-early (IE) gene promoter. VP16, a potent transcription factor

from HSV-1, binds the host cell protein HCF, which facilitates the stable complex

formation of the viral protein with Oct-1187,366. The IE gene promoter contains an Oct-1- like motif (TAATGARAT) that is important for IE gene expression, with both positive and negative effects, depending on the context of these cellular transcription factors and

VP16366.

Our laboratory group has also demonstrated that p30II co-localizes with p300 in

the nucleus and physically interacts with CBP/p300, at the highly conserved KIX domain,

the domain HTLV-1 Tax also interacts with. In addition, p30II is able to disrupt CREB-

Tax-CBP/p300 complexes bound to the viral 21-bp TRE repeats380. Since p30II and Tax

interacts with CBP/p300 through the same KIX domain, it is possible that the competitive

CBP/p300 binding between p30II and Tax might be the mechanism by which p30II attenuate the formation of these multiprotein complexes and thereby repress transcription on CREB-responsive promoters. Similarly, Tax expression has been demonstrated to interfere with the transcriptional activity of c-Myb and the binding of Tax and c-Myb to

KIX domain of CBP was found to be mutually exclusive36,53,259,260.

42 The coactivators CBP and p300 mediate transcriptional control of various cellular

and viral DNA binding transcription factors. Although these coactivators have divergent

functions, they are similar in nucleotide sequence, are evolutionarily conserved, and are

commonly referred to as CBP/p30030,134. CBP and p300 are highly related and share many functional properties, however there is evidence that these factors are not interchangeable. Several cellular and viral proteins that interact with either CBP or p300

have been identified, including steroid and retinoid hormone receptors, CREB, c-Jun, c-

Myb, Sap-1a, c-Fos, MyoD, p53, Stat-1/2, NF-κB, pp90rsk, TATA-binding protein,

TFIIB, HTLV-1 Tax, adenovirus E1A, Kaposi’s sarcoma-associated herpes virus viral

interferon regulatory factor protein, and simian virus 40 large T antigen30,134,349.

CBP/p300 protein is available only at limiting concentrations within the cell nucleus, causing an environment for competition between coactivators and transcription factors, thus providing an additional layer of regulated gene expression98,315,319. There is evidence of a functional antagonistic relationship between transcription factors, as a result of competition for binding to common regions of CBP/p30081,98,134. Under such a

condition of tight competition, relative concentrations of Tax/ p30II at various stages of

disease might be a critical factor in determining the levels of viral transcription. It is

possible that, at higher concentrations, p30II may support viral persistence by reducing

viral expression, and thus reducing immune elimination of HTLV-1 infected cells.

Additionally, p30II might also repress cellular genes necessary to maintain viral

persistence. However, at low concentrations, p30II enhances TRE/viral over

CRE/cellular mediated transcription. p30II has the potential to play a role in promoting viral transcription, cell proliferation, competitively repressing CBP/p300-dependent

43 cellular gene transcription (e.g., p53-dependent p21WAF1/CIP1 gene activity), and for

promoting the spread of the virus in vivo. This is also consistent with previous finding

from our laboratory that an infectious HTLV-1 molecular clone failed to maintain viral

loads in vivo when p30II and p13II expression was abolished 23.

Recently, it was reported that p30II modulates LTR mediated transcription, in the context of the entire provirus, by a post-transcriptional mechanism 258. However, data

presented in this thesis further confirm the role of CBP/p300 in the modulation of LTR

mediated transcription by p30II, in the context of the provirus. In the presence of

increasing concentrations of p300, we were able to rescue the p30II-mediated repression on LTR driven gene transcription, in a dose-dependent manner, irrespective of the presence or absence of the provirus. CBP and p300 bridge transcription factors to relevant promoters, has intrinsic histone acetyltransferase (HAT) activity, and form

complexes with proteins such as CBP/p300 binding protein-associated factor, which also

exhibits HAT activity 134. Recently, there is increasing knowledge of the mechanism and functional significance of the interactions between many viral proteins and CBP/p300. In the case of adenovirus oncoprotein E1A, interaction with CBP/p300 is critical for regulation of transcription, suppression of differentiation, and immortalization of cells in culture 30,134,348. The T antigen of SV40 regulates the expression of a group of cellular genes by modifying the HAT activity of CBP/p300 or by bridging the gap between DNA binding transcription factors and components of the general transcription machinery30,134.

Identifying the molecular mechanism and functional significance of the interaction between p30II and p300 is very crucial in understanding of the role of p30II in the

44 pathogenesis and replication of this important human pathogen. Therefore to further

understand the molecular mechanism and functional significance of the interaction

between p30II and p300, using N terminal and C terminal deletion mutants of p30II, we have identified the motifs within p30II that are critical in binding CBP/p300 and in

regulating LTR mediated transcription, in the presence/ absence of the provirus. Herein,

we confirm the role of p30II as a regulator viral gene transcription, in association with p300 and identified the amino acid sequence 1-132 to be critical in binding CBP/p300 and amino acids 100-179 of p30II to be critical for its function as a repressor of LTR-

mediated transcription, irrespective of the presence or absence of the HTLV-1 provirus.

Amino acid domain 100-179 contains important features such as the DNA binding

domain, serine and threonine rich residues with homology to Oct-1 and POU family of

transcription factors and five lysine residues, four of which represent potential acetylation

sites for CBP/p300. Data presented in this thesis define the roles of these four lysine

residues in p30II-mediated LTR repression. While CBP/p300-mediated acetylation has

become a common theme for regulation of protein function, herein we demonstrate that

the intrinsic histone acetyltransferase activity of CBP/p300 is critical in regulation of

p30II-mediated LTR repression.

1.8.3.3 HTLV-1 Mediated T Cell Activation /Transformation and the Role of p30II.

Mechanisms of HTLV-1 mediated T cell activation and transformation are not

completely understood. Unlike other oncogenic viruses, HTLV-1 does not contain a viral

oncogene or integrate at specific sites in the host cell genome, causing activation of

45 cellular proto-oncogenes 311. However, HTLV-1 transforms human primary CD4+CD8-

and immortalizes CD 8+CD4-, double positive, as well as double negative T cells 54 and a multistep oncogenic process is believed to occur during HTLV-1 infection, leading to cell proliferation and subsequent cell transformation. First, TCR engagement results in intracellular calcium increase and Ras/MAPK pathway activation, leading to IL-2 production and IL-2Rα accumulation at cell surface, causing induction of T cells to transit from G0 to G1 phase. Second, IL-2 and IL-2R complex interaction leads to activation of IL-2R/Jak/Stat and Shc/Ras/MAPK pathways, triggering T cell progression from G1 to S phase and eventually inducing cell proliferation. The association between these T cell activation pathways and T cell growth is not fully understood. Based on the studies on the initial stages of HTLV-1-induced T cell activation in ex vivo cultured infected T cell clones from HAM/TSP patients with HAM/TSP, infected T cell clones spontaneously proliferate in vitro up to 2 weeks after stimulation in the absence of exogenous IL-2 (involving autocrine IL-2/IL-2R stimulation), while uninfected T cell clones revert to G0 stage125. Tax is thought to increase the expression of IL-2Rα, thereby

induce this initial activation 128. Peripheral blood T cells can be immortalized and subsequently transformed by coculturing with HTLV-1-producing T cells124,125,235,293. In

vitro immortalization is thought to occur in multiple stages: At ~4 weeks after the

infection, T cell cultures reach a crisis stage with dramatic decrease in cell number,

followed by IL-2 dependent proliferation of cells surviving the crisis characterized by

upregulated expression of CD25 as well as MHC-II, polyclonal proviral integration,

scarse expression of tax/rex mRNA and transient expression of IL-2. Later, at ~14 weeks

after infection, oligoclonal proviral integration, abundant expression of tax/rex mRNA

46 and increased CD25 expression, but not IL-2, are detected158. Since transformed cells are resistant to rapamycin treatment, 365. HTLV-1 infection is believed to circumvent the rapamycin sensitive IL-2/IL-2R pathway. Interestingly, HTLV-1 transformed T cells

(but not immortalized T cells) show constitutive activation of IL-2R-associated Jak1/3 and Stat3/5 54,231. HTLV-1 Tax and p12I proteins have been linked to HTLV-1-induced

immortalization and transformation. Tax is known to upregulate p21WAF expression,

inhibit p53 42, inhibit p16INK and thus indirectly activate cyclin D/CDK4/6330, functional

disruption of mitotic checkpoint protein MAD1138 while p12 is reported to increase Stat5 activation and modestly enhance proliferation 262.

Data presented in this thesis establishes the role of p30II as a regulator of cellular

gene expression. More importantly, herein we illustrate that p30II activated many key transcription factors involved in T cell signaling/ activation, such as NFAT, NF-κB and

AP-1, and deregulated the expression of multiple genes involved in T cell activation or cell signaling and in regulation of transcription, translation, apoptosis, cell cycle, cell adhesion as well as immune response. The mechanism of function of p30II is not elucidated; however it is likely involved in T cell activation and subsequent T cell proliferation.

47 REFERENCES

1. Abeele, F. V., R. Skryma, Y. Shuba, F. Van Coppenolle, C. Slomianny, M. Roudbaraki, B. Mauroy, F. Wuytack, and N. Prevarskaya. 2002. Bcl-2- dependent modulation of Ca(2+) homeostasis and store-operated channels in prostate cancer cells. Cancer Cell 1:169-179.

2. Adedayo, O., G. Grell, and P. Bellot. 2003. Hospital admissions for human T- cell lymphotropic virus type-1 (HTLV-1) associated diseases in Dominica. Postgrad. Med. J 79:341-344.

3. Ahmed, Y. F., G. M. Gilmartin, S. M. Hanly, J. R. Nevins, and W. C. Greene. 1991. The HTLV-1 Rex Response Element Mediates a Novel Form of mRNA Polyadenylation. Cell 64:727-737.

4. Ahmed, Y. F., S. M. Hanly, M. H. Malim, B. R. Cullen, and W. C. Greene. 1990. Structure-function analyses of the HTLV-I Rex and HIV-1 Rev RNA response elements: insights into the mechanism of Rex and Rev action. Genes Dev. 4:1014-1022.

5. Akagi, T., H. Ono, N. Tsuchida, and K. Shimotohno. 1997. Aberrant expression and function of p53 in T-cells immortalized by HTLV-I Tax1. FEBS Lett. 406:263-266.

6. Alamy, A. H., F. B. Menezes, A. C. Leite, O. M. Nascimento, and A. Q. Araujo. 2001. Dysautonomia in human T-cell lymphotrophic virus type I- associated myelopathy/tropical spastic paraparesis. Ann. Neurol. 50:681-685.

7. Albrecht, B., N. D. Collins, M. T. Burniston, J. W. Nisbet, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J. Virol. 74:9828-9835.

8. Albrecht, B., C. D. D'Souza, W. Ding, S. Tridandapani, K. M. Coggeshall, and M. D. Lairmore. 2002. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J. Virol. 76:3493- 3501.

9. Albrecht, B. and M. D. Lairmore. 2002. Critical role of human T-lymphotropic virus type 1 accessory proteins in viral replication and pathogenesis. Microbiol. Mol. Biol. Rev. 66:396-406.

48 10. Alexandre, C. and B. Verrier. 1991. Four regulatory elements in the human c- fos promoter mediate transactivation by HTLV-1 Tax protein. Oncogene 6:543- 551.

11. Aramburu, J., M. B. Yaffe, C. Lopez-Rodriguez, L. C. Cantley, P. G. Hogan, and A. Rao. 1999. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A. Science 285:2129-2133.

12. Ariumi, Y., A. Kaida, J. Y. Lin, M. Hirota, O. Masui, S. Yamaoka, Y. Taya, and K. Shimotohno. 2000. HTLV-1 tax oncoprotein represses the p53-mediated trans-activation function through CBP sequestration. Oncogene 19:1491-1499.

13. Asher, D. M., J. Goudsmit, K. Pomeroy, R. M. Garruto, M. Bakker, S. Ono, N. Elliot, K. Harris, H. Askins, Z. Eldadah, A. L. Goldstein, and D. C. Gajdusek. 1988. Antibodies to HTLV-I in populations of the southwestern Pacific. J. Med. Virol. 26:339.

14. Asquith, B., E. Hanon, G. P. Taylor, and C. R. Bangham. 2000. Is human T- cell lymphotropic virus type I really silent? Philos. Trans. R. Soc Lond B Biol. Sci 355:1013-1019.

15. Azimi, N., J. Mariner, S. Jacobson, and T. A. Waldmann. 2000. How does interleukin 15 contribute to the pathogenesis of HTLV type 1- associated Myelopathy/Tropical spastic paraparesis? AIDS Res. Hum. Retroviruses 16:1717- 1722.

16. Ballard, D. W. 2001. Molecular mechanisms in lymphocyte activation and growth. Immunol. Res. 23:157-166.

17. Ballaun, C., G. K. Farrington, M. Dobrovnik, J. Rusche, J. Hauber, and E. Bohnlein. 1991. Functional Analysis of Human T-Cell Leukemia Virus Type 1 rex-Response Element: Direct RNA Binding of Rex Protein Correlates with In Vivo Activity. J. Virol. 65:4408-4413.

18. Bangham, C. R. 2000. The immune response to HTLV-I. Curr. Opin. Immunol. 12:397-402.

19. Bangham, C. R. 2003. Human T-lymphotropic virus type 1 (HTLV-1): persistence and immune control. Int. J Hematol. 78:297-303.

20. Bangham, C. R. 2003. The immune control and cell-to-cell spread of human T- lymphotropic virus type 1. J Gen. Virol 84:3177-3189.

21. Barmak, K., E. Harhaj, C. Grant, T. Alefantis, and B. Wigdahl. 2003. Human T cell leukemia virus type I-induced disease: pathways to cancer and neurodegeneration. Virology 308:1-12. 49 22. Barshira, A., A. Panet, and A. Honigman. 1991. An RNA Secondary Structure Juxtaposes Two Remote Genetic Signals for Human T-Cell Leukemia Virus Type 1 RNA 3'-End Processing. J. Virol. 65:5165-5173.

23. Bartoe, J. T., B. Albrecht, N. D. Collins, M. D. Robek, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J. Virol. 74:1094-1100.

24. Berneman, Z. N., R. B. Gartenhaus, M. S. Reitz, W. A. Blattner, A. Manns, B. Hanchard, O. Ikehara, R. C. Gallo, and M. E. Klotman. 1992. Expression of alternatively spliced human T-lymphotropic virus type 1 pX mRNA in infected cell lines and in primary uncultured cells from patients with adult T-cell leukemia/lymphoma and healthy carriers. Proc. Natl. Acad. Sci. USA 89:3005- 3009.

25. Bex, F., M. J. Yin, A. Burny, and R. B. Gaynor. 1998. Differential transcriptional activation by human T-cell leukemia virus type 1 Tax mutants is mediated by distinct interactions with CREB binding protein and p300. Mol. Cell Biol. 18:2392-2405.

26. Black, A. C., I. S. Y. Chen, S. Arrigo, C. T. Ruland, T. Allogiamento, E. Chin, and J. D. Rosenblatt. 1991. Regulation of HTLV-II gene expretion by rex involves positive and negative cis-Acting elements in the 5 long terminal repeat. Virology 181:433-444.

27. Black, A. C., C. Ruland, M. T. Yip, J. Luo, B. Tran, A. Kalssi, E. Quan, M. Aboud, I. S. Y. Chen, and J. D. Rosenblatt. 1991. Human T-Cell Leukemia Virus Type II Rex Binding and Activity Require an Intact splice Donor Site and a Specific RNA Secondary Structure. J. Virol. 65:6645-6653.

28. Black, A. C., C. T. Ruland, J. Luo, A. Bakker, J. K. Fraser, and J. D. Rosenblatt. 1994. Binding of Nuclear Proteins to HTLV-II Cis-Acting Repressive Sequence (CRS) RNA Correlates with CRS Function. Virology 200:29-41.

29. Blattner, W., V. Kalyanaraman, M. Robert-Guroff, T. Lister, D. Galton, P. S. Sarin, M. Crawford, D. Catovsky, M. Greaves, and R. C. Gallo. 1982. The human type-C retrovirus, HTLV, in blacks from the Caribbean region and relationchip to adult T-cell leukemia/lymphoma. Int. J. Cancer 30:257.

30. Blobel, G. A. 2002. CBP and p300: versatile coregulators with important roles in hematopoietic gene expression. J Leukoc. Biol. 71:545-556.

31. Bogerd, H. P., A. Echarri, T. M. Ross, and B. R. Cullen. 1998. Inhibition of human immunodeficiency virus Rev and human T- cell leukemia virus Rex

50 function, but not Mason-Pfizer monkey virus constitutive transport element activity, by a mutant human nucleoporin targeted to Crm1. J Virol 72:8627-8635.

32. Bosselut, R., J. F. Duvall, A. Gegonne, M. Bailly, A. Hemar, J. Brady, and J. Ghysdael. 1990. The product of the c-ets-1 proto-oncogene and the related Ets2 protein act as transcriptional activators of the long terminal repeat of human T cell leukemia virus HTLV-1. EMBO J. 9:3137-3144.

33. Bosselut, R., F. Lim, P. C. Romond, J. Frampton, J. Brady, and J. Ghysdael. 1992. Myb protein binds to multiple sites in the human T cell lymphotropic virus type 1 long terminal repeat and transactivates LTR-mediated expression. Virology 186:764-769.

34. Brady, J. N., K. T. Jeang, J. Duvall, and G. Khoury. 1987. Identification of p40x-Responsive Regulatory Sequences within the Human T-Cell Leukemia Virus Type 1 Long Terminal Repeat. J. Virol. 61:2175-2181.

35. Brauweiler, A., J. E. Garrus, J. C. Reed, and J. K. Nyborg. 1997. Repression of Bax gene expression by the HTLV-I tax protein: Implications for suppression of apoptosis in virally infected cells. Virology 231:135-140.

36. Campbell, K. M. and K. J. Lumb. 2002. Structurally distinct modes of recognition of the KIX domain of CBP by Jun and CREB. Biochemistry 41:13956-13964.

37. Cann, A. J., J. D. Rosenblatt, W. Wachsman, N. P. Shah, and I. S. Chen. 1985. Identification of the gene responsible for human T-cell leukaemia virus transcriptional regulation. Nature 318:571-574.

38. Cantrell, D. 1996. T cell antigen receptor signal transduction pathways. Annu. Rev. Immunol. 14:259-274.

39. Castillo, L. C., F. Gracia, G. C. Roman, P. Levine, W. C. Reeves, and J. Kaplan. 2000. Spinocerebellar syndrome in patients infected with human T- lymphotropic virus types I and II (HTLV-I/HTLV-II): report of 3 cases from Panama. Acta Neurol. Scand. 101:405-412.

40. Cavrois, M., I. Leclercq, O. Gout, A. Gessain, S. Wainhobson, and E. Wattel. 1998. Persistent oligoclonal expansion of human T-cell leukemia virus type 1 infected circulating cells in patients with Tropical spastic paraparesis/HTLV-1 associated myelopathy. Oncogene 17:77-82.

41. Cereseto, A., Z. Berneman, I. Koralnik, J. Vaughn, G. Franchini, and M. E. Klotman. 1997. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia 11:866-870.

51 42. Cereseto, A., F. Diella, J. C. Mulloy, A. Cara, P. Michieli, R. Grassmann, G. Franchini, and M. E. Klotman. 1996. P53 functional impairment and high p21(waf1/cip1) expression in human T-cell lymphotropic/leukemia virus type I- transformed T cells. Blood 88:1551-1560.

43. Chen, Y. M. A., S. H. Chen, C. Y. Fu, J. Y. Chen, and M. Osame. 1997. Antibody reactivities to tumor-suppressor protein p53 and HTLV-I Tof Rex and Tax in HTLV-I-infected people with differing clinical status. Int. J. Cancer 71:196-202.

44. Chou, K. S., A. Okayama, N. Tachibana, T. H. Lee, and M. Essex. 1995. Nucleotide sequence analysis of a full-length human T-cell leukemia virus type I from adult T-cell leukemia cells: A prematurely terminated pX open reading frame II. Int. J. Cancer 60:701-706.

45. Chrivia, J. C., R. P. Kwok, N. Lamb, M. Hagiwara, M. R. Montminy, and R. H. Goodman. 1993. Phosphorylated CREB binds specifically to the nuclear protein CBP. Nature 365(6449):855-859.

46. Ciminale, V., D. D'Agostino, L. Zotti, G. Franchini, B. K. Felber, and L. Chieco-Bianchi. 1995. Expression and characterization of proteins produced by mRNAs spliced into the X region of the human T-cell leukemia/lymphotropic virus type II. Virology 209:445-456.

47. Ciminale, V., G. N. Pavlakis, D. Derse, C. P. Cunningham, and B. K. Felber. 1992. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J. Virol. 66:1737-1745.

48. Ciminale, V., L. Zotti, D. M. Dagostino, T. Ferro, L. Casareto, G. Franchini, P. Bernardi, and L. Chiecobianchi. 1999. Mitochondrial targeting of the p13(II) protein coded by the x-II ORF of human T-cell leukemia/lymphotropic virus type I (HTLV-I). Oncogene 18:4505-4514.

49. Clark, N. M., M. J. Smith, J. M. Hilfinger, and D. M. Markovitz. 1993. Activation of the human T-cell leukemia virus type I enhancer is mediated by binding sites for Elf-1 and the pets factor. J. Virol. 67:5522-5528.

50. Cleghorn, F. R., A. Manns, R. Falk, P. Hartge, B. Hanchard, N. Jack, E. Williams, E. Jaffe, F. White, C. Bartholomew, and W. Blattner. 1995. Effect of human T-lymphotropic virus type I infection on non- Hodgkin's lymphoma incidence. J. Nat. Cancer Inst. 87:1009-1014.

51. Coffin, J. M. 1992. Structure and Classification of Retroviruses. The Retroviridae. 1:19-49.

52 52. Coffin, J. M. Retroviridae: The Viruses and Their Replication. 1767-1835. 1996. Phildaelphia, PA, Lippinocott-Raven Publishers.

53. Colgin, M. A. and J. K. Nyborg. 1998. The human T-cell leukemia virus type 1 oncoprotein Tax inhibits the transcriptional activity of c-Myb through competition for the CREB binding protein. J. Virol. 72:9396-9399.

54. Collins, N. D., C. D'Souza, B. Albrecht, M. D. Robek, L. Ratner, W. Ding, P. L. Green, and M. D. Lairmore. 1999. Proliferation response to interleukin-2 and Jak/Stat activation of T cells immortalized by human T-cell lymphotropic virus type 1 is independent of open reading frame I expression. J. Virol. 73:9642-9649.

55. Collins, N. D., G. C. Newbound, B. Albrecht, J. L. Beard, L. Ratner, and M. D. Lairmore. 1998. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood 91:4701-4707.

56. Coscoy, L. and D. Ganem. 2003. PHD domains and E3 ubiquitin ligases: viruses make the connection. Trends Cell Biol. 13:7-12.

57. Courouce, A. M., J. Pillonel, J. M. Lemaire, and C. Saura. 1998. HTLV testing in blood transfusion. Vox Sang. 74 Suppl 2:165-169.

58. Cowan, E. P., R. K. Alexander, S. Daniel, F. Kashanchi, and J. N. Brady. 1997. Induction of tumor necrosis factor alpha in human neuronal cells by extracellular human T-cell lymphotropic virus type 1 Tax. J. Virol. 71:6982-6989.

59. Crabtree, G. R. 2001. Calcium, calcineurin, and the control of transcription. J. Biol. Chem. 276:2313-2316.

60. Craig, H. M., T. R. Reddy, N. L. Riggs, P. P. Dao, and J. C. Guatelli. 2000. Interactions of HIV-1 nef with the mu subunits of adaptor protein complexes 1, 2, and 3: role of the dileucine-based sorting motif. Virology 271:9-17.

61. Cullen, B. R. 1998. Retroviruses as model systems for the study of nuclear RNA export pathways. Virology 249:203-210.

62. D'Agostino, D. M., L. Ranzato, G. Arrigoni, I. Cavallari, F. Belleudi, M. R. Torrisi, M. Silic-Benussi, T. Ferro, V. Petronilli, O. Marin, L. Chieco- Bianchi, P. Bernardi, and V. Ciminale. 2002. Mitochondrial alterations induced by the p13II protein of human T-cell leukemia virus type 1. Critical role of arginine residues. J Biol. Chem. 277:34424-34433.

63. D'Agostino, D. M., L. Zotti, T. Ferro, I. Cavallori, M. Silic-Benussi, L. Chieco-Bianchi, and V. Ciminale. 2001. Expression and functional properties of proteins encoded in the x-II ORF of HTLV-I. Virus Res. 78:35-43.

53 64. D'Agostino, D. M., L. Zotti, T. Ferro, G. Franchini, L. Chieco-Bianchi, and V. Ciminale. 2000. The p13II protein of HTLV type 1: comparison with mitochondrial proteins coded by other human viruses. AIDS Res. Hum. Retroviruses 16:1765-1770.

65. Daenke, S., S. Nightingale, J. K. Cruickshank, and C. R. M. Bangham. 1990. Sequence variants of human T-cell lymphotropic virus type I from patients with tropical spastic paraparesis and adult T-cell leukemia do not distinguish neurological from leukemic isolates. J. Virol. 64:1278-1282.

66. Dagostino, D. M., V. Ciminale, L. Zotti, A. Rosato, and L. Chiecobianchi. 1997. The human T-cell lymphotropic virus type 1 Tof protein contains a bipartite nuclear localization signal that is able to functionally replace the amino-terminal domain of Rex. J. Virol. 71:75-83.

67. Daly, T. J., R. C. Doten, P. Rennert, M. Auer, H. Jaksche, A. Donner, G. Fisk, and J. R. Rusche. 1993. Biochemical characterization of binding of multiple HIV-1 Rev monomeric proteins to the Rev responsive element. Biochemistry 32:10497-10505.

68. Dantuma, N. P. and M. G. Masucci. 2003. The ubiquitin/proteasome system in Epstein-Barr virus latency and associated malignancies. Semin. Cancer Biol. 13:69-76.

69. Datta, S., N. H. Kothari, and H. Fan. 2000. In vivo genomic footprinting of the human T-cell leukemia virus type 1 (HTLV-1) long terminal repeat enhancer sequences in HTLV-1-infected human T-cell lines with different levels of Tax I activity. J. Virol. 74:8277-8285.

70. Dekaban, G. A., A. A. Peters, J. C. Mulloy, J. M. Johnson, R. Trovato, E. Rivadeneira, and G. Franchini. 2000. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology 274:86-93.

71. Delamarre, L., C. Pique, D. Pham, T. Tursz, and M. C. Dokhelar. 1994. Identification of functional regions in the human T-cell leukemia virus type I SU glycoprotein. J. Virol. 68:3544-3549.

72. Delamarre, L., C. Pique, A. R. Rosenberg, V. Blot, M. P. Grange, B. Le, I, and M. C. Dokhelar. 1999. The Y-S-L-I tyrosine-based motif in the cytoplasmic domain of the human T-cell leukemia virus type 1 envelope is essential for cell- to-cell transmission. J Virol 73:9659-9663.

73. Derse, D. 1987. Bovine leukemia virus transcription is controlled by a virus- encoded trans-acting factor and by cis-acting response elements. J Virol 61:2462- 2471.

54 74. Derse, D., J. Mikovits, and F. Ruscetti. 1997. X-I and X-II open reading frames of HTLV-I are not required for virus replication or for immortalization of primary T-cells in vitro. Virology 237:123-128.

75. Desgranges, C., S. Souche, J. C. Vernant, D. Smadja, A. Vahlne, and P. Horal. 1994. Identification of novel neutralization-inducing regions of the human T cell lymphotropic virus type I envelope glycoproteins with human HTLV-I- seropositive sera. AIDS Res. Hum. Retroviruses 10:163-173.

76. Ding, W., Albrecht B., R. E. Kelley, N. Muthusamy, S. Kim, R. A. Altschuld, and M. D. Lairmore. 2002. Human T lymphotropic virus type 1 p12(I) expression increases cytoplasmic calcium to enhance the activation of nuclear factor of activated T cells. J Virol 76:10374-10382.

77. Ding, W., B. Albrecht, R. Luo, W. Zhang, J. R. Stanley, G. C. Newbound, and M. D. Lairmore. 2001. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J. Virol. 75:7672-7682.

78. Ding, W., S. J. Kim, A. M. Nair, B. Michael, K. Boris-Lawrie, A. Tripp, G. Feuer, and M. D. Lairmore. 2003. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol 77:11027-11039.

79. Duyao, M. P., D. J. Kessler, D. B. Spicer, C. Bartholomew, J. L. Cleveland, M. Siekevitz, and G. E. Sonenshein. 1992. Transactivation of the c-myc Promoter by Human T Cell Leukemia virus Type 1 tax is Mediated by NFkB. J. Biol. Chem. 267:16288-16291.

80. Edlich, R. F., J. A. Arnette, and F. M. Williams. 2000. Global epidemic of human T-cell lymphotropic virus type-I (HTLV-I). J. Emerg. Med. 18:109-119.

81. Ego, T., Y. Ariumi, and K. Shimotohno. 2002. The interaction of HTLV-1 Tax with HDAC1 negatively regulates the viral gene expression. Oncogene 21:7241- 7246.

82. Evangelista, A., S. Maroushek, H. Minnigan, A. Larson, E. Retzel, A. Haase, D. Gonzalez-Dunia, D. McFarlin, E. Mingioli, S. Jacobson, and . 1990. Nucleotide sequence analysis of a provirus derived from an individual with tropical spastic paraparesis. Microb. Pathog. 8:259-278.

83. Felber, B. K., H. Paskalis, and C. Kleinmanewing. 1985. The pX Protein of HTLV-1 is a Transcriptional Activator of its Long Terminal Repeats. Science 229:675-679.

55 84. Ferreira, O. C., V. Planelles, and J. D. Rosenblatt. 1997. Human T-cell leukemia viruses: Epidemiology, biology and pathogenesis. Blood Rev. 11:91- 104.

85. Franchini, G. 1995. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood 86:3619-3639.

86. Franchini, G., J. C. Mulloy, I. J. Koralnik, M. A. Lo, J. J. Sparkowski, T. Andresson, D. J. Goldstein, and R. Schlegel. 1993. The human T-cell leukemia/lymphotropic virus type I p12I protein cooperates with the E5 oncoprotein of bovine papillomavirus in cell transformation and binds the 16- kilodalton subunit of the vacuolar H+ ATPase. J Virol 67:7701-7704.

87. Franchini, G., F. Wong-Staal, and R. C. Gallo. 1984. Human T-cell leukemia virus (HTLV-I) transcripts in fresh and cultured cells of patients with adult T-cell leukemia. Proc. Natl. Acad. Sci. U. S. A 81:6207-6211.

88. Franchini, G., F. Wong-Staal, and R. C. Gallo. 1984. Molecular studies of human T-cell leukemia virus and adult T-cell leukemia. J Invest Dermatol. 83:63s-66s.

89. Fujii, M., P. Sassone-Corsi, and I. Verma. 1988. c-fos promotor trans-activation of cAMP on IL-2 stimulated gene expression. Proc. Natl. Acad. Sci. U. S. A.8526-8530.

90. Fujii, M., H. Tsuchiya, T. Chuhjo, T. Akizawa, and M. Seiki. 1992. Interaction of HTLV-1 Tax1 with p67SRF causes the aberrant induction of cellular immediate early genes through CArG boxes. Genes Dev. 6:2066-2076.

91. Fujii, M., H. Tsuchiya, X. B. Meng, and M. Seiki. 1995. c-Jun, c-Fos and their family members activate the transcription mediated by three 21-bp repetitive sequences in the HTLV-I long terminal repeat. Intervirology 38:221-228.

92. Fujisawa, J., M. Seiki, T. Kiyokawa, and M. Yoshida. 1985. Functional activation of the long terminal repeat of human T-cell leukemia virus type 1 by a trans-acting factor. Proc. Natl. Acad. Sci. USA 82:2277-2281.

93. Fukushima, T., T. Ikeda, E. Uyama, M. Uchino, H. Okabe, and M. Ando. 1994. Cognitive event-related potentials and brain magnetic resonance imaging in HTLV-1 associated myelopathy (HAM). J. Neurol. Sci. 126:30-39.

94. Furukawa, K., K. Furukawa, and H. Shiku. 1991. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 295:141-145.

95. Furukawa, Y., R. Kubota, N. Eiraku, M. Nakagawa, K. Usuku, S. Izumo, and M. Osame. 2003. Human T-cell lymphotropic virus type I (HTLV-I)-related 56 clinical and laboratory findings for HTLV-I-infected blood donors. J Acquir. Immune. Defic. Syndr. 32:328-334.

96. Gartenhaus, R. B. and P. Wang. 1995. Functional inactivation of wild-type p53 protein correlates with loss of IL-2 dependence in HTLV-I transformed human T lymphocytes. Leukemia 9:2082-2086.

97. Gatza, M. L., J. C. Watt, and S. J. Marriott. 2003. Cellular transformation by the HTLV-I Tax protein, a jack-of-all-trades. Oncogene 22:5141-5149.

98. Georges, S. A., H. A. Giebler, P. A. Cole, K. Luger, P. J. Laybourn, and J. K. Nyborg. 2003. Tax recruitment of CBP/p300, via the KIX domain, reveals a potent requirement for acetyltransferase activity that is dependent and histone tail independent. Mol. Cell Biol. 23:3392-3404.

99. Gerritsen, M. E., A. J. Williams, A. S. Neish, S. Moore, Y. Shi, and T. Collins. 1997. CREB-binding protein/p300 are transcriptional coactivators of p65. Proc. Natl. Acad. Sci. U. S. A 94:2927-2932.

100. Gessain, A. 1996. Virological aspects of tropical spastic paraparesis/HTLV-I associated myelopathy and HTLV-I infection. J. Neurovirology. 2:299-306.

101. Gessain, A., F. Barin, J. Vernant, O. Gout, L. Maurs, A. Calender, and G. Dethe. 1985. Antibodies to human T lymphotropic virus type 1 in patients with tropical spastic paresis. Lancet 2:407-410.

102. Gessain, A., R. C. Gallo, and G. Franchini. 1992. Low degree of human T-cell leukemia/lymphoma virus type I genetic drift in vivo as means of monitoring viral transmission and movement of ancient human populations. J. Virol. 66:2288- 2295.

103. Gessain, A. and R. Mahieux. 2000. [A virus called HTLV-1. Epidemiological aspects]. Presse Med. 29:2233-2239.

104. Godoy, A. J., J. Kira, K. Hasuo, and I. Goto. 1995. Characterization of cerebral white matter lesions of HTLV- I- associated myelopathy tropical spastic paraparesis in comparison with multiple sclerosis and collagen-vasculitis: A semiquantitative MRI study. J. Neurol. Sci. 133:102-111.

105. Gonzalez, G. A., K. K. Yamamoto, W. H. Fischer, D. Karr, P. Menzel, W. Biggs, III, W. W. Vale, and M. R. Montminy. 1989. A cluster of phosphorylation sites on the cyclic AMP-regulated nuclear factor CREB predicted by its sequence. Nature 337:749-752.

106. Gould, K. G. and C. R. Bangham. 1998. Virus variation, escape from cytotoxic T lymphocytes and human retroviral persistence. Semin Cell Dev Biol 9:321-328.

57 107. Grant, C., K. Barmak, T. Alefantis, J. Yao, S. Jacobson, and B. Wigdahl. 2002. Human T cell leukemia virus type I and neurologic disease: events in bone marrow, peripheral blood, and central nervous system during normal immune surveillance and neuroinflammation. J. Cell Physiol 190:133-159.

108. Green P.L. and Chen, B. K. Molecular features of the human T-cell leukemia virus: Mechanisms of transfromation and leukemogenicity. The Retroviridae. Vol. 3 (Levy J, ed.), 277-298. 1994. Plenum Publishing Corp. New York.

109. Green P.L. and I. S. Y. Chen. 2000. Human T-cell leukemia virus types 1 and 2, p. 1941-1969. Lippincott, Williams & Wilkins.

110. Green, P. L. and I. S. Y. Chen. 1994. Molecular features of the human T-cell leukemia virus. Mechanisms of transformation and leukemogenicity., p. 277-311. In J. A. Levy (ed.), The Retroviridae. Plenum Press, New York.

111. Hai, T. W., F. Liu, W. J. Coukos, and M. R. Green. 1989. Transcription factor ATF cDNA clones: an extensive family of leucine zipper proteins able to selectively form DNA-binding heterodimers. Genes Dev. 3:2083-2090.

112. Hakata, Y., T. Umemoto, S. Matsushita, and H. Shida. 1998. Involvement of human CRM1 (exportin 1) in the export and multimerization of the Rex protein of human T-cell leukemia virus type 1. J Virol 72:6602-6607.

113. Hakata, Y., M. Yamada, and H. Shida. 2003. A multifunctional domain in human CRM1 (exportin 1) mediates RanBP3 binding and multimerization of human T-cell leukemia virus type 1 Rex protein. Mol. Cell Biol. 23:8751-8761.

114. Hall, W. W., C. R. Liu, O. Schneewind, H. Takahashi, M. H. Kaplan, G. Roupe, and A. Vahlne. 1991. Deleted HTLV-1 provirus in blood and cutaneous lesions of patients with mycosis fungoides. Science 253:317-319.

115. Hanchard, B. 1996. Adult T-cell leukemia/lymphoma in Jamaica: 1986-1995. J Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S20-S25.

116. Hara, H., M. Morita, T. Iwaki, T. Hatae, Y. Itoyama, T. Kitamoto, S. Akizuki, I. Goto, and T. Watanabe. 1994. Detection of human T lymphotrophic virus type I (HTLV-I) proviral DNA and analysis of T cell receptor V beta CDR3 sequences in spinal cord lesions of HTLV-I-associated myelopathy/tropical spastic paraparesis. J. Exp. Med. 180:831-839.

117. Harrod, R., Y. L. Kuo, Y. Tang, Y. Yao, A. Vassilev, Y. Nakatani, and C. Z. Giam. 2000. p300 and p300/cAMP-responsive element-binding protein associated factor interact with human T-cell lymphotropic virus type-1 Tax in a multi- histone acetyltransferase/activator-enhancer complex. J. Biol. Chem. 275:11852-11857.

58 118. Harrod, R., Y. Tang, C. Nicot, H. S. Lu, A. Vassilev, Y. Nakatani, and C. Z. Giam. 1998. An exposed KID-like domain in human T-cell lymphotropic virus type 1 Tax is responsible for the recruitment of coactivators CBP/p300. Mol. Cell Biol. 18:5052-5061.

119. Hattori, S., T. Kiyokawa, K. Imagawa, F. Shimizu, E. Hashimura, M. Seiki, and M. Yoshida. 1984. Identification of gag and env gene products of human T- cell leukemia virus (HTLV). Virology 136:338-347.

120. Hino, S. 1989. Milk-borne transmission of HTLV-I as a major route in the endemic cycle. Acta Paediatr. Jpn. 31:428-435.

121. Hino, S., K. Yamaguchi, S. Katamine, H. Sugiyama, T. Amagasaki, K. Kinoshita, Y. Yoshida, H. Doi, Y. Tsuji, and T. Miyamoto. 1985. Mother-to- child transmission of human T-cell leukemia virus type-I. Jpn. J. Cancer Res. 76:474-480.

122. Hinuma, Y., H. Komoda, T. Chosa, T. Kondo, M. Kohakura, T. Takenaka, M. Kikuchi, M. Ichimaru, K. Yunoki, I. Sato, R. Matsuo, Y. Takiuchi, H. Uchino, and M. Hanaoka. 1982. Antibodies to adult T-cell leukemia virus- associated antigen (ATLA) in sera from patients with ATL and controls in Japan: a nation-wide seroepidemiological study. Int. J. Cancer 29:631.

123. Hinuma, Y., K. Nagata, M. Hanaoka, M. Nakai, T. Matsumoto, K. Shirakawa, and I. Miyoshi. 1981. Adult T Cell Leukemia: Antigens in an ATL cell line and detection of antibodies to antigen in human sera. Proc. Natl. Acad. Sci. U. S. A. 78:6476-6480.

124. Hollsberg, P. 1999. Mechanisms of T-cell activation by human T-cell lymphotropic virus type I. Microbiol. Mol. Biol. Rev. 63:308-333.

125. Hollsberg, P., K. W. Wucherpfennig, L. J. Ausubel, V. Calvo, B. E. Bierer, and D. A. Hafler. 1992. Characterization of HTLV-1 In Vivo Infected T Cell Clones IL-2-independent Growth of Nontransformed T Cells. J. Immunol. 148:3256-3263.

126. Hou, X., S. Foley, M. Cueto, and M. A. Robinson. 2000. The Human T-Cell Leukemia Virus Type I (HTLV-I) X Region Encoded Protein p13(II) Interacts with Cellular Proteins. Virology 277:127-135.

127. Ibrahim, F., G. Dethe, and A. Gessain. 1995. Isolation and characterization of a new simian T-cell leukemia virus type 1 from naturally infected celebes macaques (Macaca tonkeana): Complete nucleotide sequence and phylogenetic relationship with the australo-melanesian human T-cell leukemia virus type 1. J. Virol. 69:6980-6993.

59 128. Inoue, J., M. Seiki, T. Taniguchi, S. Tsuru, and M. Yoshida. 1986. Induction of interleukin 2 receptor gene expression by p40x encoded by human T-cell leukemia virus type 1. EMBO J. 5:2883-2888.

129. Inoue, J., M. Seiki, and M. Yoshida. 1986. The second pX product p27 chi-III of HTLV-1 is required for gag gene expression. FEBS Lett. 209:187-190.

130. Inoue, J., M. Yoshida, and M. Seiki. 1987. Transcriptional (p40x) and post- transcriptional (p27x-III) regulators are required for the expression and replication of human t-cell leukemia virus type I genes. Proc. Natl. Acad. Sci. USA 84:3653- 3657.

131. Itoh, M., J. Inoue, H. Toyoshima, T. Akizawa, M. Higashi, and M. Yoshida. 1989. HTLV-1 rex and HIV-1 rev act through similar mechanisms to relieve suppression of unspliced RNA expression. Oncogene 4:1275-1279.

132. Izumo, S., F. Umehara, and M. Osame. 2000. HTLV-I-associated myelopathy. Neuropathology. 20 Suppl:S65-S68.

133. Jacobson, S. 2002. Immunopathogenesis of human T cell lymphotropic virus type I-associated neurologic disease. J Infect. Dis. 186 Suppl 2:S187-S192.

134. Janknecht, R. 2002. The versatile functions of the transcriptional coactivators p300 and CBP and their roles in disease. Histol. Histopathol. 17:657-668.

135. Jeang, K. T., R. Chiu, E. Santos, and S. J. Kim. 1991. Induction of the HTLV-I LTR by Jun occurs through the Tax-responsive 21- bp elements. Virology 181:218-227.

136. Jeang, K. T., S. G. Widen, O. J. Semmes, and S. H. Wilson. 1990. HTLV-1 Trans-Activator Protein, Tax, Is a Trans-Repressor of the Human B-Ploymerase Gene. Science 247:1082-1083.

137. Jiang, H., H. Lu, R. L. Schiltz, C. A. Pise-Masison, V. V. Ogryzko, Y. Nakatani, and J. N. Brady. 1999. PCAF interacts with tax and stimulates tax transactivation in a histone acetyltransferase-independent manner [published erratum appears in Mol Cell Biol 2000 Mar;20(5):1897]. Mol. Cell Biol. 19:8136- 8145.

138. Jin, D. Y., F. Spencer, and K. T. Jeang. 1998. Human T cell leukemia virus type 1 oncoprotein Tax targets the human mitotic checkpoint protein MAD1. Cell 93:81-91.

139. Jinno, A., Y. Haraguchi, H. Shiraki, and H. Hoshino. 1999. Inhibition of cell- free human T-cell leukemia virus type 1 infection at a postbinding step by the synthetic peptide derived from an ectodomain of the gp21 transmembrane glycoprotein. J Virol 73:9683-9689. 60 140. Johnson, C., D. Van Antwerp, and T. J. Hope. 1999. An N-terminal nuclear export signal is required for the nucleocytoplasmic shuttling of IkappaBalpha. EMBO J 18:6682-6693.

141. Johnson, J. M. and G. Franchini. 2002. Retroviral proteins that target the major histocompatibility complex class I. Virus Res. 88:119-127.

142. Johnson, J. M., J. C. Mulloy, V. Ciminale, J. Fullen, C. Nicot, and G. Franchini. 2000. The MHC class I heavy chain is a common target of the small proteins encoded by the 3' end of HTLV type 1 and HTLV type 2. AIDS Res. Hum. Retroviruses 16:1777-1781.

143. Johnson, J. M., C. Nicot, J. Fullen, V. Ciminale, L. Casareto, J. C. Mulloy, S. Jacobson, and G. Franchini. 2001. Free Major Histocompatibility Complex Class I Heavy Chain Is Preferentially Targeted for Degradation by Human T-Cell Leukemia/Lymphotropic Virus Type 1 p12(I) Protein. J. Virol. 75:6086-6094.

144. Kamb, A., N. A. Gruis, J. Weaver-Feldhaus, Q. Liu, K. Harshman, S. V. Tavtigian, E. Stockert, R. S. Day, III, B. E. Johnson, and M. H. Skolnick. 1994. A cell cycle regulator potentially involved in genesis of many tumor types. Science 264:436-440.

145. Kamihira, S., S. Nakasima, Y. Oyakawa, Y. Moriuti, M. Ichimaru, H. Okuda, M. Kanamura, and T. Oota. 1987. Transmission of human T cell lymphotropic virus type I by blood transfusion before and after mass screening of sera from seropositive donors. Vox Sang. 52:43-44.

146. Kaminuma, O., M. Deckert, C. Elly, Y. C. Liu, and A. Altman. 2001. Vav- Rac1-mediated activation of the c-Jun N-terminal kinase/c-Jun/AP-1 pathway plays a major role in stimulation of the distal NFAT site in the interleukin-2 gene promoter. Mol. Cell Biol. 21:3126-3136.

147. Kanamori, H., N. Suzuki, H. Siomi, T. Nosaka, A. Sato, H. Sabe, M. Hatanaka, and T. Honjo. 1990. HTLV-1 p27rex stabilizes human interleukin-2 receptor alpha chain mRNA. EMBO J. 9:4161-4166.

148. Kaplan, J. E., M. Osame, H. Kubota, A. Igata, H. Nishitani, Y. Maeda, R. F. Khabbaz, and R. S. Janssen. 1990. The risk of development of HTLV-I- Associated myelopathy/tropical spastic paraparesis among persons infected with HTLV-I. J. Acq. Immun. Defic. Synd. 3:1096-1101.

149. Kasahata, N., J. Shiota, Y. Miyazawa, I. Nakano, and S. Murayama. 2003. Acute human T-lymphotropic virus type 1-associated myelopathy: a clinicopathologic study. Arch. Neurol. 60:873-876.

61 150. Kawakami, K., A. Miyazato, Y. Iwakura, and A. Saito. 1999. Induction of lymphocytic inflammatory changes in lung interstitium by human T lymphotropic virus type I. AMER J RESPIR CRIT CARE MED 160:995-1000.

151. Kawano, F., K. Yamaguchi, H. Nishimura, H. Tsuda, and K. Takatsuki. 1985. Variation in the clinical courses of adult T-cell leukemia. Cancer 55:851- 856.

152. Kerkhofs, P., H. Heremans, A. Burny, R. Kettmann, and L. Willems. 1998. In vitro and in vivo oncogenic potential of bovine leukemia virus G4 protein. J Virol 72:2554-2559.

153. Khabbaz, R. F., J. M. Douglas, F. N. Judson, R. A. Spiegel, M. E. St Louis, W. Whittington, T. M. Hartley, M. Lairmore, and J. E. Kaplan. 1990. Seroprevalence of human T-lymphotropic virus type I or II in sexually transmitted disease clinic patients in the USA. J. Infect. Dis. 162:241-244.

154. Khabbaz, R. F., I. M. Onorato, R. O. Cannon, T. M. Hartley, B. Roberts, B. Hosein, and J. E. Kaplan. 1992. Seroprevalence of HTLV-1 and HTLV-2 among intravenous drug users and persons in clinics for sexually transmitted diseases. N. Engl. J. Med. 326:375-380.

155. Khan, Z. M., M. Sebenik, and D. Zuckerfranklin. 1996. Localization of human T-cell lymphotropic virus-1 tax proviral sequences in skin biopsies of patients with mycosis fungoides by in situ polymerase chain reaction. J. Invest. Dermatol. 106:667-672.

156. Kim, J. H., P. A. Kaufman, S. M. Hanly, L. T. Rimsky, and W. C. Greene. 1991. Rex transregulation of human T-cell leukemia virus type II gene expression. J. Virol. 65:405-414.

157. Kim, S. J., W. Ding, B. Albrecht, P. L. Green, and M. D. Lairmore. 2003. A conserved calcineurin-binding motif in human T lymphotropic virus type 1 p12I functions to modulate nuclear factor of activated T cell activation. J Biol. Chem. 278:15550-15557.

158. Kimata, J. T. and L. Ratner. 1991. Temporal regulation of viral and cellular gene expression during HTLV-I mediated lymphocyte immortalization. J. Virol. 65:4398-4407.

159. Kimura, I., T. Tsubota, S. Tada, and J. Sogawa. 1986. Presence of antibodies against adult T cell leukemia antigen in the patients with chronic respiratory diseases. Acta Med. Okayama 40:281-284.

160. Kimzey, A. L. and W. S. Dynan. 1998. Specific regions of contact between human T-cell leukemia virus type I Tax protein and DNA identified by photocross- linking. J Biol Chem 273:13768-13775. 62 161. Kimzey, A. L. and W. S. Dynan. 1999. Identification of a human T-cell leukemia virus type I tax peptide in contact with DNA. J Biol Chem 274:34226- 34232.

162. Kira, J., I. Goto, M. Otsuka, and Y. Ichiya. 1993. Chronic Progressive Spinocerebellar Syndrome Associated with Antibodies to Human T- Lymphotropic Virus Type-I - Clinico- Virological and Magnetic Resonance Imaging Studies. J. Neurol. Sci. 115:111-116.

163. Kira, J. I., K. Fujihara, Y. Itoyama, I. Goto, and K. Hasuo. 1991. Leukoencephalopathy in HTLV-I-associated mylopathy/ tropical spastic paraparesis: MRI analysis and a two year follow up study after corticosteroid therapy. J. Neurol. Sci. 106:41-49.

164. Kitado, H., I. S. Chen, N. P. Shah, A. J. Cann, K. Shimotohno, and H. Fan. 1987. U3 sequences from HTLV-I and -II LTRs confer pX protein response to a murine leukemia virus LTR. Science 235:901-904.

165. Kitajima, I., K. Yamamoto, K. Sato, Y. Nakajima, T. Nakajima, I. Maruyama, M. Osame, and K. Nishioka. 1991. Detection of human T cell lymphotropic virus type I proviral DNA and its gene expression in synovial cells in chronic inflammatory arthropathy. J. Clin. Invest 88:1315-1322.

166. Kitze, B. and K. Usuku. 2002. HTLV-1-mediated immunopathological CNS disease. Curr. Top. Microbiol. Immunol. 265:197-211.

167. Kiwaki, T., F. Umehara, Y. Arimura, S. Izumo, K. Arimura, K. Itoh, and M. Osame. 2003. The clinical and pathological features of peripheral neuropathy accompanied with HTLV-I associated myelopathy. J Neurol. Sci. 206:17-21.

168. Kiyokawa, T., M. Seiki, S. Iwashita, K. Imagawa, F. Shimizu, and M. Yoshida. 1985. p27x-III and p21x-III, proteins encoded by the pX sequence of human T-cell leukemia virus type 1. Proc. Natl. Acad. Sci. USA 82:8359-8363.

169. Kobe, B., R. J. Center, B. E. Kemp, and P. Poumbourios. 1999. Crystal structure of human T cell leukemia virus type 1 gp21 ectodomain crystallized as a maltose-binding protein chimera reveals structural evolution of retroviral transmembrane proteins. PROC NAT ACAD SCI USA 96:4319-4324.

170. Kondo, T., H. Kono, N. Miyamoto, R. Yoshida, H. Toki, I. Matsumoto, M. Hara, H. Inoue, A. Inatsuki, T. Funatsu, and . 1989. Age- and sex-specific cumulative rate and risk of ATLL for HTLV-I carriers. Int. J Cancer 43:1061- 1064.

171. Koralnik, I. J., E. Boeri, W. C. Saxinger, A. L. Monico, J. Fullen, A. Gessain, H. G. Guo, R. C. Gallo, P. Markham, V. Kalyanaraman, V. Hirsch, J. Allan, K. Murthy, P. Alford, J. P. Slattery, S. J. O'brien, and G. Franchini. 1994. 63 Phylogenetic association of human and Simian T-cell leukemia/lymphotropic virus type I strains: evidence for interspecies transmission/. J. Virol. 68:2693- 2707.

