Transcript

STUDIES OF HTLV-1 p12(I) IN CALCINEURIN BINDING, CALCIUM-

MEDIATED CELL SIGNALLING AND VIRAL TRANSMISSION

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the Graduate

School of The Ohio State University

By

Seung-jae Kim, D.V.M., M.S.

* * * * *

The Ohio State University

2006

Dissertation Committee:

Professor Michael D. Lairmore, Adviser Approved by

Professor Patrick Green

Professor Kathleen Boris-Lawrie ________________________

Professor Natarajan Muthusamy Adviser

Veterinary Biosciences Graduate program

ii

ABSTRACT

Human T- lymphotropic virus type 1 (HTLV-1) causes a variety of

lymphoproliferative and neurodegenerative diseases. The mechanisms of early viral

infection, such as virus-mediated T cell activation, cell-to-cell transmission, and early

regulation of viral and cellular gene transcription are incompletely understood. The pX

ORF I encoded protein p12I is critically required for productive infection in a rabbit

model and viral infectivity in non-stimulated PBMC. HTLV-1 p12I regulates calcium-

mediated signaling in T cells, and induces activation of NFAT and enhancement of IL-2

production and p300 expression.

We identified a PSLPI/LTmotif in p12I, which is highly homologous to the

PxIxIT calcineurin-binding motif of NFAT. Full-length p12I and PSLPI/LT motif

containing mutants bound calcineurin in both immunoprecipitation and calmodulin bead

pull-down assays. In contrast, serial mutations of p12I that lacked the PSLPI/LT motif

or had selective alanine substitutions of the motif (p12I AxAxAA) exhibited abolished

or decreased binding affinity with calcineurin. We then tested if p12I binding to

calcineurin affected NFAT activity. p12I competed with NFAT for calcineurin binding

in calmodulin bead pull-down experiments. Furthermore, the p12I AxAxAA mutant

enhanced NFAT nuclear translocation compared to wild type p12I and increased NFAT

transcriptional activity two fold greater than wild type p12I. Thus the reduced

iii

binding of p12I to calcineurin allows enhanced nuclear translocation and transcription

mediated by NFAT, suggesting that HTLV-1 p12I modulates NFAT activation to

promote early virus infection of T lymphocytes.

We further tested the role of p12I on early HTLV-1 cell-to-cell transmission,

particularly, the role of p12I on LFA-1-mediated T cell adhesion. Our data indicated

that abrogation of pX ORF I mRNA expression in HTLV-1 infected cells (ACH.p12I)

resulted in reduction of LFA-1-mediated adhesion compared to wild-type HTLV-1

expressing cells (ACH). Furthermore, expression of p12I in Jurkat T-cells using

lentiviral vectors, enhanced LFA-1-mediated cell adhesion, which was inhibited by the

inhibitors of calcium-mediated signaling such as BAPTA-AM, SK&F 96365 and

calpeptin. Similar to the intracellular calcium mobilizer, thapsigargin, the expression of

p12I in Jurkat T-cells induced cell surface clustering of LFA-1 without changing the

level of integrin expression. Our data indicated that HTLV-1 p12I promotes cell-to-cell

spread by inducing LFA-1 clustering on T-cells via calcium-dependent signaling.

Lastly, we investigated the role of expression of HTLV-1 pX ORF I in early

HTLV-1 transmission. We compared the activation status between wild type ACH cells

and pX ORF I expression abrogated ACH.p12 cells. We then tested if the expression of

ORF I is required to induce bystander activation of uninfected target cells by

performing both proliferation assays and flow cytometric analysis of target cells.

iv

We found that abrogation of pX ORF I message decreased CD69 expression in HTLV-1

infected cells, but did not affect the HTLV-1- induced bystander target cell activation.

Furthermore, we performed real-time RT-PCR to detect pX ORF I and pX ORF III/IV

mRNA after coculture of both wild type ACH and ACH.p12 cells with pre-stimulated

or non-stimulated T cells. By measuring Tax/Rex mRNA in target cells, we were able to

compare viral infectivity of wild type ACH to ACH.p12 immortalized T cells and viral

mRNA expression in newly infected target cells. Our data indicated that expression of

pX ORF I is required for efficient HTLV-1 transmission to target cells and identified

differential expression patterns of Tax/Rex and pX ORF I mRNA during early cell-to-

cell transmission.

In conclusion, HTLV-1 p12I modulates calcium-mediated signaling and induces

enhancement of viral cell-to-cell transmission by inducing LFA-1 clustering and

facilitating the early establishment of infection by regulating viral gene expression,

providing further evidence that p12I is critically required for early HTLV-1 infection.

v

Dedicated to Rokyoun, Thomas, and my parents

vi

ACKNOWLEDGMENTS

First and foremost, I would like to thank my advisor, Dr. Michael Lairmore for

his valuable guidance and support throughout my graduate training. His offer to join his

laboratory provided me opportunity to start my carrer as a scientist. I really appreciate

his support and commitment to instructing me to allow me to fufill my academic goal.

He has always supported my imperfect ideas and helped me to develop the ability to

conduct scientific independent research that will unlimitedly benefit my future career in

every aspect.

I would like to thank my committee members Drs. Patrick Green, Kathleen

Boris-Lawrie and Natarajan Muthusamy, who have provided valuable guidance and

discussion on my thesis. I would like to express my gratitude to other members of the

Center for Retrovirus Research, in particular Dr. Larry Mathes, faculty members and

students, who have shared helpful ideas and discussion on my projects.

My studies could not have been succeful without the support of my calleagues

who worked together to aid in the completion of my research. I would like to thank all

my previous and current lab members, Wei Ding, Bjoern Albrecht, James Stanely,

Jesica Alcorn, Hajime Hiraragi, Lee Silverman, Amrithraj Nair, Bindhu Michael,

vii

Andrew Phipps, Antara Datta, Chris Premandandan, Rashade Haynes, Andy

Montgomery, John Nisbet, Laurie Millward,and Bevin Zimmerman. In particular, I

thank Wei Ding, Bjoern Albrecht, and Amrithraj Nair who helped me to perform many

experiments and shared their skills and ideas for my projects. I appreciate Min Li who

helped me to carry out real time PCR and shared her reagents. I am indebted to the

technical assistance provided by Rick Meister and Bryan McElwain, who helped flow

cytometrical anlalysis. I also thank Soledad Fernandez for statiscal analysis and Tim

Vojt for helping figure preparation.

Without continous prays and steadfast devotes of my wife, Rokyoun and my

parents, I would not be where I am today. They have always encouraged me to do my

best. Lastly, I appreciate my son, Thomas, who has always made me happy.

viii

VITA

June 19, 1972 ------------------------------ Born – Gwangju, South Korea 1995----------------------------------------- B.S. & D.V.M. Chonnam National University, Gwangju, South Korea 1995 - 1997 -------------------------------- Research Associate, Chonnam National University, Gwangju, South Korea 1997----------------------------------------- M.S. (Veterinary Pathobiology) Chonnam National University, Gwangju, South Korea 1998 - 1999 -------------------------------- Research Internship

Korean Science and Engineering Foundatoin Chonnam National University,

Gwangju, South Korea 1999 – present ----------------------------- Graduate Research Associate Department of Veterinary Biosciences The Ohio State University, Columbus, Ohio

PUBLICATIONS

Research Publications 1. Kim SJ, Nair AM, Fernandez S, Mathes L, and Lairmore MD. (2006) Enhancement of LFA-1-mediated T cell adhesion by Human T lymphotropic virus type 1 p12I Journal of Immunology, 176(9):5463-5470. 2. Hiraragi H, Kim SJ, Phipps AJ, Silic-Benussi M, Ciminale V, Ratner L, Green PL, and Lairmore MD. (2005) Human T-lymphotropic virus type 1 mitochondria localizing protein p13II is required for viral Infectivity in vivo. Journal of Virology, 80(7):3469-76.

ix

3. Kim SJ, Ding W, Albrecht B, Green PL, and Lairmore MD. (2003) A conserved calcineurin-binding motif in human T lymphotropic virus type 1 p12I functions to modulate nuclear factor of activated T cell activation. Journal of Biological Chemistry, 278(18):15550-15557. 4. Ding W, Kim SJ, Nair AM, Michael B, Boris-Lawrie K, Tripp A, Feuer G, and Lairmore MD. (2003) Human T-cell lymphotropic virus type 1 p12I enhances interleukin-2 production during T-cell activation. Journal of Virology, 77(20):11027-11039. 5. Ding W, Albrecht B, Kelley RE, Muthusamy N, Kim SJ, Altschuld RA, and Lairmore MD. (2002). Human T-cell lymphotropic virus type 1 p12(I) expression increases cytoplasmic calcium to enhance the activation of nuclear factor of activated T cells. Journal of Virology, 76(20):10374-82 6. Kim SJ and Park NY. (1997) Application of In situ hybridization for diagnosis of porcine reproductive and respiratory syndrome. Korean Journal of Veterinary Research (in Korean) 37(4), 793-807.

FIELDS OF STUDY

Major Field: Veterinary Biosciences

x

TABLE OF CONTENTS

Page Abstract--------------------------------------------------------------------------------------------- ii

Dedication ----------------------------------------------------------------------------------------- v

Acknowledgments --------------------------------------------------------------------------------vi

Vita------------------------------------------------------------------------------------------------ viii

List of Figures ----------------------------------------------------------------------------------- xiii

Chapters:

1. Literature Review: Molecular mechanism of Human T-Lymphotropic Virus Type-1 Infection, T Cell Activation and Transmission ----------------------------- 1

1.1 HTLV-1 Epidemiology--------------------------------------------------------- 1 1.2 HTLV-1- Associated Diseases ------------------------------------------------ 3

1.2.1 Adult T cell Leukemia/Lymphoma ------------------------------------ 3 1.2.2 HTLV-1- Associated Myelopathy/Tropical Spastic Paraparesis --- 5 1.2.3 Other Diseases Associated with HTLV-1-Infection ----------------- 6

1.3 HTLV-1 Viral Structure and Genome Organization ----------------------- 7 1.4 Structural and Enzymatic proteins of HTLV-1------------------------------ 9 1.5 Replication Cycle of HTLV-1 ----------------------------------------------- 10 1.6 HTLV-1 Regulatory Proteins: Tax and Rex ------------------------------- 12

1.6.1 Tax ---------------------------------------------------------------------- 12 1.6.2 Rex ---------------------------------------------------------------------- 16

1.7 HTLV-1 Non-Structural Proteins ------------------------------------------- 18 1.7.1 Minus Strand Encoded HBZ ------------------------------------------- 20

xi

1.7.2 pX ORF II Protein: p13II and p30II ------------------------------------ 21

1.7.2.1 p13II – Role in Viral Replication and Cell Survival --------- 22 1.7.2.2 p30II - A Selective Repressor of Transcription --------------- 24 1.7.2 pX ORF I Protein: p12I ------------------------------------------------ 26 1.7.3.1 Structure of p12I-------------------------------------------------- 26 1.7.3.2 p12I Subcellular Localization and Protein Interaction ------ 28 1.7.3.3 Role of p12I in Viral Infectivity-------------------------------- 31 1.7.3.4 Role p12I in Calcium- Mediated T Cell Signaling Pathway 32