172. Koralnik, I. J., J. Fullen, and G. Franchini. 1993. The p12I, p13II, and p30II proteins encoded by human T-cell leukemia/lymphotropic virus type I open reading frames I and II are localized in three different cellular compartments. J Virol 67:2360-2366.

173. Koralnik, I. J., A. Gessain, M. E. Klotman, M. A. Lo, Z. N. Berneman, and G. Franchini. 1992. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type I. Proc. Natl. Acad. Sci U. S. A 89:8813-8817.

174. Koralnik, I. J., J. C. Mulloy, T. Andresson, J. Fullen, and G. Franchini. 1995. Mapping of the intermolecular association of human T cell leukaemia/lymphotropic virus type I p12I and the vacuolar H+-ATPase 16 kDa subunit protein. J Gen. Virol 76 ( Pt 8):1909-1916.

175. Korber, B., A. Okayama, R. Donnelly, N. Tachibana, and M. Essex. 1991. Polymerase chain reaction analysis of defective human T-cell leukemia virus type 1 proviral genomes in leukemic cells of patients with adult T-cell leukemia. J. Virol. 65:5471-5476.

176. Kubota, R., S. S. Soldan, R. Martin, and S. Jacobson. 2002. Selected cytotoxic T lymphocytes with high specificity for HTLV-I in cerebrospinal fluid from a HAM/TSP patient. J. Neurovirol. 8:53-57.

177. Kubota, S., M. Hatanaka, and R. J. Pomerantz. 1996. Nucleo-cytoplasmic redistribution of the HTLV-I rex protein: Alterations by coexpression of the HTLV-I p21(x) protein. Virology 220:502-507.

178. Kuklina, E. M. and S. V. Shirshev. 2001. Role of transcription factor NFAT in the immune response. Biochemistry (Mosc. ) 66:467-475.

179. Kuroda, Y. and M. Matsui. 1993. Cerebrospinal Fluid Interferon-Gamma Is Increased in HTLV- I- Associated Myelopathy. J. Neuroimmunol. 42:223-226.

180. Kwok, R. P., J. R. Lundblad, J. C. Chrivia, J. P. Richards, H. P. Bachinger, R. G. Brennan, S. G. Roberts, M. R. Green, and R. H. Goodman. 1994. Nuclear protein CBP is a coactivator for the transcription factor. Nature 370(6486):223-226.

181. Kwok, R. P. S., M. E. Laurance, J. R. Lundblad, P. S. Goldman, H. M. Shih, L. M. Connor, S. J. Marriott, and R. H. Goodman. 1996. Control of cAMP- regulated enhancers by the viral transactivator Tax through CREB and the co- activator CBP. Nature 380:642-646.

64 182. Lagrenade, L., B. Hanchard, V. Fletcher, B. Cranston, and W. Blattner. 1990. Infective dermatitis of Jamaican children: a marker for HTLV-I infection. Lancet 336:1345-1347.

183. Lagrenade, L., A. Manns, V. Fletcher, C. Carberry, B. Hanchard, E. M. Maloney, B. Cranston, N. P. Williams, R. Wilks, E. C. Kang, and W. A. Blattner. 1998. Clinical, pathologic, and immunologic features of human T- lymphotrophic virus type I-associated infective dermatitis in children. Arch Dermatol 134:439-444.

184. Lagrenade, L., C. Morgan, C. Carberry, B. Hanchard, V. Fletcher, R. Gray, B. Cranston, P. Rodgers-Johnson, and A. Manns. 1995. Tropical spastic paraparesis occurring in HTLV-1 associated infective dermatitis. Report of two cases. West Indian Med J 44:34-35.

185. Lagrenade, L., S. Sonoda, W. Miller, E. Pate, P. Rodgersjohnson, B. Hanchard, B. Cranston, T. Fujiyoshi, S. Yashiki, M. Blank, C. J. Gibbs, and A. Manns. 1996. HLA DRB1(*)DQB1(*) haplotype in HTLV-I-associated familial infective dermatitis may predict development of HTLV-I- associated myelopathy tropical spastic paraparesis. Am. J. Med. Genet. 61:37-41.

186. Lairmore, M. D., B. Albrecht, C. D'Souza, J. W. Nisbet, W. Ding, J. T. Bartoe, P. L. Green, and W. Zhang. 2000. In vitro and in vivo functional analysis of human T cell lymphotropic virus type 1 pX open reading frames I and II. AIDS Res. Hum. Retroviruses 16:1757-1764.

187. Latchman, D. S. 1999. Regulation of DNA virus transcription by cellular POU family transcription factors. Rev Med Virol 9:31-38.

188. Le, B., I, M. P. Grange, L. Delamarre, A. R. Rosenberg, V. Blot, C. Pique, and M. C. Dokhelar. 2001. HTLV-1 structural proteins. Virus Res. 78:5-16.

189. Le, B., I, A. R. Rosenberg, and M. C. Dokhelar. 1999. Multiple functions for the basic amino acids of the human T-cell leukemia virus type 1 matrix protein in viral transmission. J Virol 73:1860-1867.

190. Lefebvre, L., V. Ciminale, A. Vanderplasschen, D. D'Agostino, A. Burny, L. Willems, and R. Kettmann. 2002. Subcellular localization of the bovine leukemia virus R3 and G4 accessory proteins. J Virol 76:7843-7854.

191. Lefebvre, L., A. Vanderplasschen, V. Ciminale, H. Heremans, O. Dangoisse, J. C. Jauniaux, J. F. Toussaint, V. Zelnik, A. Burny, R. Kettmann, and L. Willems. 2002. Oncoviral Bovine Leukemia Virus G4 and Human T-Cell Leukemia Virus Type 1 p13(II) Accessory Proteins Interact with Farnesyl Pyrophosphate Synthetase. J. Virol. 76:1400-1414.

65 192. Lemasson, I., V. Roberthebmann, S. Hamaia, M. D. Dodon, L. Gazzolo, and C. Devaux. 1997. Transrepression of lck gene expression by human T-cell leukemia virus type 1-encoded p40(tax). J. Virol. 71:1975-1983.

193. Lenzmeier, B. A., E. E. Baird, P. B. Dervan, and J. K. Nyborg. 1999. The Tax protein-DNA interaction is essential for HTLV-I transactivation in vitro. J Mol Biol 291:731-744.

194. Leonmonzon, M., I. Illa, and M. C. Dalakas. 1994. Polymyositis in patients infected with human T-cell leukemia virus type I: The role of the virus in the cause of the disease. Ann. Neurol. 36:643-649.

195. Levin, M. C., S. M. Lee, F. Kalume, Y. Morcos, F. C. Dohan, Jr., K. A. Hasty, J. C. Callaway, J. Zunt, D. Desiderio, and J. M. Stuart. 2002. Autoimmunity due to molecular mimicry as a cause of neurological disease. Nat. Med. 8:509- 513.

196. Levin, M. C., S. M. Lee, Y. Morcos, J. Brady, and J. Stuart. 2002. Cross- reactivity between immunodominant human T lymphotropic virus type I tax and neurons: implications for molecular mimicry. J Infect. Dis. 186:1514-1517.

197. Lewis, R. S. 2001. Calcium signaling mechanisms in T lymphocytes. Annu. Rev. Immunol. 19:497-521.:497-521.

198. Li, C. C., F. W. Ruscetti, N. R. Rice, E. Chen, N. S. Yang, J. Mikovits, and D. L. Longo. 1993. Differential expression of Rel family members in human T-cell leukemia virus type I-infected cells: transcriptional activation of c-rel by Tax protein. J. Virol. 67:4205-4213.

199. Li, G. Q., B. R. Vowels, B. M. Benoit, A. H. Rook, and S. R. Lessin. 1996. Failure to detect human T-lymphotropic virus type-I proviral DNA in cell lines and tissues from patients with cutaneous T-cell lymphoma. J. Invest. Dermatol. 107:308-313.

200. Li, H. C., T. Fujiyoshi, H. Lou, S. Yashiki, S. Sonoda, L. Cartier, L. Nunez, I. Munoz, S. Horai, and K. Tajima. 1999. The presence of ancient human T-cell lymphotropic virus type I provirus DNA in an Andean mummy. NATURE MED 5:1428-1432.

201. Li, X. H., K. M. Murphy, K. T. Palka, R. M. Surabhi, and R. B. Gaynor. 1999. The human T-cell leukemia virus type-1 Tax protein regulates the activity of the I kappa B kinase complex. J Biol Chem 274:34417-34424.

202. Li-Weber, M., M. Giaisi, K. Chlichlia, K. Khazaie, and P. H. Krammer. 2001. Human T cell leukemia virus type I Tax enhances IL-4 gene expression in T cells. Eur. J Immunol. 31:2623-2632.

66 203. Lindholm, P. F., R. L. Reid, and J. N. Brady. 1992. Extracellular Tax Protein Stimulates Tumor Necrosis Factor-B and Immunoglobulin Kappa Light Chain Expression in Lymphoid Cells. J. Virol. 66:1294-1302.

204. Liu, H. F., P. Goubau, M. Vanbrussel, K. Vanlaethem, Y. C. Chen, J. Desmyter, and A. M. Vandamme. 1996. The three human T-lymphotropic virus type I subtypes arose from three geographically distinct simian reservoirs. J. Gen. Virol. 77:359-368.

205. Lombard Platet, G. and P. Jalinot. 1993. Purification by DNA affinity precipitation of the cellular factors HEB1-p67 and HEB1-p94 which bind specifically to the human T-cell leukemia virus type-I 21 bp enhancer. Nucleic. Acids. Res. 21:3935-3942.

206. Lu, X., H. Yu, S. H. Liu, F. M. Brodsky, and B. M. Peterlin. 1998. Interactions between HIV1 Nef and vacuolar ATPase facilitate the internalization of CD4. Immunity. 8:647-656.

207. Lundblad, J. R., R. P. Kwok, M. E. Laurance, M. S. Huang, J. P. Richards, R. G. Brennan, and R. H. Goodman. 1998. The human T-cell leukemia virus-1 transcriptional activator Tax enhances cAMP-responsive element-binding protein (CREB) binding activity through interactions with the DNA minor groove. J Biol Chem 273:19251-19259.

208. Madeleine, M. M., S. Z. Wiktor, J. J. Goedert, A. Manns, P. H. Levine, R. J. Biggar, and W. A. Blattner. 1993. HTLV-I and HTLV-II World-Wide Distribution - Reanalysis of 4,832 Immunoblot Results. Int. J. Cancer 54:255-260.

209. Madore, S. J., L. S. Tiley, M. H. Malim, and B. R. Cullen. 1994. Sequence requirements for Rev multimerization in vivo. Virology 202:186-194.

210. Maggirwar, S. B., E. Harhaj, and S. C. Sun. 1995. Activation of NF-kappa B/Rel by Tax involves degradation of I kappa B alpha and is blocked by a proteasome inhibitor. Oncogene 11:993-998.

211. Mahana, W., T. M. Zhao, R. Teller, M. A. Robinson, and T. J. Kindt. 1998. Genes in the pX region of human T cell leukemia virus I influence Vav phosphorylation in T cells. Proc. Natl. Acad. Sci. U. S. A 95:1782-1787.

212. Mahieux, R., F. Ibrahim, P. Mauclere, V. Herve, P. Michel, F. Tekaia, C. Chappey, B. Garin, E. Vanderryst, B. Guillemain, E. Ledru, E. Delaporte, G. Dethe, and A. Gessain. 1997. Molecular epidemiology of 58 new African human T-cell leukemia virus type 1 (HTLV-1) strains: Identification of a new and distinct HTLV-1 molecular subtype in central Africa and in pygmies. J. Virol. 71:1317-1333.

67 213. Manca, N., E. Piacentini, M. Gelmi, P. Calzavara, M. A. Manganoni, A. Glukhov, F. Gargiulo, M. Defrancesco, F. Pirali, G. Depanfilis, and A. Turano. 1994. Persistence of human T cell lymphotropic virus type 1 (HTLV-1) sequences in peripheral blood mononuclear cells from patients with mycosis fungoides. J. Exp. Med. 180:1973-1978.

214. Mandic, R., O. T. Fackler, M. Geyer, T. Linnemann, Y. H. Zheng, and B. M. Peterlin. 2001. Negative factor from SIV binds to the catalytic subunit of the V- ATPase to internalize CD4 and to increase viral infectivity. Mol. Biol. Cell 12:463-473.

215. Manel, N., F. J. Kim, S. Kinet, N. Taylor, M. Sitbon, and J. L. Battini. 2003. The ubiquitous glucose transporter GLUT-1 is a receptor for HTLV. Cell 115:449-459.

216. Manninen, A., R. G. Herma, and K. Saksela. 2000. Synergistic activation of NFAT by HIV-1 nef and the Ras/MAPK pathway. J. Biol. Chem. 275:16513- 16517.

217. Manninen, A., P. Huotari, M. Hiipakka, G. H. Renkema, and K. Saksela. 2001. Activation of NFAT-dependent gene expression by Nef: conservation among divergent Nef alleles, dependence on SH3 binding and membrane association, and cooperation with protein kinase C-theta. J. Virol. 75:3034-3037.

218. Manninen, A. and K. Saksela. 2002. HIV-1 Nef Interacts with Inositol Trisphosphate Receptor to Activate Calcium Signaling in T Cells. J. Exp. Med. 195:1023-1032.

219. Manninen, A. and K. Saksela. 2002. HIV-1 Nef interacts with inositol trisphosphate receptor to activate calcium signaling in T cells. J Exp. Med. 195:1023-1032.

220. Manns, A., M. Hisada, and L. Lagrenade. 1999. Human T-lymphotropic virus type I infection. Lancet 353:1951-1958.

221. Manns, A., R. J. Wilks, E. L. Murphy, G. Haynes, J. P. Figueroa, M. Barnett, B. Hanchard, and W. A. Blattner. 1992. A prospective study of transmission by transfusion of HTLV-I and risk factors associated with seroconversion. Int. J. Cancer 51:886-891.

222. Mariette, X., F. Agbalika, M. T. Daniel, M. Bisson, P. Lagrange, J. C. Brouet, and F. Morinet. 1993. Detection of human T lymphotropic virus type I tax gene in salivary gland epithelium from two patients with Sjogren's syndrome. Arthritis Rheum. 36:1423-1428.

223. Mariette, X., P. Cherot, D. Cazals, C. Brocheriou, J. C. Brouet, and F. Agbalika. 1995. Antibodies to HTLV-I in Sjogren's syndrome. Lancet 345:71. 68 224. Marriott, S. J., P. F. Lindholm, K. M. Brown, S. D. Gitlin, J. F. Duvall, M. F. Radonovich, and J. N. Brady. 1990. A 36-kilodalton cellular transcription factor mediates an indirect interaction of human T-cell leukemia/lymphoma virus type 1 Tax1 with a responsive element in the viral long terminal repeat. Mol. Cell Biol. 10:4192-4201.

225. Martins, M. L., B. C. Soares, J. G. Ribas, G. W. Thorun, J. Johnson, E. G. Kroon, A. B. Carneiro-Prioetti, and C. A. Bonjardim. 2002. Frequency of p12K and p12R Alleles of HTLV Type 1 in HAM/TSP Patients and in Asymptomatic HTLV Type 1 Carriers. AIDS Res. Hum. Retroviruses 18:899- 902.

226. Maruyama, M., H. Shibuya, H. Harada, M. Hatakeyama, M. Seiki, T. Fujita, J. Inoue, M. Yoshida, and T. Taniguchi. 1987. Evidence for aberrant activation of the interleukin-2 autocrine loop by HTLV-1-encoded p40x and T3/Ti complex triggering. Cell 48:343-350.

227. Matsui, M., F. Nagumo, J. Tadano, and Y. Kuroda. 1995. Characterization of humoral and cellular immunity in the central nervous system of HAM/TSP. J. Neurol. Sci. 130:183-189.

228. Mellstrom, B. and J. R. Naranjo. 2001. Mechanisms of Ca(2+)-dependent transcription. Curr. Opin. Neurobiol. 11:312-319.

229. Merino, F., M. Robert-Guroff, J. Clark, M. Biondo-Bracho, W. Blattner, and R. C. Gallo. 1984. Natural antibodies to human T-cell leukemia/lymphoma virus in healthy Venezuelan populations. Int. J. Cancer 34:501.

230. Mesaeli, N., K. Nakamura, E. Zvaritch, P. Dickie, E. Dziak, K. H. Krause, M. Opas, D. H. MacLennan, and M. Michalak. 1999. Calreticulin is essential for cardiac development. J. Cell Biol. 144:857-868.

231. Migone, T. S., J. X. Lin, A. Cereseto, J. C. Mulloy, J. J. Oshea, G. Franchini, and W. J. Leonard. 1995. Constitutively activated Jak-STAT pathway in T cells transformed with HTLV-I. Science 269:79-81.

232. Miyake, H., T. Suzuki, H. Hirai, and M. Yoshida. 1999. Trans-activator Tax of human T-cell leukemia virus type 1 enhances mutation frequency of the cellular genome. Virology 253:155-161.

233. Miyatake, S., M. Seiki, R. D. Malefijt, T. Heike, J. Fujisawa, Y. Takebe, J. Nishida, J. Shlomai, T. Yokota, M. Yoshida, and . 1988. Activation of T cell- derived lymphokine genes in T cells and fibroblasts: effects of human T cell leukemia virus type I p40x protein and bovine papilloma virus encoded E2 protein. Nucleic Acids Res. 16:6547-6566.

69 234. Miyoshi, I., I. Kubonishi, M. Sumida, S. Hiraki, T. Tsubota, I. Kimura, K. Miyamoto, and J. Sato. 1980. A novel T-cell line derived from adult T-cell leukemia. Gann 71:155-156.

235. Miyoshi, I., I. Kubonishi, S. Yoshimoto, T. Akagi, Y. Ohtsuki, Y. Shiraishi, K. Nagata, and Y. Hinuma. 1981. Type C virus particles in a cord T-cell line derived by co-cultivating normal human cord leukocytes and human leukaemic T cells. Nature 294:770-771.

236. Miyoshi, I., I. Kubonishi, S. Yoshimoto, and Y. Shiraishi. 1981. A T-cell line derived from normal human cord leukocytes by co-culturing with human leukemic T-cells. Gann 72:978-981.

237. Miyoshi, I., S. Yoshimoto, I. Kubonishi, H. Taguchi, Y. Shiraishi, Y. Ohtsuki, and T. Akagi. 1981. Transformation of normal human cord lymphocytes by co- cultivation with a lethally irradiated human T-cell line carrying type C virus particles. Gann 72:997-998.

238. Mizokami, T., K. Okamura, T. Kohno, K. Sato, H. Ikenoue, T. Kuroda, K. Inokuchi, and M. Fujishima. 1995. Human T-lymphotropic virus type I- associated uveitis in patients with Graves' disease treated with methylmercaptoimidazole. J. Clin. Endocrinol. Metab 80:1904-1907.

239. Mochizuki, M., T. Watanabe, K. Yamaguchi, K. Takatsuki, K. Yoshimura, M. Shirao, S. Nakashima, S. Mori, S. Araki, and N. Miyata. 1992. HTLV-I Uveitis: A distinct clinical entity caused by HTLV-I. Jpn. J. Cancer Res. 83:236- 239.

240. Mori, N., N. Mukaida, D. W. Ballard, K. Matsushima, and N. Yamamoto. 1998. Human T-cell leukemia virus type I Tax transactivates human interleukin 8 gene through acting concurrently on AP- 1 and nuclear factor-kappa B-like sites. Cancer Res 58:3993-4000.

241. Mori, N. and D. Prager. 1996. Transactivation of the interleukin-1 alpha promoter by human T- cell leukemia virus type I and type II tax proteins. Blood 87:3410-3417.

242. Mortreux, F., A. S. Gabet, and E. Wattel. 2003. Molecular and cellular aspects of HTLV-1 associated leukemogenesis in vivo. Leukemia 17:26-38.

243. Moura, I. C., Y. Lepelletier, B. Arnulf, P. England, C. Baude, C. Beaumont, A. Bazarbachi, M. Benhamou, R. C. Monteiro, and O. Hermine. 2003. A neutralizing monoclonal antibody (mAb A24) directed against the transferrin receptor induces apoptosis of tumor T lymphocytes from ATL patients. Blood.

70 244. Muchardt, C., J. S. Seeler, A. Nirula, S. Gong, and R. Gaynor. 1992. Transcription factor AP-2 activates gene expression of HTLV-I. EMBO Journal. 11:2573-2581.

245. Mueller, N., A. Okayama, S. Stuver, and N. Tachibana. 1996. Findings from the Miyazaki Cohort Study. J Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S2-S7.

246. Mulloy, J. C., R. W. Crownley, J. Fullen, W. J. Leonard, and G. Franchini. 1996. The human T-cell leukemia/lymphotropic virus type 1 p12I proteins bind the interleukin-2 receptor beta and gammac chains and affects their expression on the cell surface. J Virol 70:3599-3605.

247. Mulloy, J. C., T. Kislyakova, A. Cereseto, L. Casareto, A. Lomonico, J. Fullen, M. V. Lorenzi, A. Cara, C. Nicot, C. Giam, and G. Franchini. 1998. Human T-cell lymphotropic/leukemia virus type 1 Tax abrogates p53- induced cell cycle arrest and apoptosis through its CREB/ATF functional domain. J. Virol. 72:8852-8860.

248. Murakami, T., H. Hirai, T. Suzuki, J. I. Fujisawa, and M. Yoshida. 1995. HTLV-1 tax enhances NF-kappa B2 expression and binds to the products p52 and p100, but does not suppress the inhibitory function of p100. Virology 206:1066- 1074.

249. Muraoka, O., T. Kaisho, M. Tanabe, and T. Hirano. 1993. Transcriptional activation of the interleukin-6 gene by HTLV-1. Immunology Letters. 37(2- 3):159-165.

250. Murphy, E. L., M. P. Busch, M. Tong, P. Cornett, and G. N. Vyas. 1998. A prospective study of the risk of transfusion-acquired viral infections. Transfus. Med. 8:173-178.

251. Murphy, E. L., J. P. Figueroa, W. N. Gibbs, A. Brathwaite, M. Holding- Cobham, D. Waters, B. Cranston, B. Hanchard, and W. A. Blattner. 1989. Sexual transmission of human Y-lymphotropic virus type I (HTLV-I). Ann. Intern. Med. 111:555-560.

252. Murphy, E. L., S. A. Glynn, J. Fridey, J. W. Smith, R. A. Sacher, C. C. Nass, H. E. Ownby, D. J. Wright, and G. J. Nemo. 1999. Increased incidence of infectious diseases during prospective follow-up of human T-lymphotropic virus type II- and I-infected blood donors. ARCH INTERN MED 159:1485-1491.

253. Nagai, M. and M. Osame. 2003. Human T-cell lymphotropic virus type I and neurological diseases. J Neurovirol. 9:228-235.

254. Nagai, M., K. Usuku, W. Matsumoto, D. Kodama, N. Takenouchi, T. Moritoyo, S. Hashiguchi, M. Ichinose, C. R. Bangham, S. Izumo, and M. 71 Osame. 1998. Analysis of HTLV-I proviral load in 202 HAM/TSP patients and 243 asymptomatic HTLV-I carriers: high proviral load strongly predisposes to HAM/TSP. J Neurovirol. 4:586-593.

255. Nam, S. H., T. D. Copeland, M. Hatanaka, and S. Oroszlan. 1993. Characterization of ribosomal frameshifting for expression of pol gene products of human T-cell leukemia virus type I. J. Virol. 67:196-203.

256. Nam, S. H., M. Kidokoro, H. Shida, and M. Hatanaka. 1988. Processing of gag Precursor Polyprotein of Human T-Cell Leukemia Virus Type 1 by Virus- Encoded Protease. J. Virol. 62:3718-3728.

257. Newbound, G. C., J. P. O'Rourke, N. D. Collins, J. M. Andrews, J. Dewille, and M. D. Lairmore. 2000. Repression of tax-mediated human t-lymphotropic virus type 1 transcription by inducible cAMP early repressor (ICER) protein in peripheral blood mononuclear cells. J. Med. Virol. 62:286-292.

258. Nicot, C., M. Dundr, J. M. Johnson, J. R. Fullen, N. Alonzo, R. Fukumoto, G. L. Princler, D. Derse, T. Misteli, and G. Franchini. 2004. HTLV-1-encoded p30(II) is a post-transcriptional negative regulator of viral replication. Nat. Med.

259. Nicot, C., R. Mahieux, R. Opavsky, A. Cereseto, L. Wolff, J. N. Brady, and G. Franchini. 2000. HTLV-I Tax transrepresses the human c-Myb promoter independently of its interaction with CBP or p300. Oncogene 19:2155-2164.

260. Nicot, C., R. Mahieux, C. Pise-Masison, J. Brady, A. Gessain, S. Yamaoka, and G. Franchini. 2001. Human T-Cell Lymphotropic Virus Type 1 Tax Represses c-Myb-Dependent Transcription through Activation of the NF-kappaB Pathway and Modulation of Coactivator Usage. Mol. Cell Biol. 21:7391-7402.

261. Nicot, C., R. Mahieux, S. Takemoto, and G. Franchini. 2000. Bcl-X(L) is up- regulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL samples. Blood 96:275-281.

262. Nicot, C., J. C. Mulloy, M. G. Ferrari, J. M. Johnson, K. Fu, R. Fukumoto, R. Trovato, J. Fullen, W. J. Leonard, and G. Franchini. 2001. HTLV-1 p12(I) protein enhances STAT5 activation and decreases the interleukin-2 requirement for proliferation of primary human peripheral blood mononuclear cells. Blood 98:823-829.

263. Nishioka, K., I. Maruyama, K. Sato, I. Kitajima, Y. Nakajima, and M. Osame. 1989. Chronic inflammatory arthropathy associated with HTLV-I [letter]. Lancet i:441.

264. Nitta, T., M. Tanaka, B. Sun, S. Hanai, and M. Miwa. 2003. The genetic background as a determinant of human T-cell leukemia virus type 1 proviral load. Biochem. Biophys. Res. Commun. 309:161-165. 72 265. Nosaka, T., H. Siomi, Y. Adachi, M. Ishibashi, S. Kubota, M. Maki, and M. Hatanaka. 1989. Nucleolar targeting signal of human T-cell leukemia virus type I rex- encoded protein is essential for cytoplasmic accumulation of unspliced viral mRNA. Proc. Natl. Acad. Sci. U. S. A 86:9798-9802.

266. Nyborg, J. K. and W. S. Dynan. 1990. Interaction of cellular proteins with the human T-cell leukemia virus type I transcriptional control region. Purification of cellular proteins that bind the 21-base pair repeat elements. J. Biol. Chem. 265:8230-8236.

267. Ohshima, K., M. Kikuchi, Y. Masuda, Y. Sumiyoshi, F. Eguchi, H. Mohtai, M. Takeshita, and N. Kimura. 1992. Human T-cell leukemia virus type I associated lymphadenitis. Cancer 69:239-248.

268. Okochi, K., H. Sato, and Y. Hinuma. 1984. A retrospective study on transmission of adult T cell leukemia virus by blood transfusion: seroconversion in recipients. Vox Sang. 46:245-253.

269. Ono, A., E. Ikeda, M. Mochizuki, M. Matsuoka, K. Yamaguchi, T. Sawada, S. Yamane, S. Tokudome, and T. Watanabe. 1998. Provirus load in patients with human T-cell leukemia virus type 1 uveitis correlates with precedent Graves' disease and disease activities. Jpn J Cancer Res 89:608-614.

270. Ono, A., M. Mochizuki, K. Yamaguchi, N. Miyata, and T. Watanabe. 1997. Immunologic and virologic characterization of the primary infiltrating cells in the aqueous humor of human T-cell leukemia virus type-1 uveitis. Accumulation of the human T-cell leukemia virus type-1- infected cells and constitutive expression of viral and interleukin-6 messenger ribonucleic acids. Invest Ophthalmol. Vis. Sci. 38:676-689.

271. Ootsuyama, Y., K. Shimotohno, M. Miwa, S. Oroszlan, and T. Sugimura. 1985. Myristylation of gag protein in human T-cell leukemia virus type-I and type-II. Jpn. J Cancer Res. 76:1132-1135.

272. Orita, S., A. Saiga, S. Takagi, T. Tanaka, K. Okumura, Y. Aono, Y. Hinuma, and H. Igarashi. 1991. A novel alternatively spliced viral mRNA transcribed in cells infected with human T cell leukemia virus type 1 is mainly responsible for expressing p21X protein. FEBS Lett. 295:127-134.

273. Orita, S., S. Sato, Y. Aono, N. Minoura, T. Yamashita, Y. Hinuma, and H. Igarashi. 1993. Identification of novel singly spliced pX mRNA transcripts common to. Virus Genes 7(2):197-204.

274. Orita, S., S. Takagi, A. Saiga, N. Minoura, K. Araki, K. I. Kinoshita, T. Kondo, Y. Hinuma, and H. Igarashi. 1992. Human T cell leukaemia virus type 1 p21X mRNA: constitutive expression peripheral blood mononuclear cells of patients with adult T cell leukaemia. J. Gen. Virol. 73:2283-2289. 73 275. Oroszlan, S. 1984. Chemical Analysis of HTLV-1 Structural proteins, p. 101- 110. In R. C. Gallo, M. E. Essex, and L. Gross (eds.), Human T-cell Leukemia Lymphoma Viruses. Cold Spring Harbor Laboratory Press.

276. Osame, M. 2002. Pathological mechanisms of human T-cell lymphotropic virus type I-associated myelopathy (HAM/TSP). J Neurovirol. 8:359-364.

277. Osame, M., R. Janssen, H. Kubota, H. Nishitani, A. Igata, S. Nagataki, M. Mori, I. Goto, H. Shimabukuro, R. Khabbaz, and J. Kaplan. 1990. Nationwide survey of HTLV-I-associated Myelopathy in Japan: Association with blood transfusion. Ann. Neurol. 28:50-56.

278. Osame, M., K. Usuku, S. Izumo, N. Ijichi, H. Amitani, A. Igata, M. Matsumoto, and M. Tara. 1986. HTLV-I associated myelopathy, a new clinical entity. Lancet i:1031-1032.

279. Overbaugh, J. 2004. HTLV-1 sweet-talks its way into cells. Nat. Med. 10:20-21.

280. Ozawa, Y., M. Migita, T. Watanabe, I. Okuda, A. Takeshita, A. Takagi, and Y. Shishiba. 1994. Development of Graves' ophthalmopathy and uveitis after radioiodine therapy for Graves' disease in a patient with HTLV-I associated myelopathy (HAM). Intern. Med. 33:564-568.

281. Pancake, B. A., D. Z. Franklin, and E. E. Coutavas. 1995. The cutaneous T cell lymphoma, mycosis fungoides, is a human T cell lymphotropic virus-associated disease - A study of 50 patients. J. Clin. Invest. 95:547-554.

282. Perini, G., S. Wagner, and M. R. Green. 1995. Recognition of bZIP proteins by the human T-cell leukaemia virus transactivator Tax. Nature 376:602-605.

283. Perkins, N. D., L. K. Felzien, J. C. Betts, K. Leung, D. H. Beach, and G. J. Nabel. 1997. Regulation of NF-kappaB by cyclin-dependent kinases associated with the p300 coactivator. Science 275:523-527.

284. Petit, C., F. Buseyne, C. Boccaccio, J. P. Abastado, J. M. Heard, and O. Schwartz. 2001. Nef is required for efficient -1 replication in cocultures of dendritic cells and lymphocytes. Virology 286:225-236.

285. Petrini, M., E. Pelosi-Testa, N. M. Sposi, G. Mastroberardino, A. Camagna, L. Bottero, F. Mavilio, U. Testa, and C. Peschle. 1989. Constitutive expression and abnormal glycosylation of transferrin receptor in acute T-cell leukemia. Cancer Res. 49:6989-6996.

286. Philpott, S. M. and G. C. Buehring. 1999. Defective DNA repair in cells with human T-cell leukemia/bovine leukemia viruses: role of tax gene. J Natl. Cancer Inst. 91:933-942.

74 287. Piguet, V., O. Schwartz, S. Legall, and D. Trono. 1999. The downregulation of CD4 and MHC-I by primate lentiviruses: a paradigm for the modulation of cell surface receptors. Immunol Rev 168:51-63.

288. Pinton, P., D. Ferrari, E. Rapizzi, F. D. Di Virgilio, T. Pozzan, and R. Rizzuto. 2001. The Ca2+ concentration of the endoplasmic reticulum is a key determinant of ceramide-induced apoptosis: significance for the molecular mechanism of Bcl-2 action. EMBO J 20:2690-2701.

289. Pique, C., D. Pham, T. Tursz, and M. C. Dokhelar. 1992. Human T-Cell Leukemia Virus Type I Envelope Protein Maturation Process: Requirements for Syncytium Formation. J. Virol. 66:906-913.

290. Pique, C., A. Uretavidal, A. Gessain, B. Chancerel, O. Gout, R. Tamouza, F. Agis, and M. C. Dokhelar. 2000. Evidence for the chronic in vivo production of human T cell leukemia virus type I Rof and Tof proteins from cytotoxic T lymphocytes directed against viral peptides. J. Exp. Med. 191:567-572.

291. Pise-Masison, C. A., R. Mahieux, H. Jiang, M. Ashcroft, M. Radonovich, J. Duvall, C. Guillerm, and J. N. Brady. 2000. Inactivation of p53 by human T- cell lymphotropic virus type 1 Tax requires activation of the NF-kappaB pathway and is dependent on p53 phosphorylation. Mol. Cell Biol. 20:3377-3386.

292. Pisemasison, C. A., M. Radonovich, K. Sakaguchi, E. Appella, and J. N. Brady. 1998. Phosphorylation of p53: a novel pathway for p53 inactivation in human T-cell lymphotropic virus type 1- transformed cells. J Virol 72:6348-6355.

293. Poiesz, B. J., F. W. Ruscetti, A. F. Gazdar, P. A. Bunn, J. D. Minna, and R. C. Gallo. 1980. Detection and isolation of type C retrovirus particles from fresh and cultured lymphocytes of a patient with cutaneous T-cell lymphoma. Proc. Natl. Acad. Sci. USA 77:7415-7419.

294. Pollard, V. W. and M. H. Malim. 1998. The HIV-1 Rev protein. Annu. Rev. Microbiol. 52:491-532.

295. Pombo de Oliveira, M. S., E. Matutes, T. Schulz, S. M. Carvalho, H. Noronha, J. D. Reaves, P. Loureiro, C. Machado, and D. Catovsky. 1995. T- cell malignancies in Brazil. Clinico-pathological and molecular studies of HTLV- I-positive and -negative cases. Int. J Cancer 60:823-827.

296. Princler, G. L., J. G. Julias, S. H. Hughes, and D. Derse. 2003. Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology 317:136-145.

297. Ratner, L. 1989. Regulation of expression of the c-sis proto-oncogene. Nucleic Acids Res. 17:4101-4115.

75 298. Reeves, W., C. Saxinger, M. Brenes, E. Quinoz, M. Hoh, J. Clark, C. Holt, R. C. Gallo, and W. Blattner. 1988. Human T-cell lymphotropic virus type I (HTLV-I) and risk factors in metropolitan Panama. Am. J. Epidemiol. 127:539.

299. Reid, R. L., P. F. Lindholm, A. Mireskandari, J. Dittmer, and J. N. Brady. 1993. Stabilization of wild-type p53 in human T-lymphocytes transformed by HTLV-I. Oncogene 8:3029-3036.

300. Ressler, S., G. F. Morris, and S. J. Marriott. 1997. Human T-cell leukemia virus type 1 Tax transactivates the human proliferating cell nuclear antigen promoter. J. Virol. 71:1181-1190.

301. Rimsky, L., J. Hauber, M. Dukovich, M. H. Malim, A. Langlois, B. R. Cullen, and W. C. Greene. 1988. Functional replacement of the HIV-1 rev protein by the HTLV-1 rex protein. Nature 335:738-740.

302. Robek, M. D., F. H. Wong, and L. Ratner. 1998. Human T-Cell leukemia virus type 1 pX-I and pX-II open reading frames are dispensable for the immortalization of primary lymphocytes. J. Virol. 72:4458-4462.

303. Roithmann, S., C. Pique, A. Le Cesne, L. Delamarre, D. Pham, T. Tursz, and M. C. Dokhelar. 1994. The open reading frame I (ORF I)/ORF II part of the human T-cell leukemia virus type 1 X region is dispensable for p40tax, p27rex, or envelope expression. J. Virol. 68:3448-3451.

304. Rosenberg, J. R., L. Delamarre, C. Pique, D. Pham, and M. C. Dokhelar. 1997. The ectodomain of the human T-cell leukemia virus type 1 TM glycoprotein is involved in postfusion events. J. Virol. 71:7180-7186.

305. Sagawa, K., M. Mochizuki, K. Masuoka, K. Katagiri, T. Katayama, T. Maeda, A. Tanimoto, S. Sugita, T. Watanabe, and K. Itoh. 1995. Immunopathological mechanisms of human T cell lymphotropic virus type 1 (HTLV-I) uveitis - Detection of HTLV-I- infected T cells in the eye and their constitutive cytokine production. J. Clin. Invest. 95:852-858.

306. Saito, M., Y. Furukawa, R. Kubota, K. Usuku, S. Sonoda, S. Izumo, M. Osame, and M. Yoshida. 1995. Frequent mutation in pX region of HTLV-1 is observed in HAM/TSP patients, but is not specifically associated with the central nervous system lesions. J. Neurovirol. 1:286-294.

307. Sakai, J. A., M. Nagai, M. B. Brennan, C. A. Mora, and S. Jacobson. 2001. In vitro spontaneous lymphoproliferation in patients with human T-cell lymphotropic virus type I-associated neurologic disease: predominant expansion of CD8+ T cells. Blood 98:1506-1511.

308. Saxinger, W. C., W. Blattner, P. Levine, J. Clark, R. J. Biggar, M. Hoh, J. Moghissi, P. Jacobs, L. Wilson, R. Jacobson, R. Crookes, M. Strong, A. A. 76 Ansari, A. Dean, F. Nkrumah, N. Mourali, and R. C. Gallo. 1984. Human T- cell leukemia virus (HTLV-I) antibodies in Africa. Science 225:1473.

309. Scheffner, M. and N. J. Whitaker. 2003. Human papillomavirus-induced carcinogenesis and the ubiquitin-proteasome system. Semin. Cancer Biol. 13:59- 67.

310. Schreiber, G. B., E. L. Murphy, J. A. Horton, D. J. Wright, R. Garfein, H. C. Chien, and C. C. Nass. 1997. Risk factors for human T-cell lymphotropic virus types I and II (HTLV-I and -II) in blood donors: the Retrovirus Epidemiology Donor Study. NHLBI Retrovirus Epidemiology Donor Study. J Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 14:263-271.

311. Seiki, M., R. Eddy, T. B. Shows, and M. Yoshida. 1984. Nonspecific integration of the HTLV provirus genome into adult T-cell leukaemia cells. Nature 309:640-642.

312. Seiki, M., S. Hattori, Y. Hirayama, and M. Yoshida. 1983. Human adult T-cell leukemia virus: Complete nucleotide sequence of the provirus genome integrated in leukemia cell DNA. Proc. Natl. Acad. Sci. 80:3618-3622.

313. Seiki, M., A. Hikikoshi, T. Taniguchi, and M. Yoshida. 1985. Expression of the pX Gene of HTLV-1: General Splicing Mechanism in the HTLV Family. Science 228:1532-1534.

314. Seiki, M., J. Inoue, M. Hidaka, and M. Yoshida. 1988. Two cis-acting elements responsible for posttranscriptional trans-regulation of gene expression of human T-cell leukemia virus type I. Proc. Natl. Acad. Sci. U. S. A 85:7124-7128.

315. Shikama, N., J. Lyon, and N. B. Lathangue. 1997. The p300/CBP family: Integrating signals with transcription factors and chromatin. Tr. Cell Biol. 7:230- 236.

316. Shimotohno, K., M. Takano, T. Teruuchi, and M. Miwa. 1986. Requirement of multiple copies of a 21-nucleotide sequence in the U3 regions of human T-cell leukemia virus type I and type II long terminal repeats for trans-acting activation of transcription. Proc. Natl. Acad. Sci. U. S. A 83:8112-8116.

317. Shimoyama, M. 1991. Diagnostic criteria and classification of clinical subtypes of adult T- cell leukaemia-lymphoma. A report from the Lymphoma Study Group (1984- 87). Br. J. Haematol. 79:428-437.

318. Siomi, H., H. Shida, S. H. Nam, T. Nosaka, M. Maki, and M. Hatanaka. 1988. Sequence requirements for nucleolar localization of human T cell leukemia virus type I pX protein, which regulates viral RNA processing. Cell 55:197-209.

77 319. Snowden, A. W. and N. D. Perkins. 1998. Cell cycle regulation of the transcriptional coactivators p300 and CREB binding protein. Biochem. Pharmacol. 55:1947-1954.

320. Sodroski, J., C. Rosen, W. C. Goh, and W. Haseltine. 1985. A transcriptional activator protein encoded by the x-lor region of the human T-cell leukemia virus. Science 228:1430-1434.

321. Sodroski, J. G., C. A. Rosen, and W. A. Haseltine. 1984. Trans-Acting Transcriptional Activation of the Long Terminal Repeat of Human T Lymphotropic Viruses in Infected Cells. Science 223:381-385.

322. Song, K. J., V. R. Nerurkar, N. Saitou, A. Lazo, J. R. Blakeslee, I. Miyoshi, and R. Yanagihara. 1994. Genetic Analysis and Molecular Phylogeny of Simian T-Cell Lymphotropic Virus Type I: Evidence for Independent Virus Evolution in Asia and Africa. Virology 199:56-66.

323. Stuver, S. O., N. Tachibana, A. Okayama, S. Shioiri, Y. Tsunetoshi, K. Tsuda, and N. E. Mueller. 1993. Heterosexual transmission of human T cell leukemia/lymphoma virus type I among married couples in southwestern Japan: an initial report from the Miyazaki Cohort Study. J. Infect. Dis. 167:57-65.

324. Sugimoto, M., S. Mita, M. Tokunaga, K. Yamaguchi, I. Cho, M. Matsumoto, M. Mochizuki, S. Araki, K. Takatsuki, and M. Ando. 1993. Pulmonary involvement in human T-cell lymphotropic virus type-I uveitis: T-lymphocytosis and high proviral DNA load in bronchoalveolar lavage fluid. Eur. Respir. J. 6:938-943.

325. Sumida, T., F. Yonaha, T. Maeda, Y. Kita, I. Iwamoto, T. Koike, and S. Yoshida. 1994. Expression of Sequences Homologous to HTLV-I Tax Gene in the Labial Salivary Glands of Japanese Patients with Sjogrens Syndrome. Arthritis Rheum. 37:545-550.

326. Suzuki, T., J. Fujisawa, M. Toita, and M. Yoshida. 1993. The trans-activator Tax of human T-cell leukemia virus type 1 (HTLV-1) interacts with cAMP- responsive element (CRE) binding and CRE modulator proteins that bind to the 21-base-pair enhancer of HTLV-1. Proc. Natl. Acad. Sci. U. S. A. 90:610-614.

327. Suzuki, T., H. Hirai, J. Fujisawa, T. Fujita, and M. Yoshida. 1993. A trans- activator Tax of human T-cell leukemia virus type 1 binds to NF- kappa B p50 and serum response factor (SRF) and associates with enhancer of the NF- kappa B site and CArG box. Oncogene 8:2391-2397.

328. Suzuki, T., H. Hirai, T. Murakami, and M. Yoshida. 1995. Tax protein of HTLV-1 destabilizes the complexes of NF- kappa B and I kappa B-alpha and induces nuclear translocation of NF-kappa B for transcriptional activation. Oncogene 10:1199-1207. 78 329. Suzuki, T., H. Hirai, and M. Yoshida. 1994. Tax protein of HTLV-1 interacts with the Rel homology domain of NF-kappa B p65 and c-Rel proteins bound to the NF-kappa B binding site and activates transcription. Oncogene 9:3099-3105.

330. Suzuki, T., S. Kitao, H. Matsushime, and M. Yoshida. 1996. HTLV-1 Tax protein interacts with cyclin-dependent kinase inhibitor p16(INK4A) and counteracts its inhibitory activity towards CDK4. EMBO J. 15:1607-1614.

331. Suzuki, T., T. Narita, M. Uchidatoita, and M. Yoshida. 1999. Down-regulation of the INK4 family of cyclin-dependent kinase inhibitors by tax protein of HTLV- 1 through two distinct mechanisms. Virology 259:384-391.

332. Suzuki, T., M. Uchida-Toita, T. Andoh, and M. Yoshida. 2000. HTLV-1 tax oncoprotein binds to DNA topoisomerase I and inhibits its catalytic activity. Virology 270:291-298.

333. Suzuki, T., M. Uchidatoita, and M. Yoshida. 1999. Tax protein of HTLV-1 inhibits CBP/p300-mediated transcription by interfering with recruitment of CBP/p300 onto DNA element of E-box or p53 binding site. Oncogene 18:4137- 4143.

334. Takatsuki, K., T. Uchiyama, K. Sagawa, and J. Yodoi. 1977. Adult T-cell leukemia in Japan, p. 73-77. In S. Seno, F. Takaku, and S. Irino (eds.), Topics in Hematology. Excerpta Medica, Amsterdam.

335. Tan, T. H., R. Jia, and R. G. Roeder. 1989. Utilization of Signal Transduction Pathway by the Human T-Cell Leukemia Virus Type 1 Transcriptional Activator tax. J. Virol. 63:3761-3768.

336. Tanimura, A., H. Teshima, J. I. Fujisawa, and M. Yoshida. 1993. A New Regulatory Element that Augments the Tax-Dependent Enhancer of Human T- Cell Leukemia Virus Type 1 and Cloning of cDNAs Encoding Its Binding Proteins. J. Virol. 67:5375-5382.

337. Terada, K., S. Katamine, K. Eguchi, R. Moriuchi, M. Kita, H. Shimada, I. Yamashita, K. Iwata, Y. Tsuji, S. Nagataki, and T. Miyamoto. 1994. Prevalence of serum and salivary antibodies to HTLV-1 in Sjogren's syndrome. Lancet 344:1116-1119.

338. Teshigawara, K., M. Maeda, K. Nishino, T. Nikaido, T. Uchiyama, M. Tsudo, Y. Wano, and J. Yodoi. 1985. Adult T leukemia cells produce a lymphokine that augments interleukin 2 receptor expression. J Mol. Cell Immunol. 2:17-26.

339. Tillmann, M., R. Wessner, and B. Wigdahl. 1994. Identification of human T- cell lymphotropic virus type I 21-base- pair repeat-specific and glial cell-specific DNA- protein complexes. J. Virol. 68:4597-4608.

79 340. Torgeman, A., N. Morvaknin, and M. Aboud. 1999. Sp1 is involved in a protein kinase C-independent activation of human T cell leukemia virus type I long terminal repeat by 12-O-tetradecanoylphorbol-13-acetate. Virology 254:279- 287.

341. Trovato, R., J. C. Mulloy, J. M. Johnson, S. Takemoto, M. P. de Oliveira, and G. Franchini. 1999. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic/leukemia virus type 1 p12(I) protein greatly influences its stability. J Virol 73:6460-6467.

342. Tsukahara, T., M. Kannagi, T. Ohashi, H. Kato, M. Arai, G. Nunez, Y. Iwanaga, N. Yamamoto, K. Ohtani, M. Nakamura, and M. Fujii. 1999. Induction of Bcl-x(L) expression by human T-cell leukemia virus type 1 tax through NF-kappa B in apoptosis-resistant T-cell transfectants with tax. J Virol 73:7981-7987.

343. Uchiyama, T., J. Yodoi, K. Sagawa, K. Takatsuki, and H. Uchino. 1977. Adult T-cell leukemia: clinical and hematologic features of 16 cases. Blood 50:481-492.

344. Van Dooren, S., M. Salemi, and A. M. Vandamme. 2001. Dating the origin of the African human T-cell lymphotropic virus type-i (HTLV-I) subtypes. Mol. Biol. Evol. 18:661-671.

345. Van Orden, K., H. A. Giebler, I. Lemasson, M. Gonzales, and J. K. Nyborg. 1999. Binding of p53 to the KIX domain of CREB binding protein. A potential link to human T-cell leukemia virus, type I-associated leukemogenesis. J. Biol. Chem. 274:26321-26328.

346. Vanbrussel, M., M. Salemi, H. F. Liu, J. Gabriels, P. Goubau, J. Desmyter, and A. M. Vandamme. 1998. The simian T-lymphotropic virus STLV-PP1664 from Pan paniscus is distinctly related to HTLV-2 but differs in genomic organization. Virology 243:366-379.

347. Vandamme, A. M., M. Salemi, and J. Desmyter. 1998. The simian origins of the pathogenic human T-cell lymphotropic virus type I. Trends Microbiol. 6:477- 483.

348. Vandel, L. and D. Trouche. 2001. Physical association between the histone acetyl transferase CBP and a histone methyl transferase. EMBO rep. 2:21-26.

349. Vo, N. and R. H. Goodman. 2001. CREB-binding protein and p300 in transcriptional regulation. J. Biol. Chem. 276:13505-13508.

350. Vogel, K. W., R. Briesewitz, T. J. Wandless, and G. R. Crabtree. 2001. Calcineurin inhibitors and the generalization of the presenting protein strategy. Adv. Protein Chem. 56:253-291.

80 351. von Bulow, G. U. and R. J. Bram. 1997. NF-AT activation induced by a CAML-interacting member of the tumor necrosis factor receptor superfamily. Science 278:138-141.

352. Wang, J. K., E. Kiyokawa, E. Verdin, and D. Trono. 2000. The Nef protein of HIV-1 associates with rafts and primes T cells for activation. Proc. Natl. Acad. Sci. U. S. A 97:394-399.

353. Wattel, E., J. P. Vartanian, C. Pannetier, and S. Wainhobson. 1995. Clonal expansion of human T-Cell leukemia virus type I- infected cells in asymptomatic and symptomatic carriers without malignancy. J. Virol. 69:2863-2868.

354. Weichselbraun, I., J. Berger, M. Dobrovnik, H. Bogerd, R. Grassmann, W. C. Greene, J. Hauber, and E. Bohnlein. 1992. Dominant-negative mutants are clustered in a domain of the human T-cell leukemia virus type I Rex protein: implications for trans dominance. J Virol 66:4540-4545.

355. Wessner, R., M. Tillmann-Bogush, and B. Wigdahl. 1995. Characterization of a glial cell-specific DNA-protein complex formed with the human T cell lymphotropic virus type I (HTLV-I) enhancer. J Neurovirol. 1:62-77.

356. Wessner, R. and B. Wigdahl. 1997. AP-1 derived from mature monocytes and astrocytes preferentially interacts with the HTLV-I promoter central 21 bp repeat. Leukemia 11 Suppl 3:21-4.:21-24.

357. White, J. D., S. L. Zaknoen, C. Kastensportes, L. E. Top, L. Navarroroman, D. L. Nelson, and T. A. Waldmann. 1995. Infectious complications and immunodeficiency in patients with human T-cell lymphotropic virus I-associated adult T- cell leukemia/lymphoma. Cancer 75:1598-1607.

358. Wiktor, S. Z., E. J. Pate, E. L. Murphy, T. J. Palker, E. Champegnie, A. Ramlal, B. Cranston, B. Hanchard, and W. A. Blattner. 1993. Mother-to-child transmission of human T-cell lymphotropic virus type I (HTLV-I) in Jamaica: association with antibodies to envelope glycoprotein (gp46) epitopes. J. Acquir. Immune. Defic. Syndr. 6:1162-1167.

359. Wilks, R., B. Hanchard, O. Morgan, E. Williams, B. Cranston, M. L. Smith, P. Rodgersjohnson, and A. Manns. 1996. Patterns of HTLV-I infection among family members of patients with adult T-cell leukemia/lymphoma and HTLV-I associated myelopathy tropical spastic paraparesis. Int. J. Cancer 65:272-273.

360. Willems, L., P. Kerkhofs, F. Dequiedt, D. Portetelle, M. Mammerickx, A. Burny, and R. Kettmann. 1994. Attenuation of bovine leukemia virus by deletion of R3 and G4 open reading frames. Proc. Natl. Acad. Sci U. S. A 91:11532-11536.

81 361. Wolin, M., M. Kornuc, C. Hong, S. K. Shin, F. Lee, R. Lau, and S. Nimer. 1993. Differential effect of HTLV infection and HTLV Tax on interleukin 3 expression. Oncogene 8:1905-1911.

362. Wood, G. S., A. Salvekar, J. Schaffer, C. F. Crooks, W. Henghold, D. P. Fivenson, Y. H. Kim, and B. R. Smoller. 1996. Evidence against a role for human T-cell lymphotrophic virus type I (HTLV-I) in the pathogenesis of American cutaneous T-cell lymphoma. J. Invest. Dermatol. 107:301-307.

363. Wood, G. S., J. M. Schaffer, R. Boni, R. Dummer, G. Burg, M. Takeshita, and M. Kikuchi. 1997. No evidence of HTLV-I proviral integration in lymphoproliferative disorders associated with cutaneous T- cell lymphoma. Am. J. Pathol. 150:667-673.

364. Wu, Y. and J. W. Marsh. 2001. Selective transcription and modulation of resting T cell activity by preintegrated HIV DNA. Science 293:1503-1506.

365. Wucherpfennig, K. W., P. Hollsberg, J. H. Richardson, D. Benjamin, and D. A. Hafler. 1992. T-cell activation by autologous human T-cell leukemia virus type I- infected T-cell clones. Proc. Natl. Acad. Sci. U. S. A 89:2110-2114.

366. Wysocka, J. and W. Herr. 2003. The herpes simplex virus VP16-induced complex: the makings of a regulatory switch. Trends Biochem. Sci 28:294-304.

367. Xu, R. Z., Q. K. Gao, S. J. Wang, H. H. Kan, L. Y. Sheng, C. Q. Li, X. H. Zhang, G. B. Xu, and K. Q. Zhang. 1996. Human acute myeloid leukemias may be etiologically associated with new human retroviral infection. Leuk. Res. 20:449-455.

368. Yajima, A., A. Kawada, Y. Aragane, and T. Tezuka. 2001. Detection of HTLV-I proviral DNA in sarcoidosis. Dermatology 203:53-56.

369. Yamaguchi, K. 1994. Human T-lymphotropic virus type I in Japan. Lancet 343:213-216.

370. Yamaguchi, K., H. Nishimura, H. Kohrogi, M. Jono, Y. Miyamoto, and K. Takatsuki. 1983. A proposal for smoldering adult T-cell leukemia: a clinicopathologic study of five cases. Blood 62:758-766.

371. Yamaguchi, K. and K. Takatsuki. 1993. Adult T cell leukaemia-lymphoma. Clin. Haematol. 6:899-915.

372. Yamaguchi, K. and K. Takatsuki. 1993. Adult T cell leukaemia-lymphoma. Clin. Haematol. 6:899-915.

82 373. Yamamoto, N., M. Okada, Y. Koyanagi, M. Kannagi, and Y. Hinuma. 1982. Transformation of human Leukocytes by cocultivation with an adult T cell leukemia virus producer cell line. Science 217:737-739.

374. Yamano, Y., M. Nagai, M. Brennan, C. A. Mora, S. S. Soldan, U. Tomaru, N. Takenouchi, S. Izumo, M. Osame, and S. Jacobson. 2002. Correlation of human T-cell lymphotropic virus type 1 (HTLV-1) mRNA with proviral DNA load, virus-specific CD8(+) T cells, and disease severity in HTLV-1-associated myelopathy (HAM/TSP). Blood 99:88-94.

375. Yamato, K., T. Oka, M. Hiroi, Y. Iwahara, S. Sugito, N. Tsuchida, and I. Miyoshi. 1993. Aberrant Expression of the p53 Tumor Suppressor Gene in Adult T- Cell Leukemia and HTLV-I-Infected Cells. Jpn. J. Cancer Res. 84:4-8.

376. Ye, J., L. Silverman, M. D. Lairmore, and P. L. Green. 2003. HTLV-1 Rex is required for viral spread and persistence in vivo but is dispensable for cellular immortalization in vitro. Blood.

377. Yoshida, M., I. Miyoshi, and Y. Hinuma. 1982. Isolation and characterization of retrovirus from cell lines of human T-cell leukemia and its implication in the disease. Proc. Natl. Acad. Sci. USA 79:2031-2035.

378. Yoshida, M., M. Seiki, K. Yamaguchi, and K. Takatsuki. 1984. Monoclonal integration of human T-cell leukemia provirus in all primary tumors of adult T- cell leukemia suggests causative role of human T-cell leukemia virus in the disease. Proc. Natl. Acad. Sci. U. S. A. 81:2534-2537.

379. Yoshimura, T., J. Fujisawa, and M. Yoshida. 1990. Multiple cDNA clones encoding nuclear proteins that bind to the tax-dependent enhancer of HTLV-I: all contain a leucine zipper structure and basic amino acid domain. EMBO J. 9:2537- 2542.

380. Zhang, W., J. W. Nisbet, B. Albrecht, W. Ding, F. Kashanchi, J. T. Bartoe, and M. D. Lairmore. 2001. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J. Virol. 75:9885-9895.

381. Zhang, W., J. W. Nisbet, J. T. Bartoe, W. Ding, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J. Virol. 74:11270-11277.

382. Zhao, L. J. and C. Z. Giam. 1992. Human T-cell lymphotropic virus type I (HTLV-I) transcriptional activator, Tax, enhances CREB binding to HTLV-I 21- base-pair repeats by protein-protein interaction. Proceedings. of. the. National. Academy. of. Sciences. of. the. United. States. of. America. 89:7070-7074.

83 383. Zhong, H., R. E. Voll, and S. Ghosh. 1998. Phosphorylation of NF-kappa B p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with the coactivator CBP/p300. Mol. Cell 1:661-671.

384. Zhou, J. and C. Aiken. 2001. Nef enhances human immunodeficiency virus type 1 infectivity resulting from intervirion fusion: evidence supporting a role for Nef at the virion envelope. J Virol 75:5851-5859.

385. Zhu, J. and F. Mckeon. 2000. Nucleocytoplasmic shuttling and the control of NF-AT signaling. Cell Mol. Life Sci. 57:411-420.

84

Figure 1.1. Diagram of HTLV-1 genome and alternatively spliced mRNA and protein species from pX ORFs. Numbers correspond to nucleotide positions of exon splice acceptor and donor sites with respect to the full length HTLV-1 genome. Dotted lines indicate the mRNA while the boxes below the dotted lines indicate coding regions of each protein. Open reading frame from which each protein is produced, is indicated as superscript, on the right of each protein.

85

Figure 1.2. Diagram of p12I, p30II, and p13II with predicted functional motifs. Abbreviations: TM, transmembrane region; LZip, leucine zipper motif; DxxxLL, dileucine motif; PxxP, SH3 binding motif; K/R, lysine-to-arginine variant at position 88 (arginine at this position increases stability of protein). Boxes below p30II indicate regions sharing sequence homology with the DNA binding domain and homeodomain of Oct-1. Numbers below bars indicate amino acid position numbers.

86

CHAPTER 2

HUMAN T LYMPHOTROPIC VIRUS TYPE-1 P30II INFLUENCES CELLULAR GENE EXPRESSION AND ENHANCES TRANSCRIPTION ACTIVATION OF T LYMPHOCYTES

INTRODUCTION

Human T-lymphotropic virus type 1 (HTLV-1), the first characterized human

retrovirus, causes adult T cell leukemia/lymphoma (ATL) and is associated with several

lymphocyte-mediated disorders such as HTLV-1-associated myelopathy/tropical spastic

paraparesis (HAM/TSP)7. Mature CD4+ T lymphocytes are the primary targets of HTLV-

1 infection25. Although the mechanism by which the virus causes oncogenic

transformation of host T lymphocytes is incompletely understood, altered gene

expression has been associated with the initiation or progression of ATL 47. This complex

retrovirus encodes structural and enzymatic gene products, as well as regulatory and

accessory proteins from open reading frames (ORF) in the pX region between env and the

3’ long terminal repeat (LTR) of the provirus19. The well characterized Rex and Tax

proteins are encoded in the ORF III and IV respectively. Rex is a nucleolus-localizing

87 phosphoprotein, involved in nuclear export of unspliced or singly spliced viral RNA 24.

Tax is a nuclear localizing phosphoprotein that interacts with cellular transcription factors and activates transcription from the viral promoter, Tax-responsive element (TRE) and enhancer elements of various cellular genes associated with host cell proliferation 22.

Emerging evidence has documented the role of pX ORF I and II gene products in the replication of HTLV-13. There are four proteins expressed from these ORFs- p12I, p27I, p13II, and p30II. pX ORFs I and II mRNAs are present in infected cell lines and freshly isolated cells from HTLV-1-infected subjects36 as well as in ATL and HAM/TSP patients12. Antibodies13,16 and cytotoxic T cells46 that recognize recombinant proteins or

peptides of the pX ORF I and II proteins are present in HTLV-1 infected patients and

asymptomatic carriers.

Using molecular clones of HTLV-1 with selective mutations of ORF I and II, we have tested the requirement of p12I and p13II/p30II in the establishment of infection and

maintenance of viral loads in a rabbit model of infection 1 9. ORF II protein p30II contains a highly conserved bipartite nuclear localization signal (NLS) and localizes within the nucleus of cells 35,58. In addition, p30II contains serine- and threonine-rich regions with

distant homology to transcription factors Oct-1 and -2, Pit-1, and POU-M114. p30II also co-localizes with p300 in the nucleus and physically interacts with CREB binding protein

(CBP)/p300 and differentially modulates cAMP responsive element (CRE) and TRE mediated transcription 57 58. Based on these reports, we hypothesized that p30II functions as a regulator of cellular and viral gene expression to promote HTLV-1 replication.

88 Gene arrays have primarily been employed to study the changes in gene

expression profile of HTLV-1-immortalized and transformed cell lines or in cells from

ATL patients and attempts to test the influence of individual HTLV-1 viral proteins on

cellular gene expression have been limited to Tax 15,23,32,33,33,44,47. Herein we used the

Affymterix U133A human gene chip to confirm the role of p30II as a regulator of gene

expression and identified several novel and important alterations in gene expression

profiles, unique to cell cycle regulation, apoptosis and T cell signaling/ activation. In

addition, using semi-quantitative RT-PCR, we have confirmed the expression of multiple

genes modulated by p30II in primary CD4+ T lymphocytes. Moreover, we have validated

the function of p30II in T cell signaling/ activation, using reporter assays. This is the first

report that demonstrates the role of p30II as an activator of many key transcription factors

involved in T cell signaling/ activation, such as Nuclear Factor of Activated T cells

(NFAT), Nuclear Factor-Kappa B (NF-κB) and Activator Protein-1 (AP-1). Together, our data indicates that HTLV-1, a complex retrovirus associated with lymphoproliferative disorders, uses accessory genes to promote lymphocyte activation to enhance clonal expansion of infected cells and maintain proviral loads in vivo.

89 MATERIALS AND METHODS

Lentiviral vectors and other plasmids

The plasmid pWPT-IRES-GFP was generated by cloning the internal ribosome entry site (IRES) sequence from pHR’CMV/Tax1/eGFP 54 (Gerald Feuer, SUNY,

Syracuse) into pWPT-GFP plasmid (Didier Trono, University of Geneva). Subsequently, the plasmid pWPT-p30IIHA-IRES-GFP was created by cloning p30II sequence from

ACH30 with the downstream influenza hemagglutinin (HA1) tag (Fig. 2.1). Sanger sequencing confirmed both the plasmids to have the correct sequence and were in frame.

GFP and p30IIHA expression were confirmed by fluorescence activated cell sorting

(FACS) analysis (Beckman Coulter, Miami, FL) and western blot respectively. GFP expression from each of the plasmids was confirmed by flow cytometry (Beckman

Coulter) and the p30IIHA expression from pWPT-p30IIHA-IRES-GFP plasmid was confirmed by western blot using mouse monoclonal anti- hemagglutinin antibody

(1:1000) (Covance, Princeton, NJ) as described previously57,58. The plasmid pME- p30IIHA was created by cloning p30II sequence from ACH with HA1 tag, into pME-18S

(G. Franchini, NIH). Other plasmids used include previously reported pRSV-βGal 58and

AP-1, NF-κB and NFAT-luciferase reporter plasmids 2.

Recombinant lentivirus production and infection of Jurkat T lymphocytes and primary CD4+ T lymphocytes

Recombinant lentiviruses was produced by transfecting pHCMV-G, pCMV∆R8.2 and pWPT-p30IIHA-IRES-GFP (sample) or pWPT-IRES-GFP (control) as described

90 previously18. Briefly, 293T cells (5 × 106) were seeded in a 10-cm dish and transfected the following day with 2 µg of pHCMV-G, 10 µg of pCMV∆R8.2 and 10 µg of pWPT- p30IIHA-IRES-GFP or pWPT-IRES-GFP using the calcium phosphate method.

Supernatant from 10 to 20 dishes was collected at 24, 48 and 72 h post transfection, cleared of cellular debris by centrifugation at 1000 rpm for 10 min at room temp and then filtered through a 0.2 µm filter. The resulted supernatant was then centrifuged at 6,500 g for 16 h at 4° C. The viral pellet was suspended in cDMEM (DMEM containing 10%

FBS and 10% streptomycin and penicillin) overnight at 4° C and the concentrated virus was aliquoted and stored at –80° C. To determine the virus titer, serial dilutions of the virus stock were used to spin infect 293T cells and 48 h post infection, eGFP expression and p30II expression was measured by flow cytometry and RT-PCT respectively. Briefly,

on the day before infection, 293T cells (1 × 105) were seeded in a 6 well plate. The medium was removed the following day and the cells were then incubated with the

diluted virus containing 8 µg/ml polybrene (Sigma, St. Louis, MO). Cells were then spin-

infected by centrifugation at 2700 rpm for 1 h at 30o C, supplied with fresh medium and

cultured for 48 h. Then cells were treated with trypsin (Invitrogen, Carlsbad, CA),

pelleted and resuspended in D-PBS (Invitrogen) for fluorescence activity cell sorting

(FACS) analysis on an ELITE ESP flow cytometer (Beckman Coulter). 1 X 106 cells were used to perform western blot to detect the expression of p30 II HA. Jurkat T

lymphocytes (clone E6.1, American Type Culture Collection) were transduced with

recombinant virus at multiplicity of infection of 4 in the presence of 8 µg/ml polybrene

(Sigma) and spin-infected at 2700 rpm for 1 h at 25o C. Primary CD4+ T cells were

extracted using dynabead CD4 positive isolation kit (Dynal Biotech, Lake Success, NY) 91 according to manufacturer’s instructions. Primary CD4+ T cells were stimulated with

Phytohemagglutinin (PHA) for 48 hours, transduced with recombinant virus at

multiplicity of infection of 20 in the presence of 8 µg/ml polybrene (Sigma) and spin- infected at 2700 rpm for 1 h at 25o C. At 10 days post-transduction, GFP expression of

controls and samples were verified to be above 90% by FACS analysis and the presence of p30II mRNA expression in samples (and absence in controls) was verified by RT-PCR

(Fig. 2.2).

Probe preparation and microarray analysis

According to the instructions of manufacturers, total cellular RNA was isolated

from transduced Jurkat T lymphocytes using RNAqueous (Ambion, Austin, TX). To test

the concentration and purity of the RNA samples, absorbance at 260 nm and 280 nm

were measured and the 260/280 ratio was calculated using a spectrophotometer

(Genequant, Amersham Pharmacia, Piscataway, NJ). The 260/280 ratio of all the RNA

samples were between the range of 1.9-2.1. The probe preparation for GeneChip was

performed according to the Affymetrix GeneChip Expression Analysis Technical Manual

(Affymetrix, Santa Clara, CA). Briefly, cDNA was synthesized using genechip T7-Oligo

(dT) promoter primer kit (Affymetrix) and superscript double stranded cDNA synthesis

kit (Invitrogen), according to the manufacturers instructions. cDNA cleanup was done

using Genechip Sample Cleanup module (Affymetrix). In vitro transcription was

performed on the cDNA to produce biotin-labeled cRNA with ENZO RNA Transcript

labeling kit (Affymetrix), according to the manufacturers instructions. cRNA cleanup was

92 done using Genechip Sample Cleanup module (Affymetrix). The quality of total RNA and biotin-labeled cRNA of all the samples and controls were checked by calculating the

ratio of absorbance at 260 nm and 280 nm (between 1.9 to 2.1) using a spectrophotometer

(Genequant) and agarose gel electrophoresis. The labeled cRNA was fragmented to 50–

200 nucleotides, and hybridized to U133A arrays (Affymetrix) using GeneChip®

Hybridization Oven (Affymetrix). Arrays were washed and stained using GeneChip®

Fluidics Station 400 (Affymetrix) and scanned by GeneArray Scanner (Affymetrix).

Quality control criteria evaluations done as part of the basic analysis include (1)

The ratios of 3’ signal to 5’ signal of two housekeeping genes, beta-actin and GAPDH

were between 0 and 3. (2) The hybridization controls BioB, BioC, BioD, and Cre were all

present and in a linear relationship of intensity. (3) The scale factors between arrays did

not vary by 3 fold. (4) The background intensity was not significantly higher than

expected. (5) The percent of gene present was monitored and found to be not less than

the standard 30%. To determine the quantitative RNA level, the average differences representing the perfectly matched minus the mismatched for each gene-specific probe set

was calculated. Differential gene expression and comparative analysis was done using

Data Mining Tool (Microarray suite 5) to identify probes with at least 1.5 fold

difference in expression between control and p30II and verified for cluster formation by

dCHIP software39. The biological and molecular functional grouping of these probes was done using Gene Ontology Mining Tool (Affymetrix)6. The selected probe lists are presented in the supplemental data.

93 RT-PCR

One µg of RNA was converted to cDNA (Reverse Transcription system,

Promega, Madison, WI) as described by the manufacturer. cDNA from 100 ng of total

RNA was amplified with AmpliTaq DNA polymerase (Perkin Elmer, Boston, MA), PCR

products were separated by agarose gel electrophoresis, normalized to GAPDH and

quantified using alpha imager spot densitometry (Alpha Innotech, San Leandro, CA).

DNA contamination was tested by performing a control with no reverse transcriptase.

The PCR primers for p30II were as follows: TAG CAA ACC GTC AAG CAC AG

(forward) and CGA ACA TAG TCC CCC AGA GA (reverse). The PCR primers for CHP

were as follows: CCC ACA GTC AAA TCA CTC GCC (forward) and ATG GTC CTG

TCT GCG ATG CTG (reverse). The PCR primers for JUN were as follows: CTC TCA

GTG CTT CTT ACT ATT AAG CAG (forward) and TTA TCT AGG AAT TGT CAA

AGA GAA GATT (reverse). The PCR primers for NFATc were as follows: TTG GGA

GAG ACA TGT CCC AGA TT (forward) and TCA TTT CCC CAA AGC TCA AAC A

(reverse). The results were expressed as a graph. Statistical analysis was performed using

Student’s t test, P < 0.05.

Transient transfection and reporter gene assay

Analysis of AP-1, NF-κB, and NFAT transcriptional activity in pME- and pME- p30II-transfected Jurkat T lymphocytes was performed as described previously (25).

Briefly, transient transfection of Jurkat T lymphocytes was done by electroporating 107

94 cells in cRPMI (RPMI 1640 containing 10% fetal bovine serum (FBS) and 10%

streptomycin and penicillin) at 350V and 975 µF using Bio-Rad Gene Pulser II (Bio-

Rad,Laboratories, Hercules, CA) with 30 µg of pME-p30 or pME empty plasmid, 10 µg

of reporter plasmid (NFAT-Luc, AP-1 Luc or NF-κB Luc), and 1 µg of pRSV-Gal

plasmid or 1 ug pWPT-IRES-GFP plasmid. The transfected cells were seeded in six-well

plates at a density of 5 X 105/ml and were either left untreated or stimulated with 20

ng/ml of phorbol myristate acetate (PMA) (Sigma) or with 2 µM ionomycin (Sigma), or both at 6 h post-transfection, followed by incubation for 18 h prior to lysis for analysis of luciferase activity. Stimulations with anti-CD3 and/or anti-CD28 antibodies (each at 3

µg/ml) (BD Pharmingen, Heidelberg, Germany) were carried out 18 h post-transfection.

Following 8 h of stimulation, to measure luciferase activity, the cells were lysed with Cell

Culture Lysis Reagent (Promega), and the cell lysates were tested for luciferase activity according to the manufacturer’s protocol. Transfection efficiency was normalized by staining with 5-bromo-4-chloro-3-indolyl-beta-D-galactopyranoside (X-Gal) (Sigma) and counting β-Gal expressing cells. Transfection efficiency was also normalized by counting

GFP positive cells under the fluorescence microscope. Results were expressed as mean of optimized luciferase activity (luciferase activity / percentage cells stained positive for β-

Gal expression) in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed using Student’s t test, P <

0.05.

95 RESULTS

At 10 days post-transduction, GFP expression of Jurkat T lymphocytes transduced

with recombinant lentivirus expressing GFP alone (controls) or p30II and GFP (samples)

was confirmed to be > 95%. Presence of expression of p30II mRNA in the sample cells

and absence of p30II mRNA expression in control cells were confirmed by RT-PCR.

Differential gene expression and comparative analysis was done to identify probes with at least 1.5 fold difference in expression between control and p30II and verified for cluster

formation39. We categorized genes deregulated by HTLV-1 p30II into those upregulated

or downregulated in expression. We further grouped genes deregulated by p30II based on

their functions, such as apoptosis, cell cycle, cell adhesion, transcription / translation

factors and T cell activation or cell signaling. In all the categories, p30II appeared to be an

overall repressor of cellular gene expression, while selectively increasing the expression

of certain key regulatory genes (Table 2.1).

Apoptosis

Based on changes in gene expression in p30II expressing cells (Table 2.1), p30II would be predicted to modulate apoptosis. These include Bcl-2 related/interacting genes such as pro-apoptotic Bcl-2-related protein A1, anti-apoptotic MCL1, cell-death regulator

Harakiri, apoptotic protector BNIP1 (downregulated) and pro-apoptotic BIK

(upregulated). In addition, p30II expression correlated with downregulation of genes

associated with Fas mediated apoptosis pathway such as tumor suppressing

subtransferable candidate 3 and TNF receptor superfamily member 25. p30II expression

96 was also associated with decrease in caspases (2 and 4) and increase in genes associated

with the DNA fragmentation pathway (CIDE-B and CIDE-3). In addition, p30II expression correlated with decreased expression of many other apoptosis related genes including CD28, Lck, cyclin B1, Cullin 5, Adenosine A2a receptor, TAF4B and NCK- associated protein 1.

Cell cycle

p30II expression altered the expression of multiple genes involved in regulation of

different stages of cell cycle. These include checkpoint suppressor 1, cytosolic branched- chain amino acid transaminase 1, histone deacetylase 6, cyclin B1, WEE1 kinase,

CDC14A, Lck, JAK2, GAS7, BZAP45, Cullin, Rab6 GTPase activating protein

(downregulated) and TERF1, AKAP8, DDX11, MSH2 and JUN (upregulated). Another gene downregulated by p30II expression was MDM2, which, interestingly is

overexpressed in certain leukemias34 and capable of enhancing the tumorigenic potential

of cells by inhibiting p300 / PCAF mediated p53 acetylation 40.

Cell adhesion

p30II expression was associated with altered expression of several genes involved

in cell adhesion. These include decrease in integrin (integrin β8) immunoglobulin

(MADCAM1), a counter-receptor for P-selectin (SELPLG), cadherin (desmocollin 3),

protocadherin (PC-LKC) liprin (PPF1BP1), CD84 / Ly-9, CD58, CD43 / sialophorin and

glycosyl-phosphatidyl-inositol phospholipase D1. Expression of p30II correlated with increase in integrin receptor α1 subunit and KIT ligand.

97 Transcription and translation

Expression of p30II was associated with decrease in expression of transcriptional

control genes such as TATA-binding protein associated factor 4 (TAF4), two co-

repressors (Enolase-1 and Chromosome 19 ORF2 protein), a novel specific coactivator

for mammalian TEFs, namely TONDU 53, homeo box genes (mesenchyme homeo box 1,

homeobox A1), T-box genes (T-box 21) and proteins containing helix-loop-helix domain,

which are known to be critical in cell growth/differentiation and tumorigenesis (neuronal

PAS domain protein 2, Myc-associated factor protein, inhibitor of DNA binding-3).

Additionally, p30II expression correlated with downregulation of proteins

(zinc finger protein 36), a group of transcription regulators proposed to be candidates in

malignant disorders 51 and coiled coil proteins (JEM-1). p30II was also associated with downregulation of many genes with positive transcriptional effects (including SEC14- like 2, Nurr 1, CITED2 / MRG1, LXR alpha and SMARCA2). Downregulation of

HDAC6, a histone deacetylase and nuclear receptor coactivator 3 (CBP interacting protein) with histone acetyltransferase and pCAF/CBP recruiting abilities 50 are

particularly interesting, since p30II contains multiple highly conserved lysines, which

could play a role in acetylation 57,58. Expression of p30II was associated with decrease in

GAS 7, which has sequence homology to Oct and POU family of transcription factors27.

Genes upregulated by p30II also included HTLV enhancer factor, c-Jun, TAF1C,

Kruppel-type zinc finger, PQBP1, AF4 and SOX4. Based on this, p30II appears to

regulate transcription through various mechanisms and different levels, some of which

could be attributed to the interaction between p30II and CBP/p30057. Additionally, p30II

98 expression was associated with decrease in translation initiation factor 2 (IF2) and

eukaryotic translation elongation factor 1δ (EEF1D) and increase in eukaryotic

translation elongation factor 1α (EEF1A2), a putative oncogene 4

p30II alters the expression of genes involved in T cell signaling / activation

Expression of p30II was associated with decrease in mRNA levels of CD28, a co- stimulatory molecule with a distinct role in T lymphocyte activation 10. Cells expressing

p30II had upregulated levels of Vav-2 and CD72 and downregulated levels of CD46 and

Lck tyrosine kinase, a member of the Src family protein tyrosine kinases that participates

in signal transduction pathways initiated by T cell surface receptors such as TCR/CD3,

CD2, CD4, CD8, and CD28 43. Additionally, p30II expression correlated with decrease in

the level of CHP, an endogenous calcineurin inhibitor, indicating a likely increase in

calcineurin activity, which would be predicted to correlate with upregulation of NFAT

expression by p30II. Moreover, p30II expression was associated with increased expression

of c-Jun and c-Fos, suggesting activation of AP-1 mediated transcription. p30II appeared

to downregulate the expression of protein kinase D (PKD), which negatively modulates

JNK signaling pathway 26, mediates cross-talk between different signaling systems, and is

critical in processes as diverse as cell proliferation, apoptosis, immune cell regulation and

tumor cell invasion 52. Interestingly, p30II expression was also associated with decrease in

epidermal growth factor (EGF), which stimulates calcium influx 56 and activates ETS

domain transcription factor Elk-1 48. In Jurkat T lymphocytes expressing p30II, there were

no detectable levels of IKKγ, which is important for NF-κB signaling in response to both

99 T cell activation signals and Tax 49. However p30II expression increased HPK-1

(Hematopoetic Progenitor Kinase-1), a known NF-κB activator5. p30II expression was

associated with decreased Ras GRP2, a guanyl nucleotide exchange factor that increases

Ras-GTP, suggesting a decrease in the level of activated Ras (Ras-GTP).

We performed semiquantitative RT-PCR analysis in Jurkat T lymphocytes and

primary CD4+ T lymphocytes to confirm the altered expression of genes involved in T cell activation / signaling pathway (Fig. 2.3A, 2.3B, 2.3C and 2.3D). Overall, HTLV-1 p30II downregulated the expression of multiple upstream regulators of T cell activation / signaling pathway to enhance NFAT, NF-κB and AP-1 driven transcription, key factors in T cell proliferation and leukemogenesis.

HTLV-1 p30II enhances transcription mediated by NFAT, NF-κB and AP-1 in

Jurkat T lymphocytes.

Using luciferase reporter assays, we directly tested the ability of p30II to enhance

NFAT, NF-κB and AP-1 driven transcription. NF-κB, AP-1, and NFAT luciferase

reporter plasmids and p30II expression plasmid were transiently transfected into Jurkat T

lymphocytes, and stimulated with PMA or ionomycin or both, CD3 or CD28 or both.

p30II increased the NFAT driven luciferase reporter gene activity from 2.2 to 10.7 fold

depending on co-stimulatory treatment e.g., PMA, ionomycin, CD3, CD28 etc. (Fig.

2.4A), indicating that p30II effectively enhanced NFAT driven transcription, when

stimulated with ionomycin or CD3. HTLV-1 p30II increased NF-κB driven luciferase

reporter gene activity from 3.1 to 11.4 fold, depending on co-stimulation (Fig. 2.4B),

100 indicating that p30II is able to effectively enhanced NF-κB driven transcription,

significantly even under unstimulated conditions. HTLV-1 p30II modestly increased the

AP-1 driven luciferase reporter gene activity from 1.2 to 5.2 fold in the presence of co-

stimulator treatments (Fig. 2.4C), indicating that p30II enhanced AP-1 driven

transcription, significantly when stimulated with PMA.

DISCUSSION

Our study is the most comprehensive analysis of changes in gene expression patterns caused by an individual retrovirus accessory protein. Our approach included methods to strengthen the reliability of our data by (a) use of triplicate samples and appropriate controls (b) use of multiple software for data analysis (c) minimization of

nonspecific hybridization and background signals by using Affymetrix chip 42 (d) use of a

well-characterized T lymphocyte system (Jurkat) and (e) verification of microarray data

by semiquantitative RT-PCR in Jurkat T lymphocytes and primary CD4+ T lymphocytes

(f) validation of microarray data by reporter assays, all of which were consistent with our

micro array findings. Some of these findings are consistent with previous studies using

gene arrays to test HTLV-1-transformed cell lines. For example, HTLV-1 infected cell

lines contain low levels of caspase-4 and high levels of JUN 47 and cyclin B1 levels are

low in HTLV-1 leukemic T cells 28. Overall, this study not only confirmed that p30II is a

regulator of cellular genes, but also identified several potential new functional roles for

p30II, in T cell activation or cell signaling and in regulation of apoptosis and cell cycle.

101 Others have used gene array approaches to study HTLV-1-related changes in gene

expression. Harhaj et al 23 studied the gene expression in HTLV-1 mediated oncogenesis

using human cDNA array analysis of normal and HTLV-1 immortalized T cells and

found that the expression of a large number of genes involved in apoptosis were deregulated in HTLV-1 immortalized T cells. Subsequently, the same type of cDNA arrays were employed by De La Fuente et al 15 to study upregulation of a number of

transcription factors in HTLV-1-infected cells, including zinc fingers, paired domains,

and basic helix-loop-helix (bHLH) proteins. Gene expression profiles of fresh peripheral

blood mononuclear cells (PBMC) from acute and chronic ATL patients were used to

identify the genes associated with progression of ATL including a T cell differentiated

antigen (MAL), a lymphoid specific member of the G-protein-coupled receptor family

(EBI-1/CCR7) and a novel human homolog to a subunit (MNLL) of the bovine

ubiquinone oxidoreductase complex31. Using NIH OncoChip cDNA arrays containing

2304 cancer related cDNA elements, Ng et al, 2001 44 compared normal and Tax-

expressing Jurkat T lymphocytes and identified Tax induced changes in gene expression,

associated with apoptosis, cell cycle, DNA repair, signaling factors, immune modulators,

cytokines, growth factors, and adhesion molecules. Recently, Affymetrix, GeneChip

microarrays containing oligonucleotide hybridization probes representative of ~7000

genes were used to compare the expression profiles of normal activated peripheral blood

lymphocytes to HTLV-I-immortalized and transformed cell lines 47. In this study, by

employing a gene chip representing ~33000 genes, we tested the role of p30II on cellular

gene expression profile of a larger number of genes.

102 Expression from the IL-2 promoter requires binding of several transcription factors, including NFAT, AP-1 and NF-κB. NFAT is vital to proliferation of peripheral lymphocytes for HTLV-1 infection 55 while AP-1 is linked to the dysregulated

phenotypes of HTLV-1 infected T cells 20 and malignant transformation 41. Activation of

AP-1 occurs through Tax-dependent and independent mechanisms in HTLV-1-infected T cells in vitro and in leukemia cells in vivo 20. NF-κB is highly activated in many

hematopoietic malignancies, HTLV-1 infected T cell lines and in primary ATL cells, even when Tax expression levels are low 41 and due to its anti-apoptotic activity, it is

considered to be a key survival factor for several types of cancer. Ours is the first report

demonstrating the ability of a retroviral accessory protein to have broad modulating

activities on the transcriptional activity of NF-κB, NFAT and AP-1.

Mechanisms employed by p30II to enhance NFAT, AP-1 and NF-κB mediated

transcription will require additional research. We have previously reported that another

HTLV-1 accessory protein p12I stimulates NFAT mediated transcription, when

stimulated with PMA, indicating that p12I acts synergistically with Ras/ MAPK pathway

to promote NFAT activation and thus may facilitate host cell activation and establishment

of persistent HTLV-1 infection 2 However p30II enhances NFAT driven transcription

significantly when stimulated with ionomycin or CD3, and therefore likely uses a

different mechanism than p12I. To modulate NFAT driven transcription and subsequent T

cell activation / signaling, it is possible that these two accessory proteins act

synergistically. AP-1 is able to interact with transcriptional coactivator CBP/ p300 8 as well as viral CREs and mediate HTLV-I gene expression (50). Intriguingly, we have 103 previously reported that p30II interacts with CBP/p300 at KIX domain of CBP 11, influences CRE and TRE mediated transcription 58 and disrupts CREB-Tax-p300

complexes on TRE probes 57. NF-κB (52) and NFAT 21 are also known to interact with

transcriptional coactivator CBP/p300. Therefore it is possible that p30II regulates the transcriptional activity of NFAT, NF-κB and AP-1, at least in part, by its interaction with

CBP/p300.

HTLV-1 mediated interference with normal T-cell apoptosis is thought to be a mechanism of tumorigenicity 25, but specific mechanisms by which HTLV-1 infection or

any particular HTLV-1 gene products influence on T-cell survival are not fully

understood 25. Similar to the effect of HTLV-1 Tax on apoptosis related genes23,44, we

found that p30II also deregulates multiple genes resulting in possible pro-apoptotic and

anti-apoptotic effects. Since apoptosis is a well-known mechanism of cellular defense

against viral infection, a possible role of p30II in lymphocyte apoptosis might correlate with the requirement of p30II in maintaining proviral loads in vivo 9. Previous studies

indicate that several members of the cell cycle machinery have altered expression in

HTLV-1 infected cells 47. Several studies examined the aberrations in cell cycle caused

by HTLV-1 Tax 25; however, not much is known about the role of other HTLV-1 proteins

in causing abnormalities in cell cycle.

p30II appears to regulate viral gene expression and modulate immune response.

We have previously reported that, p30II activated HTLV-1 LTR at lower concentrations

and repressed at higher concentrations. 58. Interestingly, p30II expression was associated

104 with downregulation of lck (p56), which suppresses the HTLV-1 promoter 45 and upregulate HTLV enhancer factor, which is known to bind to LTR at a region involved in regulation of gene expression by the ets family of transcription factors38. Additionally, p30II expression was associated with altered expression of cellular genes involved in

immune modulation such as CD46, CD43, CD58, IFNγ and CD72.

Overall, this study supports our earlier reports 2,9,57,58 and sheds light on the

mechanisms by which p30II functions in HTLV-1 pathogenesis and in leukemogenesis.

Many of the effects of p30II are similar to that of other HTLV-1 proteins like Tax or p12I.

It is possible that these proteins act coordinately or synergistically. We postulate that, by modulating the expression of various HTLV-1 proteins, the virus employs selective use of different viral proteins during different stages of the infection. However, since information on the expression profile of HTLV-1 proteins during different stages of the infection is limited, additional studies are required to explore this possibility. Further studies will also be required to verify our data in primary T lymphocytes. Such future studies might provide new directions in the development of therapeutic interventions against HTLV-1 disorders, which are associated with immune mediated mechanisms.

HTLV-1 accessory proteins are functionally homologous to the accessory proteins of bovine leukemia virus (BLV), a deltaretrovirus with conserved organizational structure similar to HTLV-1 that is associated with a naturally occurring lymphoma in cattle 17,29,37.

BLV-wild type and mutant proviruses that contained deletions in the G4 or R3 genes infected B lymphocytes and permitted the infected cell to resist apoptotic signals 17. G4

105 protein interacts with farnesyl pyrophosphate synthetase (FPPS), an enzyme in the

mevalonate/squalene pathway that is critical for synthesis of FPP, a substrate required for

prenylation of Ras 37. Infectious molecular clones of BLV with mutations in gene regions

encoding G4 and R3 were limited in their ability to maintain proviral loads and producing

lymphosarcomas in infected sheep29. These reports taken together with our data indicates

that complex retroviruses associated with lymphoproliferative diseases rely upon accessory gene products to modify their cellular environment to enhance clonal expansion of the virus genome and thus maintain proviral loads in vivo.

106 REFERENCES

1. Albrecht, B., N. D. Collins, M. T. Burniston, J. W. Nisbet, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J. Virol. 74:9828-9835.

2. Albrecht, B., C. D. D'Souza, W. Ding, S. Tridandapani, K. M. Coggeshall, and M. D. Lairmore. 2002. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J. Virol. 76:3493-3501.

3. Albrecht, B. and M. D. Lairmore. 2002. Critical role of human T-lymphotropic virus type 1 accessory proteins in viral replication and pathogenesis. Microbiol. Mol. Biol. Rev. 66:396-406.

4. Anand, N., S. Murthy, G. Amann, M. Wernick, L. A. Porter, I. H. Cukier, C. Collins, J. W. Gray, J. Diebold, D. J. Demetrick, and J. M. Lee. 2002. Protein elongation factor EEF1A2 is a putative oncogene in ovarian cancer. Nat. Genet. 31:301-305.

5. Arnold, R., J. Liou, H. C. Drexler, A. Weiss, and F. Kiefer. 2001. Caspase- mediated cleavage of hematopoietic progenitor kinase 1 (HPK1) converts an activator of NFkappaB into an inhibitor of NFkappaB. J Biol. Chem. 276:14675- 14684.

6. Ashburner, M., C. A. Ball, J. A. Blake, D. Botstein, H. Butler, J. M. Cherry, A. P. Davis, K. Dolinski, S. S. Dwight, J. T. Eppig, M. A. Harris, D. P. Hill, L. Issel-Tarver, A. Kasarskis, S. Lewis, J. C. Matese, J. E. Richardson, M. Ringwald, G. M. Rubin, and G. Sherlock. 2000. Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nat. Genet. 25:25-29.

7. Bangham, C. R. 2003. Human T-lymphotropic virus type 1 (HTLV-1): persistence and immune control. Int. J Hematol. 78:297-303.

8. Bannister, A. J. and T. Kouzarides. 1996. The CBP co-activator is a histone acetyltransferase. Nature 384:641-643.

9. Bartoe, J. T., B. Albrecht, N. D. Collins, M. D. Robek, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J. Virol. 74:1094-1100.

10. Bour-Jordan, H. and J. A. Blueston. 2002. CD28 function: a balance of costimulatory and regulatory signals. J Clin. Immunol. 22:1-7.

107 11. Campbell, K. M. and K. J. Lumb. 2002. Structurally distinct modes of recognition of the KIX domain of CBP by Jun and CREB. Biochemistry 41:13956- 13964.

12. Cereseto, A., Z. Berneman, I. Koralnik, J. Vaughn, G. Franchini, and M. E. Klotman. 1997. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia 11:866-870.

13. Chen, Y. M. A., S. H. Chen, C. Y. Fu, J. Y. Chen, and M. Osame. 1997. Antibody reactivities to tumor-suppressor protein p53 and HTLV-I Tof Rex and Tax in HTLV-I-infected people with differing clinical status. Int. J. Cancer 71:196- 202.

14. Ciminale, V., G. N. Pavlakis, D. Derse, C. P. Cunningham, and B. K. Felber. 1992. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J. Virol. 66:1737-1745.

15. de La, F. C., L. Deng, F. Santiago, L. Arce, L. Wang, and F. Kashanchi. 2000. Gene expression array of HTLV type 1-infected T cells: Up-regulation of transcription factors and cell cycle genes. AIDS Res. Hum. Retroviruses 16:1695- 1700.

16. Dekaban, G. A., A. A. Peters, J. C. Mulloy, J. M. Johnson, R. Trovato, E. Rivadeneira, and G. Franchini. 2000. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology 274:86-93.

17. Dequiedt, F., E. Hanon, P. Kerkhofs, P. P. Pastoret, D. Portetelle, A. Burny, R. Kettmann, and L. Willems. 1997. Both wild-type and strongly attenuated bovine leukemia viruses protect peripheral blood mononuclear cells from apoptosis. J Virol 71:630-639.

18. Ding, W., S. J. Kim, A. M. Nair, B. Michael, K. Boris-Lawrie, A. Tripp, G. Feuer, and M. D. Lairmore. 2003. Human T-cell lymphotropic virus type 1 p12I enhances interleukin-2 production during T-cell activation. J Virol 77:11027-11039.

19. Franchini, G. 1995. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood 86:3619-3639.

20. Fujii, M., K. Iwai, M. Oie, M. Fukushi, N. Yamamoto, M. Kannagi, and N. Mori. 2000. Activation of oncogenic transcription factor AP-1 in T cells infected with human T cell leukemia virus type 1. AIDS Res. Hum. Retroviruses 16:1603- 1606.

108 21. Garcia-Rodriguez, C. and A. Rao. 1998. Nuclear factor of activated T cells (NFAT)-dependent transactivation regulated by the coactivators p300/CREB- binding protein (CBP). J. Exp. Med. 187:2031-2036.

22. Giebler, H. A., J. E. Loring, K. Vanorden, M. A. Colgin, J. E. Garrus, K. W. Escudero, A. Brauweiler, and J. K. Nyborg. 1997. Anchoring of CREB binding protein to the human T-cell leukemia virus type 1 promoter: A molecular mechanism of tax transactivation. Mol. Cell Biol. 17:5156-5164.

23. Harhaj, E. W., L. F. Good, G. T. Xiao, and S. C. Sun. 1999. Gene expression profiles in HTLV-I-immortalized T cells: deregulated expression of genes involved in apoptosis regulation. Oncogene 18:1341-1349.

24. Heger, P., O. Rosorius, J. Hauber, and R. H. Stauber. 1999. Titration of cellular export factors, but not heteromultimerization, is the molecular mechanism of trans- dominant HTLV-1 Rex mutants. Oncogene 18:4080-4090.

25. Hollsberg, P. 1999. Mechanisms of T-cell activation by human T-cell lymphotropic virus type I. Microbiol. Mol. Biol. Rev. 63:308-333.

26. Hurd, C. and E. Rozengurt. 2003. Uncoupling of protein kinase D from suppression of EGF-dependent c-Jun phosphorylation in cancer cells. Biochem. Biophys. Res. Commun. 302:800-804.

27. Ju, Y. T., A. C. Chang, B. R. She, M. L. Tsaur, H. M. Hwang, C. C. Chao, S. N. Cohen, and S. Lin-Chao. 1998. gas7: A gene expressed preferentially in growth- arrested fibroblasts and terminally differentiated Purkinje neurons affects neurite formation. Proc. Natl. Acad. Sci U. S. A 95:11423-11428.

28. Kasai, T., Y. Iwanaga, H. Iha, and K. T. Jeang. 2002. Prevalent loss of mitotic spindle checkpoint in adult T-cell leukemia confers resistance to microtubule inhibitors. J. Biol. Chem. 277:5187-5193.

29. Kerkhofs, P., H. Heremans, A. Burny, R. Kettmann, and L. Willems. 1998. In vitro and in vivo oncogenic potential of bovine leukemia virus G4 protein. J Virol 72:2554-2559.

30. Kimata, J. T., F. Wong, J. Wang, and L. Ratner. 1994. Construction and characterization of infectious human T-cell leukemia virus type 1 molecular clones. Virology 204:656-664.

31. Kohno, T., R. Moriuchi, S. Katamine, Y. Yamada, M. Tomonaga, and T. Matsuyama. 2000. Identification of genes associated with the progression of adult T cell leukemia (ATL). Jpn. J Cancer Res. 91:1103-1110.

109 32. Kohno, T., R. Moriuchi, S. Katamine, Y. Yamada, M. Tomonaga, and T. Matsuyama. 2000. Identification of genes associated with the progression of adult T cell leukemia (ATL). Jpn. J Cancer Res. 91:1103-1110.

33. Kohno, T., R. Moriuchi, S. Katamine, Y. Yamada, M. Tomonaga, and T. Matsuyama. 2000. Identification of genes associated with the progression of adult T cell leukemia (ATL). Jpn. J Cancer Res. 91:1103-1110.

34. Konikova, E. and J. Kusenda. 2003. Altered expression of p53 and MDM2 proteins in hematological malignancies. Neoplasma 50:31-40.

35. Koralnik, I. J., J. Fullen, and G. Franchini. 1993. The p12, p13 and p30 proteins encoded by human T-cell leukemia/lymphotropic virus type-1 open reading frames I and II are localized in three different cellular compartments. J. Virol. 67:2360- 2366.

36. Koralnik, I. J., A. Gessain, M. E. Klotman, A. Lo Monico, Z. N. Berneman, and G. Franchini. 1992. Protein isoforms encoded by the pX region of human T- cell leukemia/lymphotropic virus type 1. Proc. Natl. Acad. Sci. USA 89:8813-8817.

37. Lefebvre, L., A. Vanderplasschen, V. Ciminale, H. Heremans, O. Dangoisse, J. C. Jauniaux, J. F. Toussaint, V. Zelnik, A. Burny, R. Kettmann, and L. Willems. 2002. Oncoviral Bovine Leukemia Virus G4 and Human T-Cell Leukemia Virus Type 1 p13(II) Accessory Proteins Interact with Farnesyl Pyrophosphate Synthetase. J. Virol. 76:1400-1414.

38. Li, C., A. J. Lusis, R. Sparkes, S. M. Tran, and R. Gaynor. 1992. Characterization and chromosomal mapping of the gene encoding the cellular DNA binding protein HTLF. Genomics 13:658-664.

39. Li, C. and W. H. Wong. 2001. Model-based analysis of oligonucleotide arrays: expression index computation and outlier detection. Proc. Natl. Acad. Sci U. S. A 98:31-36.

40. McCoy, M. A., J. J. Gesell, M. M. Senior, and D. F. Wyss. 2003. Flexible lid to the p53-binding domain of human Mdm2: implications for p53 regulation. Proc. Natl. Acad. Sci U. S. A 100:1645-1648.

41. Mori, N., A. Ueda, S. Ikeda, Y. Yamasaki, Y. Yamada, M. Tomonaga, S. Morikawa, R. Geleziunas, T. Yoshimura, and N. Yamamoto. 2000. Human T- cell leukemia virus type I tax activates transcription of the human monocyte chemoattractant protein-1 gene through two nuclear factor-kappaB sites. Cancer Res. 60:4939-4945.

42. Murphy, D. 2002. Gene expression studies using microarrays: principles, problems, and prospects. Adv. Physiol Educ. 26:256-270.

110 43. Nemorin, J. G. and P. Duplay. 2000. Evidence that Llck-mediated phosphorylation of p56dok and p62dok may play a role in CD2 signaling. J Biol. Chem. 275:14590-14597.

44. Ng, P. W., H. Iha, Y. Iwanaga, M. Bittner, Y. Chen, Y. Jiang, G. Gooden, J. M. Trent, P. Meltzer, K. T. Jeang, and S. L. Zeichner. 2001. Genome-wide expression changes induced by HTLV-1 Tax: evidence for MLK- 3 mixed lineage kinase involvement in Tax-mediated NF-kappaB activation. Oncogene 20:4484- 4496.

45. Ohta, M., H. Nagai, H. Nyunoya, and K. Shimotohno. 1996. Src-homology domain 2 is responsible for transcriptional suppression induced by expression of Lck. Oncogene 12:989-997.

46. Pique, C., A. Uretavidal, A. Gessain, B. Chancerel, O. Gout, R. Tamouza, F. Agis, and M. C. Dokhelar. 2000. Evidence for the chronic in vivo production of human T cell leukemia virus type I Rof and Tof proteins from cytotoxic T lymphocytes directed against viral peptides. J. Exp. Med. 191:567-572.

47. Pise-Masison, C. A., M. Radonovich, R. Mahieux, P. Chatterjee, C. Whiteford, J. Duvall, C. Guillerm, A. Gessain, and J. N. Brady. 2002. Transcription profile of cells infected with human T-cell leukemia virus type I compared with activated lymphocytes. Cancer Res. 62:3562-3571.

48. Pusl, T., J. J. Wu, T. L. Zimmerman, L. Zhang, B. E. Ehrlich, M. W. Berchtold, J. B. Hoek, S. J. Karpen, M. H. Nathanson, and A. M. Bennett. 2002. Epidermal growth factor-mediated activation of the ETS domain transcription factor Elk-1 requires nuclear calcium. J Biol. Chem. 277:27517-27527.

49. Sun, S. C., E. W. Harhaj, G. Xiao, and L. Good. 2000. Activation of I-kappaB kinase by the HTLV type 1 tax protein: mechanistic insights into the adaptor function of IKKgamma [In Process Citation]. AIDS Res. Hum. Retroviruses 16:1591-1596.

50. Takeshita, A., G. R. Cardona, N. Koibuchi, C. S. Suen, and W. W. Chin. 1997. TRAM-1, A novel 160-kDa thyroid hormone receptor activator molecule, exhibits distinct properties from steroid receptor coactivator-1. J Biol. Chem. 272:27629- 27634.

51. Tommerup, N. and H. Vissing. 1995. Isolation and fine mapping of 16 novel human zinc finger-encoding cDNAs identify putative candidate genes for developmental and malignant disorders. Genomics 27:259-264.

52. Van Lint, J., A. Rykx, Y. Maeda, T. Vantus, S. Sturany, V. Malhotra, J. R. Vandenheede, and T. Seufferlein. 2002. Protein kinase D: an intracellular traffic regulator on the move. Trends Cell Biol. 12:193-200.

111 53. Vaudin, P., R. Delanoue, I. Davidson, J. Silber, and A. Zider. 1999. TONDU (TDU), a novel human protein related to the product of vestigial (vg) gene of Drosophila melanogaster interacts with vertebrate TEF factors and substitutes for Vg function in wing formation. Development 126:4807-4816.

54. Wrzesinski, S., R. Seguin, Y. Liu, S. Domville, V. Planelles, P. Massa, E. Barker, J. Antel, and G. Feuer. 2000. HTLV type 1 Tax transduction in microglial cells and astrocytes by lentiviral vectors. AIDS Res. Hum. Retroviruses 16:1771-1776.

55. Yoshida, H., H. Nishina, H. Takimoto, L. E. Marengere, A. C. Wakeham, D. Bouchard, Y. Y. Kong, T. Ohteki, A. Shahinian, M. Bachmann, P. S. Ohashi, J. M. Penninger, G. R. Crabtree, and T. W. Mak. 1998. The transcription factor NF-ATc1 regulates lymphocyte proliferation and Th2 cytokine production. Immunity 8:115-124.

56. Zhang, B. X., X. Ma, C. K. Yeh, M. D. Lifschitz, M. X. Zhu, and M. S. Katz. 2002. Epidermal growth factor-induced depletion of the intracellular Ca2+ store fails to activate capacitative Ca2+ entry in a human salivary cell line. J Biol. Chem. 277:48165-48171.

57. Zhang, W., J. W. Nisbet, B. Albrecht, W. Ding, F. Kashanchi, J. T. Bartoe, and M. D. Lairmore. 2001. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J. Virol. 75:9885-9895.

58. Zhang, W., J. W. Nisbet, J. T. Bartoe, W. Ding, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J. Virol. 74:11270-11277.

112

Genbank ID Description / Title Apoptosis NM021960 myeloid cell leukemia sequence 1 AF017061 cullin 5 NM000675 adenosine A2a receptor NM004049 BCL2-related protein A1 NM003806 Harakiri NM001197 BCL2/adeno E1B interacting protein 1 BC005034 TSSC 3 AF314174 caspase 2 AL050391 caspase 4 Y09321 TAF4b RNA polymerase II AK001291 NCK-associated protein 1 NM003790 TNF receptor superfamily, member 25 U70056 seven in absentia homolog 1(Drosophila) AF222341 CD 28 antigen U07236 lymphocyte specific PTK (lck) AJ276888 Mdm2 *NM001197 BCL2-interacting killer, BIK *NM014430 cell death-inducing DFFA-like effector b *NM022094 cell death activator CIDE-3 Cell Adhesion BC002630 integrin, beta 8 AI335208 integrin, alpha 6 NM007164 MADCAM1 U02297 selectin P ligand AI797281 desmocollin 3 NM003622 PPFIBP1 (liprin beta 1) AF244129 CD84 (lymphocyte antigen 9) D28586 CD58 antigen X60502 CD43 (sialophorin) *NM000899 KIT ligand

Continued

Table 2.1. Genes modulated by HTLV-1 p30II *Genes upregulated by p30II, indicating that the signal intensity was either increased by a minimum of 1.5 fold or turned on (not expressed in controls) in at least 2 of the 3 p30II expressing samples. Unmarked genes are downregulated by p30II, indicating that the signal intensity was either decreased by a minimum of 1.5 fold or completely shut down in at least 2 of the 3 p30II expressing samples.