1.8 HTLV-1 Cell-to-cell Transmission ----------------------------------------- 36 1.9 Regulation of Integrin --------------------------------------------------------- 38 1.10 Bystander Cell Activation by HTLV-1-------------------------------------- 41 1.11 References----------------------------------------------------------------------- 43

2. A Conserved Calcineurin-binding Motif in Human T Lymphotropic Virus Type 1 p12I Functions to Modulate NFAT Activation. ----------------------------------- 78

2.1 Introduction --------------------------------------------------------------------- 78 2.2 Materials and Methods -------------------------------------------------------- 81 2.3 Results Discussion ------------------------------------------------------------- 87 2.4 Discussion----------------------------------------------------------------------- 94 2.5 References----------------------------------------------------------------------- 97

3. Enhancement of LFA-1-Mediated T-Cell Adhesion by Human T-Lymphotropic Virus type 1 p12I --------------------------------------------------------------------- 111

3.1 Introduction ------------------------------------------------------------------- 111 3.2 Materials and Methods ------------------------------------------------------ 114 3.3 Results ------------------------------------------------------------------------- 118 3.4 Discussion--------------------------------------------------------------------- 124 3.5 References--------------------------------------------------------------------- 126

xii

4. Expression of HTLV-1 pX openreading frame I enhances early viral infectivity during cell-to-cell transmission in T-lymphocytes-------------------------------- 138

4.1 Introduction ------------------------------------------------------------------- 138 4.2 Materials and Methods ------------------------------------------------------ 142 4.3 Results ------------------------------------------------------------------------- 146 4.4 Discussion--------------------------------------------------------------------- 154 4.5 References--------------------------------------------------------------------- 159

5. Synopsis and Future Directions ----------------------------------------------------- 177

5.1 Studies to test the role of p12I binding to calcineurin on HTLV-1-mediated T cell activation and in vivo and in vitro viral infectivity --- 178

5.2 Further investigation of the role of p12I in MTOC polarization and virological synapse formation ---------------------------------------------- 179

5.3 Studies to explore the mechanisms of LFA-1 affinity regulation influenced by p12I expression ---------------------------------------------- 180

5.4 Studies required to test the role of p12I in HTLV-1 envelope glycoprotein folding and cell surface expression ------------------------ 181

5.5 Possible role of p12I in regulating CD69 expression during HTLV-1 infection ----------------------------------------------------------------------- 182

5.6 The potential role of p12I in regulation of viral gene expression during HTLV-1 cell-to-cell transmission ------------------------------------------ 182

5.7 Studies to test the influence of pX OFR I or p12I in early expression of HTLV-1 regulatory proteins ------------------------------------------------ 184

5.8 References--------------------------------------------------------------------- 185

Bibliography ------------------------------------------------------------------------------------ 188

xiii

LIST OF FIGURES

Figure Page 1.1 Schematic illustration of HTLV-1 proviral genome, mRNA and protein

species ------------------------------------------------------------------------------------ 75

1.2 Schematic Illustration of HTLV-1 non-structural protein p12I-------------------- 76

1.3 A model of p12I role in calcium-mediated signaling ------------------------------- 77

2.1 HTLV-1 p12I contains a highly conserved putative calcineurin-binding motif in p12I --------------------------------------------------------------------------- 102

2.2 The C-terminal half of p12I containing the PSLPI/LT motif is required for calcineurin binding --------------------------------------------------------------- 103

2.3 Alanine substitution mutant of p12I (AxAxAA) decreases binding affinity for calcineurin---------------------------------------------------------------- 104

2.4 p12I and calcineurin binding is inhibited by calcium chelators but not inhibited by cyclosporin A ---------------------------------------------------------- 105

2.5 p12I binding to calcineurin decreases the amount of NFAT binding to calcineurin ----------------------------------------------------------------------------- 106

2.6 The substitution mutation in calcineurin binding sequence in p12I induces increased NFAT transcription activity and did not compete against wild type p12I for NFAT transcription activity ------------------------- 107

2.7 Reduced p12I binding to calcineurin induces increased NFAT nuclear translocation --------------------------------------------------------------------------- 108

xiv

2.8 Calcineurin phosphatase activity is not affected by p12I binding to calcineurin ----------------------------------------------------------------------------- 110

3.1 LFA-1 mediated adhesion is reduced in ACH.p12I cell lines ------------------- 132

3.2 ACH.p12I cells have decreased binding of sICAM-1 without alteration of LFA-1 expression -------------------------------------------------------------------- 133

3.3 p12I stable expression in Jurkat T-cells did not alter LFA-1 affinity ---------- 134

3.4 Expression of p12I in Jurkat T-cells induced LFA-1-mediated cell adhesion ------------------------------------------------------------------------------- 135

3.5 HTLV-1 p12I-mediated LFA-1 activation is inhibited by calcium signal inhibitors ------------------------------------------------------------------------------- 136

3.6 Expression of p12I in Jurkat T-cells modulated surface distribution of LFA-1 on the cell membrane -------------------------------------------------------- 137

4.1 Cell surface expression of early activation markers and adhesion molecules on wild type HTLV-1 infected cells and T cells lacking pX ORF I expression --------------------------------------------------------------------- 165

4.2 Deletion of pX ORF I expression in HTLV-1 infected cells did not affect bystander proliferation of uninfected target PBMC ------------------------------ 167

4.3 Abrogation of pX ORF I expression in HTLV-1 infected cells did not affect bystander activation of target PBMC----------------------------------------------- 168

4.4 HTLV-1 pX ORF I and pX ORF III/IV mRNAs were differentially expressed during the early HTLV-1 transmission --------------------------------------------- 171

4.5 Target PBMC were selectively sorted from HTLV-1 infected cells------------ 173

xv

4.6 Expression of pX ORF I enhanced HTLV-1 viral transmission to target PBMC-- 174

4.7 Expression of HTLV-1 pX ORF I and pX ORF III/IV mRNAs in de novo HTLV--1 infected target PBMC --------------------------------------------------------- 175

1

CHAPTER 1

LITERATURE REVIEW

MOLECULAR MECHANISM OF HTLV-1 INFECTION, T CELL

ACTIVATION, AND TRANSMISSION

1.1 HTLV-1 Epidemiology

The first identified human retrovirus, human T lymphotropic virus type 1

(HTLV-1) was detected as type C particles propagated from T cell lymphoblastoid cell

lines and peripheral blood lymphocytes of a cutaneous T cell lymphoma patient by

Poiesz, Gallo and their colleagues in 1980 1. Subsequent epidemiologic, immunologic,

and molecular biologic studies 2-5 during the early 1980’s demonstrated that HTLV-1

was the etiological agent of adult T-cell leukemia/lymphoma (ATLL) that was

previously reported in Japan in 1977 6,7. By the mid-1980’s, HTLV-1 was also

confirmed as a causative agent of a degenerative neurologic disease, HTLV-1-

associated myelopathy/tropical spastic paraparesis (HAM/TSP) 8,9.

2

HTLV-1 infections have been reported throughout the world and it is estimated

that 15 to 25 million people are infected 10,11. Infected individuals however are primarily

confined in highly endemic areas of southern Japan 3, the Caribbean 12, central Africa 13,

Central and South America 14,15, Melanesian Islands in the Pacific basin 16 and among

certain high risk groups within Europe and the United states of America (USA) 17.

Within these endemic areas, the seroprevalence rate among general population varies

between 0.1 % and 30 %, and the rate increases with age and is higher in females than

males 11,18.

Unlike human immunodeficiency virus 1 (HIV-1), HTLV-1 is poorly infectious

as cell-free virions 19. Cell-cell contact between virus infected and target cells is

required for efficient transmission 20. Breast feeding and perinatal contamination of the

infant with blood of an infected mother are major routes of transmission in endemic

areas 21,22, but transmission of HTLV-1 via trans placenta route is considered extremely

rare 23. The risk of HTLV-1 infection via these routes are correlated to maternal factors

such as high HTLV-1 antibody titer, prolonged ruptured membranes during delivery,

and low socioeconomic status 24. Exposure to infected blood or blood products is

another major route of HTLV-1 transmission 25. The most common cause of blood-to-

blood transmission occurs among intravenous drug users by sharing needles 26,27.

Because transmission through infected blood products has been a major public health

concern, all blood products are screened in Japan, the USA and Brazil 11,28-33. Sexual

transmission is a less efficient mode of HTLV-1 transmission 34. Male to female

transmission via semen is about four times as frequent as female to male transmission 34.

3

For diagnosis of HTLV-1 infection, enzyme linked immunosorbent assays

(ELISA) 35-37 in USA, and particle agglutination assays in Japan are widely used for

detection of anti-HTLV-1 antibodies. After these prototype screening assays,

polymerase chain reaction (PCR) or western blot assays are performed as confirmation

tests 38.

1.2 HTLV-1 Associated Diseases

Approximately 5% of HTLV-1 infected people develop HTLV-1- mediated

disease and the risk increases to 8-10 % when the patient has other illnesses 11. Most

HTLV-1 carriers remain asymptomatic throughout their lives and only develop disease

after months to years of infection 39. HTLV-1 infection is strongly associated with

ATLL and HAM/HSP, and has been implicated to cause or complicate immune

mediated conditions.

1.2.1 Adult T Cell Leukemia/ Lymphoma (ATLL)

In 1977, ATLL was recognized from epidemiological studies in southwest

Japan 6,7. Geographically clustered patients with lymphoid neoplasms were just

identified because of their unique clinical features 6,7. Characteristics of ATLL include:

adult onset, acute or chronic leukemia with rapid progression, resistance to treatment,

peripheral, pleomorphic leukemic cells with markedly deformed nuclei, frequent

lymphadenopathy, hepatosplenomegaly and hypercalcemia, absence of mediastinal

tumors, and frequent skin lesions 2. Similar to non-Hodgkin’s lymphoma, affected

4

patients present with malaise, fever, jaundice, drowsiness, weight loss, and

opportunistic infections 2. The diagnosis of ATLL is made based on specific parameters

including sero-positivity to HTLV-1, marked leukocytosis, morphology of neoplastic T

cells (“cerbriform” or “flower cell”), T cell immunophenotyping, hypercalcemia,

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

serum lactate dehydrogenase (LDH) levels 40. The predominant phenotype of ATLL

neoplastic cells is characteristic of helper T cells: CD3+, CD4+, L-Selectin+, CD25+,

CD45RA+, HLA-DR+, CD29-, and CD45RO- in circulating peripheral blood, or CD3+,

CD4+, L-Selectin+, CD29+, CD45RO+, HLA-DR+, and CD45RA- in cutaneous and

lymphoma lesions 41,42.

ATLL can be classified into four subcategories (smoldering, chronic, acute, and

lymphoma) based on clinical and laboratory features including the percentage of

abnormal T cells in the peripheral blood, blood LDH and calcium levels, and malignant

tumors in various organs 40,43. Smoldering ATLL, which is characterized by the

presence of a few neoplastic cells (less than 5%) in the peripheral blood, has the best

prognosis among the four subtypes. Four-year survival rates vary from 5.0% for acute,

5.7 % for lymphoma, 26.9% for chronic and 62.8% for smoldering 40. Although various

therapeutic strategies 44 have been tried, treatment of ATLL is still unsatisfactory.