113 Table 2.1 continued

Apoptosis NM021960 myeloid cell leukemia sequence 1 AF017061 cullin 5 NM000675 adenosine A2a receptor NM004049 BCL2-related protein A1 NM003806 Harakiri NM001197 BCL2/adeno E1B interacting protein 1 BC005034 TSSC 3 AF314174 caspase 2 AL050391 caspase 4 Y09321 TAF4b RNA polymerase II AK001291 NCK-associated protein 1 NM003790 TNF receptor superfamily, member 25 U70056 seven in absentia homolog 1(Drosophila) AF222341 CD 28 antigen U07236 lymphocyte specific PTK (lck) AJ276888 Mdm2 *NM001197 BCL2-interacting killer, BIK *NM014430 cell death-inducing DFFA-like effector b *NM022094 cell death activator CIDE-3 Cell Adhesion BC002630 integrin, beta 8 AI335208 integrin, alpha 6 NM007164 MADCAM1 U02297 selectin P ligand AI797281 desmocollin 3 NM003622 PPFIBP1 (liprin beta 1) AF244129 CD84 (lymphocyte antigen 9) D28586 CD58 antigen X60502 CD43 (sialophorin) *NM000899 KIT ligand Cell cycle AL545982 chaperonin containing TCP1, subunit 2 β AA860806 checkpoint suppressor 1 NM003672 CDC14 cell division cycle 14 homolog A AF001362 Janus kinase 2 NM003644 growth arrest-specific 7 AI652662 cytosolic BCAT-1 AK022408 rab6 GTPase activating protein

Continued 114 Table 2.1 continued

AJ277546 WEE1 homolog BC005872 histone deacetylase 6 AL518328 basic leucine zipper & W2 domains 1 BE407516 cyclin B1 *AI347136 telomeric repeat binding factor 1 *AW341501 A kinase (PRKA) anchor protein 8 *NM004399 DEAD/H box polypeptide11 *U04045 mutS homolog 2 *BG491844 JUN Transcription NM000281 TCF1 (PCBD, PCD, DCOH) NM003666 basic leucine zipper nuclear factor 1 NM005693 nuclear receptor subfamily1-H, member 3 NM012429 SEC14-like 2 (S. cerevisiae) NM006186 nuclear receptor subfamily4-A member 2 NM000965 retinoic acid receptor, beta NM002518 neuronal PAS domain protein 2 NM004527 mesenchyme homeo box 1 NM014724 zinc finger protein 305 NM003070 SMARCA2 NM003438 zinc finger protein 137 NM002167 ID3 (HEIR-1) NM006079 CITED2 (MRG1) M64240 MAX protein U80737 nuclear receptor coactivator 3 AB006572 chromosome 19 ORF2 X79067 zinc finger protein 36 L1 S79910 homeo box A1 BC004145 trinucleotide repeat containing 4 BE542323 TONDU Y09321 TAF4b RNA polymerase II X59740 zinc finger protein, X-linked U88968 enolase 1, (alpha) NM024567 hypothetical protein FLJ21616 NM013351 T-box 21 AF016005 arginine-glutamic acid dipeptide repeats *NM003489 nuclear receptor interacting protein 1 *AW015313 TBP associated factor I C *NM005341 GLI-Kruppel family member HKR3

Continued 115 Table 2.1 continued

*NM004956 ets variant gene 1 *NM002158 HTLV enhancer factor *AB041834 polyglutamine binding protein 1 *BC000052 peroxisome proliferative activated receptor α *L22179 MLLT2 (AF-4) *AI989477 SRY box 4 *NM021020 leucine zipper, putative tumor suppressor 1 Translation NM004564 PET112-like (yeast) BE138647 translation initiation factor IF2 AI335509 eukaryotic translation elongation factor 1 δ *NM001958 eukaryotic translation elongation factor 1 α 2 T cell activation / signaling NM002389 CD46 NM001963 epidermal growth factor NM005825 RAS guanyl releasing protein 2 AF309082 Protein Kinase D AF074382 IKKγ NM007236 calcium binding protein P22 (CHP) *NM003371 vav 2 oncogene *AF283777 CD72 Other *M29383 IFNγ

116

Figure 2.1. Schematic illustration of lentiviral vectors expressing both p30HA and GFP (sample vector) as bicistronic messages and GFP alone (control vector) from elongation factor 1 alpha promoter. Abbreviations: LTR - Long Terminal Repeats; RRE - Rev Response Element; EF1 α - Elongation Factor 1 alpha promoter; IRES – Internal Ribosome Entry Site; WPRE – Woodchuck Hepatitis Post-transcriptional Regulatory Element.

117

Figure 2.2. Triplicate p30II samples express GFP and p30II while triplicate controls express only GFP. (A) Flow cytometric analysis illustrating the expression of GFP in Jurkat T cells 10 days post spin-infection with lentiviral vectors. Both sample (expressing p30II and GFP) and control (GFP alone) group contains relatively high and similar levels of GFP. (B) RT-PCR demonstrating the expression of p30II in Jurkat T cells 10 days post spin-infection with lentiviral vectors. Jurkat T cells spin-infected with sample vector express p30II while the control vector spin-infected cells do not express p30II. RT-PCR was performed with triplicate samples and controls. GAPDH was used as a control for the integrity of the message.

118

Figure 2.3. Semiquantitative RT-PCR of CHP, JUN and NFATc in controls and p30II expressing Jurkat T lymphocytes (A and B) and primary CD4+ T lymphocyte (C and D) samples. PCR products were separated by electrophoresis (A and C), normalized to GAPDH and quantified by densitometry (B and D). Black bars indicate control and grey bars indicate p30II expressing cells. Data points are mean of triplicates. CHP was downregulated while JUN and NFATc was upregulated by p30II. Fold decrease / increase in activity in the presence of p30II are indicated above each bar.

119

Figure 2.4. p30II activates NFAT, AP-1 and NF-κB transcriptional activity in Jurkat T lymphocytes. Black bars indicate control and grey bars indicate p30II. Data points are mean of triplicate experiments. Fold increase in activity in the presence of p30II is indicated above each bar. p30II increased the NFAT-luc activity from 2.2 to 10.7 fold depending on co-stimulatory treatment e.g., PMA, ionomycin, CD3, CD28 etc. (A), p30II increased NF-κB-luc activity from 3.1 to 11.4 fold (B) and modestly increased the AP-1 driven luciferase reporter gene activity from 1to 5 fold in the presence of co-stimulator treatments (C).

120

CHAPTER 3

MOTIFS OF HUMAN T LYMPHOTROPIC VIRUS TYPE 1 P30II CRITICAL

FOR ITS INTERACTION WITH P300 AND REGULATION OF LTR

TRANSCRIPTIONAL ACTIVITY.

INTRODUCTION

Human T-cell lymphotropic virus type 1 (HTLV-1) is the etiologic agent of adult

T cell leukemia/lymphoma (ATL), and HTLV-1-associated myelopathy/ tropical spastic

paraparesis (HAM/TSP) and many immune-mediated disorders48,26. Molecular

mechanisms used by the virus to circumvent immune elimination by the host and to

facilitate lymphocyte proliferation are not fully understood. The genome of HTLV-1

encodes Gag, Pol, and Env, the common structural and enzymatic proteins characteristic

of all retroviruses. Using alternative splicing mechanism and internal initiator codons,

this complex retrovirus makes several regulatory and accessory proteins encoded by four

open reading frames (ORFs) of the pX region (pX ORF I to IV) between the env gene

and the 3’ long terminal repeat16. ORF III and IV encode the well characterized Rex and

Tax proteins respectively. Rex is a nucleolar-localizing phosphoprotein, involved in

121 nuclear export of unspliced or singly spliced viral RNA21 while Tax is a nuclear-

localizing phosphoprotein, which interacts with cellular transcription factors and activates

transcription from the viral promoter containing repeats of TRE and enhancer elements of

various cellular genes associated with host cell proliferation18,34,37,38,47. pX ORFs I and II

also produce alternatively spliced forms of mRNA, which encode four accessory

proteins, p12I, p27I, p13II, and p30II7,12,17,28. Less is known about the role of pX ORF I

and ORF II in the replication or pathogenesis of HTLV-1. However, pX ORFs I and II

mRNAs are present in infected cell lines and freshly isolated cells from HTLV-1-infected

subjects 28 and in ATL and HAM/TSP patients10. Antibodies11,15 and cytotoxic T cells36 against recombinant proteins or peptides of the pX ORF I and II proteins are present in

HTLV-1 infected patients, and asymptomatic carriers.

We have reported that mutations in the ACH.p30II viral clone that insert an

artificial termination codon in the p30II mRNA prevent the virus from obtaining typical

proviral levels in the rabbit model of infection41. p30II contains a highly conserved bipartite nuclear localization signal14 as well as serine-and-threonine-rich regions with

distant homology to transcription factors Oct-1 and -2, Pit-1, and POU-M112. Taken

together, these characteristics suggest that p30II has a role in transcription of viral and

cellular gene expression. In addition, a recent report demonstrated that p30II modulates

LTR mediated transcription, in the context of the entire provirus, at the post-

transcriptional level, by retaining Tax/Rex mRNA in the nucleus33. We have previously

reported that, when provided in limiting concentrations, p30II expression stimulated

HTLV-1 LTR-driven reporter gene activity, even in the presence of Tax, whereas higher

concentrations repressed LTR and CRE-driven reporter gene activity50. 122 We have shown that p30II is co-localized with p300 in the nucleus and physically

interacts with CBP/p300, at the highly conserved KIX domain, the domain HTLV-1 Tax

also interacts with. CBP/p300 protein is available only at limiting concentrations within

the cell nucleus, causing an environment for competition between coactivators and

transcription factors, thus providing an additional layer of regulated gene expression35,43.

There is evidence of a functional antagonistic relationship between transcription factors, as a result of competition for binding to common regions of CBP/p3005,13,24. Under such

a condition of tight competition, relative concentrations of Tax/ p30II at various stages of

disease might be a critical factor in determining the levels of viral transcription. We have

previously demonstrated that p30II disrupts CREB-Tax-CBP/p300 complexes bound to

the TRE repeats49. At higher concentrations, p30II may support viral persistence by

reducing viral expression, thus reducing immune elimination of HTLV-1 infected cells

and at low concentrations, enhance TRE/viral over CRE/cellular mediated transcription,

thus promote viral transcription and spread of the virus in vivo. This is also consistent

with our previous results showing that an infectious HTLV-1 molecular clone failed to

maintain viral loads in vivo when p30II expression was abolished41.

Identifying the molecular mechanism and functional significance of the

interaction between p30II and p300 is crucial in understanding the role of p30II in the

pathogenesis and replication of HTLV-1. In this study, we have further characterized the

nature of this interaction, by identifying the motifs within p30II that are critical in binding

CBP/p300 and in regulating LTR mediated transcription. Using N terminal and C terminal deletion mutants of p30II, we have localized the binding site of p300 on p30II.

123 These serial deletion mutants were employed to identify the domain of p30II, which is

important in regulating LTR mediated transcription. Using our serial deletion mutants of

p30II, we identified the domain of p30II that is involved in regulating LTR mediated

transcription, in the context of the provirus. By dissecting this amino acid domain of p30II

(100-179) critical for repressing LTR mediated transcription further, we identified that lysine residue at amino acid position 106 (K3) of HTLV-1 p30II is critical for its

repression of TRE-mediated transcription. More importantly, by adding p300, we were

able to rescue the p30II-mediated repression on LTR driven gene transcription, irrespective of the presence or absence of the provirus. Our data confirms the role of p30II as a regulator viral gene transcription, in association with p300. This is the first report that dissects p30II and identifies the functional domains important for binding CBP/p300

and in regulating LTR mediated transcription. Furthermore, we found that HTLV-1 p30II represses Tax-mediated LTR transactivation, while HTLV-1 Tax rescues p30II-mediated repression of LTR driven transcription, indicating the ability of these proteins to compete with each other in modulating HTLV-1 LTR transcriptional activity. Taken together, our data indicates that HTLV-1, a complex retrovirus associated with lymphoproliferative disorders, uses accessory genes, in conjunction with regulatory proteins, to promote cell- to-cell transmission of the virus, clonal expansion of infected cells and maintenance of proviral loads in vivo.

124 MATERIALS AND METHODS

Cell lines.

293T cells were obtained from G. Franchini (National Cancer Institute, NIH,

Bethesda) and Hela-Tat cells were obtained from AIDS Research and Reference Reagent

Program, National Institute of Health. 293T and Hela-Tat cells were grown in modified

Dulbecco’s eagle medium containing 10% fetal bovine serum and 1% streptomycin and

penicillin at 37°C. Cells were split and cultured in six-well plates, 10 cm dishes or

chamber slides to 50-60% confluence 16-18 h before transfection.

Plasmids.

The pTRE-Luc plasmid and pRSV-B-Gal have been described previously49,50. pME-p30IIHA wildtype and serial deletion mutant plasmids were created by cloning the

p30II sequence from HTLV-1 molecular clone, ACH with downstream influenza

hemagglutinin (HA1) tag, into pME-18S plasmid (G. Franchini, National Cancer

Institute) between 5’ EcoRI and 3’ NotI sites. Site directed mutants were made by

substituting lysines in pMEp30IIHA with arginine. Fidelity of the plasmids were

confirmed by Sanger sequencing and p30II HA protein expression was confirmed by

western blot using monoclonal anti-HA antibody (Covance, Berkeley, CA). ACH.30

plasmid has a 24 bp insertion that causes an artificial termination codon in p30II40. pCMV-Tax expresses the HTLV-1 Tax protein and has been described previously32. pCMV-p300 expresses the full-length p300 protein from a CMV I/E promoter (Upstate

USA, Inc., Charlottesville, VA).

125 Cell transfection and reporter gene assay.

For each transfection, 0.3 µg of pTRE-Luc reporter plasmid was cotransfected with pME-p30IIHA wildtype or serial deletion / arginine substitution mutant plasmids and

0.4 µg ACH.30 plasmid when specified, using Lipofectamine Plus (Invitrogen,

Carlsbad, CA). When pCMV-p300 plasmid was used, it was also co-transfected. As an

internal control for transfection efficiency, 0.1 µg of pRSV-βGal (Invitrogen) was also

used in each transfection. pME 18S was used as carrier DNA to equalize DNA

concentrations for each transfection. Transfected cells were lysed with 400 µl lysis buffer

(Promega, Madison, WI) per well for 25 min. Twenty microliters of each lysate was used

to test luciferase reporter gene activity using an Enhanced Luciferase Assay kit

(Promega). To normalize transfection efficiency, cells were washed with PBS, stained

with 5-bromo-4-chloro-3-indolyl-beta-D-galactopyranoside (X-Gal) (Sigma, St. Louis,

MO) and counted βGal expressing cells using 20X objective of light microscope. Co-

transfection of pME-p30IIHA had no effect on the X-Gal staining (data not shown).

Results were expressed as mean of optimized luciferase activity (luciferase activity /

percentage cells stained positive for β-Gal expression) in arbitrary light units (ALU) with

standard error (SE) from a minimum of triplicate experiments. Statistical analysis was

done using Student’s t test, P < 0.05.

Western immunoblot assay.

Transiently transfected cells were lysed in RIPA buffer containing PBS, 1%

Nonidet P-40, 0.5% sodium deoxycholate, and 0.1% sodium dodecyl sulfate (SDS). Cell

126 lysates were prepared by centrifugation at 14,000 rpm (Beckman, Fullerton, CA) for 20

min at 4° C. Equal amounts of proteins were mixed with Laemmli buffer (62.5 mM Tris;

pH 6.8), 2 % SDS, 10 % glycerol, 0.2 % bromophenol blue, 100 mM dithiothreitol

(DTT). After boiling for 5 min, samples were electrophoresed through 6 or 10%

polyacrylamide gels. The fractionated proteins were transferred to nitrocellulose

membranes (Amersham Pharmacia Biotechnology, Piscataway, NJ) at 100 V for 1 h at 4°

C. Membranes were then blocked in Tris-buffered saline containing 5 % nonfat milk and

0.1 % Tween 20. Proteins were detected with primary monoclonal anti- hemagglutinin

(HA) antibody (Covance) followed by an anti-mouse (Cell Signaling Technology,

Beverly, MA) immunoglobulin G (IgG)-horseradish peroxidase-conjugated goat

antibody. Blots were developed using an enhanced chemiluminescence detection system

(Cell Signaling Technology).

Coimmunoprecipitation of p30II with p300.

Sixty percent confluent 293T cells were cotransfected by pME-p30IIHA, pME-

p30IIHA serial deletion or arginine substitution mutants and pCMV-p300 using

Superfect (Gibco BRL, Gaithersburg MD). After 48 h, the transfected cells were washed with PBS and resuspended in 400 µl of lysis buffer containing 50 mM Tris-HCl

(pH 8.0), 100 mM NaCl, 1 mM phenylmethylsulfonyl fluoride (PMSF), 0.5 µg of leupeptin per ml, and 1 µg of aprotinin per ml. Cell suspensions were incubated on ice for

20 min and then lysed by homogenization. The lysates were centrifuged at 14,000 rpm for 20 min at 4° C and incubated overnight at 4° C with 100 ng of a polyclonal HA antibody (Covance). After adding 40 µl of 100 % of protein A-agarose slurry, the mixture

127 was incubated at 4° C for 1 h. The immunoprecipitated complexes were washed twice

with 10 volumes of lysis buffer and three times with PBS buffer. The components of the

complexes were resolved on 6% SDS-polyacrylamide gels and detected by western

immunoblot assay using anti-p300 antibody (Santa cruz biotechnology, Santa cruz, CA).

The density of each band was measured and compared to wildtype p30II using a commercial software package (Gel-pro Analyzer software; Media Cybernatic Inc., San

Diego, CA). Normal rabbit IgG was used as a negative control in immunoprecipitation assays and equal amounts of input cell extracts were processed to determine the expression of p30II by western immunoblot assay.

Localization of p30II and co-localization with CBP/p300 by confocal microscopy.

To detect cellular colocalization of p30II and p300 by immunofluorescence, Hela

Tat cells were seeded in two chamber slides (Fisher Scientific) at approximately 50 %

confluence 18 h prior to transfection. Transfection with 4 µg of pME-p30II-HA wildtype

or pME-p30IIHA serial deletion mutants alone or with 2 µg of pCMV-p300 was performed using Superfect (Gibco BRL). At 48 h posttransfection, media were removed and cells were fixed for 15 min using 4 % paraformaldehyde. Cells were then incubated with FITC (Fluorescin isothyocyanate) conjugated anti-HA antibody alone (diluted

1:1000; Santa Cruz) or with TRITC (Texas red isothyocyanate) conjugated anti-p300

antibody (diluted 1:1000; Molecular probes) in antibody dilution buffer containing 0.01

M sodium phosphate, 0.5 M NaCl, 0.5 % Triton-X 100 and 2 % bovine serum albumin overnight at 4° C. The expression of p30II-HA and p300 was evaluated using Zeiss LSM

510 confocal microscope.

128 RESULTS

HTLV-1 p30II localizes to the nucleus and colocalizes with p300 in the nucleus.

HTLV-1 p30II contains a nuclear localization signal, characteristic of most

proteins that function as a transcription factor. Using imunofluorescence and immunoblot

methods, previous reports have shown that HTLV-1 p30II was localized in the nucleus of

transfected cells27,50. To confirm the subcellular localization of p30II, we performed

confocal microscopy on HeLa- cells transiently transfected with pMEp30IIHA. Anti-

Phalloidin was used to stain the cytokeratin of the cells (Fig 3.1A). Consistent with previous reports, p30II was detected predominantly in the nucleus of our transiently

transfected HeLa-tat cells by confocal microscopy (Fig. 3.1B). Based on nuclear

localization, as typical of proteins that play a role in transcription, and the sequence homology of p30II with the POU family of transcription factors, we hypothesized that

p30II has a role in transcriptional regulation. To test whether p30II colocalizes with p300

in the nucleus, we performed confocal microscopy of Hela-Tat cells transiently

cotransfected with pME p30IIHA and pCMV-p300, and verified the subcellular

localization of p30II (Fig. 3.1B) and p300 (Fig. 3.1C). Hoechst was used as a nuclear

marker (Fig. 3.1D) and we found that p30II-HA colocalized with p300 in the nucleus of

cells that expressed both proteins (Fig. 3.1E).

129 HTLV-1 p30II physically interacts with p300

HTLV-1 p30II contains a bipartite nuclear localization sequence and regions with

sequence homology to the DNA binding domain and homeodomain of transcription

factors such as Oct-1 (Fig. 3.2A). To test if p30II physically interacts with the cellular

transcriptional co-activator p300, we transiently transfected 293T cells with pCMV-p300

and pME p30IIHA plasmid encoding full-length p30II, prepared whole-cell extracts and

performed co-immunoprecipitation assays. Our data indicated that p300 was

immunoprecipitated as a complex with p30II (Fig. 3.2B). In contrast, p300 could not be

detected in the complexes precipitated using preimmune serum while p30IIHA was detected from the cellular lysate (data not shown).

Amino acids 1-132 of HTLV-1 p30II are critical for its physical interaction with p300

To identify the domain of p30II that is required in its interaction with p300, we

performed transient co-transfection of 293T cells with pCMV-p300 and various amino-

terminal and carboxyl-terminal serial deletion mutants of p30II, and whole-cell extracts

were used for coimmunoprecipitation assays. These mutants of p30II contained serial

deletions of both the amino-terminal and carboxyl-terminal regions of p30II (Fig. 3.2A).

Each of the p30II mutant proteins was expressed at the expected molecular weight, as indicated by immunoblot analysis (Fig. 3.2B). The subcellular localization characteristics

of these mutants were tested by confocal microscopy. Mutants containing amino acids 1-

220, 1-179, 1-132 or 71-220 localized to the nucleus while mutants containing amino acids 1-71, 100-179 and 179-241 localized to the cytoplasm (Fig. 3.3). As seen in Fig. 130 3.2B, we have confirmed that amino acids 1-132 of p30II are critical for binding p300.

The p30II mutants 1-132, 1-220 and 1-179 (in the decreasing order) appeared to interact

with p300 more intensely compared to the wildtype p30II.

HTLV-1 p30II modulates LTR driven transcription.

To test the effect of p30II on LTR mediated transcription, we transiently transfected increasing concentrations of pME p30IIHA plasmid with constant amounts of

our LTR-luc reporter plasmid and tested for luciferase reporter activity. We found that

lower concentrations of the p30II plasmid (below 0.6 µg) consistently activated the

HTLV-1 LTR reporter gene activity, but increased amounts (above 0.6 µg) of the plasmid repressed LTR reporter gene activity (Fig. 3.4A). Consistent with our previous report, this data confirms that at low concentrations, p30II has the potential to enhance LTR-

mediated transcription while at higher doses, p30II represses the LTR-mediated

transcription 50.

We tested the ability of p30II to regulate LTR mediated transcription in the presence of the HTLV-1 provirus. By transiently transfecting pME-p30IIHA at various

doses with ACH.30, provirus deleted for p30II expression, we have found that p30II represses LTR mediated transcription, in a dose responsive fashion (Fig. 3.4B).

Interestingly, at concentrations that activated LTR mediated transcription in the absence of ACH.30 (below 0.6 µg), we found that p30II indeed repress LTR mediated

transcription if the provirus is present. This difference could be due to the presence of

other HTLV-1 proteins in the provirus.

131 p300 expression reverses the p30II-dependent repression of LTR driven

transcription, in the presence or absence of provirus, in a dose-responsive manner.

HTLV-1 Tax is known to activate the LTR. This transactivation function of

HTLV-1 Tax is dependent on the co-adaptors CBP/p300. Tax is reported to bind p300 at

the KIX domain, the same domain that p30II binds to. Our previous reports suggested that

p30II and Tax serve different roles in the regulation of transcription. Based on the

hypothesis that p30II represses LTR-driven reporter gene activities as a consequence of

competition for limited basal quantities of CBP/p300, we tested if overexpressing p300

rescues p30II-mediated repression of LTR-driven reporter gene expression. By transiently

transfecting increasing concentrations of pCMV-p300 with constant concentration of

pME-p30IIHA plasmid, herein, we demonstrated that p300 expression reverses the p30II- dependent repression of LTR-luciferase reporter gene activity, in the absence of the provirus (Fig. 3.5A). This data confirms that the transcriptional regulatory effect of p30II is p300 dependent.

To test if the p30II mediated repression of the LTR-mediated transcription is p300

dependent in the context of the provirus, we transiently transfected increasing amounts of

pCMV-p300 plasmid with constant amounts of pME-p30IIHA plasmid and ACH.30, the provirus deleted for p30II expression. Interestingly, even in the presence of the provirus,

increasing doses of p300 reversed the ability of p30II to repress the LTR mediated

transcription, in a dose responsive fashion (Fig. 3.5B). Taken together, this data further

confirm that the transcriptional regulatory function of p30II on the LTR is p300

dependent, in the presence of the provirus as well.

132 HTLV-1 p30II represses Tax-mediated LTR transactivation and HTLV-1 Tax

rescues p30II-mediated repression of LTR driven transcription

Since HTLV-1 p30II and Tax bind CBP/p300 at the KIX domain and since the

transcriptional regulatory functions of p30II and Tax on the LTR are p300 dependent, we

hypothesized that HTLV-1 Tax and p30II might competitively influence the

transcriptional regulatory function of each other on the HTLV-1 LTR. To test the effect

of p30II on Tax-mediated LTR transactivation, we transiently co-transfected increasing

concentrations of pME p30IIHA plasmid with constant concentration of LTR-luc reporter

plasmid, as well as pCMV-Tax plasmid and tested the luc reporter activity. We found that p30II consistently repressed the Tax transactivation of the HTLV-1 LTR reporter gene

activity, in a dose dependent manner (Fig. 3.6A). Conversely, to test the effect of HTLV-

1 Tax on p30II-mediated repression of the LTR transcriptional activity, we transiently co-

transfected increasing concentrations of pCMV-Tax plasmid along with constant

concentration of LTR-luc reporter plasmid as well as pME p30IIHA plasmid and tested

the luc reporter activity. We found that Tax consistently rescued the p30II-mediated repression of the LTR reporter gene activity, in a dose dependent manner (Fig. 3.6B).

HTLV-1 p30II contains multiple domains critical for its repression of LTR driven

transcription

To identify the structural domain of p30II that is critical for repression of LTR

driven transcription, 293T cells were transiently co-transfected with a series of carboxyl-

terminal and amino-terminal truncation mutants of p30II and LTR-luciferase reporter 133 plasmid, and was tested for their ability to reduce the LTR-luciferase activity. The

luciferase activity elicited by LTR-luciferase construct, in the presence of the full-length

p30II was compared to luciferase activity in the presence of each of the serial deletion mutants (Fig. 3.7A). Luciferase activity elicited by the serially deleted mutants 1-220, 1-

179, 1-132 and 1-71 were higher than that by the full-length p30II (in the decreasing order

of luciferase activity), while in the presence of mutants 179-241, 100-179 and 71-220 (in

the decreasing order of luciferase activity), the luciferase activity was less than that in the

presence of the wildtype p30II. This data shows that the region between 71-220 is able to

repress the LTR-luc reporter, even more than the wildtype p30II, indicating that these

amino acid sequences are critical for the repression of LTR driven transcription by p30II.

However, mutant containing amino acid sequences 1-220 appeared to elicit the highest

LTR driven reporter activity, which suggests the existence of a domain that inhibits the transcriptional repressive function of p30II in this mutant, which is likely amino acids 1-

71. This is also consistent with the higher luciferase activity elicited by the mutant

containing only amino acids 1-71, compared to the wildtype p30II. However, it appears

that amino acids 1-71 do not inhibit the repressive function of the wildtype p30II. A

comparison of 1-220 mutant p30II and the wildtype p30II raises the possibility of another

critical domain at amino acids 220-241, which helps the wildtype overcome the

inhibitory role by amino acids 1-71. This is also consistent with the increased ability of

the mutant containing 179-241 to repress LTR mediated transcription, compared to the

wildtype. However, we cannot rule out the possibility of a change in the physical

conformation or other properties of the mutant proteins, causing the observed differences

in their transcriptional activity. These data indicate that the amino acid sequence from 71-

134 220 of p30II is important for its function as a repressor of LTR-mediated transcription.

The mutant containing amino acid sequence 100-179 was also able to repress LTR

mediated transcription at levels similar to that of 71-220. The differences in activity

between these mutant was not statistically significant (p < 0.15), and therefore, the amino

acids 100-179 appears to be the structural motif within p30II, most critical for its

repression of LTR mediated reporter activity.

To determine if 100-179 itself is the structural domain that is critical for

repressing LTR mediated transcription, in the presence of the provirus, we performed

transient transfection of 293T cells with ACH.30, LTR-luc and the carboxyl- terminal

and amino-terminal truncation mutants of p30II described above and measured the luciferase activity. The luciferase activity elicited by LTR-luciferase construct, in the presence of the full length / wildtype p30II, in the context of the provirus was then

compared to luciferase activity in the presence of each of the serial deletion mutants (Fig.

3.7B). Similar to the findings in the absence of the provirus, luciferase activity elicited by

the serially deleted mutants 1-179, 1-220, 1-71 and 1-132 were higher than that by the

full-length p30II (in the decreasing order of luciferase activity), and in the presence of

mutants 100-179, the luciferase activity was less than that in the presence of the wildtype

p30II. Taken together, these data indicate that the amino acid sequence 100-179 of p30II is

critical in repression of LTR-mediated transcription in the presence or absence of the

HTLV-1 provirus.

135 Lysine 106 (K3) of HTLV-1 p30II is critical for its repression of LTR driven

transcription.

HTLV-1 p30II contains 6 lysine residues (K1-K6) (Table 3.1). Interestingly, 5 of

these are within the critical domain comprised of amino acids 100-179. Importantly, four

(K2-K5) of the lysine residues within this domain are part of a consensus acetylation

sequence (G/SK motif),6 representing potential acetylation sites for CBP/p300. To

determine the role of these lysine residues in p30II-mediated repression of LTR driven transcription, 293T cells were transiently co-transfected with the arginine substitution

mutants for each lysine residue of p30II and LTR-luciferase reporter plasmid, and was

tested for their ability to reduce the LTR-luciferase activity. The luciferase activity

elicited by LTR-luciferase construct, in the presence of the full-length p30II was

compared to luciferase activity in the presence of each of the arginine substitution

mutants (Fig. 3.8A). Luciferase activity elicited by the arginine substitution mutant K3

(amino acid 106) was higher than that by the full-length p30II, pME empty plasmid,

arginine substitution mutant K2 (amino acid 103), arginine substitution mutant K4

(amino acid 123) and arginine substitution mutant K5 (amino acid 143) (in the decreasing

order of luciferase activity), indicating a loss of repression function with substitution of

lysine at amino acid position 106. This data shows that the lysine residue at amino acid

position 106 is critical for repressing the LTR-luc reporter.

To determine the role of these lysine residues in p30II-mediated repression of LTR

driven transcription, in the presence of the provirus, we performed transient transfection

of 293T cells with ACH.30, LTR-luc and the arginine substitution mutants of p30II 136 described above and measured the luciferase activity. The luciferase activity elicited by

LTR-luciferase construct, in the presence of the full length / wildtype p30II, in the context

of the provirus was then compared to luciferase activity in the presence of each of the

arginine substitution mutants (Fig. 3.8B). Similar to the findings in the absence of the

provirus, luciferase activity elicited by the arginine substitution mutant K3 (amino acid

106) was higher than that by the full-length p30II, pME empty plasmid, arginine

substitution mutant K2 (amino acid 103), arginine substitution mutant K4 (amino acid

123) and arginine substitution mutant K5 (amino acid 143) (in the decreasing order of

luciferase activity), indicating a loss of repression function with substitution of lysine at

amino acid position 106. This data confirms that the lysine residue at amino acid position

106 is critical for repressing the LTR-luc reporter in the presence of the provirus as well.

Taken together, these data indicate that the lysine residue at amino acid position 106

within p30II is important for its function as a repressor of LTR-mediated transcription in the presence or absence of the HTLV-1 provirus.

DISCUSSION

Our data is the first to dissect p30II and demonstrate the functional motifs of p30II that are critical in binding p300 and in repressing LTR mediated transcription.

Identification of these motifs is important for defining the molecular mechanism of p30II- mediated repression of LTR-driven transcription. This information is vital in gaining more insight into HTLV-1 gene expression and pathogenesis. We found that the motif

137 critical for binding p300 is amino acid sequence 1-132 while the motif that plays a major

role in repressing LTR-mediated transcription is amino acid sequence 100-179, which

contains the entire DNA binding region, part of the homeodomain with homology to Oct-

1 and POU family of transcription factors as well as 5 lysine residues, four of which are

part of a G/SK consensus acetylation sequence6 representing potential acetylation sites

for CBP/p300.

The coactivators CREB binding protein (CBP) and p300 mediate transcriptional

control of various cellular and viral DNA binding transcription factors. Although these coactivators have divergent functions, they are similar in nucleotide sequence, are evolutionarily conserved, and are commonly referred to as CBP/p3009,19. CBP and p300

are highly related and share many functional properties, however there is evidence that

these factors are not interchangeable45. Several cellular and viral proteins that interact

with either CBP or p300 have been identified, including steroid and retinoid hormone

receptors, CREB, c-Jun, c-Myb, Sap-1a, c-Fos, MyoD, p53, Stat-1/2, NF-κB, pp90rsk,

TATA-binding protein, TFIIB4,19,22,23, HTLV-1 Tax, adenovirus E1A, Kaposi’s sarcoma-

associated herpes virus viral interferon regulatory factor protein, and simian virus 40

large T antigen1-3,13,25,29,30,43.

CBP and p300 bridge transcription factors to relevant promoters, has intrinsic

histone acetyltransferase (HAT) activity, and form complexes with proteins such as

CBP/p300 binding protein-associated factor, which also exhibits HAT activity20.

Recently, there is increasing knowledge of the mechanism and functional significance of

138 the interactions between many viral proteins and CBP/p300. In the case of adenovirus

oncoprotein E1A, interaction with CBP/p300 is critical for regulation of transcription,

suppression of differentiation, and immortalization of cells in culture1,3,31. Simian virus T

antigen regulates the expression of a group of cellular genes by modifying the HAT activity of CBP/p300 or by bridging the gap between DNA binding transcription factors

and components of the general transcription machinery4,19. Identifying the molecular

mechanism and functional significance of the interaction between p30II and p300 is very

crucial in understanding of the role of p30II in the pathogenesis and replication of this important human pathogen.

Previously, using deletion mutants of CBP/p300 in GST pull-down assays, we localized the binding site of CBP/p300 for p30II to a highly conserved KIX region.

Interestingly, KIX domain is the binding site of CBP/p300 for HTLV-1 Tax protein as well. In addition, p30II was able to disrupt CREB-Tax-CBP/p300 complexes bound to the

viral 21-bp TRE repeats49. These results indicate that p30II might act in contrast to the role of HTLV-1 Tax. The role of CBP and p300 coactivators in HTLV-1 gene expression has been the focus of many previous reports. HTLV-1 Tax, a transactivator of LTR mediated transcription, is critical in the activation of the HTLV-1 viral genes through its interaction with the p300 and CBP coactivators8.

Although CBP and p300 mediate the activities of various transcription factors,

many earlier reports prove that it is present in the cell only at limiting concentrations.

Even small reductions in the concentrations of these coactivator are damaging in many

139 instances, like the human Rubinstein-Taybi syndrome where loss of a single CBP allele

causes developmental defects35. In experiments where transcription factors are

overexpressed, it is possible that the capacity of the endogenous CBP/p300 is overridden

and the effects due to competition may be more obvious. Many such studies have

demonstrated the competition between different molecules for CBP/p300. Even though

some of these involve two mutually antagonistic transcription factors with the same

shared binding site on CBP/p300, competition does not necessarily require having the

same binding site. In circumstances where CBP/p300 levels are limiting, there could be

selective preference of one over the other, causing an exclusion of one of the proteins. In

our study, we found that HTLV-1 p30II and Tax appear to compete with each other in

modulating the transcriptional activity from the LTR. Since HTLV-1 p30II and Tax interacts with CBP/p300 through the same KIX domain, we believe that this is through competitive CBP/p300 binding between p30II and Tax. Similar mechanism involving

Tax and CBP/p300 has been previously reported. The binding of Tax and c-Myb to KIX

domain of CBP was found to be mutually exclusive and Tax expression was shown to

interfere with transcriptional activity of c-Myb13.

An environment of coactivator competition between transcription factors provides

an additional layer of regulated gene expression. The relative amounts of Tax and p30II may be critical in determining which protein binds to CBP/p300 and the consequent regulation of LTR mediated transcription of viral genes from the HTLV-1 viral promoter.

When its levels are low, p30II also enhance transcription from the viral promoter. Under

such circumstances, p30II and Tax might act synergistically. However, at higher

140 concentrations, p30II antagonizes the Tax transactivation of the HTLV-1 promoter. There

are similar reports on positive and negative effects by the same protein in other viruses,

such as herpes simplex virus type 1 (HSV-1), in which the interaction between cellular

and viral transcription factors play a critical role in the regulation of the immediate-early

(IE) gene promoter. VP16, a potent transcription factor from HSV-1, binds the host cell

protein HCF, which facilitates the stable complex formation of the viral protein with Oct-

146. The IE gene promoter contains an Oct-1-like motif (TAATGARAT) that is important

for IE gene expression, with both positive and negative effects, depending on the context

of these cellular transcription factors and VP1644.

CBP and p300 bridge transcription factors to relevant promoters and through its

intrinsic histone acetyl transferase (HAT) activity, acetylates lysine residues at the amino

terminus of histones, which results in modification of chromatin structure to allow access

of the transcriptional machinery. CBP/p300 is known to form complexes with proteins

such as CBP/p300 binding protein-associated factor, which also exhibits HAT activity20. p300 also directly acetylate transcription factors such as p53, GATA-1 and c- and thus augment their sequence-specific DNA-binding activity42. Simian virus T antigen

regulates the expression of a group of cellular genes by modifying the HAT activity of

CBP/p300 or by bridging the gap between DNA binding transcription factors and

components of the general transcription machinery4,19. c-myb oncoprotein of the avian

retrovirus group is known to be acetylated by CBP/p300 at its multiple lysine residues13.

This is considered to be the mechanism by which CBP/p300 regulates the functions of c- myb. p30II contains six highly conserved lysine residues, 5 of which are within the amino

141 acid region 100-179, four of which are part of the consensus acetylation sequence (G/SK

motif)6 and thus represent potential acetylation sites for CBP/p300 (Table. 3.1). We

found that the lysine residue at amino acid position 106 (K3) of HTLV-1 p30II is critical

for its repression of LTR driven transcription in the presence or absence of the provirus.

Based on this, it is possible that the intrinsic HAT activity of CBP/p300 is utilized to

acetylate and potentially modulate the transcriptional regulatory function of p30II.

Recently, it was reported that p30II modulates LTR mediated transcription, in the context of the entire provirus, by a post-transcriptional mechanism 33. Previous studies

from our laboratory demonstrated that HTLV-1 p30II directly interacts with CBP/p300 to

modulate gene transcription49,50. Data presented in this report further confirms the role of

CBP/p300 in the modulation of LTR mediated transcription by p30II, in the context of the

provirus. In the presence of increasing concentrations of p300, there was a dose-

dependent rescue of LTR driven gene transcription, irrespective of the presence or absence of the provirus. Although, we cannot rule out the possibility of a post- transcriptional mechanism by HTLV-1 p30II, this report presents convincing evidence for

a p300 dependent transcriptional regulatory function. Therefore, HTLV-1 p30II appears to

be a multi-functional protein with transcriptional and post-transcriptional role in regulating gene expression.

As of yet, there is not much information regarding the relative levels of various

HTLV-1 proteins during various stages of the infection. It is possible that differences in

expression levels of various viral proteins, leading to differences in transcriptional

142 regulation as described above might be the mechanism by which HTLV-1 infection/ disease progresses though various stages39. This is likely to be in synergy with differential regulation of cellular gene regulation by Tax and/or p30II and thus changing the immune responses in accordance with different stages of progression of HTLV-1 infection and disease. Our current data can be supported by recent findings from our laboratory where an infectious HTLV-1 molecular clone failed to maintain viral loads in vivo when p30II expression was abolished41. Taken together, our data indicates that

HTLV-1, a complex retrovirus associated with lymphoproliferative disorders, uses accessory genes to promote cell-to-cell transmission of the virus, clonal expansion of infected cells and maintenance of proviral loads in vivo.

143 REFERENCES

1. Arany, Z., D. Newsome, E. Oldread, D. M. Livingston, and R. Eckner. 1995. A family of transcriptional adaptor proteins targeted by the E1A oncoprotein. Nature 374:81-84.

2. Ariumi, Y., A. Kaida, J. Y. Lin, M. Hirota, O. Masui, S. Yamaoka, Y. Taya, and K. Shimotohno. 2000. HTLV-1 tax oncoprotein represses the p53-mediated trans-activation function through coactivator CBP sequestration. Oncogene 19:1491-1499.

3. Avantaggiati, M. L., M. Carbone, A. Graessmann, Y. Nakatani, B. Howard, and A. S. Levine. 1996. The SV40 large T antigen and adenovirus E1a oncoproteins interact with distinct isoforms of the transcriptional co-activator, p300. EMBO J. 15:2236-2248.

4. Avantaggiati, M. L., V. Ogryzko, K. Gardner, A. Giordano, A. S. Levine, and K. Kelly. 1997. Recruitment of p300/CBP in p53-dependent signal pathways. Cell 89:1175-1184.

5. Bannister, A. J. and T. Kouzarides. 1995. CBP-induced stimulation of c-Fos activity is abrogated by E1A. EMBO J. 14:4758-4762.

6. Bannister, A. J. and E. A. Miska. 2000. Regulation of gene expression by transcription factor acetylation. Cell Mol. Life Sci. 57:1184-1192.

7. Berneman, Z. N., R. B. Gartenhaus, M. S. Reitz, W. A. Blattner, A. Manns, B. Hanchard, O. Ikehara, R. C. Gallo, and M. E. Klotman. 1992. Expression of alternatively spliced human T-lymphotropic virus type 1 pX mRNA in infected cell lines and in primary uncultured cells from patients with adult T-cell leukemia/lymphoma and healthy carriers. Proc. Natl. Acad. Sci. USA 89:3005- 3009.

8. Bex, F. and R. B. Gaynor. 1998. Regulation of gene expression by HTLV-I Tax protein. Methods 16:83-94.

9. Blobel, G. A. 2000. CREB-binding protein and p300: molecular integrators of hematopoietic transcription. Blood 95:745-755.

10. Cereseto, A., Z. Berneman, I. Koralnik, J. Vaughn, G. Franchini, and M. E. Klotman. 1997. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia 11:866-870.

11. Chen, Y. M., S. H. Chen, C. Y. Fu, J. Y. Chen, and M. Osame. 1997. Antibody reactivities to tumor-suppressor protein p53 and HTLV-I Tof, Rex and Tax in 144 HTLV-I-infected people with differing clinical status. International. Journal. of. Cancer 71:196-202.

12. Ciminale, V., G. N. Pavlakis, D. Derse, C. P. Cunningham, and B. K. Felber. 1992. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J. Virol. 66:1737-1745.

13. Colgin, M. A. and J. K. Nyborg. 1998. The human T-cell leukemia virus type 1 oncoprotein Tax inhibits the transcriptional activity of c-Myb through competition for the CREB binding protein. J. Virol. 72:9396-9399.

14. D'Agostino, D. M., V. Ciminale, L. Zotti, and L. Chieco-Bianchi. 1999. Influence of Rex and intronic sequences on expression of spliced mRNAs produced by human T cell leukemia virus type I. AIDS Res. Hum. Retroviruses 15:1351- 1363.

15. Dekaban, G. A., A. A. Peters, J. C. Mulloy, J. M. Johnson, R. Trovato, E. Rivadeneira, and G. Franchini. 2000. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology 274:86-93.

16. Franchini, G. 1995. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood 86:3619-3639.

17. Furukawa, K., K. Furukawa, and H. Shiku. 1991. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 295:141-145.

18. Giebler, H. A., J. E. Loring, K. Vanorden, M. A. Colgin, J. E. Garrus, K. W. Escudero, A. Brauweiler, and J. K. Nyborg. 1997. Anchoring of CREB binding protein to the human T-cell leukemia virus type 1 promoter: A molecular mechanism of tax transactivation. Mol. Cell Biol. 17:5156-5164.

19. Goodman, R. H. and S. Smolik. 2000. CBP/p300 in cell growth, transformation, and development. Genes Dev. 14:1553-1577.

20. Grunstein, M. 1997. Histone acetylation in chromatin structure and transcription. Nature 389:349-352.

21. Heger, P., O. Rosorius, J. Hauber, and R. H. Stauber. 1999. Titration of cellular export factors, but not heteromultimerization, is the molecular mechanism of trans- dominant HTLV-1 Rex mutants. Oncogene 18:4080-4090.

22. Janknecht, R. and T. Hunter. 1996. Transcription. A growing coactivator network [news; comment]. Nature 383:22-23.

145 23. Janknecht, R. and T. Hunter. 1996. Transcriptional control: Versatile molecular glue. Curr. Biol. 6:951-954.

24. Kamei, Y., L. Xu, T. Heinzel, J. Torchia, R. Kurokawa, B. Gloss, S. C. Lin, R. A. Heyman, D. W. Rose, C. K. Glass, and M. G. Rosenfeld. 1996. A CBP integrator complex mediates transcriptional activation and AP-1 inhibition by nuclear receptors. Cell 85:403-414.

25. Kashanchi, F., J. F. Duvall, R. P. Kwok, J. R. Lundblad, R. H. Goodman, and J. N. Brady. 1998. The coactivator CBP stimulates human T-cell lymphotrophic virus type I Tax transactivation in vitro. J. Biol. Chem. 273:34646-34652.

26. Kitze, B. and J. N. Brady. 1997. Human T cell lymphotropic retroviruses: Association with diseases of the nervous system. Intervirology 40:132-142.

27. Koralnik, I. J., J. Fullen, and G. Franchini. 1993. The p12I, p13II, and p30II proteins encoded by human T-cell leukemia/lymphotropic virus type I open reading frames I and II are localized in three different cellular compartments. J Virol 67:2360-2366.

28. Koralnik, I. J., A. Gessain, M. E. Klotman, A. Lo Monico, Z. N. Berneman, and G. Franchini. 1992. Protein isoforms encoded by the pX region of human T- cell leukemia/lymphotropic virus type 1. Proc. Natl. Acad. Sci. USA 89:8813-8817.

29. Kwok, R. P., M. E. Laurance, J. R. Lundblad, P. S. Goldman, H. Shih, L. M. Connor, S. J. Marriott, and R. H. Goodman. 1996. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature 380:642-646.

30. Li, M., B. Damania, X. Alvarez, V. Ogryzko, K. Ozato, and J. U. Jung. 2000. Inhibition of p300 histone acetyltransferase by viral interferon regulatory factor [In Process Citation]. Mol. Cell Biol. 20:8254-8263.

31. Lundblad, J. R., R. P. Kwok, M. E. Laurance, M. L. Harter, and R. H. Goodman. 1995. Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature 374:85-88.

32. Newbound, G. C., J. M. Andrews, J. Orourke, J. N. Brady, and M. D. Lairmore. 1996. Human T-cell lymphotropic virus type 1 tax mediates enhanced transcription in CD4(+) T lymphocytes. J. Virol. 70:2101-2106.

33. Nicot, C., M. Dundr, J. M. Johnson, J. R. Fullen, N. Alonzo, R. Fukumoto, G. L. Princler, D. Derse, T. Misteli, and G. Franchini. 2004. HTLV-1-encoded p30(II) is a post-transcriptional negative regulator of viral replication. Nat. Med.

146 34. Nyborg, J. K., M. H. Matthews, W. T. Golde, and W. S. Dynan. 1995. The identification of Cellular Proteins that Interact with the HTLV-I LTR, p. 131-140. Human Retroviruses. Alan R. Liss, Inc..

35. Petrij, F., R. H. Giles, H. G. Dauwerse, J. J. Saris, R. C. Hennekam, M. Masuno, N. Tommerup, G. J. van Ommen, R. H. Goodman, and D. J. Peters. 1995. Rubinstein-Taybi syndrome caused by mutations in the transcriptional co- activator CBP [see comments]. Nature 376:348-351.

36. Pique, C., A. Uretavidal, A. Gessain, B. Chancerel, O. Gout, R. Tamouza, F. Agis, and M. C. Dokhelar. 2000. Evidence for the chronic in vivo production of human T cell leukemia virus type I Rof and Tof proteins from cytotoxic T lymphocytes directed against viral peptides. J. Exp. Med. 191:567-572.

37. Pise-Masison, C. A., R. Mahieux, M. Radonovich, H. Jiang, and J. N. Brady. 2001. Human T-lymphotropic virus type I Tax protein utilizes distinct pathways for p53 inhibition that are cell type-dependent. J. Biol. Chem. 276:200-205.

38. Pise-Masison, C. A., R. Mahieux, M. Radonovich, H. Jiang, J. Duvall, C. Guillerm, and J. N. Brady. 2000. Insights into the molecular mechanism of p53 inhibition by HTLV type 1 Tax. AIDS Res. Hum. Retroviruses 16:1669-1675.

39. Princler, G. L., J. G. Julias, S. H. Hughes, and D. Derse. 2003. Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology 317:136-145.

40. Robek, M. D., F. H. Wong, and L. Ratner. 1998. Human T-Cell leukemia virus type 1 pX-I and pX-II open reading frames are dispensable for the immortalization of primary lymphocytes. J. Virol. 72:4458-4462.

41. Silverman, L. R., A. J. Phipps, A. Montgomery, L. Ratner, and M. D. Lairmore. 2004. Human T-cell lymphotropic virus type 1 open reading frame II- encoded p30II is required for in vivo replication: evidence of in vivo reversion. J Virol 78:3837-3845.

42. Sterner, D. E. and S. L. Berger. 2000. Acetylation of histones and transcription- related factors. Microbiol. Mol. Biol. Rev. 64:435-459.

43. Suzuki, T., M. Uchidatoita, and M. Yoshida. 1999. Tax protein of HTLV-1 inhibits CBP/p300-mediated transcription by interfering with recruitment of CBP/p300 onto DNA element of E-box or p53 binding site. Oncogene 18:4137- 4143.

44. Thomas, S., R. Coffin, P. Watts, G. Gough, and D. S. Latchman. 1998. The TAATGARAT motif in the herpes simplex virus immediate-early gene promoter can confer both positive and negative responses to cellular octomer-binding proteins when it is located within the viral genome. J. Virol. 72:3495-3500. 147 45. Vo, N. and R. H. Goodman. 2001. CREB-binding protein and p300 in transcriptional regulation. J. Biol. Chem. 276:13505-13508.

46. Wilson, A. C., R. W. Freeman, H. Goto, T. Nishimoto, and W. Herr. 1997. VP16 targets an amino-terminal domain of HCF involved in cell cycle progression. Mol. Cell. Biol. 17:6139-6146.

47. Yin, M. J. and R. B. Gaynor. 1996. Complex formation between CREB and tax enhances the binding affinity of CREB for the human T-cell leukemia virus type 1 21- base-pair repeats. Mol. Cell Biol. 16:3156-3168.

48. Yoshida, M. 1996. Molecular biology of HTLV-I: recent progress. J. Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S63-8:S63-S68.

49. Zhang, W., J. W. Nisbet, B. Albrecht, W. Ding, F. Kashanchi, J. T. Bartoe, and M. D. Lairmore. 2001. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J. Virol. 75:9885-9895.

50. Zhang, W., J. W. Nisbet, J. T. Bartoe, W. Ding, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J. Virol. 74:11270-11277.

148

Lysine residue and amino acid position within p30II Percentage conservation

K1 (50) 87.5

K2 (103) 66.7

K3 (106) 87.5

K4 (123) 87.5

K5 (143) 91.7

K6 (152) 91.7

Table 3.1. Lysine residues in p30II are conserved among various HTLV isolates. Amino acid position of each lysine is indicated in parenthesis.

149

Figure 3.1. p30II localizes to the nucleus and colocalizes with p300 in the nucleus of transiently transfected HeLa-Tat cells. Hela Tat cells were seeded in two chamber slides (Fisher Scientific) at approximately 50 % confluence 18 h prior to transfection, and co- transfected with 4 µg of pME-p30II-HA and 2 µg of pCMV-p300 using Superfect (Gibco BRL). At 48 h post-transfection, media were removed and cells were fixed for 15 min using 4 % paraformaldehyde. Cells were then incubated with FITC conjugated anti-HA antibody (diluted 1:1000; Santa Cruz) and TRITC conjugated anti-p300 antibody (diluted 1:1000; Molecular probes) overnight at 4° C. Panel A represents the cytokeratin staining of the cells using anti-Phalloidin while Panel D represents the nuclear staining of the cell using Hoechst. The expression of p30II-HA (Panel B) and p300 (Panel C) were evaluated using Zeiss LSM 510 confocal microscope (40X) and the images were overlayed using Adobe photoshop (Panel E). These results are representative of a minimum of triplicate experiments.