ATLL is characterized by clonal expansion of mature T cells, each harboring a

single copy or multiple copies of HTLV-1 sequences 45. Initial polyclonal expansion of

infected cells is followed by a progression to oligoclonal and then to monoclonal

5

proliferation in vivo 46-48. The principal mode of viral replication after initial viral spread

in vivo is by mitosis of infected cells. The low incidence and long clinical latency of

ATLL strongly suggest that accumulation of genetic mutations are required for T cell

transformation in addition to HTLV-1 infection 49 . Although molecular pathogenesis of

ATLL is not fully understood, evidences suggests that HTLV-1 oncoproteins like Tax

play a critical role in transformation of infected lymphocytes by modulating T cell

activation and death pathways 50.

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

In 1985, Gessain and colleagues 8 first reported that a group of patients in

French Martinique with a slowly progressive neurologic disorder, called Tropical

Spastic Paraparesis (TSP), had antibodies directed against HTLV-1. Subsequently,

another neurologic condition observed among HTLV-1-infected individuals in southern

Japan, termed HTLV-1-Associated Myelopathy (HAM) was reported by Osame et al 9.

Studies the established that TSP and HAM were clinically are identical diseases with

the common viral etiology of HTLV-1.

Compared to ATLL, HAM/TSP has a relatively shorter latency period ranging

from months to decades 51. HTLV-1 seropositive women are more likely to develop

disease than men 52. HAM/TSP is a chronic progressive demyelinating disease

predominantly affecting the thoracic spinal cord 28,53-56. Symptoms of HAM/TSP are

related to myelopathy including low back pain, weakness and spasm of lower

extremities, and dysfunction of the urinary bladder 57. Occasionally, a cerebellar

6

syndrome with ataxia and intention tremor is observed 58. Also, a predominant

dysfunction of descending sympathetic pathway has been reported in some cases 59.

Severe cellular destruction and inflammation both in the spinal cord and brain

characterized by multiple foci of severe demyelination and mononuclear cell infiltrates

with perivascular cuffing, parenchymal invasion and gliosis within white matter are

histopathological features of HAM/TSP 60, 61.

Risk of development and progression to HAM/TSP is increased when HTLV-1

proviral copy number is high in infected individuals’ blood leukocytes 62-66, and is

clearly associated with transfusion of HTLV-1-contaminated blood products 67,68. Other

factors influencing HAM/TSP development are host genetic factors 69, host immune

response, and perhaps viral variation especially in tax 63,70-74. Although the mechanisms

of HAM/TSP development by HTLV-1 infection are not yet clear, important role for

proinflammatory cytokines and activated lymphocytes in the development of the disease

has been suggested 61. High levels of inflammatory cytokines, such as IFN-γ, TNF-α,

IL-1 and IL-6 75-78, and large numbers of activated lymphocytes are present in the

cerebrospinal fluid (CSF) of affected individuals. Several hypotheses of HAM/TSP

pathogenesis such as CD8+ T cells mediate CNS damage 79, or autoimmune response

mediated HTLV-1- associated CNS damage 80,81 are suggested 45.

1.2.3 Other Diseases Associated with HTLV-1-Infection

A variety of autoimmune disorders and chronic inflammatory conditions are

associated with HTLV-1 infection, even though linkage between these diseases and

7

HTLV-1 infection is not as strong as those for ATLL and HAM/TSP. These include

idiopathic uveitis 82, arthropathy 83, Sjögren’s syndrome 84, infective dermatitis 85,

polymyositis 86, lymphadenitis 87, chronic respiratory disease 88, and acute myeloid

leukemia 89. Also, HTLV-1 infection is associated with severe Strongyloides stercoralis

infection 90,91. However, the role of HTLV-1 infection in the pathogenesis of these

disorders is still controversial.

1.3 HTLV-1 Virus Structure and Genome Organization

HTLV-1 is classified as the genus Deltaretrovirus, along with bovine leukemia

virus (BLV) and simian T lymphotropic virus (STLV). The mature HTLV-1 virion is

spherical and has a diameter of 110 to 140 nm with type C retroviral morphology 92.

The outer envelope is composed of host cell-derived cell membrane that contains the

viral encoded envelope glycoprotein spikes. The core of the virion consists of a highly

dense, spherical ribonucleoprotein complex of two copies of the 9 kb genomic RNA and

the host cell origin, primer tRNA-Pro, with virus encoded enzymatic and structural

proteins including reverse transcriptase and integrase, protease, nucleocapsid, capsid

and matrix 93.

The HTLV-1 genome contains elements common to other retroviruses, as well

as genes unique to HTLV-1. The proviral DNA genome is 9032 nucleotides long 94 is

flanked by long terminal repeat (LTR), which is a hallmark of retroviral genomic

structure, at each ends. Each LTR, composed of a U3 (unique 3'), R (repeated), and U5

(unique 5') region , contains sequences essential for viral reverse transcription,

8

integration, transcription, polyadenylation, splicing and viral message transport (Fig.

1.1) 95-97. A unique feature of the HTLV-1 LTR is the presence of three imperfect

tandem 21-base-pair repeats in its U3 region, collectively called Tax-responsive

elements-1 (TRE-1) upstream of the transcription start site 98. These three cis acting

regulatory elements are required for Tax-mediated trans-activation 99-104. In the middle

of the genome, gag , pol, and env genes are located, which are common to all known

retroviruses. The structural Gag precursor (later cleaved into matrix [MA], capsid [CA],

and nucleocapsid [NC] proteins) and enzymatic proteins (protease [PR], reverse

transcriptase [RT] and integrase [IN]) are translated from an unspliced RNA, and the

envelope protein (transmembrane [TM] and surface glycoprotein [SU]) is translated

from a singly spliced RNA 105.

In the pX region, which is located between env and 3’ LTR, HTLV-1 genome

contains unique genes encoded for regulatory and accessory proteins. These proteins are

generated by alternative splicing and internal initiation from the four open reading

frames (ORF). Regulatory proteins Tax and Rex, which are essential for the viral life

cycle, encoded by ORF IV and ORF III, respectively 106,107. ORF I and ORF II encode

four accessory proteins, p12I, p27I, p30II and p13II 95-97,108 109(Fig. 1.1). More recently, a

novel HTLV-1 protein, HBZ which is encoded by complementary (minus) RNA was

identified 110.

9

1.4 Structural and Enzymatic Proteins of HTLV-1

After translation into a single precursor polyprotein p55 (Gag), HTLV-1 Gag

protein is targeted to the inner side of lipid plasma membrane through post-translational

myristylation at its N-terminal end 111 and cleaved into matrix (MA, 19 kDa), capsid

(CA, 24 kDa) and nucelocapsid (NC, 15 kDa) proteins by viral protease(Fig. 1.1) 105,112.

Functionally, three types of domains, the M (membrane binding), I (interaction between

Gag proteins), and L (late budding) domains, have been identified in the Gag sequences

of different retroviruses 113,114. The p19 MA protein contains L domain (PPPY motif),

which is critical in budding of HTLV-1 virus particles 115. This budding process

requires interaction between L domain and cellular proteins Nedd4.1 and Tsg101 116.

CA proteins interact with each other and form the shell of an inner core structure.

Negatively charged NC proteins interact with viral RNA genome within the encased

capsid structure 45.

HTLV-1 protease (PR), which is generated by ribosomal frame shifting near 3'

end of gag and 5' of pol, is self-cleaved to become the active form 117. The catalytic

activities of PR are required for whole HTLV-1 life cycle, because mature or active

forms of viral proteins are generated by PR 118. Reverse transcriptase (RT) and integrase

(IN), which are produced by cleavage of the Gag/Pol precursor by the PR 118, have

enzymatic activities; Mg2+-dependent reverse transcription and proviral DNA

integration respectively. RT also has RNase H activity which specifically degrades the

RNA in the RNA-DNA duplexes during the reverse transcription process 93,118.

10

HTLV-1 envelope (ENV) protein is synthesized as a 61 to 68 kDa glycoprotein

in the endoplasmic reticulum (ER) and subsequently transported to the Golgi apparatus

through the secretory pathway. In the Golgi apparatus precursor Env protein is cleaved

into the surface (SU, gp46) and the transmembrane (TM, gp21) protein 119. SU-TM

heterodimer is essential for receptor binding and the fusion events viral entry120.

Recently, an ubiquitous glucose transporter, GLUT-1 has been identified as a HTLV-1

receptor by Manel et al 121.

1.5 Replication Cycle of HTLV-1

Major events in the retroviral replication cycle include adsorption and entry,

reverse transcription, nuclear transport and integration, viral gene expression, and viral

protein synthesis, processing, and assembly. From entry to integration step, the process

accomplished by the viral structural proteins and several enzymatic proteins occur in the

virion without de novo viral gene expression 122. The process of viral gene expression

and assembly, which is a later stage of viral replication, depends on host cellular

transcription and protein synthesis machinery, as well as viral machinery.

Viral particle attachment and entry involve an interaction between envelope (SU) and

HTLV-1 receptors. Many cell surface molecules have been implicated as HTLV-1

receptors including adhesion molecules123, heat shock cognate proteins 124, lipids 125,

lipid rafts 126, heparan sulfate 127 and transferrin receptor 128. More recently, the first

definitive identification was reported by Manel et al 121, who identified GLUT-1, a

ubiquitous glucose transport protein, as a receptor for HTLV-1. Before identification,

11

they found that HTLV-1 candidate receptor is poorly expressed on resting T cells but is

up-regulated after T-cell activation 129,130. Expression of full-length HTLV envelope and

envelope driven peptide resulted in dramatic changes in glucose metabolism such as

reduced glucose uptake and consumption 121. Because of these findings, they suspected

GLUT-1 as a HTLV-1 receptor. Further characterization of GLUT-1 and SU interaction

is required for understanding of HTLV-1 pathogenesis, as well as for developing

strategies to prevent HTLV-1 early transmission 131.

After envelope mediated attachment and fusion, components of the virion such

as NC and CA enter into cytoplasm. In the cytoplasm, RT (RNA-dependent/DNA-

dependent polymerase) initiates the synthesis of double-stranded DNA (dsDNA) from

viral genomic RNA. Viral dsDNA associates with cellular and viral proteins to form a

pre-integration complex 132, which is then transported into the nucleus where the

integration of the proviral DNA into the host genome occur. HTLV-1 integration is

considered to take place randomly in the host genome by virally encoded IN. However,

retroviral selection of integration sites is influenced by properties of host genomic DNA

properties such as bendability or open structure, which are commonly observed in

chromatin region of actively transcribed genes 133.

After integration, HTLV-1 provirus behaves like host genomic DNA; therefore

it requires participation of cellular transcription, translation, and transport machinery,

as well as viral proteins especially Tax and Rex 134. Integrated provirus passively

spreads to daughter cells by host cell division, or actively produces progeny virions.