150

Figure 3.2. Amino acids 1-132 of HTLV-1 p30II are critical for its physical interaction with p300. (A) Schematic diagram of the known/ putative functional domains of p30II. Schematic diagram of the carboxyl terminal and amino terminal deletion mutants of p30II is indicated. Bipartite NLS indicates the nuclear localization sequence. Regions with sequence homology to the DNA binding domain and homeodomain of Oct-1 are indicated. The numbers on both sides of p30II indicate length of the peptide as number of amino acids. Letters within the boxes indicate amino acids in p30II and Oct-1. (B) Top panel represents the immunoprecipitation of p300 with p30II and serial deletion mutants of p30II. Physical binding between various deletion mutants of p30II and p300 was tested by expressing pCMV-p300 and pME-p30IIHA wildtype or serial deletion mutant plasmids in 293T cells, p300 pull-down from the cellular lysates using polyclonal anti- HA antibody, followed western immunoblot assay with anti-p300 antibody. Below each line, the binding is indicated, as density of each band measured by Gel-pro Analyzer software (middle panel). The results above are representative of three experiments. The protein expression of the carboxyl terminal and amino terminal deletion mutants of p30II was tested using monoclonal anti-HA antibody (lower panel).

151

Figure 3.3. Subcellular localization characteristics of various p30II serial deletion mutants p30II in transiently transfected HeLa-Tat cells. Hela Tat cells were seeded in two chamber slides (Fisher Scientific) at approximately 50 % confluence 18 h prior to transfection, and co-transfected with 4 µg of each pME-p30II-HA serial deletion mutant using Superfect (Gibco BRL). At 48 h post-transfection, media were removed and cells were fixed for 15 min using 4 % paraformaldehyde. Cells were then incubated with FITC conjugated anti- HA antibody (diluted 1:1000; Santa Cruz) overnight at 4° C. Cytokeratin staining of the cells was performed using Anti-Phalloidin. Amino acid domains present in each mutant is indicated. The expression of p30II-HA was evaluated using Zeiss LSM 510 confocal microscope (40X). These results represent a minimum of triplicate experiments.

152

Figure 3.4. HTLV-1 p30II differentially modulates LTR-mediated transcription in the absence or the presence of the provirus. (A) In the absence of the provirus, p30II enhances LTR mediated transcription at low concentrations (upto 0.6 µg) and represses LTR mediated transcription at higher doses (above 0.6 µg) (B) In the presence of full-length provirus, p30II represses LTR mediated transcription in a dose responsive fashion. 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid, increasing amounts of pME-p30II HA, +/- 0.4 µg of ACH.30 proviral plasmid and the luciferase activity was measured. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed using Student’s t test. * indicates P < 0.05.

153

Figure 3.5 p300 expression reverses the p30II-dependent repression of LTR driven transcription, in the absence or the presence of the provirus. (A) In the absence of the provirus, p300 expression reverses the p30II-dependent repression of LTR driven transcription, in a dose-responsive manner. (B) p300 reverses the p30II-dependent repression of LTR driven transcription, in the presence of the provirus, in a dose- responsive manner. 293T cells were transiently cotransfected with 0.3 µg of pLTR- luciferase reporter plasmid, 1.2 µg pME-p30II-HA, increasing quantities of pCMV-p300, +/- 0.4 µg of ACH.30 proviral plasmid and the luciferase activity was measured. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed using Student’s t test. * indicates P < 0.05.

154

Figure 3.6 HTLV-1 p30II and Tax appear to compete with each other in modulating the LTR mediated transcription. (A) p30II represses Tax-mediated LTR transactivation in a dose-dependent manner. (B) Tax rescues p30II-mediated repression of LTR driven transcription in a dose-dependent fashion. 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid along with either 0.4 µg of pCMV-Tax plasmid and increasing quantities of pME-p30II HA (Fig. 6A) or 1.2 µg of pME-p30II HA plasmid and increasing quantities of pCMV-Tax (Fig. 6B). pME 18S (Fig. 6A) or a plasmid containing CMV promoter (Fig. 6B) was used as carrier DNA to equalize DNA concentrations for each transfection. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed using Student’s t test. * indicates P < 0.05.

155

Figure 3.7. Amino acids 100-179 of p30II is critical for its repression of LTR driven transcription, in the absence or the presence of the provirus. (A) In the absence of the provirus, amino acids 100-179 of p30II are critical for its repression of LTR driven transcription. (B) In the presence of the provirus, amino acids 100-179 of p30II are critical for its repression of LTR driven transcription. 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid, 1.2 µg pME-p30II-HA wildtype or serial deletion mutants, +/- 0.4 µg of ACH.30 proviral plasmid and the luciferase activity was measured. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed using Student’s t test. * indicates P < 0.05.

156

Figure 3.8. Lysine 106 (K3) of p30II is critical for its repression of LTR driven transcription in the absence or the presence of the provirus. (A) In the absence of the provirus, lysine 106 (K3) of p30II is critical for its repression of LTR driven transcription. (B) In the presence of the provirus, lysine 106 (K3) of p30II is critical for its repression of LTR driven transcription. 293T cells were transiently cotransfected with 0.3 µg of pLTR- luciferase reporter plasmid, 1.2 µg pME-p30II-HA wildtype or site directed mutants, +/- 0.4 µg of ACH.30 proviral plasmid and the luciferase activity was measured. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed using Student’s t test. * indicates P < 0.05.

157

CHAPTER 4

HISTONE ACETYLTRANSFERASE (HAT) ACTIVITY OF P300 MODULATES

HUMAN T LYMPHOTROPIC VIRUS TYPE 1 P30II-MEDIATED REPRESSION

OF LTR TRANSCRIPTIONAL ACTIVITY

INTRODUCTION

Human T-cell leukemia virus type 1 (HTLV-1) is the causative agent of an aggressive and often fatal CD4+ T cell malignancy, known as adult T-cell leukemia

(ATL), a chronic neurodegenerative disease, called HTLV-1 associated myelopathy/ tropical spastic paraparesis (HAM/TSP) and is also implicated in various immune- mediated disorders50,25. The virus infection persists in 15-20 million people worldwide, but is a particular problem in endemic regions of the Caribbean, Japan, Africa, South

America, and among at-risk populations within the United States. HTLV-1 is a complex retrovirus, which encodes the typical gag, pol and env gene products, and unique regulatory and accessory genes encoded in four open reading frames (pX ORF I-IV) between the env gene and the 3’ long terminal repeat (LTR)13. Four accessory proteins designated p12I, p27I, p13II and p30II are expressed from these highly conserved ORFs.

158 Using molecular clones of HTLV-1 with selective mutations of pX ORFs I and II, our

laboratory was the first to identify functional roles of p12I, p13II/p30II for establishment

of infection in the rabbit model and for infection of resting T-cells7.

p30II, also known as Tof, is a 241 amino acid protein that localizes in the

nucleolus and nucleus of transiently transfected cells. The viral protein has a bipartite

nuclear localization signal (NLS) located between amino acids 73-98, which contains two

arginine-rich stretches separated by a 12 amino acid spacer and serine/threonine-rich

regions. This region shares distant homology to the activation domain of the POU family

of transcription factors. We have previously demonstrated that p30II functions as a

transcriptional regulator and differentially modulates cAMP responsive element (CRE)

and Tax-responsive element (TRE)-mediated transcription through its interaction with

CBP/p30052. At low concentrations, p30II activated the transcription from the HTLV-1

LTR. However, at higher concentrations, p30II repressed HTLV-1 LTR driven

transcription52. In addition, p30II regulates CRE and TRE-mediated transcription in the

presence or absence of Tax52. We have also reported that p30II interacts with cellular

transcriptional regulator, CBP/p300, through the KIX domain, similar to the transactivating protein, Tax, encoded in pX ORF-III51. More importantly, p30II inhibits the assembly of Tax-p300-CREB multiprotein complexes on the TRE promoter51.

Recently, using serial deletion mutants of p30II, we localized the domain of p30II that is

critical for its binding to CBP/p300 and in mediating its transcriptional activity. In

addition, we also demonstrated that p30II-mediated LTR repression can be rescued by

p300 51(Michael et al, 2004, submitted). Recently, Nicot et al 34 reported a post-

159 transcriptional role of HTLV-1 p30II in modulating LTR mediated transcription.

Therefore, p30II appears to be a multi-functional protein with transcriptional and post- transcriptional role in regulating viral gene expression.

Recent studies have provided insight into the mechanism and functional significance of the interactions between many viral proteins and CBP/p300. CBP and p300 bridge transcription factors to relevant promoters and through their intrinsic histone acetyl transferase (HAT) activity, acetylates lysine residues at the amino terminus of histones, which results in modification of chromatin structure to allow access of the transcriptional machinery. p300 also directly acetylates transcription factors such as p53 and GATA-1 and thus augment their sequence-specific DNA-binding activity43.

Additionally, CBP/p300 is known to form complexes with proteins such as p300/CBP- associated factor (P/CAF), which also exhibits HAT activity17. HTLV-1 Tax also

associates with the LTR through its interaction with CREB. These Tax-CREB promoter

complexes act as a high-affinity binding site to recruit multifunctional cellular

coactivators CBP, p300 and P/CAF to activate the expression of viral genes from the

viral promoter28.

Identifying the molecular mechanism and functional significance of the

interaction between p30II and p300 is crucial in understanding the pathogenesis and

replication of HTLV-1. p30II contains six highly conserved lysine residues, four of which

are part of the consensus acetylation sequence (G/SK motif)6, and thus represent potential

acetylation sites for CBP/p300. We have previously demonstrated that lysine 106 (K3) of

160 p30II is critical in LTR repression (Michael et al, 2004, manuscript submitted). Based on

this, we hypothesized that the intrinsic HAT activity of CBP/p300 is utilized to acetylate and potentially modulate the transcriptional regulatory function of p30II. Herein, we have

characterized the interaction between p30II and p300 and demonstrate that p300 HAT

mutants do not rescue p30II-mediated LTR repression as much as wild type p300. In

addition, we also show that p30II is acetylated and that the deacetylation by histone

deacetylase-1 (HDAC-1) enhances p30II-mediated HTLV-1 LTR repression. Moreover,

inhibition of deacetylation by trichostatin A decreases p30II-mediated HTLV-1 LTR repression. This is the first report to demonstrate that the HAT activity of p300 is crucial in modulating HTLV-1 gene expression from the LTR by p30II. Overall, our study

confirms the role of p30II as a regulator of viral gene transcription, in association with

p300. Taken together, our data indicates that HTLV-1, a complex retrovirus associated

with lymphoproliferative disorders, uses accessory genes to modulate viral gene

expression, in context with cellular-mediated acetylation. Thus, HTLV-1 regulates its

gene expression in concert with cellular transcriptional regulators to avoid immune

elimination from the host, and facilitate clonal expansion of infected cells and

maintenance of proviral loads in vivo.

161 MATERIALS AND METHODS

Cell lines.

293T cells were obtained from G. Franchini (National Cancer Institute, NIH,

Bethesda) and grown in modified Dulbecco’s eagle medium containing 10% fetal bovine

serum and 1% streptomycin and penicillin at 37°C. Cells were split and cultured in six-

well plates or 10 cm dishes to 50-60 % confluence 18 h before transfection.

Plasmids.

The pTRE-Luc plasmid and pRSV-B-Gal have been described previously51,52. pME-p30IIHA plasmid contains p30II sequence from HTLV-1 molecular clone, ACH upstream of influenza hemagglutinin (HA1). Site directed mutants K2, K3, K4 and K5

harbor single arginine substitution of lysines in pMEp30IIHA at amino acid position 103,

106, 123 or 143 respectively. Fidelity of these plasmids were confirmed by Sanger

sequencing and p30II HA protein expression was confirmed by western immunoblot

assay using monoclonal anti-HA antibody (Covance, Berkeley, CA ). ACH.30 plasmid

has a 24 bp insertion that causes an artificial termination codon in p30II40. pCMV-p300

expresses the full-length p300 protein from a CMV I/E promoter (Upstate USA Inc.,

Charlottesville, VA). HDAC-1 plasmid codes for HDAC-1 and was a kind gift from S.

Schreiber, Harvard University, Cambridge, MA. Wild type and HAT mutants of P/CAF

have been described previously21.

162 Cell transfection and reporter gene assay.

For each transfection, 0.3 µg of pTRE-Luc reporter plasmid was cotransfected with pME-p30IIHA plasmid and 0.4 µg ACH.30 plasmid when specified, using

Lipofectamine Plus (Invitrogen, Carlsbad, CA). When pCMV-p300 wild type or mutant

plasmids were used, it was also co-transfected at quantities specified. When HDAC-1

plasmid was used, it was also co-transfected at 0.0, 0.5, 1.0 or 1.5 µg concentrations.

When specified, 0.0, 0.6, 1.2 or 2.4 µg of wild type P/CAF plasmid was also co- transfected. When HAT mutant P/CAF plasmid was used, 2.4 µg of this plasmid was also co-transfected. As described previously12, TSA (Sigma, St. Louis, MO) was added to the

culture medium at 100 nm concentration, at 24 h posttransfection. As an internal control

for transfection efficiency, 0.1 µg of pRSV-βGal (Invitrogen) was also used in each

transfection. pME 18S was used as carrier DNA to equalize DNA concentrations for each

transfection. Transfected cells were lysed with 400 µl lysis buffer (Promega, Madison,

WI) per well for 25 min. Twenty microliters of each lysate was used to test luciferase

reporter gene activity using an Enhanced Luciferase Assay kit (Promega). To normalize

the transfection efficiency, cells were washed with PBS, stained with 5-bromo-4-chloro-

3-indolyl-beta-D-galactopyranoside (X-Gal) (Sigma) and βGal expressing cells were

counted using the microscope. Co-transfection of pME-p30IIHA had no effect on the X-

Gal staining (data not shown). Results were expressed as mean of optimized luciferase

activity (luciferase activity / percentage cells stained positive for β-Gal expression) in

arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate

experiments. Statistical analysis was done using Student’s t test, P < 0.05.

163 Immunoprecipitation of p30II with anti-acetyl lysine antibody.

Sixty percent confluent 293T cells were cotransfected with 15 µg of either pME-

p30IIHA alone or with 10 µg pCMV-p300 wild type or mutant plasmids using Superfect

(Gibco BRL). After 48 h, the transfected cells were washed with PBS and resuspended in

1 ml of lysis buffer containing 50 mM Tris-HCl (pH 8.0), 100 mM NaCl, 1 mM phenylmethylsulfonyl fluoride (PMSF), 0.5 µg of leupeptin per ml, and 1 µg of aprotinin per ml. Cell suspensions were incubated on ice for 20 min and the lysates were centrifuged at 14,000 rpm for 20 min at 4° C and 1.5 mg of the lysate was incubated for 2 hours at 4° C with 100 ng of anti-acetyl lysine antibody (Upstate USA Inc). After adding

70 µl of 100% of protein A/G agarose slurry (Pierce, Rockford, IL), the mixture was incubated overnight at 4° C. The immunoprecipitated complexes were washed twice with

10 volumes of lysis buffer. The components of the complexes were resolved on 15 %

SDS-polyacrylamide gels and detected by western immunoblot assay using monoclonal

HA antibody (Covance).

164 RESULTS

Histone acetyltransferase (HAT) activity of p300 modulates p30II-dependent

repression of LTR driven transcription

HTLV-1 Tax mediated transactivation of the viral LTR is dependent on

transcriptional co-adaptors CBP/p300. Tax is reported to bind p300 at the KIX domain,

the same domain that p30II binds51. Our previous reports suggested that p30II and Tax serve different roles in the regulation of transcription. Recently, we also reported that overexpressing p300 rescues p30II-mediated repression of LTR-driven reporter gene

expression, indicating the possibility that p30II represses LTR-driven reporter gene

activities as a consequence of competition for limited basal quantities of CBP/p300. Our

findings that the lysine residue at amino acid position 106 (K3) of p30II is critical for

repressing TRE-mediated transcription, suggested a potential role of acetylation in p30II- mediated LTR repression (Michael et al, 2004, manuscript submitted). To test the role of p300 HAT domain on p30II-dependent repression of LTR driven transcription, we

transiently co-transfected pME-p30IIHA and an LTR-luciferase reporter plasmid along

with either wild type or HAT mutant of p300 plasmids. In contrast to the wild type p300,

p300 HAT mutants only partially rescued the p30II-dependent repression of LTR-

luciferase reporter gene activity (Fig. 4.1A).

We then tested the role of histone acetyltransferase activity mediated by p300

plasmids on p30II-dependent repression of LTR driven transcription, in the presence of

the HTLV-1 provirus. We transiently co-transfected pME-p30IIHA, LTR-luc reporter 165 plasmid and ACH.30, a provirus deleted for p30II expression, along with either wild type

or HAT mutants of pCMV-p300, and tested the luciferase reporter activity. Interestingly, p300 HAT mutant plasmids failed to fully rescue p30II-dependent repression of LTR-

luciferase reporter gene activity, compared to the wildtype p300 plasmid, in the presence

of the HTLV-1 provirus as well (Fig. 4.1B). These data confirm that the HTLV-1 LTR

transcriptional regulatory effect of p30II is dependent on the HAT activity of p300.

Deacetylation by HDAC-1 enhances p30II-dependent repression of HTLV-1 LTR

driven transcription

Histone acetyl transferases and histone deacetylases (HDACs) form multiprotein

complexes with transcriptional factors and are known to regulate their transcriptional

activity12. In addition, HDAC-1 is known to physically and functionally interact with

HTLV-1 Tax and repress the LTR transactivation function of Tax12. To test if

deacetylation by HDAC-1 has a similar role on p30II-dependent repression of LTR

mediated transcription, we transiently co-transfected increasing concentrations of HDAC-

1 plasmid with constant concentration of pME-p30IIHA plasmid and LTR-luciferase

reporter plasmid and tested the luciferase reporter activity. Our data indicated that

HDAC-1 expression enhanced p30II-dependent repression of LTR-luciferase reporter gene activity, in a dose-dependent manner (Fig. 4.2A).

We then tested the role of deacetylation on p30II-dependent repression of LTR

mediated transcription, in the presence of the HTLV-1 provirus, by transiently co-

transfecting increasing concentrations of the HDAC-1 plasmid with constant 166 concentration of pME-p30IIHA plasmid, LTR-luciferase reporter plasmid and ACH.30, a

provirus deleted for p30II expression. Our data indicated that addition of HDAC-1

enhanced p30II-dependent repression of LTR-luciferase reporter gene activity, in a dose-

dependent manner, in the presence of the provirus (Fig. 4.2B). These data further support

our findings that the HTLV-1 LTR transcriptional modulatory function of p30II is dependent on deacetylation by HDAC-1.

Acetylation by trichostatin A (TSA) decreases p30II-dependent repression of HTLV-

1 LTR driven transcription

Reversible modification of the core histones is crucial in the regulation of gene

expression. While histone acetylation is typically associated with nucleosomal

decondensation necessary for transcriptional activation, histone deacetylation is

associated with nucleosomal condensation and subsequent transcriptional repression12.

Repression of Tax transactivation of HTLV-1 LTR by HDAC-1 can be restored when treated with an HDAC inhibitor, namely, trichostatin A12. Similarly, we tested the effect

of inhibition of deacetylation on p30II-dependent repression of LTR mediated

transcription. To determine the most effective concentration of TSA, we transiently co-

transfected 293T cells with 1.0 µg of HDAC-1 plasmid, 1.2 µg of pME-p30IIHA plasmid and 0.3 µg of LTR-luciferase reporter plasmid, and added either 0.0, 100, 200 or 400 nM

TSA at 24 h post-transfection. We then tested the luciferase reporter activity at 48 h post- transfection and compared the difference in luciferase activity in the presence of various concentrations of TSA and found that 100nM TSA had the highest effect in decreasing

167 p30II-dependent LTR repression (Fig. 4.3). Next, to determine the effect of TSA at

different concentrations of HDAC-1, we transiently co-transfected 293T cells with

increasing concentrations of a HDAC-1 plasmid with constant concentration of pME-

p30IIHA plasmid and LTR-luciferase reporter plasmid, and added 100 nM TSA at 24 h

post-transfection. We then tested the luciferase reporter activity at 48 h post-transfection

and compared the difference in luciferase activity in the presence or absence of TSA. Our

data demonstrated that TSA, an inhibitor of deacetylation decreases p30II-dependent

repression of LTR-luciferase reporter gene activity, at various doses of HDAC-1, in the

absence of the provirus (Fig. 4.4A).

We then tested the role of inhibition of deacetylation on p30II-dependent

repression of LTR mediated transcription, in the presence of the HTLV-1 provirus, by

transiently co-transfecting increasing concentrations of HDAC-1 plasmid with constant

concentration of pME-p30IIHA plasmid, LTR-luciferase reporter plasmid and ACH.30, a

provirus deleted for p30II expression and by adding 100 nM TSA at 24 h post-

transfection. We then tested the luciferase reporter activity at 48 h post-transfection and

compared the difference in luciferase activity in the presence or absence of TSA.

Interestingly, addition of TSA decreased p30II-dependent repression of LTR-luciferase

reporter gene activity, at various concentrations of HDAC-1, in the presence of the

provirus as well (Fig. 4.4B). These data confirm that the HTLV-1 LTR transcriptional

regulatory effect of p30II is dependent on deacetylation by HDAC-1 and inhibition of deacetylation by TSA.

168 P/CAF does not rescue p30II-dependent repression of HTLV-1 LTR driven

transcription

P/CAF is an important coactivator that mediates transcription through HAT

activity43. P/CAF acetylates various nonhistone transcription-related proteins, such as the

chromatin proteins HMG17 and HMG I(Y), activators p53, MyoD, human

immunodeficiency virus (HIV) Tat, and general transcription factors TFIIE and TFIIF43.

In addition, P/CAF is recruited to the Tax responsive element sites on the HTLV-1 LTR,

through direct interaction with Tax and enhance Tax-mediated HTLV-1 transcription, in a

HAT-independent manner21. We therefore tested if P/CAF has a role on p30II-dependent

repression of HTLV-1 LTR driven transcription. We transiently co-transfected increasing

concentrations of P/CAF plasmid with constant concentration of pME-p30IIHA plasmid

and LTR-luc reporter plasmid and tested the luciferase reporter activity. Our data

indicated that, unlike p300, P/CAF does not rescue p30II-dependent repression of HTLV-

1 LTR driven transcription (Fig. 4.5). In addition, to test if the HAT activity of P/CAF

influences p30II-dependent repression of HTLV-1 LTR driven transcription, we transiently co-transfected pME-p30IIHA plasmid and LTR-luc reporter plasmid along

with either wild type P/CAF plasmid or P/CAF HAT mutant plasmid, and tested the luc

reporter activity. We found that the P/CAF HAT mutant did not significantly alter p30II- dependent repression of HTLV-1 LTR driven transcription, compared to wild type

P/CAF (Fig. 4.5). These data confirm that the HTLV-1 LTR transcriptional regulatory effect of p30II is p300-dependent and not P/CAF-dependent.

169 HTLV-1 p30II is acetylated by transcriptional co-activator p300

To test if HTLV-1 p30II is acetylated by transcriptional co-activator p300, we

transiently transfected 293T cells with pMEp30IIHA alone or in the presence of either the

wild type p300 or p300 HAT mutants, and whole-cell extracts made at 48 h post-

transfection were used for immunoprecipitation assays. We used anti-acetyl lysine

antibody pull down and subsequent probing with monoclonal HA antibody.

Demonstration of acetylation was performed with a weak, but consistent band, presumed to be acetylation. We detected the presence of a 27 kDa protein in the immunoprecipitated samples containing p30II, which appeared more intense when

expressed in the presence of exogenous wild type p300 (Fig. 4.6). Moreover, in the

presence of p300 HAT mutants, we did not see an increase in the presence of the 27 kDa

band, suggesting that the HAT activity of p300 is critical in the acetylation of p30II(Fig.

4.6). Based on this data, it appears that HTLV-1 p30II is acetylated by the transcriptional co-activator p300. Moreover, we tested four arginine substitution mutants, carrying a mutation in the lysine residues, and found that there was a decrease in the intensity of the

27 kDa band in the K3 mutant, compared to wildtype or K2 or K4 or K5 mutants of p30II

(Fig. 4.6). Interestingly, we have previously demonstrated that K3 mutant that contains an arginine substitution of the lysine at amino acid 106 of p30II is critical for its repression

of LTR driven transcription (Michael et al, 2004, manuscript submitted). Our data

indicated that HTLV-1 p30II is acetylated at lysine 106 (K3) by the transcriptional co-

activator p300. Follow up experiments are being performed (a) using nuclear lysates (b)

at different time points and (c) using an in vitro transcription/translation system.

170 DISCUSSION

Our data is the first to demonstrate that the histone acetyl transferase activity of p300 is critical for modulation of HTLV-1 LTR mediated transcription by p30II. These findings confirm previous reports from our laboratory on the transcriptional regulatory function of HTLV-1 p30II51,52. We have previously reported that the amino acid domain

100-179 of p30II is critical for repression of HTLV-1 LTR, in the presence or absence of

the provirus. Amino acids 100-179 of HTLV-1 p30II contains 5 lysine residues, four of

which are part of a G/SK consensus acetylation sequence6 representing potential

acetylation sites for CBP/p300, and we have reported that the lysine residue at amino acid

position 106 (K3) of HTLV-1 p30II is critical for p30II-mediated repression of LTR

driven transcription in the presence or absence of the provirus (Michael et al, 2004,

manuscript submitted).

CREB binding protein (CBP) and p300 are highly related and evolutionarily

conserved transcriptional coactivators with similar nucleotide sequence, but divergent

functions. CBP and p300 are not interchangeable, however they share many functional

properties45, such as mediating the transcriptional control of various cellular and viral

DNA binding transcription factors9,16. There are several cellular and viral proteins that

interact with either CBP or p300, including steroid and retinoid hormone receptors,

CREB, c-Jun, c-Myb, Sap-1a, c-Fos, MyoD, p53, Stat-1/2, NF-κB, pp90rsk, TATA-

binding protein, TFIIB4,16,19,20, HTLV-1 Tax, adenovirus E1A, Kaposi’s sarcoma-

associated herpes virus (KSHV) viral interferon regulatory factor protein (vIRF), simian

171 virus 40 large T antigen, HIV-1 Tat, small delta antigen of hepatitis delta virus and

Epstein-Barr virus (EBV) nuclear antigen 3C (EBNA3C) and EBNA2, and herpes

simplex virion protein-16 (VP16) 1-3,10,23,27,29,44,46. Additionally, proteins such as papillomavirus E 2 protein employs novel mechanism of transcriptional activation, by

interacting with cellular protein, AMF-1 to form complexes with p30036.

Acetylation at lysine residues within nucleosomal histones is known to neutralize

the basic charge of lysine and thus loosen the integrity of nucleosomes. CBP and p300

bridge transcription factors to relevant promoters and through its intrinsic histone acetyl transferase (HAT) activity, acetylates lysine residues at the amino terminus of histones, which results in modification of chromatin structure to allow access of the transcriptional

machinery. Recently, the mechanism and functional significance of the interactions between many viral proteins and CBP/p300 are increasingly understood. For example, interaction between the adenovirus oncoprotein E1A and CBP/p300 is critical for E1A- mediated regulation of transcription, suppression of differentiation, and immortalization of cells in culture1,3,32. Modification of the HAT activity of CBP/p300 or the bridging of

the gap between DNA binding transcription factors and components of the general transcription machinery is the mechanism by which simian virus large T antigen regulates the expression of a group of cellular genes4,16. Acetylation of c-myb oncoprotein of the

avian retrovirus group by CBP/p300 at multiple lysine residues is considered to be the

mechanism by which CBP/p300 regulates the functions of c-myb10. Acetylation of

hepatitis δ virus small δ antigen by p300, at lysine 72 is thought to influence its cellular

localization and viral RNA synthesis33. HIV Tat is acetylated by p300 at lysine 50 and 51

in the TAR RNA binding domain, which promotes the dissociation of Tat from TAR 172 RNA during early transcription elongation24. Acetylation of Tat is also thought to

facilitate dissociation of the Tat cofactor cyclinT1 from TAR RNA, serving to transfer

Tat onto the elongating RNA polymerase II22. Overall, acetylation of Tat by p300/CBP is important for its transcriptional activation of the HIV promoter35. EBNA3C, which is

essential for EBV-dependent immortalization of human primary B lymphocytes, interacts

with p300 and acetylation is thought to be critical in its function11. In addition, EBV

nuclear protein protein EBNA2 and herpes simplex VP16 acidic domains utilize the

intrinsic HAT or scaffolding properties of p300 to activate transcription46. Interestingly,

KSHV vIRF interacts with and inhibits p300 HAT activity, resulting in drastic reduction of nucleosomal histone acetylation and alteration of chromatin structure and thus help the virus to circumvent the host antiviral immune response and to induce a global alteration of cellular gene expression29.

An array of transcriptional activators, coactivators, and histone deacetylases are

known to participate in the regulation of HTLV-1 transcription in infected T

lymphocytes28. Chromatin immunoprecipitation analysis investigating factor binding and

histone modification at the integrated HTLV-1 provirus in infected T lymphocytes,

demonstrated the presence of Tax, CBP, p300, histone deacetylases and a variety of

ATF/CREB and AP-1 family members such as CREB, CREB-2, ATF-1, ATF-2, c-Fos,

and c-Jun, at the HTLV-1 promoter28. CBP stimulates Tax-mediated HTLV-1 LTR

transcription initiation and reinitiation from a naked DNA template in vitro23 while p300

and P/CAF act as coactivators for Tax-dependent HTLV-1 LTR transcription21. Tax

directly interacts with both P/CAF and p300/CBP in a multi-histone acetyltransferase/

173 activator-enhancer complex 18. Histone H3 and H4 acetylation occurs at 3 regions within

the proviral genome, and histone H4 acetylation on the HTLV-1 promoter is known to

increase upon inhibition of histone deacetylases, leading to increase in viral RNA28.

Histone acetylation by CBP/p300 is critical in activation of HTLV-1 transcription from chromatin templates in vitro28. Furthermore, it is necessary that the coactivator HAT activity has to be tethered to the template by Tax and CREB, since a p300 mutant that

fails to interact with Tax was unable to facilitate transcription or acetylate histones31. p300 mediates HTLV-1 LTR expression by targeted nucleosomal acetylation of histones

H3 and H4, leading to RNAP II and TFIID recruitment to the chromatin template. CREB-

2, also known as ATF-4, is known to activate the HTLV-1 promoter, in association with

Tax15. CREB-2 is acetylated by CBP/p300, but not by P/CAF, however CREB-2 acetylation does not affect the Tax transactivation of HTLV-1 LTR14.

CBP and p300 mediate the activities of various transcription factors, however they are present only at limiting concentrations in cells. Even small reductions in the concentrations of these coactivator are damaging in many instances, such as the human

Rubinstein-Taybi syndrome where loss of a single CBP allele causes developmental defects37. Our study demonstrates that deacetylation by HDAC-1 enhances p30II- dependent repression of HTLV-1 LTR driven transcription while inhibition of deacetylation by TSA decreases p30II-dependent repression of HTLV-1 LTR driven transcription. Therefore, it appears that p30II acts as an activator of the LTR when

acetylated and as a repressor when deacetylated. Since the level of p300 is limited in

cells, we believe that the endogenous amount of p300 is sufficient to acetylate p30II when

174 provided at lower concentrations, but not at higher concentrations. This is consistent with

our earlier reports that at lower concentrations, p30II acts as an activator of HTLV-1 LTR

mediated transcription, and as a repressor of LTR mediated transcription at higher

concentrations52 (Michael et al, 2004, manuscript submitted).

CBP/p300 is known to form complexes with proteins such as P/CAF, which is

important for transcription as a coactivator and due to its ability to activate selective

promoters via intrinsic HAT activity8,26,38,39,17. P/CAF interacts with p300 and CBP49at

the same site as adenoviral oncoprotein E1A in a competitive manner, and acetylates

either free histones or nucleosomes49, primarily on lysine-14 of histone H3, and more

weakly on lysine-8 of histone H441. P/CAF is important for transcription and is required

as a HAT and coactivator43. Importantly, P/CAF has the ability to acetylate various nonhistone transcription-related proteins, such as the chromatin proteins HMG17 and

HMG I(Y), activators p53, MyoD and general transcription factors TFIIE and TFIIF43. In addition, P/CAF acetylates HIV-1 Tat at Lys28 in the activation domain,and enhances

Tat binding to the Tat-associated kinase, CDK9/P-TEFb24. P/CAF also interacts with and

acetylates adenoviral E1 B protein, and this is known to interfere with the physical

interaction between P/CAF and p53, preventing its activation and thus contributing to

viral transformation30. P/CAF interacts with and acetylates polyomavirus large T antigen, thus stimulate DNA replication when tethered near the polyomavirus origin47. Besides, interaction between human papillomavirus E7 oncoprotein and the acetyl transferase domain of P/CAF, result in reduction of its acetyltransferase activity and thereby, alteration of cellular gene expression and growth5. In EBV-infected cells that do not

175 express the EBV encoded oncoproteins EBNA2 or LMP1, p300 expression enhances the

ability of EBNA2 to up-regulate LMP1 expression and through its intrinsic HAT activity,

P/CAF can further potentiate this p300 effect46. Interestingly, P/CAF is recruited to the

Tax responsive element sites on the HTLV-1 LTR, through direct interaction with Tax

and enhance Tax-mediated HTLV-1 transcription, in a HAT-independent manner21.

However, based on our findings, HTLV-1 p30II-mediated repression of the LTR mediated

transcription appears to be dependent on p300, but independent of P/CAF.

Recently, it was reported that p30II modulates LTR mediated transcription, in the context of the entire provirus, by a post-transcriptional mechanism 34. However, we have

previously reported that p300 rescues LTR driven gene transcription, in a dose-dependent

manner, irrespective of the presence or absence of the provirus. This report provides

additional evidence and further elucidates the mechanism of p300-dependent LTR-

mediated transcriptional modulation by HTLV-1 p30II. Therefore, HTLV-1 p30II appears to be a multi-functional protein with transcriptional and post-transcriptional role in regulating gene expression.

Using deletion mutants of CBP/p300 in GST pull-down assays, previously, we localized the binding site of CBP/p300 for p30II to a highly conserved KIX region51.

Interestingly, KIX domain is the binding site of CBP/p300 for HTLV-1 Tax protein as well48. In addition, p30II was able to disrupt CREB-Tax-CBP/p300 complexes bound to

the viral 21-bp TRE repeats51. In circumstances where CBP/p300 levels are limiting,

proteins are known to compete for binding CBP/p300 and there could be selective

preference of one over the other, causing an exclusion of one of the proteins10. We have 176 previously reported that HTLV-1 p30II and Tax compete with each other in modulating

the transcriptional activity from the LTR, possibly through competitive CBP/p300

binding between p30II and Tax 52(Michael et al, 2004, manuscript submitted). Such an

environment of coactivator competition between transcription factors provides an

additional layer of regulated gene expression. Overall, we believe that the relative amounts of Tax and p30II may be critical in determining which protein binds to

CBP/p300 and the consequent regulation of LTR mediated transcription of viral genes

from the HTLV-1 viral promoter. At low concentrations, p30II and Tax might act

synergistically to enhance transcription from the viral promoter, but at higher

concentrations, we believe that p30II might antagonize the Tax transactivation of the

HTLV-1 promoter (Fig. 4.7). Information on the relative levels of various HTLV-1

proteins during various stages of the infection is lacking and it is possible that differences

in expression levels of various viral proteins, leading to differences in transcriptional

regulation might be the mechanism by which HTLV-1 infection/ disease progresses though various stages. This is likely to be in synergy with differential regulation of cellular gene regulation by Tax and/or p30II and thus changing the immune responses in

accordance with different stages of progression of HTLV-1 infection and disease. Our

present data can be supported by recent findings from our laboratory where an infectious

HTLV-1 molecular clone failed to maintain viral loads in vivo when p30II expression was

abolished42. Overall, our data indicates that HTLV-1, a complex retrovirus associated

with lymphoproliferative disorders, uses accessory genes to regulate viral gene expression from the LTR, avoid immune elimination by the host and facilitate clonal

expansion of infected cells and maintenance of proviral loads in vivo.

177 REFERENCES

1. Arany, Z., D. Newsome, E. Oldread, D. M. Livingston, and R. Eckner. 1995. A family of transcriptional adaptor proteins targeted by the E1A oncoprotein. Nature 374:81-84.

2. Ariumi, Y., A. Kaida, J. Y. Lin, M. Hirota, O. Masui, S. Yamaoka, Y. Taya, and K. Shimotohno. 2000. HTLV-1 tax oncoprotein represses the p53-mediated trans-activation function through coactivator CBP sequestration. Oncogene 19:1491-1499.

3. Avantaggiati, M. L., M. Carbone, A. Graessmann, Y. Nakatani, B. Howard, and A. S. Levine. 1996. The SV40 large T antigen and adenovirus E1a oncoproteins interact with distinct isoforms of the transcriptional co-activator, p300. EMBO J. 15:2236-2248.

4. Avantaggiati, M. L., V. Ogryzko, K. Gardner, A. Giordano, A. S. Levine, and K. Kelly. 1997. Recruitment of p300/CBP in p53-dependent signal pathways. Cell 89:1175-1184.

5. Avvakumov, N., J. Torchia, and J. S. Mymryk. 2003. Interaction of the HPV E7 proteins with the pCAF acetyltransferase. Oncogene 22:3833-3841.

6. Bannister, A. J. and E. A. Miska. 2000. Regulation of gene expression by transcription factor acetylation. Cell Mol. Life Sci. 57:1184-1192.

7. Bartoe, J. T., B. Albrecht, N. D. Collins, M. D. Robek, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J. Virol. 74:1094-1100.

8. Blanco, J. C., S. Minucci, J. Lu, X. J. Yang, K. K. Walker, H. Chen, R. M. Evans, Y. Nakatani, and K. Ozato. 1998. The histone acetylase PCAF is a nuclear receptor coactivator. Genes Dev. 12:1638-1651.

9. Blobel, G. A. 2000. CREB-binding protein and p300: molecular integrators of hematopoietic transcription. Blood 95:745-755.

10. Colgin, M. A. and J. K. Nyborg. 1998. The human T-cell leukemia virus type 1 oncoprotein Tax inhibits the transcriptional activity of c-Myb through competition for the CREB binding protein. J. Virol. 72:9396-9399.

11. Cotter, M. A. and E. S. Robertson. 2000. Modulation of histone acetyltransferase activity through interaction of epstein-barr nuclear antigen 3C with prothymosin alpha. Mol. Cell Biol. 20:5722-5735.

178 12. Ego, T., Y. Ariumi, and K. Shimotohno. 2002. The interaction of HTLV-1 Tax with HDAC1 negatively regulates the viral gene expression. Oncogene 21:7241- 7246.

13. Franchini, G. 1995. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood 86:3619-3639.

14. Gachon, F., C. Devaux, and J. M. Mesnard. 2002. Activation of HTLV-I transcription in the presence of Tax is independent of the acetylation of CREB-2 (ATF-4). Virology 299:271-278.

15. Gachon, F., A. Peleraux, S. Thebault, J. Dick, I. Lemasson, C. Devaux, and J. M. Mesnard. 1998. CREB-2, a cellular CRE-Dependent transcription repressor, functions in association with tax as an activator of the human T-cell leukemia virus type 1 promoter. J Virol 72:8332-8337.

16. Goodman, R. H. and S. Smolik. 2000. CBP/p300 in cell growth, transformation, and development. Genes Dev. 14:1553-1577.

17. Grunstein, M. 1997. Histone acetylation in chromatin structure and transcription. Nature 389:349-352.

18. Harrod, R., Y. L. Kuo, Y. Tang, Y. Yao, A. Vassilev, Y. Nakatani, and C. Z. Giam. 2000. p300 and p300/cAMP-responsive element-binding protein associated factor interact with human T-cell lymphotropic virus type-1 Tax in a multi- histone acetyltransferase/activator-enhancer complex. J. Biol. Chem. 275:11852-11857.

19. Janknecht, R. and T. Hunter. 1996. Transcription. A growing coactivator network [news; comment]. Nature 383:22-23.

20. Janknecht, R. and T. Hunter. 1996. Transcriptional control: Versatile molecular glue. Curr. Biol. 6:951-954.

21. Jiang, H., H. Lu, R. L. Schiltz, C. A. Pise-Masison, V. V. Ogryzko, Y. Nakatani, and J. N. Brady. 1999. PCAF interacts with tax and stimulates tax transactivation in a histone acetyltransferase-independent manner [published erratum appears in Mol Cell Biol 2000 Mar;20(5):1897]. Mol. Cell Biol. 19:8136- 8145.

22. Kaehlcke, K., A. Dorr, C. Hetzer-Egger, V. Kiermer, P. Henklein, M. Schnoelzer, E. Loret, P. A. Cole, E. Verdin, and M. Ott. 2003. Acetylation of Tat defines a cyclinT1-independent step in HIV transactivation. Mol. Cell 12:167-176.

23. Kashanchi, F., J. F. Duvall, R. P. Kwok, J. R. Lundblad, R. H. Goodman, and J. N. Brady. 1998. The coactivator CBP stimulates human T-cell lymphotrophic virus type I Tax transactivation in vitro. J. Biol. Chem. 273:34646-34652.

179 24. Kiernan, R. E., C. Vanhulle, L. Schiltz, E. Adam, H. Xiao, F. Maudoux, C. Calomme, A. Burny, Y. Nakatani, K. T. Jeang, M. Benkirane, and C. Van Lint. 1999. HIV-1 tat transcriptional activity is regulated by acetylation. EMBO J. 18:6106-6118.

25. Kitze, B. and J. N. Brady. 1997. Human T cell lymphotropic retroviruses: Association with diseases of the nervous system. Intervirology 40:132-142.

26. Korzus, E., J. Torchia, D. W. Rose, L. Xu, R. Kurokawa, E. M. Mcinerney, T. M. Mullen, C. K. Glass, and M. G. Rosenfeld. 1998. Transcription factor-specific requirements for coactivators and their acetyltransferase functions. Science 279:703-707.

27. Kwok, R. P., M. E. Laurance, J. R. Lundblad, P. S. Goldman, H. Shih, L. M. Connor, S. J. Marriott, and R. H. Goodman. 1996. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature 380:642-646.

28. Lemasson, I., N. J. Polakowski, P. J. Laybourn, and J. K. Nyborg. 2002. Transcription factor binding and histone modifications on the integrated proviral promoter in human T-cell leukemia virus-I-infected T-cells. J Biol. Chem. 277:49459-49465.

29. Li, M., B. Damania, X. Alvarez, V. Ogryzko, K. Ozato, and J. U. Jung. 2000. Inhibition of p300 histone acetyltransferase by viral interferon regulatory factor. Mol. Cell Biol. 20:8254-8263.

30. Liu, Y., A. L. Colosimo, X. J. Yang, and D. Liao. 2000. Adenovirus E1B 55- kilodalton oncoprotein inhibits p53 acetylation by PCAF. Mol. Cell Biol. 20:5540- 5553.

31. Lu, H., C. A. Pise-Masison, T. M. Fletcher, R. L. Schiltz, A. K. Nagaich, M. Radonovich, G. Hager, P. A. Cole, and J. N. Brady. 2002. Acetylation of nucleosomal histones by p300 facilitates transcription from tax-responsive human T-cell leukemia virus type 1 chromatin template. Mol. Cell Biol. 22:4450-4462.

32. Lundblad, J. R., R. P. Kwok, M. E. Laurance, M. L. Harter, and R. H. Goodman. 1995. Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature 374:85-88.

33. Mu, J. J., Y. G. Tsay, L. J. Juan, T. F. Fu, W. H. Huang, D. S. Chen, and P. J. Chen. 2004. The small delta antigen of hepatitis delta virus is an acetylated protein and acetylation of lysine 72 may influence its cellular localization and viral RNA synthesis. Virology 319:60-70.

180 34. Nicot, C., M. Dundr, J. M. Johnson, J. R. Fullen, N. Alonzo, R. Fukumoto, G. L. Princler, D. Derse, T. Misteli, and G. Franchini. 2004. HTLV-1-encoded p30(II) is a post-transcriptional negative regulator of viral replication. Nat. Med.

35. Ott, M., M. Schnolzer, J. Garnica, W. Fischle, S. Emiliani, H. R. Rackwitz, and E. Verdin. 1999. Acetylation of the HIV-1 Tat protein by p300 is important for its transcriptional activity. Curr. Biol. 9:1489-1492.

36. Peng, Y. C., D. E. Breiding, F. Sverdrup, J. Richard, and E. J. Androphy. 2000. AMF-1/Gps2 binds p300 and enhances its interaction with papillomavirus E2 proteins. J. Virol. 74:5872-5879.

37. Petrij, F., R. H. Giles, H. G. Dauwerse, J. J. Saris, R. C. Hennekam, M. Masuno, N. Tommerup, G. J. van Ommen, R. H. Goodman, and D. J. Peters. 1995. Rubinstein-Taybi syndrome caused by mutations in the transcriptional co- activator CBP [see comments]. Nature 376:348-351.

38. Puri, P. L., V. Sartorelli, X. J. Yang, Y. Hamamori, V. V. Ogryzko, B. H. Howard, L. Kedes, J. Y. Wang, A. Graessmann, Y. Nakatani, and M. Levrero. 1997. Differential roles of p300 and PCAF acetyltransferases in muscle differentiation. Mol. Cell 1:35-45.

39. Reid, J. L., A. J. Bannister, P. Zegerman, M. A. Martinez-Balbas, and T. Kouzarides. 1998. E1A directly binds and regulates the P/CAF acetyltransferase. EMBO J 17:4469-4477.

40. Robek, M. D., F. H. Wong, and L. Ratner. 1998. Human T-Cell leukemia virus type 1 pX-I and pX-II open reading frames are dispensable for the immortalization of primary lymphocytes. J. Virol. 72:4458-4462.

41. Schiltz, R. L., C. A. Mizzen, A. Vassilev, R. G. Cook, C. D. Allis, and Y. Nakatani. 1999. Overlapping but distinct patterns of histone acetylation by the human coactivators p300 and PCAF within nucleosomal substrates. J Biol. Chem. 274:1189-1192.

42. Silverman, L. R., A. J. Phipps, A. Montgomery, L. Ratner, and M. D. Lairmore. 2004. Human T-cell lymphotropic virus type 1 open reading frame II- encoded p30II is required for in vivo replication: evidence of in vivo reversion. J Virol 78:3837-3845.

43. Sterner, D. E. and S. L. Berger. 2000. Acetylation of histones and transcription- related factors. Microbiol. Mol. Biol. Rev. 64:435-459.

44. Suzuki, T., M. Uchidatoita, and M. Yoshida. 1999. Tax protein of HTLV-1 inhibits CBP/p300-mediated transcription by interfering with recruitment of CBP/p300 onto DNA element of E-box or p53 binding site. Oncogene 18:4137- 4143. 181 45. Vo, N. and R. H. Goodman. 2001. CREB-binding protein and p300 in transcriptional regulation. J. Biol. Chem. 276:13505-13508.

46. Wang, L., S. R. Grossman, and E. Kieff. 2000. Epstein-Barr virus nuclear protein 2 interacts with p300, CBP, and PCAF histone acetyltransferases in activation of the LMP1 promoter. Proc. Natl. Acad. Sci. U. S. A 97:430-435.

47. Xie, A. Y., V. P. Bermudez, and W. R. Folk. 2002. Stimulation of DNA replication from the polyomavirus origin by PCAF and GCN5 acetyltransferases: acetylation of large T antigen. Mol. Cell Biol. 22:7907-7918.

48. Yan, J. P., J. E. Garrus, H. A. Giebler, L. A. Stargell, and J. K. Nyborg. 1998. Molecular interactions between the coactivator CBP and the human T-cell leukemia virus Tax protein. J. Mol. Biol. 281:395-400.

49. Yang, X. J., V. V. Ogryzko, J. Nishikawa, B. H. Howard, and Y. Nakatani. 1996. A p300/CBP-associated factor that competes with the adenoviral oncoprotein E1A. Nature 382:319-324.

50. Yoshida, M. 1996. Molecular biology of HTLV-I: recent progress. J. Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S63-8:S63-S68.

51. Zhang, W., J. W. Nisbet, B. Albrecht, W. Ding, F. Kashanchi, J. T. Bartoe, and M. D. Lairmore. 2001. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J. Virol. 75:9885-9895.

52. Zhang, W., J. W. Nisbet, J. T. Bartoe, W. Ding, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J. Virol. 74:11270-11277.

182

Figure 4.1. p300 HAT mutants only partially rescue p30II-dependent repression of LTR driven transcription in the absence of the provirus (A) or the presence of the provirus (B). 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid, 1.2 µg pME-p30II HA, 2.4 µg of either the wild type p300 plasmid or one of the p300 HAT mutant plasmids, M3 or M7, +/- 0.4 µg of ACH.30 proviral plasmid and the luciferase activity was measured. A plasmid containing CMV promoter was used as carrier DNA to equalize DNA concentrations for each transfection. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed, comparing to the values obtained in the presence of p30II and CMV empty plasmid, using Student’s t test. * indicates P < 0.05.

183

Figure 4.2. Addition of HDAC-1 enhances p30II-dependent repression of HTLV-1 LTR driven transcription in the absence of the provirus (A) or the presence of the provirus (B). 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid, 1.2 µg pME-p30II HA, 0.0, 0.5, 1.0 or 1.5 µg HDAC-1 plasmid and +/- 0.4 µg of ACH.30 proviral plasmid and the luciferase activity was measured. A plasmid containing the same promoter as the HDAC-1 plasmid was used as carrier DNA to equalize DNA concentrations for each transfection. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed, comparing to the values obtained in the presence of p30II and no HDAC-1 plasmid, using Student’s t test. * indicates P < 0.05.

184

Figure 4.3. Addition of 100 nM Trichostatin A (TSA) causes a decrease in the enhancement of p30II-dependent repression of HTLV-1 LTR driven transcription by HDAC-1. 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid, 1.2 µg pME-p30II HA and 1.0 µg HDAC-1 plasmid. At 24 h post- transfection, 100 nM, 200 nM or 400 nM TSA (Sigma) was added to the culture medium, as described previously12 and the luciferase activity was measured. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed, comparing to the values obtained in the absence of TSA, using Student’s t test. * indicates P < 0.05. In the presence of 200 nM TSA, P = 0.23. In the presence of 400 nM TSA, P = 0.34

185

Figure 4.4. Addition of Trichostatin A causes a decrease in the enhancement of p30II- dependent repression of HTLV-1 LTR driven transcription by HDAC-1 in the absence (A) or the presence of the provirus (B). 293T cells were transiently cotransfected with 0.3 µg of pLTR-luciferase reporter plasmid, 1.2 µg pME-p30II HA, 0.0, 0.5, 1.0 or 1.5 µg HDAC-1 plasmid and +/- 0.4 µg of ACH.30 proviral plasmid. A plasmid containing the same promoter as the HDAC-1 plasmid was used as carrier DNA to equalize DNA concentrations for each transfection. At 24 h post-transfection, 100 nM TSA (Sigma) was added to the culture medium, when specified, as described previously12 and the luciferase activity was measured. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed, comparing between values obtained in the presence or absence of TSA, at the same concentration of HDAC-1, using Student’s t test. * indicates P < 0.05.

186

Figure 4.5. Expression of P/CAF does not rescue p30II-dependent repression of HTLV-1 LTR driven transcription. 293T cells were transiently cotransfected with 0.3 µg of pLTR- luciferase reporter plasmid, 1.2 µg pME-p30II HA and 0.0, 0.6, 1.2 or 2.4 µg wildtype P/CAF or 2.4 µg P/CAF HAT mutant plasmid, and the luciferase activity was measured. A plasmid containing the same promoter as the P/CAF plasmids (CMV promoter) was used as carrier DNA to equalize DNA concentrations for each transfection. Results represent the mean of optimized luciferase activity in arbitrary light units (ALU) with standard error (SE) from a minimum of triplicate experiments. Statistical analysis was performed, comparing to the values obtained in the presence of p30II alone, using Student’s t test. Statistical significance was also performed, to compare between pCAF wildtype and HAT mutant, using Student’s t test. * indicates P < 0.05.

187

Figure 4.6. HTLV-1 p30II appears to be acetylated by the transcriptional co-activator p300 at lysine 106 (K3). (A) p30II appears to be acetylated by the transcriptional co- activator p300. Demonstration of acetylation was performed with a weak, but consistent band, presumed to be acetylation. 293T cells were transiently cotransfected with 15 µg of pME-p30II HA wildtype alone or in the presence of 10 µg of either the wildtype p300 expression plasmid or p300 HAT mutant. Cell lysates were pulled down using anti-acetyl lysine antibody (Upstate USA Inc.) and subsequently immunopreciptated with monoclonal anti-HA antibody (Covance). Follow up experiments are being performed (a) using nuclear lysates (b) at different time points and (c) using an in vitro transcription/translation system (B). p30II appears to be acetylated at lysine 106 (K3). Demonstration of acetylation was performed with a weak, but consistent band, presumed to be acetylation. 293T cells were transiently cotransfected with 15 µg of pME-p30II HA wildtype or arginine substitution mutants K2, K3, K4 or K5. Cell lysates were pulled down using anti-acetyl lysine antibody (Upstate USA Inc.) and subsequently immunopreciptated with monoclonal anti-HA antibody (Covance). K2, K3, K4 and K5 represent single amino acid mutations in p30II, where arginine is substituted for lysine at amino acid position 103, 106, 123 or 143 respectively.