The HTLV-1 replication cycle is completed when virions are assembled and released by

12

budding. However, HTLV-1 infectious virions are not efficiently released from the

infected cells, therefore viral transfer by cell-to-cell contact is a more efficient way for

HTLV-1 transmission 20.

1.6 Regulatory Proteins: Tax and Rex

1.6.1 Tax

Tax (the transcriptional activator of pX region) regulatory proteins are unique to

deltaretroviruses and are important for productive viral replication and proliferation of

host cells. HTLV-1 Tax, a 40 kDa phosphoprotein, was identified as a trans-activator of

viral gene transcription 135-139. Tax ,which is translated from a doubly spliced

mRNA(Fig. 1.1) 135-139, mainly accumulates in the nuclear region of HTLV-1-infected

cells and shuttles into the cytoplasm using a nuclear export protein 140,141. HTLV-1 Tax,

which has a pleiotropic function, is associated with transcriptional activation or

repression of various viral or cellular genes and alteration of cell cycle and apoptosis in

infected cells 142. It activates transcription by recruiting or modifying the activity of

cellular transcription factors such as cyclic AMP-responsive element binding protein

(CREB), serum-responsive factor (SRF) and NF-κB 143-147. Tax mediated transcriptional

activity can be achieved by various mechanism. For example, Tax activates CREB and

SRF mediated transcription by stabilizing the complexes of the transcription machinery,

or it enhances NF-κB mediated transcription by destabilizing IκB indirectly 148.

13

HTLV-1 Tax activates expression of viral genes via the LTR. Three highly

conserved 21-bp repeat elements collectively referred to as Tax-response element 1

(TRE-1) 99-104 are located in the U3 region of 5’ LTR 101-104. Similar to cAMP-

responsive elements (CRE) of host genomic sequence, HTLV-1 TRE-1 elements are

binding sites for multiple cellular transcription factors including CREB 149,150, cAMP

response element modulator (CREM) 149, activating transcription factors (ATFs) 151,

Tax-responsive element binding proteins (TREB) 152, activator protein-1 (AP-1)

102,153,154, and activator protein-2 (AP-2) 155. Also, a second enhancer element called

Tax-responsive element 2 (TRE-2) located between the central and proximal TRE-1 is

also important for viral transcription 156,157. Ets family transcription factors (Ets-1, Ets-

2, Elf-1 and TIF-1) and c-Myb transcription factors bind to TRE-2 158-162. Tax binds

directly to GC-rich sequences flanking the TRE-1 elements 163-165, and interacts with

basic region of host cellular basic leucine zipper (bZIP) transcription factors resulting in

enhanced bZIP dimerization and DNA binding activity 166.

A bZIP transcription factor, CREB requires a protein kinase A (PKA)-

mediated phosphorylation to bind with the transcriptional cofactor CBP/p300 during the

cell activation signaling 167,168. By interacting with both CREB and CBP/p300, Tax

eliminates the requirement of PKA activation, CREB phosphorylation, and recruitment

of CBP/p300 to the transcriptional complex. Therefore, CREB-Tax-CBP/p300 complex

results in constitutive activation of this pathway in HTLV-1 infected cells 169. Histone

acetylation by coactivators CBP/ p300 has been shown to play a major role in activating

HTLV-1 transcription 170. Also, CBP/p300 associated factor (P/CAF) is required for

Tax- mediated constitutive transcriptional activation 171,172.

14

In addition to viral gene expression, Tax affects a variety of cellular genes. In

fact, a gene array study has demonstrated that Tax modulates expression of hundreds of

cellular genes173. These includes cytokine genes, IL-2 174, IL-6 175, IL-8 176, IL-2Rα 174,

IL-1 177, GM-CSF 178, TNFα 179 and TNFβ 180, genes of transcription factors c-myc 181,

c-fos 182, c-sis 183, erg-1 184, c-rel 185, and genes involved in apoptosis Bcl-xL 186,187 and

DNA repair (PCNA) 188. Tax modulates cellular gene expression via at least four

distinct cellular signaling pathways: CREB, NF-κB, AP-1, and SRF 146,154,184,189,190.

Activation of NF-κB signaling by Tax is thought be critical for ATLL

pathogenesis, because ATLL cells express numerous cytokines and their receptors,

which are known to be regulated by the NF-κB pathway190. Over-expression of IL-2Rα

191, which is one of key features of ATLL cells, is induced by Tax-mediated NF-κB

activation 192. However, constitutive NF-κB activity is observed within ATLL cells of

patients, even though expression of Tax is lacking in these cells 193. The mechanism

underlying this Tax-independent pathway of NF-B activation remains poorly

understood.

Activation of NF-κB by Tax occurs in a different manner both in nuclear and

cytoplasmic components. Tax binds to multiple NF-κB family proteins such as p50, p52,

p65, c-Rel 147,194 and lyt10 195. Within the nucleus, Tax has been shown to bind to the

p50 and p52 subunits of NF-κB 147,195. Because both NF-κB proteins and Tax bind to

CBP/p300 196-198, Tax may complex with NF-κB and CBP/p300 to enhance the stability

of transcription machinery. However, the predominant mode of Tax mediated NF-κB

activation occurs by interacting with cytoplasmic IκB proteins, the inhibitor of NF-κB

15

190. The NF-κB heterodimer is usually retained in the cytoplasm by interacting with IκB

proteins 199. When cell signal is activated, IκB proteins are phosphorylated by IκB

kinase (IKK) and subsequently ubiquitinated and degraded in the proteasome 190.