188

Figure 4.7. Schematic model diagram of the potential mechanism by which HTLV-1 p30II modulates LTR-driven transcription. Tax and p30II compete with each other in modulating the LTR-driven transcription, possibly through competitive CBP/p300 binding, since both the viral proteins bind CBP/p300 through the KIX domain. At low concentrations, p30II and Tax might act synergistically to enhance transcription from the viral promoter, but at higher concentrations, we believe that p30II antagonizes Tax transactivation of the LTR, by sequestering p300, and thus disrupting CREB-Tax- CBP/p300 complexes at the TRE repeats on the viral promoter.

189

CHAPTER 5

SYNOPSIS AND FUTURE DIRECTIONS

HTLV-1 is able to evade immune elimination by the host and facilitate

lymphocyte proliferation, the molecular mechanisms of which are still not completely

understood. Distinct features of HTLV-1 pathogenesis include HTLV-1 mediated T cell

activation and transformation, which has been extensively investigated with specific

focus on the role of HTLV-1 Tax12,14. Tax, is the viral transcriptional transactivator

protein, which is considered to be the major viral oncoprotein with critical role in HTLV-

1 mediated T cell activation and lymphocyte transformation12,14. Animal model studies and in vitro transformation assays have established the oncogenic potential of Tax 12,20,30.

Although Tax appears to be responsible for multiple events necessary for HTLV-1- mediated lymphocyte immortalization, it is not certain whether Tax aids the virus to establish persistent infection. Recent studies from our laboratory, using molecular clones of HTLV-1 with selective mutations of ORF I and II, demonstrated that the expression of the accessory proteins encoded by pX ORFs I and II, p12I and p13II/p30II are critical for

the efficient HTLV-1 infection and maintenance of viral loads in vivo in a rabbit model of infection, although potentially dispensable for viral replication under activation conditions in vitro1,4,34. Additionally, previous reports from our laboratory 42,43and data 190 presented in this thesis demonstrate that HTLV-1 pX ORF II encoding protein, p30II regulates viral gene expression from HTLV-1 LTR and may thus be critical for the virus to evade immune recognition and thereby facilitate the establishment of persistent infection in vivo. In addition, data presented in this thesis demonstrate that p30II plays an important role in regulating cellular gene expression and activates many key transcription factors involved in T cell activation, such as NFAT, NFκB and AP-1. These results not only add to our understanding of the role of HTLV-1 p30II in HTLV-1 pathogenesis, but

also shed light on the potential mechanism by which p30II modulates HTLV-1 viral gene

expression and cellular gene expression in T lymphocytes. Taken together, data presented herein, indicates that HTLV-1, a complex retrovirus associated with lymphoproliferative disorders, uses accessory genes products such as p30II to promote lymphocyte activation

to enhance clonal expansion of infected cells and maintain proviral loads in vivo.

5.1 Role of p30II in T cell signaling / activation and transformation

Gene array data presented in this thesis shed light on the possible role of p30II in

T cell signaling / activation. Expression of p30II was associated with decrease in mRNA

levels of CD28, a co-stimulatory molecule with a distinct role in T lymphocyte activation

5. Besides, p30II expression appeared to upregulate Vav-2 and CD72 while

downregulating CD46 and Lck tyrosine kinase, a member of the Src family protein

tyrosine kinases that participates in signal transduction pathways initiated by T cell

surface receptors such as TCR/CD3, CD2, CD4, CD8, and CD28 29. Additionally, p30II

expression correlated with decrease in the level of CHP, an endogenous calcineurin

inhibitor, indicating a likely increase in calcineurin activity, which would be predicted to

191 correlate with upregulation of NFAT expression by p30II. Moreover, p30II expression was associated with increased expression of c-Jun and c-Fos, suggesting activation of AP-1 mediated transcription. In our gene array studies, p30II expression was associated with decreased expression of protein kinase D (PKD), which negatively modulates JNK signaling pathway 18, mediates cross-talk between different signaling systems, and is critical in processes as diverse as cell proliferation, apoptosis, immune cell regulation and tumor cell invasion 37. Interestingly, p30II expression was also associated with decrease in epidermal growth factor (EGF), which stimulates calcium influx 41 and activates ETS domain transcription factor Elk-1 33. In Jurkat T lymphocytes expressing p30II, there were no detectable levels of IKKγ, which is important for NF-κB signaling in response to both

T cell activation signals and Tax 35. However p30II expression increased HPK-1

(Hematopoetic Progenitor Kinase-1), a known NF-κB activator3. p30II expression was associated with decreased Ras GRP2, a guanyl nucleotide exchange factor that increases

Ras-GTP, suggesting a decrease in the level of activated Ras (Ras-GTP). In addition, using luciferase reporter assays, we found that p30II enhanced NFAT, NF-κB and AP-1 driven transcription. Therefore, it is possible that p30II increases IL-2 production, cause T cell activation and lymphocyte transformation.

5.2 Identify the mechanism by which p30II activates NF-κB, NFAT and AP-1 mediated transcription

Herein, we demonstrate that p30II enhanced NFAT, NF-κB and AP-1 driven transcription. p30II increased the NFAT driven luciferase reporter gene activity by 2.2,

4.5, 7.7 and 10.7 fold, when left unstimulated, stimulated with PMA or ionomycin alone 192 or both respectively and 5.2, 2.4 and 5 fold in the presence of CD3 or CD28 alone or both

respectively, indicating that p30II effectively enhanced NFAT driven transcription, when

stimulated with ionomycin or CD3. HTLV-1 p30II increased the AP-1 driven luciferase reporter gene activity by 2.6, 5.2, 1.3, 4.9 fold, when left unstimulated, stimulated with

PMA or ionomycin alone or both respectively and only 1.2, 1 and 1 fold in the presence of CD3 or CD28 alone or both repectively, indicating that p30II effectively enhanced AP-

1 driven transcription, significantly when stimulated with PMA. HTLV-1 p30II increased

the NF-κB driven luciferase reporter gene activity by 11.4, 3.7, 3.1 and 3.8 fold, when

left unstimulated, stimulated with PMA or ionomycin alone or both respectively and 5.6,

6 and 3.6 fold in the presence of CD3 or CD28 alone or both respectively, indicating that

p30II is able to effectively enhance NF-κB driven transcription, significantly even under

unstimulated conditions.

Binding of transcription factors such as, NFAT, AP-1 and NF-κB is required for

the expression from the IL-2 promoter. While NFAT is vital to proliferation of peripheral

lymphocytes for HTLV-1 infection 38, AP-1 is linked to the dysregulated phenotypes of

HTLV-1 infected T cells 11 and malignant transformation 28. Activation of AP-1 occurs

through Tax-dependent and independent mechanisms in HTLV-1-infected T cells in vitro

and in leukemia cells in vivo 11. NF-κB is highly activated in many hematopoietic

malignancies, HTLV-1 infected T cell lines and in primary ATL cells, even when Tax expression levels are low 28 and due to its anti-apoptotic activity, it is considered to be a

key survival factor for several types of cancer.

193 In parallel, HIV-1 accessory proteins Nef and Vpr have been implicated in

regulation of cellular gene expression during HIV-1 infection. Vpr-induced modest

increase in NFAT, AP-1, NF-κB and LTR-dependent transcription is dependent on its

ability to cause G2 arrest25. While Nef stimulates calcium signaling to activate NFAT,

Vpr potentiates Nef-induced activation of NFAT-dependent transcription by functioning farther downstream25. Previous studies from our laboratory have indicated that p12I mediated NFAT activation is dependent on cytosolic calcium, similar to HIV-1 Nef8.

Similar to HIV-1 Nef and Vpr, it is possible that p12I and p30II act synergistically, to modulate NFAT driven transcription and subsequent T cell activation / signaling.

Data presented in this thesis demonstrates the first HTLV-1 accessory protein

with broad modulating activities on the transcriptional activity of NF-κB, NFAT and AP-

1. However, additional research is crucial in elucidating the mechanisms employed by

p30II to enhance NFAT, AP-1 and NF-κB mediated transcription. It will also be valuable

to determine the functional domains and specific amino acids within p30II, critical for the

activation of NF-kB, NFAT and AP-1 mediated transcription.

5.3 Determine if p30II has an effect on IL-2 production and lymphocyte proliferation

Although T lymphocyte proliferation is highly regulated, HTLV-1 has developed

numerous mechanisms to destabilize this control and induce T lymphocyte proliferation

in the absence of appropriate signals. High levels of IL-2 are necessary to promote the

progression of resting T cells through cell cycle, leading to T lymphocyte

194 proliferation16,17. Most of the HTLV-1 infected T lymphocytes require IL-2 to proliferate

during early stages of the infection, although the T lymphocyte proliferation becomes IL-

2 independent later in the infection 39. The role of Tax on the IL-2 pathway and

lymphocyte proliferation has been extensively investigated12. In addition, previous

studies from our laboratory showed that another HTLV-1 protein, namely, p12I enhances

IL-2 production in T lymphocytes9. Data presented in this thesis demonstrates that p30II enhanced NFAT, NF-κB and AP-1 driven transcription. Based on this, we hypothesize that p30II has a major effect on IL-2 production and lymphocyte proliferation.

5.4 Does p30II affect cell cycle progression in T lymphocytes?

Dysregulation of the cell cycle is a key concept in HTLV-1 leukemogenesis and

the ability of Tax to dysregulate the cell-cycle machinery to regulate DNA replication

and cell division has been a focus of a multitude of investigations. Interestingly, as

described in this thesis, we found that p30II expression altered the expression of multiple genes involved in the regulation of different stages of cell cycle. These include checkpoint suppressor 1, cytosolic branched-chain amino acid transaminase 1, histone deacetylase 6, cyclin B1, WEE1 kinase, CDC14A, Lck, JAK2, GAS7, BZAP45, Cullin,

Rab6 GTPase activating protein (downregulated) and TERF1, AKAP8, DDX11, MSH2

and JUN (upregulated). Another gene downregulated by p30II expression was MDM2,

which, interestingly is overexpressed in certain leukemias24 and capable of enhancing the

tumorigenic potential of cells by inhibiting p300 / PCAF mediated p53 acetylation 27.

195 In parallel, HIV-1 Vpr is known to cause G2/M cell cycle arrest, characterized by

low levels of Cyclin B1, by directly binding and inhibiting the in vitro activity of a phosphatase, Cdc25C13. In addition, the cell cycle regulatory kinase, WEE-1, is also

depleted following prolonged G2 arrest induced by Vpr40. Interestingly, overexpression

of WEE-1 suppresses Vpr-mediated apoptosis and is thought to regulate Vpr and γ

irradiation-mediated apoptosis and possibly serve as a general regulator linking the cell

cycle to some pathways of apoptosis40. Our gene array findings that p30II expression was

associated with decreased expression of cyclin B1 and WEE-1, indicate another

functional similarity between p30II and Vpr. Based on the clues available from our gene

array study, future studies are essential not only to verify the findings, but also to test the

functional and biological significance of p30II in HTLV-1 pathogenesis.

5.5 Does p30II affect cell-to-cell adhesion between T lymphocytes?

Lymphocytes naturally infected with HTLV-1 produce virtually no cell-free

infectious HTLV-1 particles and cell-to-cell contact is necessary for efficient

transmission of the virus between cells and between individuals7,10. However, until recently, there was little known about the mechanism of cell-to-cell spread of HTLV-1.

Recently, Igakura et al19 demonstrated that cell contact rapidly induces polarization of the

cytoskeleton of the HTLV-1 infected cell to the cell-cell junction. The molecules that

mediate this process are not elucidated, and HTLV-1 Env protein is thought to be

involved in this, because it is the only HTLV-1 protein expressed intact on the outside of

the virus infected cell. Nonetheless, HTLV-1 is known to upregulate the expression levels

of certain adhesion molecules such as integrins32,36, which would increase the chances of cell to cell adhesion. 196 Gene array data presented in this thesis shed light on the possible mechanisms by

which p30II functions in HTLV-1 pathogenesis and in leukemogenesis, including in cell- to-cell adhesion between T lymphocytes. Based on our findings, p30II expression

appeared to alter the expression of several genes involved in cell adhesion. These include

decrease in integrin (integrin β8) immunoglobulin (MADCAM1), a counter-receptor for

P-selectin (SELPLG), cadherin (desmocollin 3), protocadherin (PC-LKC) liprin

(PPF1BP1), CD84 / Ly-9, CD58, CD43 / sialophorin and glycosyl-phosphatidyl-inositol phospholipase D1. Expression of p30II correlated with increase in integrin receptor α1

subunit and KIT ligand. Additional studies are essential to confirm these findings and to test their functional and biological significance in HTLV-1 pathogenesis.

5.6 Determine the significance of critical functional domains and specific amino acids within p30II, in vivo in rabbit model of infection

Data presented in this thesis demonstrates that amino acids 100-179 of p30II are

critical in regulating LTR mediated transcription, irrespective of the presence or absence

of the provirus, in our LTR-luciferase reporter assays. We have also identified the

domain of p30II, which is critical in the interaction between p30II and Tax. By dissecting

the amino acid domain (100-179) of p30II critical for repressing LTR mediated transcription further, we demonstrated that the lysine residue at amino acid position 106

(K3) of HTLV-1 p30II is critical for its repression of TRE-mediated transcription. In vivo

experiments using infectious molecular clones with mutations in the motifs or specific

amino acids, critical for the functions in vitro are necessary to determine the biological

significance of these findings. Therefore, the findings from the deletion and site directed

197 mutational analysis, presented in this thesis, are valuable in designing experiments to

reintroduce these mutations into infectious molecular clones to test their functional

significance in vivo.

5.7 Does HTLV-1 p30II form complexes at the HTLV-1 promoter, to modulate LTR mediated transcription?

An array of transcriptional activators, coactivators, and histone deacetylases are known to participate in the regulation of HTLV-1 transcription in infected T cells26.

Chromatin immunoprecipitation analysis investigating factor binding and histone modification at the integrated HTLV-1 provirus in infected T cells demonstrated the presence of Tax, CBP, p300, histone deacetylases and a variety of ATF/CREB and AP-1 family members such as CREB, CREB-2, ATF-1, ATF-2, c-Fos, and c-Jun, at the HTLV-

1 promoter26. CBP stimulates Tax-mediated HTLV-1 LTR transcription initiation and

reinitiation from a naked DNA template in vitro22 while p300 and pCAF act as

coactivators for Tax-dependent HTLV-1 LTR transcription21. Tax directly interacts with

both pCAF and p300/CBP in a multi-histone acetyltransferase/ activator-enhancer

complex 15 and histone acetylation by CBP/p300 is critical in the activation of HTLV-1

transcription from chromatin templates in vitro26. p300 mediates HTLV-1 LTR

expression by targeted nucleosomal acetylation of histones H3 and H4, leading to RNAP

II and TFIID recruitment to the chromatin template26.

Our laboratory has previously demonstrated that p30II binds p300 at the KIX domain, which is the same domain that HTLV-1 Tax also binds to42. Data presented in this thesis demonstrates that p30II and Tax compete to modulate the HTLV-1 LTR 198 mediated transcription. In addition, our data suggests that p30II is acetylated and that the

deacetylation by Histone Deacetylase-1 (HDAC-1) enhances p30II-mediated HTLV-1

LTR repression while inhibition of deacetylation by trichostatin A decreases p30II-

mediated HTLV-1 LTR repression. Herein, we also demonstrate that the HAT domain of

p300 is crucial in modulating HTLV-1 gene expression from the LTR by p30II. It will be

important to identify if p30II interacts with proteins other than CBP/p300, present at the

HTLV-1 promoter, and thus form complexes at the HTLV-1 promoter, to modulate the

LTR mediated transcription.

5.8 Identify the functional domains and specific amino acids within p30II, which are

critical in nuclear retention of tax/rex mRNA and protein-protein interactions

Recently, Nicot et al 31 reported that p30II modulates LTR mediated transcription,

in the context of the entire provirus, by a post-transcriptional mechanism, involving the

nuclear retention of tax/rex mRNA. Identifying the functional domains and specific

amino acids within p30II, which are critical in its ability to retain tax/rex mRNA in the nucleus, will be of use in elucidating the mechanism of function of p30II further. On-

going studies in our laboratory are aimed at identifying various proteins that interact with

p30II. To elucidate the functional role of p30II in HTLV-1 pathogenesis, it will be useful

to identify the functional domains and specific amino acids within p30II, which are

critical in such protein-protein interactions. Various serial deletion and site directed

mutants of p30II, described in this thesis, can be valuable tools in such investigations.

199 5.9 Temporal expression pattern and interaction between HTLV-1 regulatory and

accessory proteins during various stages of the HTLV-1 infection

Based on recent reports from our own laboratory and data presented in this thesis,

molecular functions for pX ORF I-encoded p12I and ORF II encoded p30II appears to be

associated with HTLV-1-induced T cell activation, particularly in NFAT, NF-κB and

AP-1 pathways2,8,9,23. While the mechanism of p30II function awaits additional research,

p12I function has been established to be calcium-dependent8, independent of Tax and

possibly occurs before Tax is expressed during a natural infection. However, highly

activated T cells mediated by expression of p12I and/or p30II likely allow HTLV-1

provirus to integrate into host cell genome and permit the early viral infection. Since Tax

is also able to cause T cell activation, it appears to be redundant for HTLV-1 to use

multiple proteins to activate T lymphocytes. It is possible that these proteins act

coordinately or synergistically. By modulating the expression of various HTLV-1

proteins, the virus might employ selective use of different viral proteins during different

stages of the infection. Particularly, it will be important to test if p12I and p30II act synergistically or compete with each other, in modulating NFAT mediated transcription.

Similar studies to elucidate the interaction of p30II and Tax in modulating NF-κB

mediated transcription may also be valuable in further understanding HTLV-1

pathogenesis.

In addition, previous results from our laboratory demonstrated that HTLV-1 Tax

and p30II both bind cellular transcriptional co-activator CBP/p300 at the KIX domain42.

Recent studies from our laboratory indicated that p30II activate the HTLV-1 LTR 200 mediated transcription at lower concentrations, and repress the LTR mediated

transcription at higher concentrations43. Based on the data presented herein, HTLV-1 Tax

and p30II appears to compete with each other in modulating the HTLV-1 LTR mediated transcription. When its levels are low, p30II also enhances transcription from the viral

promoter. Under such circumstances, p30II and Tax might act synergistically. However, at higher concentrations, p30II antagonizes the Tax transactivation of the HTLV-1

promoter. In addition, the data presented in this thesis indicate that p30II modulates the

LTR mediated transcription in a p300 dependent manner. The amount of p300 is limited in cells and such an environment of coactivator competition between transcription factors

is known to provide an additional layer of regulated gene expression6. Interestingly, recent unpublished findings from our laboratory indicate that HTLV-1 p12I increases the

levels of p300. Therefore, it appears that the relative amounts of Tax, p30II and p12I may be critical in modulating LTR mediated transcription of viral genes from the HTLV-1 viral promoter. It is possible that differences in expression levels of various viral proteins leading to differences in transcriptional regulation, might be the mechanism by which

HTLV-1 infection/ disease progresses though various stages. This is likely to be in synergy with differential regulation of cellular gene regulation by Tax, p30II and p12I, and thus changing the immune responses in accordance with different stages of progression of HTLV-1 infection and disease. However, since information on the expression profile of HTLV-1 proteins during different stages of the infection is limited, additional studies designed to test the temporal expression patterns of HTLV-1 regulatory and accessory proteins during viral infection are required to explore this possibility.

201 REFERENCES

1. Albrecht, B., N. D. Collins, M. T. Burniston, J. W. Nisbet, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J. Virol. 74:9828-9835.

2. Albrecht, B., C. D. D'Souza, W. Ding, S. Tridandapani, K. M. Coggeshall, and M. D. Lairmore. 2002. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J. Virol. 76:3493-3501.

3. Arnold, R., J. Liou, H. C. Drexler, A. Weiss, and F. Kiefer. 2001. Caspase- mediated cleavage of hematopoietic progenitor kinase 1 (HPK1) converts an activator of NFkappaB into an inhibitor of NFkappaB. J Biol. Chem. 276:14675- 14684.

4. Bartoe, J. T., B. Albrecht, N. D. Collins, M. D. Robek, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J. Virol. 74:1094-1100.

5. Bour-Jordan, H. and J. A. Blueston. 2002. CD28 function: a balance of costimulatory and regulatory signals. J Clin. Immunol. 22:1-7.

6. Colgin, M. A. and J. K. Nyborg. 1998. The human T-cell leukemia virus type 1 oncoprotein Tax inhibits the transcriptional activity of c-Myb through competition for the CREB binding protein. J. Virol. 72:9396-9399.

7. Derse, D., S. A. Hill, P. A. Lloyd, H. Chung, and B. A. Morse. 2001. Examining human t-lymphotropic virus type 1 infection and replication by cell-free infection with recombinant virus vectors. J. Virol. 75:8461-8468.

8. Ding, W., Albrecht B., R. E. Kelley, N. Muthusamy, S. Kim, R. A. Altschuld, and M. D. Lairmore. 2002. Human T lymphotropic virus type 1 p12(I) expression increases cytoplasmic calcium to enhance the activation of nuclear factor of activated T cells. J Virol 76:10374-10382.

9. Ding, W., S. J. Kim, A. M. Nair, B. Michael, K. Boris-Lawrie, A. Tripp, G. Feuer, and M. D. Lairmore. 2003. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol 77:11027-11039.

10. Fan, N., J. Gavalchin, B. Paul, K. H. Wells, M. J. Lane, and B. J. Poiesz. 1992. Infection of Peripheral Blood Mononuclear Cells and Cell Lines by Cell-Free Human T-Cell Lymphoma/Leukemia Virus Type 1. J. Clin. Microbiol. 30:905-910. 202 11. Fujii, M., K. Iwai, M. Oie, M. Fukushi, N. Yamamoto, M. Kannagi, and N. Mori. 2000. Activation of oncogenic transcription factor AP-1 in T cells infected with human T cell leukemia virus type 1. AIDS Res. Hum. Retroviruses 16:1603- 1606.

12. Gatza, M. L., J. C. Watt, and S. J. Marriott. 2003. Cellular transformation by the HTLV-I Tax protein, a jack-of-all-trades. Oncogene 22:5141-5149.

13. Goh, W. C., N. Manel, and M. Emerman. 2004. The human immunodeficiency virus Vpr protein binds Cdc25C: implications for G2 arrest. Virology 318:337-349.

14. Grassmann, R., S. Berchtold, I. Radant, M. Alt, B. Fleckenstein, J. G. Sodroski, W. A. Haseltine, and U. Ramstedt. 1992. Role of Human T-Cell Leukemia Virus Type 1 X Region Proteins in Immortalization of Primary Human Lymphocytes in Culture. J. Virol. 66:4570-4575.

15. Harrod, R., Y. L. Kuo, Y. Tang, Y. Yao, A. Vassilev, Y. Nakatani, and C. Z. Giam. 2000. p300 and p300/cAMP-responsive element-binding protein associated factor interact with human T-cell lymphotropic virus type-1 Tax in a multi- histone acetyltransferase/activator-enhancer complex. J. Biol. Chem. 275:11852-11857.

16. Hollsberg, P. 1999. Mechanisms of T-cell activation by human T-cell lymphotropic virus type I. Microbiol. Mol. Biol. Rev. 63:308-333.

17. Hollsberg, P., L. J. Ausubel, and D. A. Hafler. 1994. Human T Cell Lymphotropic Virus Type 1-Induced T Cell Activation. J. Immunol. 153:566-573.

18. Hurd, C. and E. Rozengurt. 2003. Uncoupling of protein kinase D from suppression of EGF-dependent c-Jun phosphorylation in cancer cells. Biochem. Biophys. Res. Commun. 302:800-804.

19. Igakura, T., J. C. Stinchcombe, P. K. Goon, G. P. Taylor, J. N. Weber, G. M. Griffiths, Y. Tanaka, M. Osame, and C. R. Bangham. 2003. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science 299:1713-1716.

20. Jeang, K. T. 2001. Functional activities of the human T-cell leukemia virus type I Tax oncoprotein: cellular signaling through NF-kappa B. Cytokine Growth Factor Rev. 12:207-217.

21. Jiang, H., H. Lu, R. L. Schiltz, C. A. Pise-Masison, V. V. Ogryzko, Y. Nakatani, and J. N. Brady. 1999. PCAF interacts with tax and stimulates tax transactivation in a histone acetyltransferase-independent manner [published erratum appears in Mol Cell Biol 2000 Mar;20(5):1897]. Mol. Cell Biol. 19:8136- 8145.

203 22. Kashanchi, F., J. F. Duvall, R. P. Kwok, J. R. Lundblad, R. H. Goodman, and J. N. Brady. 1998. The coactivator CBP stimulates human T-cell lymphotrophic virus type I Tax transactivation in vitro. J. Biol. Chem. 273:34646-34652.

23. Kim, S. J., W. Ding, B. Albrecht, P. L. Green, and M. D. Lairmore. 2003. A conserved calcineurin-binding motif in human T lymphotropic virus type 1 p12I functions to modulate nuclear factor of activated T cell activation. J Biol. Chem. 278:15550-15557.

24. Konikova, E. and J. Kusenda. 2003. Altered expression of p53 and MDM2 proteins in hematological malignancies. Neoplasma 50:31-40.

25. Lahti, A. L., A. Manninen, and K. Saksela. 2003. Regulation of T cell activation by HIV-1 accessory proteins: Vpr acts via distinct mechanisms to cooperate with Nef in NFAT-directed gene expression and to promote transactivation by CREB. Virology 310:190-196.

26. Lemasson, I., N. J. Polakowski, P. J. Laybourn, and J. K. Nyborg. 2002. Transcription factor binding and histone modifications on the integrated proviral promoter in human T-cell leukemia virus-I-infected T-cells. J Biol. Chem. 277:49459-49465.

27. McCoy, M. A., J. J. Gesell, M. M. Senior, and D. F. Wyss. 2003. Flexible lid to the p53-binding domain of human Mdm2: implications for p53 regulation. Proc. Natl. Acad. Sci U. S. A 100:1645-1648.

28. Mori, N., A. Ueda, S. Ikeda, Y. Yamasaki, Y. Yamada, M. Tomonaga, S. Morikawa, R. Geleziunas, T. Yoshimura, and N. Yamamoto. 2000. Human T- cell leukemia virus type I tax activates transcription of the human monocyte chemoattractant protein-1 gene through two nuclear factor-kappaB sites. Cancer Res. 60:4939-4945.

29. Nemorin, J. G. and P. Duplay. 2000. Evidence that Llck-mediated phosphorylation of p56dok and p62dok may play a role in CD2 signaling. J Biol. Chem. 275:14590-14597.

30. Neuveut, C. and K. T. Jeang. 2002. Cell cycle dysregulation by HTLV-I: role of the tax oncoprotein. Front Biosci. 7:157-163.

31. Nicot, C., M. Dundr, J. M. Johnson, J. R. Fullen, N. Alonzo, R. Fukumoto, G. L. Princler, D. Derse, T. Misteli, and G. Franchini. 2004. HTLV-1-encoded p30(II) is a post-transcriptional negative regulator of viral replication. Nat. Med.

32. Pise-Masison, C. A., M. Radonovich, R. Mahieux, P. Chatterjee, C. Whiteford, J. Duvall, C. Guillerm, A. Gessain, and J. N. Brady. 2002. Transcription profile of cells infected with human T-cell leukemia virus type I compared with activated lymphocytes. Cancer Res. 62:3562-3571. 204 33. Pusl, T., J. J. Wu, T. L. Zimmerman, L. Zhang, B. E. Ehrlich, M. W. Berchtold, J. B. Hoek, S. J. Karpen, M. H. Nathanson, and A. M. Bennett. 2002. Epidermal growth factor-mediated activation of the ETS domain transcription factor Elk-1 requires nuclear calcium. J Biol. Chem. 277:27517-27527.

34. Silverman, L. R., A. J. Phipps, A. Montgomery, L. Ratner, and M. D. Lairmore. 2004. Human T-cell lymphotropic virus type 1 open reading frame II- encoded p30II is required for in vivo replication: evidence of in vivo reversion. J Virol 78:3837-3845.

35. Sun, S. C., E. W. Harhaj, G. Xiao, and L. Good. 2000. Activation of I-kappaB kinase by the HTLV type 1 tax protein: mechanistic insights into the adaptor function of IKKgamma [In Process Citation]. AIDS Res. Hum. Retroviruses 16:1591-1596.

36. Valentin, H., I. Lemasson, S. Hamaia, H. Casse, S. Konig, C. Devaux, and L. Gazzolo. 1997. Transcriptional activation of the vascular cell adhesion molecule-1 gene in T lymphocytes expressing human T-Cell leukemia virus type 1 Tax protein. J. Virol. 71:8522-8530.

37. Van Lint, J., A. Rykx, Y. Maeda, T. Vantus, S. Sturany, V. Malhotra, J. R. Vandenheede, and T. Seufferlein. 2002. Protein kinase D: an intracellular traffic regulator on the move. Trends Cell Biol. 12:193-200.

38. Yoshida, H., H. Nishina, H. Takimoto, L. E. Marengere, A. C. Wakeham, D. Bouchard, Y. Y. Kong, T. Ohteki, A. Shahinian, M. Bachmann, P. S. Ohashi, J. M. Penninger, G. R. Crabtree, and T. W. Mak. 1998. The transcription factor NF-ATc1 regulates lymphocyte proliferation and Th2 cytokine production. Immunity 8:115-124.

39. Yssel, H., M. R. De Waal, D. Duc, D. Blanchard, L. Gazzolo, J. E. de Vries, and H. Spits. 1989. Human T cell leukemia/lymphoma virus type I infection of a CD4+ proliferative/cytotoxic T cell clone progresses in at least two distinct phases based on changes in function and phenotype of the infected cells. J Immunol. 142:2279- 2289.

40. Yuan, H., Y. M. Xie, and I. S. Chen. 2003. Depletion of Wee-1 kinase is necessary for both human immunodeficiency virus type 1 Vpr- and gamma irradiation-induced apoptosis. J Virol 77:2063-2070.

41. Zhang, B. X., X. Ma, C. K. Yeh, M. D. Lifschitz, M. X. Zhu, and M. S. Katz. 2002. Epidermal growth factor-induced depletion of the intracellular Ca2+ store fails to activate capacitative Ca2+ entry in a human salivary cell line. J Biol. Chem. 277:48165-48171.

205 42. Zhang, W., J. W. Nisbet, B. Albrecht, W. Ding, F. Kashanchi, J. T. Bartoe, and M. D. Lairmore. 2001. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J. Virol. 75:9885-9895.

43. Zhang, W., J. W. Nisbet, J. T. Bartoe, W. Ding, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J. Virol. 74:11270-11277.

206

BIBLIOGRAPHY

Abeele, F. V., R. Skryma, Y. Shuba, F. Van Coppenolle, C. Slomianny, M. Roudbaraki, B. Mauroy, F. Wuytack, and N. Prevarskaya. 2002. Bcl-2-dependent modulation of Ca(2+) homeostasis and store-operated channels in prostate cancer cells. Cancer Cell 1:169-179.

Adedayo, O., G. Grell, and P. Bellot. 2003. Hospital admissions for human T-cell lymphotropic virus type-1 (HTLV-1) associated diseases in Dominica. Postgrad. Med. J 79:341-344.

Ahmed, Y. F., G. M. Gilmartin, S. M. Hanly, J. R. Nevins, and W. C. Greene. 1991. The HTLV-1 Rex Response Element Mediates a Novel Form of mRNA Polyadenylation. Cell 64:727-737.

Ahmed, Y. F., S. M. Hanly, M. H. Malim, B. R. Cullen, and W. C. Greene. 1990. Structure-function analyses of the HTLV-I Rex and HIV-1 Rev RNA response elements: insights into the mechanism of Rex and Rev action. Genes Dev. 4:1014-1022.

Akagi, T., H. Ono, N. Tsuchida, and K. Shimotohno. 1997. Aberrant expression and function of p53 in T-cells immortalized by HTLV-I Tax1. FEBS Lett. 406:263-266.

Alamy, A. H., F. B. Menezes, A. C. Leite, O. M. Nascimento, and A. Q. Araujo. 2001. Dysautonomia in human T-cell lymphotrophic virus type I-associated myelopathy/tropical spastic paraparesis. Ann. Neurol. 50:681-685.

Albrecht, B., N. D. Collins, M. T. Burniston, J. W. Nisbet, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J. Virol. 74:9828- 9835.

Albrecht, B., C. D. D'Souza, W. Ding, S. Tridandapani, K. M. Coggeshall, and M. D. Lairmore. 2002. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J. Virol. 76:3493-3501.

Albrecht, B. and M. D. Lairmore. 2002. Critical role of human T-lymphotropic virus type 1 accessory proteins in viral replication and pathogenesis. Microbiol. Mol. Biol. Rev. 66:396-406.

207 Alexandre, C. and B. Verrier. 1991. Four regulatory elements in the human c-fos promoter mediate transactivation by HTLV-1 Tax protein 5288. Oncogene 6:543-551.

Anand, N., S. Murthy, G. Amann, M. Wernick, L. A. Porter, I. H. Cukier, C. Collins, J. W. Gray, J. Diebold, D. J. Demetrick, and J. M. Lee. 2002. Protein elongation factor EEF1A2 is a putative oncogene in ovarian cancer. Nat. Genet. 31:301- 305.

Aramburu, J., M. B. Yaffe, C. Lopez-Rodriguez, L. C. Cantley, P. G. Hogan, and A. Rao. 1999. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A. Science 285:2129-2133.

Arany, Z., D. Newsome, E. Oldread, D. M. Livingston, and R. Eckner. 1995. A family of transcriptional adaptor proteins targeted by the E1A oncoprotein. Nature 374:81-84.

Ariumi, Y., A. Kaida, J. Y. Lin, M. Hirota, O. Masui, S. Yamaoka, Y. Taya, and K. Shimotohno. 2000. HTLV-1 tax oncoprotein represses the p53-mediated trans-activation function through coactivator CBP sequestration. Oncogene 19:1491-1499.

Arnold, R., J. Liou, H. C. Drexler, A. Weiss, and F. Kiefer. 2001. Caspase-mediated cleavage of hematopoietic progenitor kinase 1 (HPK1) converts an activator of NFkappaB into an inhibitor of NFkappaB. J Biol. Chem. 276:14675-14684.

Ashburner, M., C. A. Ball, J. A. Blake, D. Botstein, H. Butler, J. M. Cherry, A. P. Davis, K. Dolinski, S. S. Dwight, J. T. Eppig, M. A. Harris, D. P. Hill, L. Issel- Tarver, A. Kasarskis, S. Lewis, J. C. Matese, J. E. Richardson, M. Ringwald, G. M. Rubin, and G. Sherlock. 2000. Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nat. Genet. 25:25-29.

Asher, D. M., J. Goudsmit, K. Pomeroy, R. M. Garruto, M. Bakker, S. Ono, N. Elliot, K. Harris, H. Askins, Z. Eldadah, A. L. Goldstein, and D. C. Gajdusek. 1988. Antibodies to HTLV-I in populations of the southwestern Pacific. J. Med. Virol. 26:339.

Asquith, B., E. Hanon, G. P. Taylor, and C. R. Bangham. 2000. Is human T-cell lymphotropic virus type I really silent? Philos. Trans. R. Soc Lond B Biol. Sci 355:1013- 1019.

Avantaggiati, M. L., M. Carbone, A. Graessmann, Y. Nakatani, B. Howard, and A. S. Levine. 1996. The SV40 large T antigen and adenovirus E1a oncoproteins interact with distinct isoforms of the transcriptional co-activator, p300. EMBO J. 15:2236-2248.

Avantaggiati, M. L., V. Ogryzko, K. Gardner, A. Giordano, A. S. Levine, and K. Kelly. 1997. Recruitment of p300/CBP in p53-dependent signal pathways. Cell 89:1175- 1184.

208 Avvakumov, N., J. Torchia, and J. S. Mymryk. 2003. Interaction of the HPV E7 proteins with the pCAF acetyltransferase. Oncogene 22:3833-3841.

Azimi, N., J. Mariner, S. Jacobson, and T. A. Waldmann. 2000. How does interleukin 15 contribute to the pathogenesis of HTLV type 1- associated Myelopathy/Tropical spastic paraparesis? AIDS Res. Hum. Retroviruses 16:1717-1722.

Ballard, D. W. 2001. Molecular mechanisms in lymphocyte activation and growth. Immunol. Res. 23:157-166.

Ballaun, C., G. K. Farrington, M. Dobrovnik, J. Rusche, J. Hauber, and E. Bohnlein. 1991. Functional Analysis of Human T-Cell Leukemia Virus Type 1 rex- Response Element: Direct RNA Binding of Rex Protein Correlates with In Vivo Activity. J. Virol. 65:4408-4413.

Bangham, C. R. 2000. The immune response to HTLV-I. Curr. Opin. Immunol. 12:397- 402.

Bangham, C. R. 2003. Human T-lymphotropic virus type 1 (HTLV-1): persistence and immune control. Int. J Hematol. 78:297-303.

Bangham, C. R. 2003. The immune control and cell-to-cell spread of human T- lymphotropic virus type 1. J Gen. Virol 84:3177-3189.

Bannister, A. J. and T. Kouzarides. 1995. CBP-induced stimulation of c-Fos activity is abrogated by E1A. EMBO J. 14:4758-4762.

Bannister, A. J. and T. Kouzarides. 1996. The CBP co-activator is a histone acetyltransferase. Nature 384:641-643.

Bannister, A. J. and E. A. Miska. 2000. Regulation of gene expression by transcription factor acetylation. Cell Mol. Life Sci. 57:1184-1192.

Barmak, K., E. Harhaj, C. Grant, T. Alefantis, and B. Wigdahl. 2003. Human T cell leukemia virus type I-induced disease: pathways to cancer and neurodegeneration. Virology 308:1-12.

Barshira, A., A. Panet, and A. Honigman. 1991. An RNA Secondary Structure Juxtaposes Two Remote Genetic Signals for Human T-Cell Leukemia Virus Type 1 RNA 3'-End Processing. J. Virol. 65:5165-5173.

Bartoe, J. T., B. Albrecht, N. D. Collins, M. D. Robek, L. Ratner, P. L. Green, and M. D. Lairmore. 2000. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J. Virol. 74:1094-1100.

Berneman, Z. N., R. B. Gartenhaus, M. S. Reitz, W. A. Blattner, A. Manns, B. Hanchard, O. Ikehara, R. C. Gallo, and M. E. Klotman. 1992. Expression of

209 alternatively spliced human T-lymphotropic virus type 1 pX mRNA in infected cell lines and in primary uncultured cells from patients with adult T-cell leukemia/lymphoma and healthy carriers. Proc. Natl. Acad. Sci. USA 89:3005-3009.

Bex, F. and R. B. Gaynor. 1998. Regulation of gene expression by HTLV-I Tax protein. Methods 16:83-94.

Bex, F., M. J. Yin, A. Burny, and R. B. Gaynor. 1998. Differential transcriptional activation by human T-cell leukemia virus type 1 Tax mutants is mediated by distinct interactions with CREB binding protein and p300. Mol. Cell Biol. 18:2392-2405.

Black, A. C., I. S. Y. Chen, S. Arrigo, C. T. Ruland, T. Allogiamento, E. Chin, and J. D. Rosenblatt. 1991. Regulation of HTLV-II gene expretion by rex involves positive and negative cis-Acting elements in the 5 long terminal repeat. Virology 181:433-444.

Black, A. C., C. Ruland, M. T. Yip, J. Luo, B. Tran, A. Kalssi, E. Quan, M. Aboud, I. S. Y. Chen, and J. D. Rosenblatt. 1991. Human T-Cell Leukemia Virus Type II Rex Binding and Activity Require an Intact splice Donor Site and a Specific RNA Secondary Structure. J. Virol. 65:6645-6653.

Black, A. C., C. T. Ruland, J. Luo, A. Bakker, J. K. Fraser, and J. D. Rosenblatt. 1994. Binding of Nuclear Proteins to HTLV-II Cis-Acting Repressive Sequence (CRS) RNA Correlates with CRS Function. Virology 200:29-41.

Blanco, J. C., S. Minucci, J. Lu, X. J. Yang, K. K. Walker, H. Chen, R. M. Evans, Y. Nakatani, and K. Ozato. 1998. The histone acetylase PCAF is a nuclear receptor coactivator. Genes Dev. 12:1638-1651.

Blattner, W., V. Kalyanaraman, M. Robert-Guroff, T. Lister, D. Galton, P. S. Sarin, M. Crawford, D. Catovsky, M. Greaves, and R. C. Gallo. 1982. The human type-C retrovirus, HTLV, in blacks from the Caribbean region and relationchip to adult T-cell leukemia/lymphoma. Int. J. Cancer 30:257.

Blobel, G. A. 2000. CREB-binding protein and p300: molecular integrators of hematopoietic transcription. Blood 95:745-755.

Blobel, G. A. 2002. CBP and p300: versatile coregulators with important roles in hematopoietic gene expression. J Leukoc. Biol. 71:545-556.

Bogerd, H. P., A. Echarri, T. M. Ross, and B. R. Cullen. 1998. Inhibition of human immunodeficiency virus Rev and human T- cell leukemia virus Rex function, but not Mason-Pfizer monkey virus constitutive transport element activity, by a mutant human nucleoporin targeted to Crm1. J Virol 72:8627-8635.

Bosselut, R., J. F. Duvall, A. Gegonne, M. Bailly, A. Hemar, J. Brady, and J. Ghysdael. 1990. The product of the c-ets-1 proto-oncogene and the related Ets2 protein

210 act as transcriptional activators of the long terminal repeat of human T cell leukemia virus HTLV-1 5254. EMBO J. 9:3137-3144.

Bosselut, R., F. Lim, P. C. Romond, J. Frampton, J. Brady, and J. Ghysdael. 1992. Myb protein binds to multiple sites in the human T cell lymphotropic virus type 1 long terminal repeat and transactivates LTR-mediated expression 5258. Virology 186:764-769.

Bour-Jordan, H. and J. A. Blueston. 2002. CD28 function: a balance of costimulatory and regulatory signals. J Clin. Immunol. 22:1-7.

Brady, J. N., K. T. Jeang, J. Duvall, and G. Khoury. 1987. Identification of p40x- Responsive Regulatory Sequences within the Human T-Cell Leukemia Virus Type 1 Long Terminal Repeat. J. Virol. 61:2175-2181.

Brauweiler, A., J. E. Garrus, J. C. Reed, and J. K. Nyborg. 1997. Repression of Bax gene expression by the HTLV-I tax protein: Implications for suppression of apoptosis in virally infected cells. Virology 231:135-140.

Campbell, K. M. and K. J. Lumb. 2002. Structurally distinct modes of recognition of the KIX domain of CBP by Jun and CREB. Biochemistry 41:13956-13964.

Cann, A. J., J. D. Rosenblatt, W. Wachsman, N. P. Shah, and I. S. Chen. 1985. Identification of the gene responsible for human T-cell leukaemia virus transcriptional regulation. Nature 318:571-574.

Cantrell, D. 1996. T cell antigen receptor signal transduction pathways. Annu. Rev. Immunol. 14:259-274.

Castillo, L. C., F. Gracia, G. C. Roman, P. Levine, W. C. Reeves, and J. Kaplan. 2000. Spinocerebellar syndrome in patients infected with human T-lymphotropic virus types I and II (HTLV-I/HTLV-II): report of 3 cases from Panama. Acta Neurol. Scand. 101:405-412.

Cavrois, M., I. Leclercq, O. Gout, A. Gessain, S. Wainhobson, and E. Wattel. 1998. Persistent oligoclonal expansion of human T-cell leukemia virus type 1 infected circulating cells in patients with Tropical spastic paraparesis/HTLV-1 associated myelopathy. Oncogene 17:77-82.

Cereseto, A., Z. Berneman, I. Koralnik, J. Vaughn, G. Franchini, and M. E. Klotman. 1997. Differential expression of alternatively spliced pX mRNAs in HTLV-I- infected cell lines. Leukemia 11:866-870.

Cereseto, A., F. Diella, J. C. Mulloy, A. Cara, P. Michieli, R. Grassmann, G. Franchini, and M. E. Klotman. 1996. P53 functional impairment and high

211 p21(waf1/cip1) expression in human T-cell lymphotropic/leukemia virus type I- transformed T cells. Blood 88:1551-1560.

Chen, Y. M., S. H. Chen, C. Y. Fu, J. Y. Chen, and M. Osame. 1997. Antibody reactivities to tumor-suppressor protein p53 and HTLV-I Tof, Rex and Tax in HTLV-I- infected people with differing clinical status. International. Journal. of. Cancer 71:196- 202.

Chen, Y. M. A., S. H. Chen, C. Y. Fu, J. Y. Chen, and M. Osame. 1997. Antibody reactivities to tumor-suppressor protein p53 and HTLV-I Tof Rex and Tax in HTLV-I- infected people with differing clinical status. Int. J. Cancer 71:196-202.

Chou, K. S., A. Okayama, N. Tachibana, T. H. Lee, and M. Essex. 1995. Nucleotide sequence analysis of a full-length human T-cell leukemia virus type I from adult T-cell leukemia cells: A prematurely terminated pX open reading frame II. Int. J. Cancer 60:701-706.

Chrivia, J. C., R. P. Kwok, N. Lamb, M. Hagiwara, M. R. Montminy, and R. H. Goodman. 1993. Phosphorylated CREB binds specifically to the nuclear protein CBP. Nature 365(6449):855-859.

Ciminale, V., D. D'Agostino, L. Zotti, G. Franchini, B. K. Felber, and L. Chieco- Bianchi. 1995. Expression and characterization of proteins produced by mRNAs spliced into the X region of the human T-cell leukemia/lymphotropic virus type II. Virology 209:445-456.

Ciminale, V., G. N. Pavlakis, D. Derse, C. P. Cunningham, and B. K. Felber. 1992. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J. Virol. 66:1737-1745.

Ciminale, V., L. Zotti, D. M. Dagostino, T. Ferro, L. Casareto, G. Franchini, P. Bernardi, and L. Chiecobianchi. 1999. Mitochondrial targeting of the p13(II) protein coded by the x-II ORF of human T-cell leukemia/lymphotropic virus type I (HTLV-I). Oncogene 18:4505-4514.

Clark, N. M., M. J. Smith, J. M. Hilfinger, and D. M. Markovitz. 1993. Activation of the human T-cell leukemia virus type I enhancer is mediated by binding sites for Elf-1 and the pets factor. J. Virol. 67:5522-5528.

Cleghorn, F. R., A. Manns, R. Falk, P. Hartge, B. Hanchard, N. Jack, E. Williams, E. Jaffe, F. White, C. Bartholomew, and W. Blattner. 1995. Effect of human T- lymphotropic virus type I infection on non- Hodgkin's lymphoma incidence. J. Nat. Cancer Inst. 87:1009-1014.

Coffin, J. M. 1992. Structure and Classification of Retroviruses. The Retroviridae. 1:19- 49.

212 Coffin, J. M. Retroviridae: The Viruses and Their Replication. 1767-1835. 1996. Phildaelphia, PA, Lippinocott-Raven Publishers.

Colgin, M. A. and J. K. Nyborg. 1998. The human T-cell leukemia virus type 1 oncoprotein Tax inhibits the transcriptional activity of c-Myb through competition for the CREB binding protein. J. Virol. 72:9396-9399.

Collins, N. D., C. D'Souza, B. Albrecht, M. D. Robek, L. Ratner, W. Ding, P. L. Green, and M. D. Lairmore. 1999. Proliferation response to interleukin-2 and Jak/Stat activation of T cells immortalized by human T-cell lymphotropic virus type 1 is independent of open reading frame I expression. J. Virol. 73:9642-9649.

Collins, N. D., G. C. Newbound, B. Albrecht, J. L. Beard, L. Ratner, and M. D. Lairmore. 1998. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood 91:4701-4707.

Coscoy, L. and D. Ganem. 2003. PHD domains and E3 ubiquitin ligases: viruses make the connection. Trends Cell Biol. 13:7-12.

Cotter, M. A. and E. S. Robertson. 2000. Modulation of histone acetyltransferase activity through interaction of epstein-barr nuclear antigen 3C with prothymosin alpha. Mol. Cell Biol. 20:5722-5735.

Courouce, A. M., J. Pillonel, J. M. Lemaire, and C. Saura. 1998. HTLV testing in blood transfusion. Vox Sang. 74 Suppl 2:165-169.

Cowan, E. P., R. K. Alexander, S. Daniel, F. Kashanchi, and J. N. Brady. 1997. Induction of tumor necrosis factor alpha in human neuronal cells by extracellular human T-cell lymphotropic virus type 1 Tax(1). J. Virol. 71:6982-6989.

Crabtree, G. R. 2001. Calcium, calcineurin, and the control of transcription. J. Biol. Chem. 276:2313-2316.

Craig, H. M., T. R. Reddy, N. L. Riggs, P. P. Dao, and J. C. Guatelli. 2000. Interactions of HIV-1 nef with the mu subunits of adaptor protein complexes 1, 2, and 3: role of the dileucine-based sorting motif. Virology 271:9-17.

Cullen, B. R. 1998. Retroviruses as model systems for the study of nuclear RNA export pathways. Virology 249:203-210.

D'Agostino, D. M., V. Ciminale, L. Zotti, and L. Chieco-Bianchi. 1999. Influence of Rex and intronic sequences on expression of spliced mRNAs produced by human T cell leukemia virus type I. AIDS Res. Hum. Retroviruses 15:1351-1363.

213 D'Agostino, D. M., L. Ranzato, G. Arrigoni, I. Cavallari, F. Belleudi, M. R. Torrisi, M. Silic-Benussi, T. Ferro, V. Petronilli, O. Marin, L. Chieco-Bianchi, P. Bernardi, and V. Ciminale. 2002. Mitochondrial alterations induced by the p13II protein of human T-cell leukemia virus type 1. Critical role of arginine residues. J Biol. Chem. 277:34424- 34433.

D'Agostino, D. M., L. Zotti, T. Ferro, I. Cavallori, M. Silic-Benussi, L. Chieco- Bianchi, and V. Ciminale. 2001. Expression and functional properties of proteins encoded in the x-II ORF of HTLV-I. Virus Res. 78:35-43.

D'Agostino, D. M., L. Zotti, T. Ferro, G. Franchini, L. Chieco-Bianchi, and V. Ciminale. 2000. The p13II protein of HTLV type 1: comparison with mitochondrial proteins coded by other human viruses. AIDS Res. Hum. Retroviruses 16:1765-1770.

Daenke, S., S. Nightingale, J. K. Cruickshank, and C. R. M. Bangham. 1990. Sequence variants of human T-cell lymphotropic virus type I from patients with tropical spastic paraparesis and adult T-cell leukemia do not distinguish neurological from leukemic isolates. J. Virol. 64:1278-1282.

Dagostino, D. M., V. Ciminale, L. Zotti, A. Rosato, and L. Chiecobianchi. 1997. The human T-cell lymphotropic virus type 1 Tof protein contains a bipartite nuclear localization signal that is able to functionally replace the amino-terminal domain of Rex. J. Virol. 71:75-83.

Daly, T. J., R. C. Doten, P. Rennert, M. Auer, H. Jaksche, A. Donner, G. Fisk, and J. R. Rusche. 1993. Biochemical characterization of binding of multiple HIV-1 Rev monomeric proteins to the Rev responsive element. Biochemistry 32:10497-10505.

Dantuma, N. P. and M. G. Masucci. 2003. The ubiquitin/proteasome system in Epstein- Barr virus latency and associated malignancies. Semin. Cancer Biol. 13:69-76.

Datta, S., N. H. Kothari, and H. Fan. 2000. In vivo genomic footprinting of the human T-cell leukemia virus type 1 (HTLV-1) long terminal repeat enhancer sequences in HTLV-1-infected human T-cell lines with different levels of Tax I activity. J. Virol. 74:8277-8285. de La, F. C., L. Deng, F. Santiago, L. Arce, L. Wang, and F. Kashanchi. 2000. Gene expression array of HTLV type 1-infected T cells: Up-regulation of transcription factors and cell cycle genes. AIDS Res. Hum. Retroviruses 16:1695-1700.

Dekaban, G. A., A. A. Peters, J. C. Mulloy, J. M. Johnson, R. Trovato, E. Rivadeneira, and G. Franchini. 2000. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology 274:86-93.

214 Delamarre, L., C. Pique, D. Pham, T. Tursz, and M. C. Dokhelar. 1994. Identification of functional regions in the human T-cell leukemia virus type I SU glycoprotein. J. Virol. 68:3544-3549.

Delamarre, L., C. Pique, A. R. Rosenberg, V. Blot, M. P. Grange, B. Le, I, and M. C. Dokhelar. 1999. The Y-S-L-I tyrosine-based motif in the cytoplasmic domain of the human T-cell leukemia virus type 1 envelope is essential for cell-to-cell transmission 4. J Virol 73:9659-9663.

Dequiedt, F., E. Hanon, P. Kerkhofs, P. P. Pastoret, D. Portetelle, A. Burny, R. Kettmann, and L. Willems. 1997. Both wild-type and strongly attenuated bovine leukemia viruses protect peripheral blood mononuclear cells from apoptosis. J Virol 71:630-639.

Derse, D. 1987. Bovine leukemia virus transcription is controlled by a virus-encoded trans-acting factor and by cis-acting response elements. J Virol 61:2462-2471.

Derse, D., S. A. Hill, P. A. Lloyd, H. Chung, and B. A. Morse. 2001. Examining human t-lymphotropic virus type 1 infection and replication by cell-free infection with recombinant virus vectors. J. Virol. 75:8461-8468.

Derse, D., J. Mikovits, and F. Ruscetti. 1997. X-I and X-II open reading frames of HTLV-I are not required for virus replication or for immortalization of primary T-cells in vitro. Virology 237:123-128.

Desgranges, C., S. Souche, J. C. Vernant, D. Smadja, A. Vahlne, and P. Horal. 1994. Identification of novel neutralization-inducing regions of the human T cell lymphotropic virus type I envelope glycoproteins with human HTLV-I- seropositive sera 5. AIDS Res. Hum. Retroviruses 10:163-173.

Ding, W., Albrecht B., R. E. Kelley, N. Muthusamy, S. Kim, R. A. Altschuld, and M. D. Lairmore. 2002. Human T lymphotropic virus type 1 p12(I) expression increases cytoplasmic calcium to enhance the activation of nuclear factor of activated T cells. J Virol 76:10374-10382.

Ding, W., B. Albrecht, R. Luo, W. Zhang, J. R. Stanley, G. C. Newbound, and M. D. Lairmore. 2001. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J. Virol. 75:7672-7682.

Ding, W., S. J. Kim, A. M. Nair, B. Michael, K. Boris-Lawrie, A. Tripp, G. Feuer, and M. D. Lairmore. 2003. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol 77:11027-11039.

Ding, W., S. J. Kim, A. M. Nair, B. Michael, K. Boris-Lawrie, A. Tripp, G. Feuer, and M. D. Lairmore. 2003. Human T-cell lymphotropic virus type 1 p12I enhances interleukin-2 production during T-cell activation. J Virol 77:11027-11039. 215 Duyao, M. P., D. J. Kessler, D. B. Spicer, C. Bartholomew, J. L. Cleveland, M. Siekevitz, and G. E. Sonenshein. 1992. Transactivation of the c-myc Promoter by Human T Cell Leukemia virus Type 1 tax is Mediated by NFkB. J. Biol. Chem. 267:16288-16291.

Edlich, R. F., J. A. Arnette, and F. M. Williams. 2000. Global epidemic of human T- cell lymphotropic virus type-I (HTLV-I). J. Emerg. Med. 18:109-119.

Ego, T., Y. Ariumi, and K. Shimotohno. 2002. The interaction of HTLV-1 Tax with HDAC1 negatively regulates the viral gene expression. Oncogene 21:7241-7246.

Evangelista, A., S. Maroushek, H. Minnigan, A. Larson, E. Retzel, A. Haase, D. Gonzalez-Dunia, D. McFarlin, E. Mingioli, S. Jacobson, and . 1990. Nucleotide sequence analysis of a provirus derived from an individual with tropical spastic paraparesis. Microb. Pathog. 8:259-278.

Fan, N., J. Gavalchin, B. Paul, K. H. Wells, M. J. Lane, and B. J. Poiesz. 1992. Infection of Peripheral Blood Mononuclear Cells and Cell Lines by Cell-Free Human T- Cell Lymphoma/Leukemia Virus Type 1. J. Clin. Microbiol. 30:905-910.

Felber, B. K., H. Paskalis, and C. Kleinmanewing. 1985. The pX Protein of HTLV-1 is a Transcriptional Activator of its Long Terminal Repeats. Science 229:675-679.

Ferreira, O. C., V. Planelles, and J. D. Rosenblatt. 1997. Human T-cell leukemia viruses: Epidemiology, biology, and pathogenesis. Blood Rev. 11:91-104.

Franchini, G. 1995. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood 86:3619-3639.

Franchini, G., J. C. Mulloy, I. J. Koralnik, M. A. Lo, J. J. Sparkowski, T. Andresson, D. J. Goldstein, and R. Schlegel. 1993. The human T-cell leukemia/lymphotropic virus type I p12I protein cooperates with the E5 oncoprotein of bovine papillomavirus in cell transformation and binds the 16-kilodalton subunit of the vacuolar H+ ATPase. J Virol 67:7701-7704.

Franchini, G., F. Wong-Staal, and R. C. Gallo. 1984. Human T-cell leukemia virus (HTLV-I) transcripts in fresh and cultured cells of patients with adult T-cell leukemia 5438. Proc. Natl. Acad. Sci. U. S. A 81:6207-6211.

Franchini, G., F. Wong-Staal, and R. C. Gallo. 1984. Molecular studies of human T- cell leukemia virus and adult T-cell leukemia. J Invest Dermatol. 83:63s-66s.

Fujii, M., K. Iwai, M. Oie, M. Fukushi, N. Yamamoto, M. Kannagi, and N. Mori. 2000. Activation of oncogenic transcription factor AP-1 in T cells infected with human T cell leukemia virus type 1. AIDS Res. Hum. Retroviruses 16:1603-1606.

216 Fujii, M., P. Sassone-Corsi, and I. Verma. 1988. c-fos promotor trans-activation of cAMP on IL-2 stimulated gene expression. Proc. Natl. Acad. Sci. U. S. A.8526-8530.

Fujii, M., H. Tsuchiya, T. Chuhjo, T. Akizawa, and M. Seiki. 1992. Interaction of HTLV-1 Tax1 with p67SRF causes the aberrant induction of cellular immediate early genes through CArG boxes. Genes Dev. 6:2066-2076.

Fujii, M., H. Tsuchiya, X. B. Meng, and M. Seiki. 1995. c-Jun, c-Fos and their family members activate the transcription mediated by three 21-bp repetitive sequences in the HTLV-I long terminal repeat. Intervirology 38:221-228.

Fujisawa, J., M. Seiki, T. Kiyokawa, and M. Yoshida. 1985. Functional activation of the long terminal repeat of human T-cell leukemia virus type 1 by a trans-acting factor. Proc. Natl. Acad. Sci. USA 82:2277-2281.

Fukushima, T., T. Ikeda, E. Uyama, M. Uchino, H. Okabe, and M. Ando. 1994. Cognitive event-related potentials and brain magnetic resonance imaging in HTLV-1 associated myelopathy (HAM). J. Neurol. Sci. 126:30-39.

Furukawa, K., K. Furukawa, and H. Shiku. 1991. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 295:141- 145.

Furukawa, Y., R. Kubota, N. Eiraku, M. Nakagawa, K. Usuku, S. Izumo, and M. Osame. 2003. Human T-cell lymphotropic virus type I (HTLV-I)-related clinical and laboratory findings for HTLV-I-infected blood donors. J Acquir. Immune. Defic. Syndr. 32:328-334.

Gachon, F., C. Devaux, and J. M. Mesnard. 2002. Activation of HTLV-I transcription in the presence of Tax is independent of the acetylation of CREB-2 (ATF-4). Virology 299:271-278.

Gachon, F., A. Peleraux, S. Thebault, J. Dick, I. Lemasson, C. Devaux, and J. M. Mesnard. 1998. CREB-2, a cellular CRE-Dependent transcription repressor, functions in association with tax as an activator of the human T-cell leukemia virus type 1 promoter. J Virol 72:8332-8337.

Garcia-Rodriguez, C. and A. Rao. 1998. Nuclear factor of activated T cells (NFAT)- dependent transactivation regulated by the coactivators p300/CREB-binding protein (CBP). J. Exp. Med. 187:2031-2036.

Gartenhaus, R. B. and P. Wang. 1995. Functional inactivation of wild-type p53 protein correlates with loss of IL-2 dependence in HTLV-I transformed human T lymphocytes. Leukemia 9:2082-2086.

Gatza, M. L., J. C. Watt, and S. J. Marriott. 2003. Cellular transformation by the HTLV-I Tax protein, a jack-of-all-trades. Oncogene 22:5141-5149. 217 Georges, S. A., H. A. Giebler, P. A. Cole, K. Luger, P. J. Laybourn, and J. K. Nyborg. 2003. Tax recruitment of CBP/p300, via the KIX domain, reveals a potent requirement for acetyltransferase activity that is chromatin dependent and histone tail independent. Mol. Cell Biol. 23:3392-3404.

Gerritsen, M. E., A. J. Williams, A. S. Neish, S. Moore, Y. Shi, and T. Collins. 1997. CREB-binding protein/p300 are transcriptional coactivators of p65. Proc. Natl. Acad. Sci. U. S. A 94:2927-2932.

Gessain, A. 1996. Virological aspects of tropical spastic paraparesis/HTLV-I associated myelopathy and HTLV-I infection. J. Neurovirology. 2:299-306.

Gessain, A., F. Barin, J. Vernant, O. Gout, L. Maurs, A. Calender, and G. Dethe. 1985. Antibodies to human T lymphotropic virus type 1 in patients with tropical spastic paresis. Lancet 2:407-410.

Gessain, A., R. C. Gallo, and G. Franchini. 1992. Low degree of human T-cell leukemia/lymphoma virus type I genetic drift in vivo as means of monitoring viral transmission and movement of ancient human populations. J. Virol. 66:2288-2295.

Gessain, A. and R. Mahieux. 2000. [A virus called HTLV-1. Epidemiological aspects] 5232. Presse Med. 29:2233-2239.

Giebler, H. A., J. E. Loring, K. Vanorden, M. A. Colgin, J. E. Garrus, K. W. Escudero, A. Brauweiler, and J. K. Nyborg. 1997. Anchoring of CREB binding protein to the human T-cell leukemia virus type 1 promoter: A molecular mechanism of tax transactivation. Mol. Cell Biol. 17:5156-5164.

Godoy, A. J., J. Kira, K. Hasuo, and I. Goto. 1995. Characterization of cerebral white matter lesions of HTLV- I- associated myelopathy tropical spastic paraparesis in comparison with multiple sclerosis and collagen-vasculitis: A semiquantitative MRI study. J. Neurol. Sci. 133:102-111.

Goh, W. C., N. Manel, and M. Emerman. 2004. The human immunodeficiency virus Vpr protein binds Cdc25C: implications for G2 arrest. Virology 318:337-349.

Gonzalez, G. A., K. K. Yamamoto, W. H. Fischer, D. Karr, P. Menzel, W. Biggs, III, W. W. Vale, and M. R. Montminy. 1989. A cluster of phosphorylation sites on the cyclic AMP-regulated nuclear factor CREB predicted by its sequence 8. Nature 337:749-752.

Goodman, R. H. and S. Smolik. 2000. CBP/p300 in cell growth, transformation, and development. Genes Dev. 14:1553-1577.

Gould, K. G. and C. R. Bangham. 1998. Virus variation, escape from cytotoxic T lymphocytes and human retroviral persistence. Semin Cell Dev Biol 9:321-328.

218 Grant, C., K. Barmak, T. Alefantis, J. Yao, S. Jacobson, and B. Wigdahl. 2002. Human T cell leukemia virus type I and neurologic disease: events in bone marrow, peripheral blood, and central nervous system during normal immune surveillance and neuroinflammation. J. Cell Physiol 190:133-159.

Grassmann, R., S. Berchtold, I. Radant, M. Alt, B. Fleckenstein, J. G. Sodroski, W. A. Haseltine, and U. Ramstedt. 1992. Role of HTLV-1 X Region Proteins in Immortalization of Primary Human Lymphocytes in Culture. J. Virol. 66:4570-4575.

Green P.L. and Chen, B. K. Molecular features of the human T-cell leukemia virus: Mechanisms of transfromation and leukemogenicity. The Retroviridae. Vol. 3 (Levy J, ed.), 277-298. 1994. Plenum Publishing Corp. New York.

Green P.L. and I. S. Y. Chen. 2000. Human T-cell leukemia virus types 1 and 2, p. 1941-1969. Lippincott, Williams & Wilkins.

Green, P. L. and I. S. Y. Chen. 1994. Molecular features of the human T-cell leukemia virus. Mechanisms of transformation and leukemogenicity., p. 277-311. In J. A. Levy (ed.), The Retroviridae. Plenum Press, New York.

Grunstein, M. 1997. Histone acetylation in chromatin structure and transcription. Nature 389:349-352.

Hai, T. W., F. Liu, W. J. Coukos, and M. R. Green. 1989. Transcription factor ATF cDNA clones: an extensive family of leucine zipper proteins able to selectively form DNA-binding heterodimers. Genes Dev. 3:2083-2090.

Hakata, Y., T. Umemoto, S. Matsushita, and H. Shida. 1998. Involvement of human CRM1 (exportin 1) in the export and multimerization of the Rex protein of human T-cell leukemia virus type 1. J Virol 72:6602-6607.

Hakata, Y., M. Yamada, and H. Shida. 2003. A multifunctional domain in human CRM1 (exportin 1) mediates RanBP3 binding and multimerization of human T-cell leukemia virus type 1 Rex protein. Mol. Cell Biol. 23:8751-8761.

Hall, W. W., C. R. Liu, O. Schneewind, H. Takahashi, M. H. Kaplan, G. Roupe, and A. Vahlne. 1991. Deleted HTLV-1 provirus in blood and cutaneous lesions of patients with mycosis fungoides. Science 253:317-319.

Hanchard, B. 1996. Adult T-cell leukemia/lymphoma in Jamaica: 1986-1995. J Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S20-S25.

Hara, H., M. Morita, T. Iwaki, T. Hatae, Y. Itoyama, T. Kitamoto, S. Akizuki, I. Goto, and T. Watanabe. 1994. Detection of human T lymphotrophic virus type I (HTLV-I) proviral DNA and analysis of T cell receptor V beta CDR3 sequences in spinal cord lesions of HTLV-I-associated myelopathy/tropical spastic paraparesis. J. Exp. Med. 180:831-839. 219 Harhaj, E. W., L. F. Good, G. T. Xiao, and S. C. Sun. 1999. Gene expression profiles in HTLV-I-immortalized T cells: deregulated expression of genes involved in apoptosis regulation. Oncogene 18:1341-1349.

Harrod, R., Y. L. Kuo, Y. Tang, Y. Yao, A. Vassilev, Y. Nakatani, and C. Z. Giam. 2000. p300 and p300/cAMP-responsive element-binding protein associated factor interact with human T-cell lymphotropic virus type-1 Tax in a multi- histone acetyltransferase/activator-enhancer complex. J. Biol. Chem. 275:11852-11857.

Harrod, R., Y. Tang, C. Nicot, H. S. Lu, A. Vassilev, Y. Nakatani, and C. Z. Giam. 1998. An exposed KID-like domain in human T-cell lymphotropic virus type 1 Tax is responsible for the recruitment of coactivators CBP/p300. Mol. Cell Biol. 18:5052-5061.

Hattori, S., T. Kiyokawa, K. Imagawa, F. Shimizu, E. Hashimura, M. Seiki, and M. Yoshida. 1984. Identification of gag and env gene products of human T-cell leukemia virus (HTLV). Virology 136:338-347.

Heger, P., O. Rosorius, J. Hauber, and R. H. Stauber. 1999. Titration of cellular export factors, but not heteromultimerization, is the molecular mechanism of trans- dominant HTLV-1 Rex mutants. Oncogene 18:4080-4090.

Hino, S. 1989. Milk-borne transmission of HTLV-I as a major route in the endemic cycle. Acta Paediatr. Jpn. 31:428-435.

Hino, S., K. Yamaguchi, S. Katamine, H. Sugiyama, T. Amagasaki, K. Kinoshita, Y. Yoshida, H. Doi, Y. Tsuji, and T. Miyamoto. 1985. Mother-to-child transmission of human T-cell leukemia virus type-I. Jpn. J. Cancer Res. 76:474-480.

Hinuma, Y., H. Komoda, T. Chosa, T. Kondo, M. Kohakura, T. Takenaka, M. Kikuchi, M. Ichimaru, K. Yunoki, I. Sato, R. Matsuo, Y. Takiuchi, H. Uchino, and M. Hanaoka. 1982. Antibodies to adult T-cell leukemia virus-associated antigen (ATLA) in sera from patients with ATL and controls in Japan: a nation-wide seroepidemiological study. Int. J. Cancer 29:631.

Hinuma, Y., K. Nagata, M. Hanaoka, M. Nakai, T. Matsumoto, K. Shirakawa, and I. Miyoshi. 1981. Adult T Cell Leukemia: Antigens in an ATL cell line and detection of antibodies to antigen in human sera. Proc. Natl. Acad. Sci. U. S. A. 78:6476-6480.

Hollsberg, P. 1999. Mechanisms of T-cell activation by human T-cell lymphotropic virus type I. Microbiol. Mol. Biol. Rev. 63:308-333.

Hollsberg, P., L. J. Ausubel, and D. A. Hafler. 1994. Human T Cell Lymphotropic Virus Type 1-Induced T Cell Activation. J. Immunol. 153:566-573.

Hollsberg, P., K. W. Wucherpfennig, L. J. Ausubel, V. Calvo, B. E. Bierer, and D. A. Hafler. 1992. Characterization of HTLV-1 In Vivo Infected T Cell Clones IL-2- independent Growth of Nontransformed T Cells. J. Immunol. 148:3256-3263. 220 Hou, X., S. Foley, M. Cueto, and M. A. Robinson. 2000. The Human T-Cell Leukemia Virus Type I (HTLV-I) X Region Encoded Protein p13(II) Interacts with Cellular Proteins. Virology 277:127-135.

Hurd, C. and E. Rozengurt. 2003. Uncoupling of protein kinase D from suppression of EGF-dependent c-Jun phosphorylation in cancer cells. Biochem. Biophys. Res. Commun. 302:800-804.

Ibrahim, F., G. Dethe, and A. Gessain. 1995. Isolation and characterization of a new simian T-cell leukemia virus type 1 from naturally infected celebes macaques (Macaca tonkeana): Complete nucleotide sequence and phylogenetic relationship with the australo-melanesian human T-cell leukemia virus type 1. J. Virol. 69:6980-6993.

Igakura, T., J. C. Stinchcombe, P. K. Goon, G. P. Taylor, J. N. Weber, G. M. Griffiths, Y. Tanaka, M. Osame, and C. R. Bangham. 2003. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science 299:1713-1716.

Inoue, J., M. Seiki, T. Taniguchi, S. Tsuru, and M. Yoshida. 1986. Induction of interleukin 2 receptor gene expression by p40x encoded by human T-cell leukemia virus type 1. EMBO J. 5:2883-2888.

Inoue, J., M. Seiki, and M. Yoshida. 1986. The second pX product p27 chi-III of HTLV-1 is required for gag gene expression. FEBS Lett. 209:187-190.

Inoue, J., M. Yoshida, and M. Seiki. 1987. Transcriptional (p40x) and post- transcriptional (p27x-III) regulators are required for the expression and replication of human t-cell leukemia virus type I genes. Proc. Natl. Acad. Sci. USA 84:3653-3657.

Itoh, M., J. Inoue, H. Toyoshima, T. Akizawa, M. Higashi, and M. Yoshida. 1989. HTLV-1 rex and HIV-1 rev act through similar mechanisms to relieve suppression of unspliced RNA expression. Oncogene 4:1275-1279.

Izumo, S., F. Umehara, and M. Osame. 2000. HTLV-I-associated myelopathy. Neuropathology. 20 Suppl:S65-S68.

Jacobson, S. 2002. Immunopathogenesis of human T cell lymphotropic virus type I- associated neurologic disease. J Infect. Dis. 186 Suppl 2:S187-S192.

Janknecht, R. 2002. The versatile functions of the transcriptional coactivators p300 and CBP and their roles in disease. Histol. Histopathol. 17:657-668.

Janknecht, R. and T. Hunter. 1996. Transcription. A growing coactivator network [news; comment]. Nature 383:22-23.

Janknecht, R. and T. Hunter. 1996. Transcriptional control: Versatile molecular glue. Curr. Biol. 6:951-954. 221 Jeang, K. T. 2001. Functional activities of the human T-cell leukemia virus type I Tax oncoprotein: cellular signaling through NF-kappa B. Cytokine Growth Factor Rev. 12:207-217.

Jeang, K. T., R. Chiu, E. Santos, and S. J. Kim. 1991. Induction of the HTLV-I LTR by Jun occurs through the Tax-responsive 21- bp elements 5255. Virology 181:218-227.

Jeang, K. T., S. G. Widen, O. J. Semmes, and S. H. Wilson. 1990. HTLV-1 Trans- Activator Protein, Tax, Is a Trans-Repressor of the Human B-Ploymerase Gene. Science 247:1082-1083.

Jiang, H., H. Lu, R. L. Schiltz, C. A. Pise-Masison, V. V. Ogryzko, Y. Nakatani, and J. N. Brady. 1999. PCAF interacts with tax and stimulates tax transactivation in a histone acetyltransferase-independent manner [published erratum appears in Mol Cell Biol 2000 Mar;20(5):1897]. Mol. Cell Biol. 19:8136-8145.

Jin, D. Y., F. Spencer, and K. T. Jeang. 1998. Human T cell leukemia virus type 1 oncoprotein Tax targets the human mitotic checkpoint protein MAD1. Cell 93:81-91.

Jinno, A., Y. Haraguchi, H. Shiraki, and H. Hoshino. 1999. Inhibition of cell-free human T-cell leukemia virus type 1 infection at a postbinding step by the synthetic peptide derived from an ectodomain of the gp21 transmembrane glycoprotein 1. J Virol 73:9683-9689.

Johnson, C., D. Van Antwerp, and T. J. Hope. 1999. An N-terminal nuclear export signal is required for the nucleocytoplasmic shuttling of IkappaBalpha. EMBO J 18:6682-6693.

Johnson, J. M. and G. Franchini. 2002. Retroviral proteins that target the major histocompatibility complex class I. Virus Res. 88:119-127.

Johnson, J. M., J. C. Mulloy, V. Ciminale, J. Fullen, C. Nicot, and G. Franchini. 2000. The MHC class I heavy chain is a common target of the small proteins encoded by the 3' end of HTLV type 1 and HTLV type 2. AIDS Res. Hum. Retroviruses 16:1777- 1781.

Johnson, J. M., C. Nicot, J. Fullen, V. Ciminale, L. Casareto, J. C. Mulloy, S. Jacobson, and G. Franchini. 2001. Free Major Histocompatibility Complex Class I Heavy Chain Is Preferentially Targeted for Degradation by Human T-Cell Leukemia/Lymphotropic Virus Type 1 p12(I) Protein. J. Virol. 75:6086-6094.

Ju, Y. T., A. C. Chang, B. R. She, M. L. Tsaur, H. M. Hwang, C. C. Chao, S. N. Cohen, and S. Lin-Chao. 1998. gas7: A gene expressed preferentially in growth-arrested fibroblasts and terminally differentiated Purkinje neurons affects neurite formation. Proc. Natl. Acad. Sci U. S. A 95:11423-11428.

222 Kaehlcke, K., A. Dorr, C. Hetzer-Egger, V. Kiermer, P. Henklein, M. Schnoelzer, E. Loret, P. A. Cole, E. Verdin, and M. Ott. 2003. Acetylation of Tat defines a cyclinT1- independent step in HIV transactivation. Mol. Cell 12:167-176.

Kamb, A., N. A. Gruis, J. Weaver-Feldhaus, Q. Liu, K. Harshman, S. V. Tavtigian, E. Stockert, R. S. Day, III, B. E. Johnson, and M. H. Skolnick. 1994. A cell cycle regulator potentially involved in genesis of many tumor types. Science 264:436-440.

Kamei, Y., L. Xu, T. Heinzel, J. Torchia, R. Kurokawa, B. Gloss, S. C. Lin, R. A. Heyman, D. W. Rose, C. K. Glass, and M. G. Rosenfeld. 1996. A CBP integrator complex mediates transcriptional activation and AP-1 inhibition by nuclear receptors. Cell 85:403-414.

Kamihira, S., S. Nakasima, Y. Oyakawa, Y. Moriuti, M. Ichimaru, H. Okuda, M. Kanamura, and T. Oota. 1987. Transmission of human T cell lymphotropic virus type I by blood transfusion before and after mass screening of sera from seropositive donors 5217. Vox Sang. 52:43-44.

Kaminuma, O., M. Deckert, C. Elly, Y. C. Liu, and A. Altman. 2001. Vav-Rac1- mediated activation of the c-Jun N-terminal kinase/c-Jun/AP-1 pathway plays a major role in stimulation of the distal NFAT site in the interleukin-2 gene promoter. Mol. Cell Biol. 21:3126-3136.

Kanamori, H., N. Suzuki, H. Siomi, T. Nosaka, A. Sato, H. Sabe, M. Hatanaka, and T. Honjo. 1990. HTLV-1 p27rex stabilizes human interleukin-2 receptor alpha chain mRNA. EMBO J. 9:4161-4166.

Kaplan, J. E., M. Osame, H. Kubota, A. Igata, H. Nishitani, Y. Maeda, R. F. Khabbaz, and R. S. Janssen. 1990. The risk of development of HTLV-I-Associated myelopathy/tropical spastic paraparesis among persons infected with HTLV-I. J. Acq. Immun. Defic. Synd. 3:1096-1101.

Kasahata, N., J. Shiota, Y. Miyazawa, I. Nakano, and S. Murayama. 2003. Acute human T-lymphotropic virus type 1-associated myelopathy: a clinicopathologic study. Arch. Neurol. 60:873-876.

Kasai, T., Y. Iwanaga, H. Iha, and K. T. Jeang. 2002. Prevalent loss of mitotic spindle checkpoint in adult T-cell leukemia confers resistance to microtubule inhibitors. J. Biol. Chem. 277:5187-5193.

Kashanchi, F., J. F. Duvall, R. P. Kwok, J. R. Lundblad, R. H. Goodman, and J. N. Brady. 1998. The coactivator CBP stimulates human T-cell lymphotrophic virus type I Tax transactivation in vitro. J. Biol. Chem. 273:34646-34652.

Kawakami, K., A. Miyazato, Y. Iwakura, and A. Saito. 1999. Induction of lymphocytic inflammatory changes in lung interstitium by human T lymphotropic virus type I. AMER J RESPIR CRIT CARE MED 160:995-1000. 223 Kawano, F., K. Yamaguchi, H. Nishimura, H. Tsuda, and K. Takatsuki. 1985. Variation in the clinical courses of adult T-cell leukemia. Cancer 55:851-856.

Kerkhofs, P., H. Heremans, A. Burny, R. Kettmann, and L. Willems. 1998. In vitro and in vivo oncogenic potential of bovine leukemia virus G4 protein. J Virol 72:2554- 2559.

Khabbaz, R. F., J. M. Douglas, F. N. Judson, R. A. Spiegel, M. E. St Louis, W. Whittington, T. M. Hartley, M. Lairmore, and J. E. Kaplan. 1990. Seroprevalence of human T-lymphotropic virus type I or II in sexually transmitted disease clinic patients in the USA. J. Infect. Dis. 162:241-244.

Khabbaz, R. F., I. M. Onorato, R. O. Cannon, T. M. Hartley, B. Roberts, B. Hosein, and J. E. Kaplan. 1992. Seroprevalence of HTLV-1 and HTLV-2 among intravenous drug users and persons in clinics for sexually transmitted diseases. N. Engl. J. Med. 326:375-380.

Khan, Z. M., M. Sebenik, and D. Zuckerfranklin. 1996. Localization of human T-cell lymphotropic virus-1 tax proviral sequences in skin biopsies of patients with mycosis fungoides by in situ polymerase chain reaction. J. Invest. Dermatol. 106:667-672.

Kiernan, R. E., C. Vanhulle, L. Schiltz, E. Adam, H. Xiao, F. Maudoux, C. Calomme, A. Burny, Y. Nakatani, K. T. Jeang, M. Benkirane, and C. Van Lint. 1999. HIV-1 tat transcriptional activity is regulated by acetylation. EMBO J. 18:6106- 6118.

Kim, J. H., P. A. Kaufman, S. M. Hanly, L. T. Rimsky, and W. C. Greene. 1991. Rex transregulation of human T-cell leukemia virus type II gene expression. J. Virol. 65:405- 414.

Kim, S. J., W. Ding, B. Albrecht, P. L. Green, and M. D. Lairmore. 2003. A conserved calcineurin-binding motif in human T lymphotropic virus type 1 p12I functions to modulate nuclear factor of activated T cell activation. J Biol. Chem. 278:15550-15557.

Kimata, J. T. and L. Ratner. 1991. Temporal regulation of viral and cellular gene expression during HTLV-I mediated lymphocyte immortalization. J. Virol. 65:4398- 4407.

Kimata, J. T., F. Wong, J. Wang, and L. Ratner. 1994. Construction and characterization of infectious human T-cell leukemia virus type 1 molecular clones. Virology 204:656-664.

Kimura, I., T. Tsubota, S. Tada, and J. Sogawa. 1986. Presence of antibodies against adult T cell leukemia antigen in the patients with chronic respiratory diseases 5241. Acta Med. Okayama 40:281-284.

224 Kimzey, A. L. and W. S. Dynan. 1998. Specific regions of contact between human T- cell leukemia virus type I Tax protein and DNA identified by photocross- linking. J Biol Chem 273:13768-13775.

Kimzey, A. L. and W. S. Dynan. 1999. Identification of a human T-cell leukemia virus type I tax peptide in contact with DNA. J Biol Chem 274:34226-34232.

Kira, J., I. Goto, M. Otsuka, and Y. Ichiya. 1993. Chronic Progressive Spinocerebellar Syndrome Associated with Antibodies to Human T-Lymphotropic Virus Type-I - Clinico-Virological and Magnetic Resonance Imaging Studies. J. Neurol. Sci. 115:111- 116.

Kira, J. I., K. Fujihara, Y. Itoyama, I. Goto, and K. Hasuo. 1991. Leukoencephalopathy in HTLV-I-associated mylopathy/ tropical spastic paraparesis: MRI analysis and a two year follow up study after corticosteroid therapy. J. Neurol. Sci. 106:41-49.

Kitado, H., I. S. Chen, N. P. Shah, A. J. Cann, K. Shimotohno, and H. Fan. 1987. U3 sequences from HTLV-I and -II LTRs confer pX protein response to a murine leukemia virus LTR. Science 235:901-904.

Kitajima, I., K. Yamamoto, K. Sato, Y. Nakajima, T. Nakajima, I. Maruyama, M. Osame, and K. Nishioka. 1991. Detection of human T cell lymphotropic virus type I proviral DNA and its gene expression in synovial cells in chronic inflammatory arthropathy. J. Clin. Invest 88:1315-1322.

Kitze, B. and J. N. Brady. 1997. Human T cell lymphotropic retroviruses: Association with diseases of the nervous system. Intervirology 40:132-142.

Kitze, B. and K. Usuku. 2002. HTLV-1-mediated immunopathological CNS disease. Curr. Top. Microbiol. Immunol. 265:197-211.

Kiwaki, T., F. Umehara, Y. Arimura, S. Izumo, K. Arimura, K. Itoh, and M. Osame. 2003. The clinical and pathological features of peripheral neuropathy accompanied with HTLV-I associated myelopathy. J Neurol. Sci. 206:17-21.

Kiyokawa, T., M. Seiki, S. Iwashita, K. Imagawa, F. Shimizu, and M. Yoshida. 1985. p27x-III and p21x-III, proteins encoded by the pX sequence of human T-cell leukemia virus type 1. Proc. Natl. Acad. Sci. USA 82:8359-8363.

Kobe, B., R. J. Center, B. E. Kemp, and P. Poumbourios. 1999. Crystal structure of human T cell leukemia virus type 1 gp21 ectodomain crystallized as a maltose-binding protein chimera reveals structural evolution of retroviral transmembrane proteins. PROC NAT ACAD SCI USA 96:4319-4324.

225 Kohno, T., R. Moriuchi, S. Katamine, Y. Yamada, M. Tomonaga, and T. Matsuyama. 2000. Identification of genes associated with the progression of adult T cell leukemia (ATL). Jpn. J Cancer Res. 91:1103-1110.

Kohno, T., R. Moriuchi, S. Katamine, Y. Yamada, M. Tomonaga, and T. Matsuyama. 2000. Identification of genes associated with the progression of adult T cell leukemia (ATL). Jpn. J Cancer Res. 91:1103-1110.

Kohno, T., R. Moriuchi, S. Katamine, Y. Yamada, M. Tomonaga, and T. Matsuyama. 2000. Identification of genes associated with the progression of adult T cell leukemia (ATL). Jpn. J Cancer Res. 91:1103-1110.

Kondo, T., H. Kono, N. Miyamoto, R. Yoshida, H. Toki, I. Matsumoto, M. Hara, H. Inoue, A. Inatsuki, T. Funatsu, and . 1989. Age- and sex-specific cumulative rate and risk of ATLL for HTLV-I carriers. Int. J Cancer 43:1061-1064.

Konikova, E. and J. Kusenda. 2003. Altered expression of p53 and MDM2 proteins in hematological malignancies. Neoplasma 50:31-40.

Koralnik, I. J., E. Boeri, W. C. Saxinger, A. L. Monico, J. Fullen, A. Gessain, H. G. Guo, R. C. Gallo, P. Markham, V. Kalyanaraman, V. Hirsch, J. Allan, K. Murthy, P. Alford, J. P. Slattery, S. J. O'brien, and G. Franchini. 1994. Phylogenetic association of human and Simian T-cell leukemia/lymphotropic virus type I strains: evidence for interspecies transmission/. J. Virol. 68:2693-2707.

Koralnik, I. J., J. Fullen, and G. Franchini. 1993. The p12, p13 and p30 proteins encoded by human T-cell leukemia/lymphotropic virus type-1 open reading frames I and II are localized in three different cellular compartments. J. Virol. 67:2360-2366.

Koralnik, I. J., J. Fullen, and G. Franchini. 1993. The p12I, p13II, and p30II proteins encoded by human T-cell leukemia/lymphotropic virus type I open reading frames I and II are localized in three different cellular compartments. J Virol 67:2360-2366.

Koralnik, I. J., A. Gessain, M. E. Klotman, A. Lo Monico, Z. N. Berneman, and G. Franchini. 1992. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type 1. Proc. Natl. Acad. Sci. USA 89:8813-8817.

Koralnik, I. J., A. Gessain, M. E. Klotman, M. A. Lo, Z. N. Berneman, and G. Franchini. 1992. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type I. Proc. Natl. Acad. Sci U. S. A 89:8813-8817.

Koralnik, I. J., J. C. Mulloy, T. Andresson, J. Fullen, and G. Franchini. 1995. Mapping of the intermolecular association of human T cell leukaemia/lymphotropic virus type I p12I and the vacuolar H+-ATPase 16 kDa subunit protein. J Gen. Virol 76 ( Pt 8):1909-1916.

226 Korber, B., A. Okayama, R. Donnelly, N. Tachibana, and M. Essex. 1991. Polymerase chain reaction analysis of defective human T-cell leukemia virus type 1 proviral genomes in leukemic cells of patients with adult T-cell leukemia. J. Virol. 65:5471-5476.

Korzus, E., J. Torchia, D. W. Rose, L. Xu, R. Kurokawa, E. M. Mcinerney, T. M. Mullen, C. K. Glass, and M. G. Rosenfeld. 1998. Transcription factor-specific requirements for coactivators and their acetyltransferase functions. Science 279:703-707.

Kubota, R., S. S. Soldan, R. Martin, and S. Jacobson. 2002. Selected cytotoxic T lymphocytes with high specificity for HTLV-I in cerebrospinal fluid from a HAM/TSP patient. J. Neurovirol. 8:53-57.

Kubota, S., M. Hatanaka, and R. J. Pomerantz. 1996. Nucleo-cytoplasmic redistribution of the HTLV-I rex protein: Alterations by coexpression of the HTLV-I p21(x) protein. Virology 220:502-507.

Kuklina, E. M. and S. V. Shirshev. 2001. Role of transcription factor NFAT in the immune response. Biochemistry (Mosc. ) 66:467-475.

Kuroda, Y. and M. Matsui. 1993. Cerebrospinal Fluid Interferon-Gamma Is Increased in HTLV- I- Associated Myelopathy. J. Neuroimmunol. 42:223-226.

Kwok, R. P., M. E. Laurance, J. R. Lundblad, P. S. Goldman, H. Shih, L. M. Connor, S. J. Marriott, and R. H. Goodman. 1996. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature 380:642-646.

Kwok, R. P., J. R. Lundblad, J. C. Chrivia, J. P. Richards, H. P. Bachinger, R. G. Brennan, S. G. Roberts, M. R. Green, and R. H. Goodman. 1994. Nuclear protein CBP is a coactivator for the transcription factor. Nature 370(6486):223-226.

Kwok, R. P. S., M. E. Laurance, J. R. Lundblad, P. S. Goldman, H. M. Shih, L. M. Connor, S. J. Marriott, and R. H. Goodman. 1996. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature 380:642-646.

Lagrenade, L., B. Hanchard, V. Fletcher, B. Cranston, and W. Blattner. 1990. Infective dermatitis of Jamaican children: a marker for HTLV-I infection 5233. Lancet 336:1345-1347.

Lagrenade, L., A. Manns, V. Fletcher, C. Carberry, B. Hanchard, E. M. Maloney, B. Cranston, N. P. Williams, R. Wilks, E. C. Kang, and W. A. Blattner. 1998. Clinical, pathologic, and immunologic features of human T- lymphotrophic virus type I-associated infective dermatitis in children. Arch Dermatol 134:439-444.

227 Lagrenade, L., C. Morgan, C. Carberry, B. Hanchard, V. Fletcher, R. Gray, B. Cranston, P. Rodgers-Johnson, and A. Manns. 1995. Tropical spastic paraparesis occurring in HTLV-1 associated infective dermatitis. Report of two cases. West Indian Med J 44:34-35.

Lagrenade, L., S. Sonoda, W. Miller, E. Pate, P. Rodgersjohnson, B. Hanchard, B. Cranston, T. Fujiyoshi, S. Yashiki, M. Blank, C. J. Gibbs, and A. Manns. 1996. HLA DRB1(*)DQB1(*) haplotype in HTLV-I-associated familial infective dermatitis may predict development of HTLV-I- associated myelopathy tropical spastic paraparesis. Am. J. Med. Genet. 61:37-41.

Lahti, A. L., A. Manninen, and K. Saksela. 2003. Regulation of T cell activation by HIV-1 accessory proteins: Vpr acts via distinct mechanisms to cooperate with Nef in NFAT-directed gene expression and to promote transactivation by CREB. Virology 310:190-196.

Lairmore, M. D., B. Albrecht, C. D'Souza, J. W. Nisbet, W. Ding, J. T. Bartoe, P. L. Green, and W. Zhang. 2000. In vitro and in vivo functional analysis of human T cell lymphotropic virus type 1 pX open reading frames I and II. AIDS Res. Hum. Retroviruses 16:1757-1764.

Latchman, D. S. 1999. Regulation of DNA virus transcription by cellular POU family transcription factors. Rev Med Virol 9:31-38.

Le, B., I, M. P. Grange, L. Delamarre, A. R. Rosenberg, V. Blot, C. Pique, and M. C. Dokhelar. 2001. HTLV-1 structural proteins. Virus Res. 78:5-16.

Le, B., I, A. R. Rosenberg, and M. C. Dokhelar. 1999. Multiple functions for the basic amino acids of the human T-cell leukemia virus type 1 matrix protein in viral transmission. J Virol 73:1860-1867.

Lefebvre, L., V. Ciminale, A. Vanderplasschen, D. D'Agostino, A. Burny, L. Willems, and R. Kettmann. 2002. Subcellular localization of the bovine leukemia virus R3 and G4 accessory proteins. J Virol 76:7843-7854.

Lefebvre, L., A. Vanderplasschen, V. Ciminale, H. Heremans, O. Dangoisse, J. C. Jauniaux, J. F. Toussaint, V. Zelnik, A. Burny, R. Kettmann, and L. Willems. 2002. Oncoviral Bovine Leukemia Virus G4 and Human T-Cell Leukemia Virus Type 1 p13(II) Accessory Proteins Interact with Farnesyl Pyrophosphate Synthetase. J. Virol. 76:1400- 1414.

Lemasson, I., N. J. Polakowski, P. J. Laybourn, and J. K. Nyborg. 2002. Transcription factor binding and histone modifications on the integrated proviral promoter in human T-cell leukemia virus-I-infected T-cells. J Biol. Chem. 277:49459- 49465.

228 Lemasson, I., V. Roberthebmann, S. Hamaia, M. D. Dodon, L. Gazzolo, and C. Devaux. 1997. Transrepression of lck gene expression by human T-cell leukemia virus type 1-encoded p40(tax). J. Virol. 71:1975-1983.

Lenzmeier, B. A., E. E. Baird, P. B. Dervan, and J. K. Nyborg. 1999. The Tax protein-DNA interaction is essential for HTLV-I transactivation in vitro. J Mol Biol 291:731-744.

Leonmonzon, M., I. Illa, and M. C. Dalakas. 1994. Polymyositis in patients infected with human T-cell leukemia virus type I: The role of the virus in the cause of the disease. Ann. Neurol. 36:643-649.

Levin, M. C., S. M. Lee, F. Kalume, Y. Morcos, F. C. Dohan, Jr., K. A. Hasty, J. C. Callaway, J. Zunt, D. Desiderio, and J. M. Stuart. 2002. Autoimmunity due to molecular mimicry as a cause of neurological disease. Nat. Med. 8:509-513.

Levin, M. C., S. M. Lee, Y. Morcos, J. Brady, and J. Stuart. 2002. Cross-reactivity between immunodominant human T lymphotropic virus type I tax and neurons: implications for molecular mimicry. J Infect. Dis. 186:1514-1517.

Lewis, R. S. 2001. Calcium signaling mechanisms in T lymphocytes. Annu. Rev. Immunol. 19:497-521.:497-521.

Li, C., A. J. Lusis, R. Sparkes, S. M. Tran, and R. Gaynor. 1992. Characterization and chromosomal mapping of the gene encoding the cellular DNA binding protein HTLF. Genomics 13:658-664.

Li, C. and W. H. Wong. 2001. Model-based analysis of oligonucleotide arrays: expression index computation and outlier detection. Proc. Natl. Acad. Sci U. S. A 98:31- 36.

Li, C. C., F. W. Ruscetti, N. R. Rice, E. Chen, N. S. Yang, J. Mikovits, and D. L. Longo. 1993. Differential expression of Rel family members in human T-cell leukemia virus type I-infected cells: transcriptional activation of c-rel by Tax protein. J. Virol. 67:4205-4213.

Li, G. Q., B. R. Vowels, B. M. Benoit, A. H. Rook, and S. R. Lessin. 1996. Failure to detect human T-lymphotropic virus type-I proviral DNA in cell lines and tissues from patients with cutaneous T-cell lymphoma. J. Invest. Dermatol. 107:308-313.

Li, H. C., T. Fujiyoshi, H. Lou, S. Yashiki, S. Sonoda, L. Cartier, L. Nunez, I. Munoz, S. Horai, and K. Tajima. 1999. The presence of ancient human T-cell lymphotropic virus type I provirus DNA in an Andean mummy. NATURE MED 5:1428- 1432.

229 Li, M., B. Damania, X. Alvarez, V. Ogryzko, K. Ozato, and J. U. Jung. 2000. Inhibition of p300 histone acetyltransferase by viral interferon regulatory factor. Mol. Cell Biol. 20:8254-8263.

Li, M., B. Damania, X. Alvarez, V. Ogryzko, K. Ozato, and J. U. Jung. 2000. Inhibition of p300 histone acetyltransferase by viral interferon regulatory factor [In Process Citation]. Mol. Cell Biol. 20:8254-8263.

Li, X. H., K. M. Murphy, K. T. Palka, R. M. Surabhi, and R. B. Gaynor. 1999. The human T-cell leukemia virus type-1 Tax protein regulates the activity of the I kappa B kinase complex. J Biol Chem 274:34417-34424.

Li-Weber, M., M. Giaisi, K. Chlichlia, K. Khazaie, and P. H. Krammer. 2001. Human T cell leukemia virus type I Tax enhances IL-4 gene expression in T cells. Eur. J Immunol. 31:2623-2632.

Lindholm, P. F., R. L. Reid, and J. N. Brady. 1992. Extracellular Tax Protein Stimulates Tumor Necrosis Factor-B and Immunoglobulin Kappa Light Chain Expression in Lymphoid Cells. J. Virol. 66:1294-1302.

Liu, H. F., P. Goubau, M. Vanbrussel, K. Vanlaethem, Y. C. Chen, J. Desmyter, and A. M. Vandamme. 1996. The three human T-lymphotropic virus type I subtypes arose from three geographically distinct simian reservoirs. J. Gen. Virol. 77:359-368.

Liu, Y., A. L. Colosimo, X. J. Yang, and D. Liao. 2000. Adenovirus E1B 55-kilodalton oncoprotein inhibits p53 acetylation by PCAF. Mol. Cell Biol. 20:5540-5553.

Lombard Platet, G. and P. Jalinot. 1993. Purification by DNA affinity precipitation of the cellular factors HEB1-p67 and HEB1-p94 which bind specifically to the human T-cell leukemia virus type-I 21 bp enhancer. Nucleic. Acids. Res. 21:3935-3942.

Lu, H., C. A. Pise-Masison, T. M. Fletcher, R. L. Schiltz, A. K. Nagaich, M. Radonovich, G. Hager, P. A. Cole, and J. N. Brady. 2002. Acetylation of nucleosomal histones by p300 facilitates transcription from tax-responsive human T-cell leukemia virus type 1 chromatin template. Mol. Cell Biol. 22:4450-4462.

Lu, X., H. Yu, S. H. Liu, F. M. Brodsky, and B. M. Peterlin. 1998. Interactions between HIV1 Nef and vacuolar ATPase facilitate the internalization of CD4. Immunity. 8:647-656.

Lundblad, J. R., R. P. Kwok, M. E. Laurance, M. L. Harter, and R. H. Goodman. 1995. Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature 374:85-88.

Lundblad, J. R., R. P. Kwok, M. E. Laurance, M. S. Huang, J. P. Richards, R. G. Brennan, and R. H. Goodman. 1998. The human T-cell leukemia virus-1 transcriptional

230 activator Tax enhances cAMP-responsive element-binding protein (CREB) binding activity through interactions with the DNA minor groove. J Biol Chem 273:19251-19259.

Madeleine, M. M., S. Z. Wiktor, J. J. Goedert, A. Manns, P. H. Levine, R. J. Biggar, and W. A. Blattner. 1993. HTLV-I and HTLV-II World-Wide Distribution - Reanalysis of 4,832 Immunoblot Results. Int. J. Cancer 54:255-260.

Madore, S. J., L. S. Tiley, M. H. Malim, and B. R. Cullen. 1994. Sequence requirements for Rev multimerization in vivo. Virology 202:186-194.

Maggirwar, S. B., E. Harhaj, and S. C. Sun. 1995. Activation of NF-kappa B/Rel by Tax involves degradation of I kappa B alpha and is blocked by a proteasome inhibitor. Oncogene 11:993-998.

Mahana, W., T. M. Zhao, R. Teller, M. A. Robinson, and T. J. Kindt. 1998. Genes in the pX region of human T cell leukemia virus I influence Vav phosphorylation in T cells. Proc. Natl. Acad. Sci. U. S. A 95:1782-1787.

Mahieux, R., F. Ibrahim, P. Mauclere, V. Herve, P. Michel, F. Tekaia, C. Chappey, B. Garin, E. Vanderryst, B. Guillemain, E. Ledru, E. Delaporte, G. Dethe, and A. Gessain. 1997. Molecular epidemiology of 58 new African human T-cell leukemia virus type 1 (HTLV-1) strains: Identification of a new and distinct HTLV-1 molecular subtype in central Africa and in pygmies. J. Virol. 71:1317-1333.

Manca, N., E. Piacentini, M. Gelmi, P. Calzavara, M. A. Manganoni, A. Glukhov, F. Gargiulo, M. Defrancesco, F. Pirali, G. Depanfilis, and A. Turano. 1994. Persistence of human T cell lymphotropic virus type 1 (HTLV-1) sequences in peripheral blood mononuclear cells from patients with mycosis fungoides. J. Exp. Med. 180:1973-1978.

Mandic, R., O. T. Fackler, M. Geyer, T. Linnemann, Y. H. Zheng, and B. M. Peterlin. 2001. Negative factor from SIV binds to the catalytic subunit of the V-ATPase to internalize CD4 and to increase viral infectivity. Mol. Biol. Cell 12:463-473.

Manel, N., F. J. Kim, S. Kinet, N. Taylor, M. Sitbon, and J. L. Battini. 2003. The ubiquitous glucose transporter GLUT-1 is a receptor for HTLV. Cell 115:449-459.

Manninen, A., R. G. Herma, and K. Saksela. 2000. Synergistic activation of NFAT by HIV-1 nef and the Ras/MAPK pathway. J. Biol. Chem. 275:16513-16517.

Manninen, A., P. Huotari, M. Hiipakka, G. H. Renkema, and K. Saksela. 2001. Activation of NFAT-dependent gene expression by Nef: conservation among divergent Nef alleles, dependence on SH3 binding and membrane association, and cooperation with protein kinase C-theta. J. Virol. 75:3034-3037.

Manninen, A. and K. Saksela. 2002. HIV-1 Nef Interacts with Inositol Trisphosphate Receptor to Activate Calcium Signaling in T Cells. J. Exp. Med. 195:1023-1032.

231 Manninen, A. and K. Saksela. 2002. HIV-1 Nef interacts with inositol trisphosphate receptor to activate calcium signaling in T cells. J Exp. Med. 195:1023-1032.

Manns, A., M. Hisada, and L. Lagrenade. 1999. Human T-lymphotropic virus type I infection. Lancet 353:1951-1958.

Manns, A., R. J. Wilks, E. L. Murphy, G. Haynes, J. P. Figueroa, M. Barnett, B. Hanchard, and W. A. Blattner. 1992. A prospective study of transmission by transfusion of HTLV-I and risk factors associated with seroconversion. Int. J. Cancer 51:886-891.

Mariette, X., F. Agbalika, M. T. Daniel, M. Bisson, P. Lagrange, J. C. Brouet, and F. Morinet. 1993. Detection of human T lymphotropic virus type I tax gene in salivary gland epithelium from two patients with Sjogren's syndrome 2. Arthritis Rheum. 36:1423-1428.

Mariette, X., P. Cherot, D. Cazals, C. Brocheriou, J. C. Brouet, and F. Agbalika. 1995. Antibodies to HTLV-I in Sjogren's syndrome. Lancet 345:71.

Marriott, S. J., P. F. Lindholm, K. M. Brown, S. D. Gitlin, J. F. Duvall, M. F. Radonovich, and J. N. Brady. 1990. A 36-kilodalton cellular transcription factor mediates an indirect interaction of human T-cell leukemia/lymphoma virus type 1 Tax1 with a responsive element in the viral long terminal repeat. Mol. Cell Biol. 10:4192- 4201.

Martins, M. L., B. C. Soares, J. G. Ribas, G. W. Thorun, J. Johnson, E. G. Kroon, A. B. Carneiro-Prioetti, and C. A. Bonjardim. 2002. Frequency of p12K and p12R Alleles of HTLV Type 1 in HAM/TSP Patients and in Asymptomatic HTLV Type 1 Carriers. AIDS Res. Hum. Retroviruses 18:899-902.

Maruyama, M., H. Shibuya, H. Harada, M. Hatakeyama, M. Seiki, T. Fujita, J. Inoue, M. Yoshida, and T. Taniguchi. 1987. Evidence for aberrant activation of the interleukin-2 autocrine loop by HTLV-1-encoded p40x and T3/Ti complex triggering 5267. Cell 48:343-350.

Matsui, M., F. Nagumo, J. Tadano, and Y. Kuroda. 1995. Characterization of humoral and cellular immunity in the central nervous system of HAM/TSP. J. Neurol. Sci. 130:183-189.

McCoy, M. A., J. J. Gesell, M. M. Senior, and D. F. Wyss. 2003. Flexible lid to the p53-binding domain of human Mdm2: implications for p53 regulation. Proc. Natl. Acad. Sci U. S. A 100:1645-1648.

Mellstrom, B. and J. R. Naranjo. 2001. Mechanisms of Ca(2+)-dependent transcription. Curr. Opin. Neurobiol. 11:312-319.

232 Merino, F., M. Robert-Guroff, J. Clark, M. Biondo-Bracho, W. Blattner, and R. C. Gallo. 1984. Natural antibodies to human T-cell leukemia/lymphoma virus in healthy Venezuelan populations. Int. J. Cancer 34:501.

Mesaeli, N., K. Nakamura, E. Zvaritch, P. Dickie, E. Dziak, K. H. Krause, M. Opas, D. H. MacLennan, and M. Michalak. 1999. Calreticulin is essential for cardiac development. J. Cell Biol. 144:857-868.

Migone, T. S., J. X. Lin, A. Cereseto, J. C. Mulloy, J. J. Oshea, G. Franchini, and W. J. Leonard. 1995. Constitutively activated Jak-STAT pathway in T cells transformed with HTLV-I. Science 269:79-81.

Miyake, H., T. Suzuki, H. Hirai, and M. Yoshida. 1999. Trans-activator Tax of human T-cell leukemia virus type 1 enhances mutation frequency of the cellular genome. Virology 253:155-161.

Miyatake, S., M. Seiki, R. D. Malefijt, T. Heike, J. Fujisawa, Y. Takebe, J. Nishida, J. Shlomai, T. Yokota, M. Yoshida, and . 1988. Activation of T cell-derived lymphokine genes in T cells and fibroblasts: effects of human T cell leukemia virus type I p40x protein and bovine papilloma virus encoded E2 protein 5266. Nucleic Acids Res. 16:6547-6566.

Miyoshi, I., I. Kubonishi, M. Sumida, S. Hiraki, T. Tsubota, I. Kimura, K. Miyamoto, and J. Sato. 1980. A novel T-cell line derived from adult T-cell leukemia. Gann 71:155-156.

Miyoshi, I., I. Kubonishi, S. Yoshimoto, T. Akagi, Y. Ohtsuki, Y. Shiraishi, K. Nagata, and Y. Hinuma. 1981. Type C virus particles in a cord T-cell line derived by co-cultivating normal human cord leukocytes and human leukaemic T cells. Nature 294:770-771.

Miyoshi, I., I. Kubonishi, S. Yoshimoto, and Y. Shiraishi. 1981. A T-cell line derived from normal human cord leukocytes by co-culturing with human leukemic T-cells. Gann 72:978-981.

Miyoshi, I., S. Yoshimoto, I. Kubonishi, H. Taguchi, Y. Shiraishi, Y. Ohtsuki, and T. Akagi. 1981. Transformation of normal human cord lymphocytes by co-cultivation with a lethally irradiated human T-cell line carrying type C virus particles. Gann 72:997-998.

Mizokami, T., K. Okamura, T. Kohno, K. Sato, H. Ikenoue, T. Kuroda, K. Inokuchi, and M. Fujishima. 1995. Human T-lymphotropic virus type I-associated uveitis in patients with Graves' disease treated with methylmercaptoimidazole 5231. J. Clin. Endocrinol. Metab 80:1904-1907.

Mochizuki, M., T. Watanabe, K. Yamaguchi, K. Takatsuki, K. Yoshimura, M. Shirao, S. Nakashima, S. Mori, S. Araki, and N. Miyata. 1992. HTLV-I Uveitis: A distinct clinical entity caused by HTLV-I. Jpn. J. Cancer Res. 83:236-239. 233 Mori, N., N. Mukaida, D. W. Ballard, K. Matsushima, and N. Yamamoto. 1998. Human T-cell leukemia virus type I Tax transactivates human interleukin 8 gene through acting concurrently on AP- 1 and nuclear factor-kappa B-like sites. Cancer Res 58:3993- 4000.

Mori, N. and D. Prager. 1996. Transactivation of the interleukin-1 alpha promoter by human T- cell leukemia virus type I and type II tax proteins. Blood 87:3410-3417.

Mori, N., A. Ueda, S. Ikeda, Y. Yamasaki, Y. Yamada, M. Tomonaga, S. Morikawa, R. Geleziunas, T. Yoshimura, and N. Yamamoto. 2000. Human T-cell leukemia virus type I tax activates transcription of the human monocyte chemoattractant protein-1 gene through two nuclear factor-kappaB sites. Cancer Res. 60:4939-4945.

Mortreux, F., A. S. Gabet, and E. Wattel. 2003. Molecular and cellular aspects of HTLV-1 associated leukemogenesis in vivo. Leukemia 17:26-38.

Moura, I. C., Y. Lepelletier, B. Arnulf, P. England, C. Baude, C. Beaumont, A. Bazarbachi, M. Benhamou, R. C. Monteiro, and O. Hermine. 2003. A neutralizing monoclonal antibody (mAb A24) directed against the transferrin receptor induces apoptosis of tumor T lymphocytes from ATL patients. Blood.

Mu, J. J., Y. G. Tsay, L. J. Juan, T. F. Fu, W. H. Huang, D. S. Chen, and P. J. Chen. 2004. The small delta antigen of hepatitis delta virus is an acetylated protein and acetylation of lysine 72 may influence its cellular localization and viral RNA synthesis. Virology 319:60-70.

Muchardt, C., J. S. Seeler, A. Nirula, S. Gong, and R. Gaynor. 1992. Transcription factor AP-2 activates gene expression of HTLV-I. EMBO Journal. 11:2573-2581.

Mueller, N., A. Okayama, S. Stuver, and N. Tachibana. 1996. Findings from the Miyazaki Cohort Study. J Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S2-S7.

Mulloy, J. C., R. W. Crownley, J. Fullen, W. J. Leonard, and G. Franchini. 1996. The human T-cell leukemia/lymphotropic virus type 1 p12I proteins bind the interleukin- 2 receptor beta and gammac chains and affects their expression on the cell surface. J Virol 70:3599-3605.

Mulloy, J. C., T. Kislyakova, A. Cereseto, L. Casareto, A. Lomonico, J. Fullen, M. V. Lorenzi, A. Cara, C. Nicot, C. Giam, and G. Franchini. 1998. Human T-cell lymphotropic/leukemia virus type 1 Tax abrogates p53- induced cell cycle arrest and apoptosis through its CREB/ATF functional domain. J. Virol. 72:8852-8860.

Murakami, T., H. Hirai, T. Suzuki, J. I. Fujisawa, and M. Yoshida. 1995. HTLV-1 tax enhances NF-kappa B2 expression and binds to the products p52 and p100, but does not suppress the inhibitory function of p100. Virology 206:1066-1074.

234 Muraoka, O., T. Kaisho, M. Tanabe, and T. Hirano. 1993. Transcriptional activation of the interleukin-6 gene by HTLV-1. Immunology Letters. 37(2-3):159-165.

Murphy, D. 2002. Gene expression studies using microarrays: principles, problems, and prospects. Adv. Physiol Educ. 26:256-270.

Murphy, E. L., M. P. Busch, M. Tong, P. Cornett, and G. N. Vyas. 1998. A prospective study of the risk of transfusion-acquired viral infections. Transfus. Med. 8:173-178.

Murphy, E. L., J. P. Figueroa, W. N. Gibbs, A. Brathwaite, M. Holding-Cobham, D. Waters, B. Cranston, B. Hanchard, and W. A. Blattner. 1989. Sexual transmission of human Y-lymphotropic virus type I (HTLV-I). Ann. Intern. Med. 111:555-560.

Murphy, E. L., S. A. Glynn, J. Fridey, J. W. Smith, R. A. Sacher, C. C. Nass, H. E. Ownby, D. J. Wright, and G. J. Nemo. 1999. Increased incidence of infectious diseases during prospective follow-up of human T-lymphotropic virus type II- and I-infected blood donors. ARCH INTERN MED 159:1485-1491.

Nagai, M. and M. Osame. 2003. Human T-cell lymphotropic virus type I and neurological diseases. J Neurovirol. 9:228-235.

Nagai, M., K. Usuku, W. Matsumoto, D. Kodama, N. Takenouchi, T. Moritoyo, S. Hashiguchi, M. Ichinose, C. R. Bangham, S. Izumo, and M. Osame. 1998. Analysis of HTLV-I proviral load in 202 HAM/TSP patients and 243 asymptomatic HTLV-I carriers: high proviral load strongly predisposes to HAM/TSP. J Neurovirol. 4:586-593.

Nam, S. H., T. D. Copeland, M. Hatanaka, and S. Oroszlan. 1993. Characterization of ribosomal frameshifting for expression of pol gene products of human T-cell leukemia virus type I. J. Virol. 67:196-203.

Nam, S. H., M. Kidokoro, H. Shida, and M. Hatanaka. 1988. Processing of gag Precursor Polyprotein of Human T-Cell Leukemia Virus Type 1 by Virus-Encoded Protease. J. Virol. 62:3718-3728.

Nemorin, J. G. and P. Duplay. 2000. Evidence that Llck-mediated phosphorylation of p56dok and p62dok may play a role in CD2 signaling. J Biol. Chem. 275:14590-14597.

Neuveut, C. and K. T. Jeang. 2002. Cell cycle dysregulation by HTLV-I: role of the tax oncoprotein. Front Biosci. 7:157-163.

Newbound, G. C., J. M. Andrews, J. Orourke, J. N. Brady, and M. D. Lairmore. 1996. Human T-cell lymphotropic virus type 1 tax mediates enhanced transcription in CD4(+) T lymphocytes. J. Virol. 70:2101-2106.

235 Newbound, G. C., J. P. O'Rourke, N. D. Collins, J. M. Andrews, J. Dewille, and M. D. Lairmore. 2000. Repression of tax-mediated human t-lymphotropic virus type 1 transcription by inducible cAMP early repressor (ICER) protein in peripheral blood mononuclear cells. J. Med. Virol. 62:286-292.

Ng, P. W., H. Iha, Y. Iwanaga, M. Bittner, Y. Chen, Y. Jiang, G. Gooden, J. M. Trent, P. Meltzer, K. T. Jeang, and S. L. Zeichner. 2001. Genome-wide expression changes induced by HTLV-1 Tax: evidence for MLK- 3 mixed lineage kinase involvement in Tax-mediated NF-kappaB activation. Oncogene 20:4484-4496.

Nicot, C., M. Dundr, J. M. Johnson, J. R. Fullen, N. Alonzo, R. Fukumoto, G. L. Princler, D. Derse, T. Misteli, and G. Franchini. 2004. HTLV-1-encoded p30(II) is a post-transcriptional negative regulator of viral replication. Nat. Med.

Nicot, C., R. Mahieux, R. Opavsky, A. Cereseto, L. Wolff, J. N. Brady, and G. Franchini. 2000. HTLV-I Tax transrepresses the human c-Myb promoter independently of its interaction with CBP or p300. Oncogene 19:2155-2164.

Nicot, C., R. Mahieux, C. Pise-Masison, J. Brady, A. Gessain, S. Yamaoka, and G. Franchini. 2001. Human T-Cell Lymphotropic Virus Type 1 Tax Represses c-Myb- Dependent Transcription through Activation of the NF-kappaB Pathway and Modulation of Coactivator Usage. Mol. Cell Biol. 21:7391-7402.

Nicot, C., R. Mahieux, S. Takemoto, and G. Franchini. 2000. Bcl-X(L) is up-regulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL samples. Blood 96:275-281.

Nicot, C., J. C. Mulloy, M. G. Ferrari, J. M. Johnson, K. Fu, R. Fukumoto, R. Trovato, J. Fullen, W. J. Leonard, and G. Franchini. 2001. HTLV-1 p12(I) protein enhances STAT5 activation and decreases the interleukin-2 requirement for proliferation of primary human peripheral blood mononuclear cells. Blood 98:823-829.

Nishioka, K., I. Maruyama, K. Sato, I. Kitajima, Y. Nakajima, and M. Osame. 1989. Chronic inflammatory arthropathy associated with HTLV-I [letter]. Lancet i:441.

Nitta, T., M. Tanaka, B. Sun, S. Hanai, and M. Miwa. 2003. The genetic background as a determinant of human T-cell leukemia virus type 1 proviral load. Biochem. Biophys. Res. Commun. 309:161-165.

Nosaka, T., H. Siomi, Y. Adachi, M. Ishibashi, S. Kubota, M. Maki, and M. Hatanaka. 1989. Nucleolar targeting signal of human T-cell leukemia virus type I rex- encoded protein is essential for cytoplasmic accumulation of unspliced viral mRNA 5295. Proc. Natl. Acad. Sci. U. S. A 86:9798-9802.

Nyborg, J. K. and W. S. Dynan. 1990. Interaction of cellular proteins with the human T-cell leukemia virus type I transcriptional control region. Purification of cellular proteins that bind the 21-base pair repeat elements 5257. J. Biol. Chem. 265:8230-8236. 236 Nyborg, J. K., M. H. Matthews, W. T. Golde, and W. S. Dynan. 1995. The identification of Cellular Proteins that Interact with the HTLV-I LTR, p. 131-140. Human Retroviruses. Alan R. Liss, Inc..

Ohshima, K., M. Kikuchi, Y. Masuda, Y. Sumiyoshi, F. Eguchi, H. Mohtai, M. Takeshita, and N. Kimura. 1992. Human T-cell leukemia virus type I associated lymphadenitis. Cancer 69:239-248.

Ohta, M., H. Nagai, H. Nyunoya, and K. Shimotohno. 1996. Src-homology domain 2 is responsible for transcriptional suppression induced by expression of Lck. Oncogene 12:989-997.

Okochi, K., H. Sato, and Y. Hinuma. 1984. A retrospective study on transmission of adult T cell leukemia virus by blood transfusion: seroconversion in recipients. Vox Sang. 46:245-253.

Ono, A., E. Ikeda, M. Mochizuki, M. Matsuoka, K. Yamaguchi, T. Sawada, S. Yamane, S. Tokudome, and T. Watanabe. 1998. Provirus load in patients with human T-cell leukemia virus type 1 uveitis correlates with precedent Graves' disease and disease activities. Jpn J Cancer Res 89:608-614.

Ono, A., M. Mochizuki, K. Yamaguchi, N. Miyata, and T. Watanabe. 1997. Immunologic and virologic characterization of the primary infiltrating cells in the aqueous humor of human T-cell leukemia virus type-1 uveitis. Accumulation of the human T-cell leukemia virus type-1- infected cells and constitutive expression of viral and interleukin-6 messenger ribonucleic acids. Invest Ophthalmol. Vis. Sci. 38:676-689.

Ootsuyama, Y., K. Shimotohno, M. Miwa, S. Oroszlan, and T. Sugimura. 1985. Myristylation of gag protein in human T-cell leukemia virus type-I and type-II. Jpn. J Cancer Res. 76:1132-1135.

Orita, S., A. Saiga, S. Takagi, T. Tanaka, K. Okumura, Y. Aono, Y. Hinuma, and H. Igarashi. 1991. A novel alternatively spliced viral mRNA transcribed in cells infected with human T cell leukemia virus type 1 is mainly responsible for expressing p21X protein. FEBS Lett. 295:127-134.

Orita, S., S. Sato, Y. Aono, N. Minoura, T. Yamashita, Y. Hinuma, and H. Igarashi. 1993. Identification of novel singly spliced pX mRNA transcripts common to. Virus Genes 7(2):197-204.

Orita, S., S. Takagi, A. Saiga, N. Minoura, K. Araki, K. I. Kinoshita, T. Kondo, Y. Hinuma, and H. Igarashi. 1992. Human T cell leukaemia virus type 1 p21X mRNA: constitutive expression peripheral blood mononuclear cells of patients with adult T cell leukaemia. J. Gen. Virol. 73:2283-2289.

237 Oroszlan, S. 1984. Chemical Analysis of HTLV-1 Structural proteins, p. 101-110. In R. C. Gallo, M. E. Essex, and L. Gross (eds.), Human T-cell Leukemia Lymphoma Viruses. Cold Spring Harbor Laboratory Press.

Osame, M. 2002. Pathological mechanisms of human T-cell lymphotropic virus type I- associated myelopathy (HAM/TSP). J Neurovirol. 8:359-364.

Osame, M., R. Janssen, H. Kubota, H. Nishitani, A. Igata, S. Nagataki, M. Mori, I. Goto, H. Shimabukuro, R. Khabbaz, and J. Kaplan. 1990. Nationwide survey of HTLV-I-associated Myelopathy in Japan: Association with blood transfusion. Ann. Neurol. 28:50-56.

Osame, M., K. Usuku, S. Izumo, N. Ijichi, H. Amitani, A. Igata, M. Matsumoto, and M. Tara. 1986. HTLV-I associated myelopathy, a new clinical entity. Lancet i:1031- 1032.

Ott, M., M. Schnolzer, J. Garnica, W. Fischle, S. Emiliani, H. R. Rackwitz, and E. Verdin. 1999. Acetylation of the HIV-1 Tat protein by p300 is important for its transcriptional activity. Curr. Biol. 9:1489-1492.

Overbaugh, J. 2004. HTLV-1 sweet-talks its way into cells. Nat. Med. 10:20-21.

Ozawa, Y., M. Migita, T. Watanabe, I. Okuda, A. Takeshita, A. Takagi, and Y. Shishiba. 1994. Development of Graves' ophthalmopathy and uveitis after radioiodine therapy for Graves' disease in a patient with HTLV-I associated myelopathy (HAM) 5230. Intern. Med. 33:564-568.

Pancake, B. A., D. Z. Franklin, and E. E. Coutavas. 1995. The cutaneous T cell lymphoma, mycosis fungoides, is a human T cell lymphotropic virus-associated disease - A study of 50 patients. J. Clin. Invest. 95:547-554.

Peng, Y. C., D. E. Breiding, F. Sverdrup, J. Richard, and E. J. Androphy. 2000. AMF-1/Gps2 binds p300 and enhances its interaction with papillomavirus E2 proteins. J. Virol. 74:5872-5879.

Perini, G., S. Wagner, and M. R. Green. 1995. Recognition of bZIP proteins by the human T-cell leukaemia virus transactivator Tax. Nature 376:602-605.

Perkins, N. D., L. K. Felzien, J. C. Betts, K. Leung, D. H. Beach, and G. J. Nabel. 1997. Regulation of NF-kappaB by cyclin-dependent kinases associated with the p300 coactivator. Science 275:523-527.

Petit, C., F. Buseyne, C. Boccaccio, J. P. Abastado, J. M. Heard, and O. Schwartz. 2001. Nef is required for efficient hiv-1 replication in cocultures of dendritic cells and lymphocytes. Virology 286:225-236.

238 Petrij, F., R. H. Giles, H. G. Dauwerse, J. J. Saris, R. C. Hennekam, M. Masuno, N. Tommerup, G. J. van Ommen, R. H. Goodman, and D. J. Peters. 1995. Rubinstein- Taybi syndrome caused by mutations in the transcriptional co- activator CBP [see comments]. Nature 376:348-351.

Petrini, M., E. Pelosi-Testa, N. M. Sposi, G. Mastroberardino, A. Camagna, L. Bottero, F. Mavilio, U. Testa, and C. Peschle. 1989. Constitutive expression and abnormal glycosylation of transferrin receptor in acute T-cell leukemia. Cancer Res. 49:6989-6996.

Philpott, S. M. and G. C. Buehring. 1999. Defective DNA repair in cells with human T- cell leukemia/bovine leukemia viruses: role of tax gene. J Natl. Cancer Inst. 91:933-942.

Piguet, V., O. Schwartz, S. Legall, and D. Trono. 1999. The downregulation of CD4 and MHC-I by primate lentiviruses: a paradigm for the modulation of cell surface receptors. Immunol Rev 168:51-63.

Pinton, P., D. Ferrari, E. Rapizzi, F. D. Di Virgilio, T. Pozzan, and R. Rizzuto. 2001. The Ca2+ concentration of the endoplasmic reticulum is a key determinant of ceramide- induced apoptosis: significance for the molecular mechanism of Bcl-2 action. EMBO J 20:2690-2701.

Pique, C., D. Pham, T. Tursz, and M. C. Dokhelar. 1992. Human T-Cell Leukemia Virus Type I Envelope Protein Maturation Process: Requirements for Syncytium Formation. J. Virol. 66:906-913.

Pique, C., A. Uretavidal, A. Gessain, B. Chancerel, O. Gout, R. Tamouza, F. Agis, and M. C. Dokhelar. 2000. Evidence for the chronic in vivo production of human T cell leukemia virus type I Rof and Tof proteins from cytotoxic T lymphocytes directed against viral peptides. J. Exp. Med. 191:567-572.

Pise-Masison, C. A., R. Mahieux, H. Jiang, M. Ashcroft, M. Radonovich, J. Duvall, C. Guillerm, and J. N. Brady. 2000. Inactivation of p53 by human T-cell lymphotropic virus type 1 Tax requires activation of the NF-kappaB pathway and is dependent on p53 phosphorylation. Mol. Cell Biol. 20:3377-3386.

Pise-Masison, C. A., R. Mahieux, M. Radonovich, H. Jiang, and J. N. Brady. 2001. Human T-lymphotropic virus type I Tax protein utilizes distinct pathways for p53 inhibition that are cell type-dependent. J. Biol. Chem. 276:200-205.

Pise-Masison, C. A., R. Mahieux, M. Radonovich, H. Jiang, J. Duvall, C. Guillerm, and J. N. Brady. 2000. Insights into the molecular mechanism of p53 inhibition by HTLV type 1 Tax. AIDS Res. Hum. Retroviruses 16:1669-1675.

239 Pise-Masison, C. A., M. Radonovich, R. Mahieux, P. Chatterjee, C. Whiteford, J. Duvall, C. Guillerm, A. Gessain, and J. N. Brady. 2002. Transcription profile of cells infected with human T-cell leukemia virus type I compared with activated lymphocytes. Cancer Res. 62:3562-3571.

Pisemasison, C. A., M. Radonovich, K. Sakaguchi, E. Appella, and J. N. Brady. 1998. Phosphorylation of p53: a novel pathway for p53 inactivation in human T-cell lymphotropic virus type 1- transformed cells. J Virol 72:6348-6355.

Poiesz, B. J., F. W. Ruscetti, A. F. Gazdar, P. A. Bunn, J. D. Minna, and R. C. Gallo. 1980. Detection and isolation of type C retrovirus particles from fresh and cultured lymphocytes of a patient with cutaneous T-cell lymphoma. Proc. Natl. Acad. Sci. USA 77:7415-7419.

Pollard, V. W. and M. H. Malim. 1998. The HIV-1 Rev protein. Annu. Rev. Microbiol. 52:491-532.

Pombo de Oliveira, M. S., E. Matutes, T. Schulz, S. M. Carvalho, H. Noronha, J. D. Reaves, P. Loureiro, C. Machado, and D. Catovsky. 1995. T-cell malignancies in Brazil. Clinico-pathological and molecular studies of HTLV-I-positive and -negative cases. Int. J Cancer 60:823-827.

Princler, G. L., J. G. Julias, S. H. Hughes, and D. Derse. 2003. Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology 317:136-145.

Puri, P. L., V. Sartorelli, X. J. Yang, Y. Hamamori, V. V. Ogryzko, B. H. Howard, L. Kedes, J. Y. Wang, A. Graessmann, Y. Nakatani, and M. Levrero. 1997. Differential roles of p300 and PCAF acetyltransferases in muscle differentiation. Mol. Cell 1:35-45.

Pusl, T., J. J. Wu, T. L. Zimmerman, L. Zhang, B. E. Ehrlich, M. W. Berchtold, J. B. Hoek, S. J. Karpen, M. H. Nathanson, and A. M. Bennett. 2002. Epidermal growth factor-mediated activation of the ETS domain transcription factor Elk-1 requires nuclear calcium. J Biol. Chem. 277:27517-27527.

Ratner, L. 1989. Regulation of expression of the c-sis proto-oncogene. Nucleic Acids Res. 17:4101-4115.

Reeves, W., C. Saxinger, M. Brenes, E. Quinoz, M. Hoh, J. Clark, C. Holt, R. C. Gallo, and W. Blattner. 1988. Human T-cell lymphotropic virus type I (HTLV-I) and risk factors in metropolitan Panama. Am. J. Epidemiol. 127:539.

Reid, J. L., A. J. Bannister, P. Zegerman, M. A. Martinez-Balbas, and T. Kouzarides. 1998. E1A directly binds and regulates the P/CAF acetyltransferase. EMBO J 17:4469-4477.

240 Reid, R. L., P. F. Lindholm, A. Mireskandari, J. Dittmer, and J. N. Brady. 1993. Stabilization of wild-type p53 in human T-lymphocytes transformed by HTLV-I. Oncogene 8:3029-3036.

Ressler, S., G. F. Morris, and S. J. Marriott. 1997. Human T-cell leukemia virus type 1 Tax transactivates the human proliferating cell nuclear antigen promoter. J. Virol. 71:1181-1190.

Rimsky, L., J. Hauber, M. Dukovich, M. H. Malim, A. Langlois, B. R. Cullen, and W. C. Greene. 1988. Functional replacement of the HIV-1 rev protein by the HTLV-1 rex protein. Nature 335:738-740.

Robek, M. D., F. H. Wong, and L. Ratner. 1998. Human T-Cell leukemia virus type 1 pX-I and pX-II open reading frames are dispensable for the immortalization of primary lymphocytes. J. Virol. 72:4458-4462.

Roithmann, S., C. Pique, A. Le Cesne, L. Delamarre, D. Pham, T. Tursz, and M. C. Dokhelar. 1994. The open reading frame I (ORF I)/ORF II part of the human T-cell leukemia virus type 1 X region is dispensable for p40tax, p27rex, or envelope expression. J. Virol. 68:3448-3451.

Rosenberg, J. R., L. Delamarre, C. Pique, D. Pham, and M. C. Dokhelar. 1997. The ectodomain of the human T-cell leukemia virus type 1 TM glycoprotein is involved in postfusion events. J. Virol. 71:7180-7186.

Sagawa, K., M. Mochizuki, K. Masuoka, K. Katagiri, T. Katayama, T. Maeda, A. Tanimoto, S. Sugita, T. Watanabe, and K. Itoh. 1995. Immunopathological mechanisms of human T cell lymphotropic virus type 1 (HTLV-I) uveitis - Detection of HTLV-I- infected T cells in the eye and their constitutive cytokine production. J. Clin. Invest. 95:852-858.

Saito, M., Y. Furukawa, R. Kubota, K. Usuku, S. Sonoda, S. Izumo, M. Osame, and M. Yoshida. 1995. Frequent mutation in pX region of HTLV-1 is observed in HAM/TSP patients, but is not specifically associated with the central nervous system lesions. J. Neurovirol. 1:286-294.

Sakai, J. A., M. Nagai, M. B. Brennan, C. A. Mora, and S. Jacobson. 2001. In vitro spontaneous lymphoproliferation in patients with human T-cell lymphotropic virus type I-associated neurologic disease: predominant expansion of CD8+ T cells. Blood 98:1506- 1511.

Saxinger, W. C., W. Blattner, P. Levine, J. Clark, R. J. Biggar, M. Hoh, J. Moghissi, P. Jacobs, L. Wilson, R. Jacobson, R. Crookes, M. Strong, A. A. Ansari, A. Dean, F. Nkrumah, N. Mourali, and R. C. Gallo. 1984. Human T-cell leukemia virus (HTLV-I) antibodies in Africa. Science 225:1473.

241 Scheffner, M. and N. J. Whitaker. 2003. Human papillomavirus-induced carcinogenesis and the ubiquitin-proteasome system. Semin. Cancer Biol. 13:59-67.

Schiltz, R. L., C. A. Mizzen, A. Vassilev, R. G. Cook, C. D. Allis, and Y. Nakatani. 1999. Overlapping but distinct patterns of histone acetylation by the human coactivators p300 and PCAF within nucleosomal substrates. J Biol. Chem. 274:1189-1192.

Schreiber, G. B., E. L. Murphy, J. A. Horton, D. J. Wright, R. Garfein, H. C. Chien, and C. C. Nass. 1997. Risk factors for human T-cell lymphotropic virus types I and II (HTLV-I and -II) in blood donors: the Retrovirus Epidemiology Donor Study. NHLBI Retrovirus Epidemiology Donor Study. J Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 14:263-271.

Seiki, M., R. Eddy, T. B. Shows, and M. Yoshida. 1984. Nonspecific integration of the HTLV provirus genome into adult T-cell leukaemia cells. Nature 309:640-642.

Seiki, M., S. Hattori, Y. Hirayama, and M. Yoshida. 1983. Human adult T-cell leukemia virus: Complete nucleotide sequence of the provirus genome integrated in leukemia cell DNA. Proc. Natl. Acad. Sci. 80:3618-3622.

Seiki, M., A. Hikikoshi, T. Taniguchi, and M. Yoshida. 1985. Expression of the pX Gene of HTLV-1: General Splicing Mechanism in the HTLV Family. Science 228:1532- 1534.

Seiki, M., J. Inoue, M. Hidaka, and M. Yoshida. 1988. Two cis-acting elements responsible for posttranscriptional trans-regulation of gene expression of human T-cell leukemia virus type I. Proc. Natl. Acad. Sci. U. S. A 85:7124-7128.

Shikama, N., J. Lyon, and N. B. Lathangue. 1997. The p300/CBP family: Integrating signals with transcription factors and chromatin. Tr. Cell Biol. 7:230-236.

Shimotohno, K., M. Takano, T. Teruuchi, and M. Miwa. 1986. Requirement of multiple copies of a 21-nucleotide sequence in the U3 regions of human T-cell leukemia virus type I and type II long terminal repeats for trans-acting activation of transcription 5252. Proc. Natl. Acad. Sci. U. S. A 83:8112-8116.

Shimoyama, M. 1991. Diagnostic criteria and classification of clinical subtypes of adult T- cell leukaemia-lymphoma. A report from the Lymphoma Study Group (1984- 87) 5222. Br. J. Haematol. 79:428-437.

Silverman, L. R., A. J. Phipps, A. Montgomery, L. Ratner, and M. D. Lairmore. 2004. Human T-cell lymphotropic virus type 1 open reading frame II-encoded p30II is required for in vivo replication: evidence of in vivo reversion. J Virol 78:3837-3845.

Siomi, H., H. Shida, S. H. Nam, T. Nosaka, M. Maki, and M. Hatanaka. 1988. Sequence requirements for nucleolar localization of human T cell leukemia virus type I pX protein, which regulates viral RNA processing. Cell 55:197-209. 242 Snowden, A. W. and N. D. Perkins. 1998. Cell cycle regulation of the transcriptional coactivators p300 and CREB binding protein. Biochem. Pharmacol. 55:1947-1954.

Sodroski, J., C. Rosen, W. C. Goh, and W. Haseltine. 1985. A transcriptional activator protein encoded by the x-lor region of the human T-cell leukemia virus. Science 228:1430-1434.

Sodroski, J. G., C. A. Rosen, and W. A. Haseltine. 1984. Trans-Acting Transcriptional Activation of the Long Terminal Repeat of Human T Lymphotropic Viruses in Infected Cells. Science 223:381-385.

Song, K. J., V. R. Nerurkar, N. Saitou, A. Lazo, J. R. Blakeslee, I. Miyoshi, and R. Yanagihara. 1994. Genetic Analysis and Molecular Phylogeny of Simian T-Cell Lymphotropic Virus Type I: Evidence for Independent Virus Evolution in Asia and Africa. Virology 199:56-66.

Sterner, D. E. and S. L. Berger. 2000. Acetylation of histones and transcription-related factors. Microbiol. Mol. Biol. Rev. 64:435-459.

Stuver, S. O., N. Tachibana, A. Okayama, S. Shioiri, Y. Tsunetoshi, K. Tsuda, and N. E. Mueller. 1993. Heterosexual transmission of human T cell leukemia/lymphoma virus type I among married couples in southwestern Japan: an initial report from the Miyazaki Cohort Study. J. Infect. Dis. 167:57-65.

Sugimoto, M., S. Mita, M. Tokunaga, K. Yamaguchi, I. Cho, M. Matsumoto, M. Mochizuki, S. Araki, K. Takatsuki, and M. Ando. 1993. Pulmonary involvement in human T-cell lymphotropic virus type-I uveitis: T-lymphocytosis and high proviral DNA load in bronchoalveolar lavage fluid. Eur. Respir. J. 6:938-943.

Sumida, T., F. Yonaha, T. Maeda, Y. Kita, I. Iwamoto, T. Koike, and S. Yoshida. 1994. Expression of Sequences Homologous to HTLV-I Tax Gene in the Labial Salivary Glands of Japanese Patients with Sjogrens Syndrome. Arthritis Rheum. 37:545-550.

Sun, S. C., E. W. Harhaj, G. Xiao, and L. Good. 2000. Activation of I-kappaB kinase by the HTLV type 1 tax protein: mechanistic insights into the adaptor function of IKKgamma [In Process Citation]. AIDS Res. Hum. Retroviruses 16:1591-1596.

Sun, S. C., E. W. Harhaj, G. Xiao, and L. Good. 2000. Activation of I-kappaB kinase by the HTLV type 1 tax protein: mechanistic insights into the adaptor function of IKKgamma [In Process Citation]. AIDS Res. Hum. Retroviruses 16:1591-1596.

Suzuki, T., J. Fujisawa, M. Toita, and M. Yoshida. 1993. The trans-activator Tax of human T-cell leukemia virus type 1 (HTLV-1) interacts with cAMP-responsive element (CRE) binding and CRE modulator proteins that bind to the 21-base-pair enhancer of HTLV-1. Proc. Natl. Acad. Sci. U. S. A. 90:610-614.

243 Suzuki, T., H. Hirai, J. Fujisawa, T. Fujita, and M. Yoshida. 1993. A trans-activator Tax of human T-cell leukemia virus type 1 binds to NF- kappa B p50 and serum response factor (SRF) and associates with enhancer DNAs of the NF-kappa B site and CArG box 5014. Oncogene 8:2391-2397.

Suzuki, T., H. Hirai, T. Murakami, and M. Yoshida. 1995. Tax protein of HTLV-1 destabilizes the complexes of NF- kappa B and I kappa B-alpha and induces nuclear translocation of NF-kappa B for transcriptional activation. Oncogene 10:1199-1207.

Suzuki, T., H. Hirai, and M. Yoshida. 1994. Tax protein of HTLV-1 interacts with the Rel homology domain of NF-kappa B p65 and c-Rel proteins bound to the NF-kappa B binding site and activates transcription. Oncogene 9:3099-3105.

Suzuki, T., S. Kitao, H. Matsushime, and M. Yoshida. 1996. HTLV-1 Tax protein interacts with cyclin-dependent kinase inhibitor p16(INK4A) and counteracts its inhibitory activity towards CDK4. EMBO J. 15:1607-1614.

Suzuki, T., T. Narita, M. Uchidatoita, and M. Yoshida. 1999. Down-regulation of the INK4 family of cyclin-dependent kinase inhibitors by tax protein of HTLV-1 through two distinct mechanisms. Virology 259:384-391.

Suzuki, T., M. Uchida-Toita, T. Andoh, and M. Yoshida. 2000. HTLV-1 tax oncoprotein binds to DNA topoisomerase I and inhibits its catalytic activity. Virology 270:291-298.

Suzuki, T., M. Uchidatoita, and M. Yoshida. 1999. Tax protein of HTLV-1 inhibits CBP/p300-mediated transcription by interfering with recruitment of CBP/p300 onto DNA element of E-box or p53 binding site. Oncogene 18:4137-4143.

Takatsuki, K., T. Uchiyama, K. Sagawa, and J. Yodoi. 1977. Adult T-cell leukemia in Japan, p. 73-77. In S. Seno, F. Takaku, and S. Irino (eds.), Topics in Hematology. Excerpta Medica, Amsterdam.

Takeshita, A., G. R. Cardona, N. Koibuchi, C. S. Suen, and W. W. Chin. 1997. TRAM-1, A novel 160-kDa thyroid hormone receptor activator molecule, exhibits distinct properties from steroid receptor coactivator-1. J Biol. Chem. 272:27629-27634.

Tan, T. H., R. Jia, and R. G. Roeder. 1989. Utilization of Signal Transduction Pathway by the Human T-Cell Leukemia Virus Type 1 Transcriptional Activator tax. J. Virol. 63:3761-3768.

Tanimura, A., H. Teshima, J. I. Fujisawa, and M. Yoshida. 1993. A New Regulatory Element that Augments the Tax-Dependent Enhancer of Human T-Cell Leukemia Virus Type 1 and Cloning of cDNAs Encoding Its Binding Proteins. J. Virol. 67:5375-5382.

244 Terada, K., S. Katamine, K. Eguchi, R. Moriuchi, M. Kita, H. Shimada, I. Yamashita, K. Iwata, Y. Tsuji, S. Nagataki, and T. Miyamoto. 1994. Prevalence of serum and salivary antibodies to HTLV-1 in Sjogren's syndrome. Lancet 344:1116-1119.

Teshigawara, K., M. Maeda, K. Nishino, T. Nikaido, T. Uchiyama, M. Tsudo, Y. Wano, and J. Yodoi. 1985. Adult T leukemia cells produce a lymphokine that augments interleukin 2 receptor expression. J Mol. Cell Immunol. 2:17-26.

Thomas, S., R. Coffin, P. Watts, G. Gough, and D. S. Latchman. 1998. The TAATGARAT motif in the herpes simplex virus immediate-early gene promoter can confer both positive and negative responses to cellular octomer-binding proteins when it is located within the viral genome. J. Virol. 72:3495-3500.

Tillmann, M., R. Wessner, and B. Wigdahl. 1994. Identification of human T-cell lymphotropic virus type I 21-base- pair repeat-specific and glial cell-specific DNA- protein complexes. J. Virol. 68:4597-4608.

Tommerup, N. and H. Vissing. 1995. Isolation and fine mapping of 16 novel human zinc finger-encoding cDNAs identify putative candidate genes for developmental and malignant disorders. Genomics 27:259-264.

Torgeman, A., N. Morvaknin, and M. Aboud. 1999. Sp1 is involved in a protein kinase C-independent activation of human T cell leukemia virus type I long terminal repeat by 12-O-tetradecanoylphorbol-13-acetate. Virology 254:279-287.

Trovato, R., J. C. Mulloy, J. M. Johnson, S. Takemoto, M. P. de Oliveira, and G. Franchini. 1999. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic/leukemia virus type 1 p12(I) protein greatly influences its stability. J Virol 73:6460-6467.

Tsukahara, T., M. Kannagi, T. Ohashi, H. Kato, M. Arai, G. Nunez, Y. Iwanaga, N. Yamamoto, K. Ohtani, M. Nakamura, and M. Fujii. 1999. Induction of Bcl-x(L) expression by human T-cell leukemia virus type 1 tax through NF-kappa B in apoptosis- resistant T-cell transfectants with tax. J Virol 73:7981-7987.

Uchiyama, T., J. Yodoi, K. Sagawa, K. Takatsuki, and H. Uchino. 1977. Adult T-cell leukemia: clinical and hematologic features of 16 cases. Blood 50:481-492.

Valentin, H., I. Lemasson, S. Hamaia, H. Casse, S. Konig, C. Devaux, and L. Gazzolo. 1997. Transcriptional activation of the vascular cell adhesion molecule-1 gene in T lymphocytes expressing human T-Cell leukemia virus type 1 Tax protein. J. Virol. 71:8522-8530.

Van Dooren, S., M. Salemi, and A. M. Vandamme. 2001. Dating the origin of the African human T-cell lymphotropic virus type-i (HTLV-I) subtypes. Mol. Biol. Evol. 18:661-671.

245 Van Lint, J., A. Rykx, Y. Maeda, T. Vantus, S. Sturany, V. Malhotra, J. R. Vandenheede, and T. Seufferlein. 2002. Protein kinase D: an intracellular traffic regulator on the move. Trends Cell Biol. 12:193-200.

Van Orden, K., H. A. Giebler, I. Lemasson, M. Gonzales, and J. K. Nyborg. 1999. Binding of p53 to the KIX domain of CREB binding protein. A potential link to human T-cell leukemia virus, type I-associated leukemogenesis. J. Biol. Chem. 274:26321- 26328.

Vanbrussel, M., M. Salemi, H. F. Liu, J. Gabriels, P. Goubau, J. Desmyter, and A. M. Vandamme. 1998. The simian T-lymphotropic virus STLV-PP1664 from Pan paniscus is distinctly related to HTLV-2 but differs in genomic organization. Virology 243:366-379.

Vandamme, A. M., M. Salemi, and J. Desmyter. 1998. The simian origins of the pathogenic human T-cell lymphotropic virus type I. Trends Microbiol. 6:477-483.

Vandel, L. and D. Trouche. 2001. Physical association between the histone acetyl transferase CBP and a histone methyl transferase. EMBO rep. 2:21-26.

Vaudin, P., R. Delanoue, I. Davidson, J. Silber, and A. Zider. 1999. TONDU (TDU), a novel human protein related to the product of vestigial (vg) gene of Drosophila melanogaster interacts with vertebrate TEF factors and substitutes for Vg function in wing formation. Development 126:4807-4816.

Vo, N. and R. H. Goodman. 2001. CREB-binding protein and p300 in transcriptional regulation. J. Biol. Chem. 276:13505-13508.

Vogel, K. W., R. Briesewitz, T. J. Wandless, and G. R. Crabtree. 2001. Calcineurin inhibitors and the generalization of the presenting protein strategy. Adv. Protein Chem. 56:253-291. von Bulow, G. U. and R. J. Bram. 1997. NF-AT activation induced by a CAML- interacting member of the tumor necrosis factor receptor superfamily. Science 278:138- 141.

Wang, J. K., E. Kiyokawa, E. Verdin, and D. Trono. 2000. The Nef protein of HIV-1 associates with rafts and primes T cells for activation. Proc. Natl. Acad. Sci. U. S. A 97:394-399.

Wang, L., S. R. Grossman, and E. Kieff. 2000. Epstein-Barr virus nuclear protein 2 interacts with p300, CBP, and PCAF histone acetyltransferases in activation of the LMP1 promoter. Proc. Natl. Acad. Sci. U. S. A 97:430-435.

Wattel, E., J. P. Vartanian, C. Pannetier, and S. Wainhobson. 1995. Clonal expansion of human T-Cell leukemia virus type I- infected cells in asymptomatic and symptomatic carriers without malignancy. J. Virol. 69:2863-2868. 246 Weichselbraun, I., J. Berger, M. Dobrovnik, H. Bogerd, R. Grassmann, W. C. Greene, J. Hauber, and E. Bohnlein. 1992. Dominant-negative mutants are clustered in a domain of the human T-cell leukemia virus type I Rex protein: implications for trans dominance. J Virol 66:4540-4545.

Wessner, R., M. Tillmann-Bogush, and B. Wigdahl. 1995. Characterization of a glial cell-specific DNA-protein complex formed with the human T cell lymphotropic virus type I (HTLV-I) enhancer. J Neurovirol. 1:62-77.

Wessner, R. and B. Wigdahl. 1997. AP-1 derived from mature monocytes and astrocytes preferentially interacts with the HTLV-I promoter central 21 bp repeat. Leukemia 11 Suppl 3:21-4.:21-24.

White, J. D., S. L. Zaknoen, C. Kastensportes, L. E. Top, L. Navarroroman, D. L. Nelson, and T. A. Waldmann. 1995. Infectious complications and immunodeficiency in patients with human T-cell lymphotropic virus I-associated adult T- cell leukemia/lymphoma. Cancer 75:1598-1607.

Wiktor, S. Z., E. J. Pate, E. L. Murphy, T. J. Palker, E. Champegnie, A. Ramlal, B. Cranston, B. Hanchard, and W. A. Blattner. 1993. Mother-to-child transmission of human T-cell lymphotropic virus type I (HTLV-I) in Jamaica: association with antibodies to envelope glycoprotein (gp46) epitopes. J. Acquir. Immune. Defic. Syndr. 6:1162-1167.

Wilks, R., B. Hanchard, O. Morgan, E. Williams, B. Cranston, M. L. Smith, P. Rodgersjohnson, and A. Manns. 1996. Patterns of HTLV-I infection among family members of patients with adult T-cell leukemia/lymphoma and HTLV-I associated myelopathy tropical spastic paraparesis. Int. J. Cancer 65:272-273.

Willems, L., P. Kerkhofs, F. Dequiedt, D. Portetelle, M. Mammerickx, A. Burny, and R. Kettmann. 1994. Attenuation of bovine leukemia virus by deletion of R3 and G4 open reading frames. Proc. Natl. Acad. Sci U. S. A 91:11532-11536.

Wilson, A. C., R. W. Freeman, H. Goto, T. Nishimoto, and W. Herr. 1997. VP16 targets an amino-terminal domain of HCF involved in cell cycle progression. Mol. Cell. Biol. 17:6139-6146.

Wolin, M., M. Kornuc, C. Hong, S. K. Shin, F. Lee, R. Lau, and S. Nimer. 1993. Differential effect of HTLV infection and HTLV Tax on interleukin 3 expression. Oncogene 8:1905-1911.

Wood, G. S., A. Salvekar, J. Schaffer, C. F. Crooks, W. Henghold, D. P. Fivenson, Y. H. Kim, and B. R. Smoller. 1996. Evidence against a role for human T-cell lymphotrophic virus type I (HTLV-I) in the pathogenesis of American cutaneous T-cell lymphoma. J. Invest. Dermatol. 107:301-307.

247 Wood, G. S., J. M. Schaffer, R. Boni, R. Dummer, G. Burg, M. Takeshita, and M. Kikuchi. 1997. No evidence of HTLV-I proviral integration in lymphoproliferative disorders associated with cutaneous T- cell lymphoma. Am. J. Pathol. 150:667-673.

Wrzesinski, S., R. Seguin, Y. Liu, S. Domville, V. Planelles, P. Massa, E. Barker, J. Antel, and G. Feuer. 2000. HTLV type 1 Tax transduction in microglial cells and astrocytes by lentiviral vectors. AIDS Res. Hum. Retroviruses 16:1771-1776.

Wu, Y. and J. W. Marsh. 2001. Selective transcription and modulation of resting T cell activity by preintegrated HIV DNA. Science 293:1503-1506.

Wucherpfennig, K. W., P. Hollsberg, J. H. Richardson, D. Benjamin, and D. A. Hafler. 1992. T-cell activation by autologous human T-cell leukemia virus type I- infected T-cell clones. Proc. Natl. Acad. Sci. U. S. A 89:2110-2114.

Wysocka, J. and W. Herr. 2003. The herpes simplex virus VP16-induced complex: the makings of a regulatory switch. Trends Biochem. Sci 28:294-304.

Xie, A. Y., V. P. Bermudez, and W. R. Folk. 2002. Stimulation of DNA replication from the polyomavirus origin by PCAF and GCN5 acetyltransferases: acetylation of large T antigen. Mol. Cell Biol. 22:7907-7918.

Xu, R. Z., Q. K. Gao, S. J. Wang, H. H. Kan, L. Y. Sheng, C. Q. Li, X. H. Zhang, G. B. Xu, and K. Q. Zhang. 1996. Human acute myeloid leukemias may be etiologically associated with new human retroviral infection. Leuk. Res. 20:449-455.

Yajima, A., A. Kawada, Y. Aragane, and T. Tezuka. 2001. Detection of HTLV-I proviral DNA in sarcoidosis. Dermatology 203:53-56.

Yamaguchi, K. 1994. Human T-lymphotropic virus type I in Japan. Lancet 343:213-216.

Yamaguchi, K., H. Nishimura, H. Kohrogi, M. Jono, Y. Miyamoto, and K. Takatsuki. 1983. A proposal for smoldering adult T-cell leukemia: a clinicopathologic study of five cases. Blood 62:758-766.

Yamaguchi, K. and K. Takatsuki. 1993. Adult T cell leukaemia-lymphoma. Clin. Haematol. 6:899-915.

Yamaguchi, K. and K. Takatsuki. 1993. Adult T cell leukaemia-lymphoma. Clin. Haematol. 6:899-915.

Yamamoto, N., M. Okada, Y. Koyanagi, M. Kannagi, and Y. Hinuma. 1982. Transformation of human Leukocytes by cocultivation with an adult T cell leukemia virus producer cell line. Science 217:737-739.

Yamano, Y., M. Nagai, M. Brennan, C. A. Mora, S. S. Soldan, U. Tomaru, N. Takenouchi, S. Izumo, M. Osame, and S. Jacobson. 2002. Correlation of human T-cell 248 lymphotropic virus type 1 (HTLV-1) mRNA with proviral DNA load, virus-specific CD8(+) T cells, and disease severity in HTLV-1-associated myelopathy (HAM/TSP). Blood 99:88-94.

Yamato, K., T. Oka, M. Hiroi, Y. Iwahara, S. Sugito, N. Tsuchida, and I. Miyoshi. 1993. Aberrant Expression of the p53 Tumor Suppressor Gene in Adult T- Cell Leukemia and HTLV-I-Infected Cells. Jpn. J. Cancer Res. 84:4-8.

Yan, J. P., J. E. Garrus, H. A. Giebler, L. A. Stargell, and J. K. Nyborg. 1998. Molecular interactions between the coactivator CBP and the human T-cell leukemia virus Tax protein. J. Mol. Biol. 281:395-400.

Yang, X. J., V. V. Ogryzko, J. Nishikawa, B. H. Howard, and Y. Nakatani. 1996. A p300/CBP-associated factor that competes with the adenoviral oncoprotein E1A. Nature 382:319-324.

Ye, J., L. Silverman, M. D. Lairmore, and P. L. Green. 2003. HTLV-1 Rex is required for viral spread and persistence in vivo but is dispensable for cellular immortalization in vitro. Blood.

Yin, M. J. and R. B. Gaynor. 1996. Complex formation between CREB and tax enhances the binding affinity of CREB for the human T-cell leukemia virus type 1 21- base-pair repeats. Mol. Cell Biol. 16:3156-3168.

Yoshida, H., H. Nishina, H. Takimoto, L. E. Marengere, A. C. Wakeham, D. Bouchard, Y. Y. Kong, T. Ohteki, A. Shahinian, M. Bachmann, P. S. Ohashi, J. M. Penninger, G. R. Crabtree, and T. W. Mak. 1998. The transcription factor NF-ATc1 regulates lymphocyte proliferation and Th2 cytokine production. Immunity 8:115-124.

Yoshida, M. 1996. Molecular biology of HTLV-I: recent progress. J. Acquir. Immune. Defic. Syndr. Hum. Retrovirol. 13 Suppl 1:S63-8:S63-S68.

Yoshida, M., I. Miyoshi, and Y. Hinuma. 1982. Isolation and characterization of retrovirus from cell lines of human T-cell leukemia and its implication in the disease. Proc. Natl. Acad. Sci. USA 79:2031-2035.

Yoshida, M., M. Seiki, K. Yamaguchi, and K. Takatsuki. 1984. Monoclonal integration of human T-cell leukemia provirus in all primary tumors of adult T-cell leukemia suggests causative role of human T-cell leukemia virus in the disease. Proc. Natl. Acad. Sci. U. S. A. 81:2534-2537.

Yoshimura, T., J. Fujisawa, and M. Yoshida. 1990. Multiple cDNA clones encoding nuclear proteins that bind to the tax-dependent enhancer of HTLV-I: all contain a leucine zipper structure and basic amino acid domain. EMBO J. 9:2537-2542.

Yssel, H., M. R. De Waal, D. Duc, D. Blanchard, L. Gazzolo, J. E. de Vries, and H. Spits. 1989. Human T cell leukemia/lymphoma virus type I infection of a CD4+ 249 proliferative/cytotoxic T cell clone progresses in at least two distinct phases based on changes in function and phenotype of the infected cells. J Immunol. 142:2279-2289.

Yuan, H., Y. M. Xie, and I. S. Chen. 2003. Depletion of Wee-1 kinase is necessary for both human immunodeficiency virus type 1 Vpr- and gamma irradiation-induced apoptosis. J Virol 77:2063-2070.

Zhang, B. X., X. Ma, C. K. Yeh, M. D. Lifschitz, M. X. Zhu, and M. S. Katz. 2002. Epidermal growth factor-induced depletion of the intracellular Ca2+ store fails to activate capacitative Ca2+ entry in a human salivary cell line. J Biol. Chem. 277:48165-48171.

Zhang, W., J. W. Nisbet, B. Albrecht, W. Ding, F. Kashanchi, J. T. Bartoe, and M. D. Lairmore. 2001. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J. Virol. 75:9885-9895.

Zhang, W., J. W. Nisbet, J. T. Bartoe, W. Ding, and M. D. Lairmore. 2000. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J. Virol. 74:11270-11277.

Zhao, L. J. and C. Z. Giam. 1992. Human T-cell lymphotropic virus type I (HTLV-I) transcriptional activator, Tax, enhances CREB binding to HTLV-I 21-base-pair repeats by protein-protein interaction. Proceedings. of. the. National. Academy. of. Sciences. of. the. United. States. of. America. 89:7070-7074.

Zhong, H., R. E. Voll, and S. Ghosh. 1998. Phosphorylation of NF-kappa B p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with the coactivator CBP/p300. Mol. Cell 1:661-671.

Zhou, J. and C. Aiken. 2001. Nef enhances human immunodeficiency virus type 1 infectivity resulting from intervirion fusion: evidence supporting a role for Nef at the virion envelope. J Virol 75:5851-5859.

Zhu, J. and F. Mckeon. 2000. Nucleocytoplasmic shuttling and the control of NF-AT signaling. Cell Mol. Life Sci. 57:411-420.

250