Interaction between Tax and IκBα destabilizes NF-κB and IκB complex and causes the

nuclear translocation of NF-κB 200-202. Moreover, Tax activates IKK which consists of

two catalytic subunits, IKKα and IKKβ and a regulatory IKKγ subunit (also known as

NEMO) 200,202. Tax interacts with IKKγ subunit and recruits MAP3Ks to this complex

190, and resulting in phosphorylation and degradation of IκB and subsequent NF-κB

activation 203.

HTLV-1 Tax was also demonstrated to inhibit expression of several cellular

genes, such as p18INK4c, p53 and Bax 204. Sequestration effect of Tax on p300/CBP is

thought be one mechanism of repression activity 205,206, because as a transcriptional co-

activator, p300/CBP is required for cellular gene expression by binding to multiple

transcription factors 207,208. For example, promoter of p18INK4c gene requires

CBP/p300 co-factor for activation of E-47 transcription factor 209.

In addition to the transcriptional deregulation, Tax binds to a number of protein

complexes regulating transformation, cell cycle and apoptosis. Proteins affecting cell

cycle include p16INK4a 210 and p15INK4b 211,212 that directly regulate the Rb protein

and the G1/S phase of the cell cycle. Tax also interacts with human mitotic checkpoint

protein MAD1 213. Tax can affect apoptosis by interacting with an inhibitor of cell death

protein, A20 214 which is induced by a variety of inflammatory stimuli. A variety of

16

small cytoplasmic GTPases including RhoA, Rac1 and Cdc42, which are components

of cytoskeletal regulation, are also associated with Tax 215.

Collectively, HTLV-1 Tax effects on a wide variety of cellular targets by

transactivating, repressing or interacting to promote cell proliferation and

leukemogenesis.

1.6.2 Rex

Unlike Tax, ORF III encoded protein, Rex, regulates viral gene expression post-

transcriptional. Rex is a 27 kDa RNA-binding protein that is essential for transport of

viral mRNAs. Rex facilitates transport of unspliced RNA (gag/pol/pro) and singly

spliced RNA (env) from nucleus to cytoplasm, while it may inhibit the splicing and

transport of doubly spliced RNAs that encode the regulatory and accessory proteins in

the pX region. Therefore, Rex regulates the balance between viral structural and

regulatory gene expression. HTLV-1 Rex is not required for cellular immortalization in

vitro, but it is required for viral spread and persistence in vivo106,216,217. Rex is also

reported to have a role in stabilizing unspliced transcripts in T cells 218.

Rex performs its function by interacting with a sequence called the Rex-

responsive element (RxRE) within the U3 and R regions of the 3' LTR 219. RxRE forms

a stable secondary RNA structure, consisting of four stem loops and a long stem

structure 219-224. The stem-loop structure is also critical for appropriate polyadenylation

17

of viral RNAs 225,226. Because RxRE is present in all mRNAs, i.e. unspliced, singly and

doubly spliced mRNAs, other cis-acting sequences may play a role in the determination

of Rex regulation of mRNA nuclear export. Inhibitory sequences within the viral RNA

have been identified, which are termed cis-acting repressive sequences (CRS) 227-229.

Rex has multiple domains, which are important for its function, including

RNA-binding domain, nuclear localization sequence (NLS), nuclear export sequence

(NES), and multimerization domain. A highly basic N-terminal RNA-binding domain is

located within aa 1−19 230,231. This domain also functions as an NLS and is necessary

for the transport of unspliced viral mRNAs to the cytoplasm 232,233. The NES sequence

(also known as activation domain) is located in the middle of the Rex protein. Rex

mediated RNA transport is reported to depend on CRM1/exportin pathway 234,235. The

NES is required for the shuttling of Rex between the nucleus and cytoplasm236 The

activity of Rex is affected through phosphorylation. Treatment of infected cells with

protein kinase C inhibitor resulted in accumulation of unspliced mRNA and decreased

gag protein synthesis 237.

In addition to the full length 27 kDa form of Rex, a truncated 21 kDa form of

Rex (p21Rex) has been detected in HTLV-1 infected cell lines (Fig. 1.1) 108,109,238.

Because p21Rex has a truncation in the N terminal NLS sequence, it inhibits the shuttling

function of the full-length form of Rex protein when over-expressed 239. However the

function of p21Rex in the HTLV-1 replication and pathogenesis is still unknown.

18

1.7 HTLV-1 Non-structural Proteins

In addition to the regulatory proteins Tax and Rex, which are encoded in ORF

III and IV respectively, the HTLV-1 pX region encodes four additional accessory

proteins, p12I, p27I, p13II, and p30II. These proteins are generated from alternatively

spliced mRNA encoded in ORF I and II (Fig. 1.1) 95,240-242. All of these spliced mRNA

species have a common first exon encoded from nucleotides (nt) 1-119 in the R region

of the viral 5’ LTR. Doubly spliced mRNAs encode second exons that start at either nt

4641 or 4658 and end at nt 4831. Various splice acceptor sites in the pX region

correspond to the third exon start point for doubly spliced messages or the second exon

start site for singly spliced mRNA. A splice acceptor at nt 6383 is used to generate the

pX ORF I proteins, p27I (doubly spliced) and p121 (singly spliced). Similar to pX ORF

I, pX ORF II proteins are also produced from two alternatively spliced mRNAs. A

splice acceptor site at nt 6478 is used for generating mRNA encoding the larger protein,

p30II (doubly spliced) and a site at nt 6875 creating mRNA used for the smaller protein

p13II. A splice acceptor site at nt 6950 is used to produce doubly spliced pX-tax/rex

mRNA that encodes Tax or Rex using ORF IV or ORF III 95-97,108, 109.

Although HTLV-1 accessory proteins were thought not to be required for viral

replication 243, recent findings indicated that HTLV-1 accessory proteins play a critical

role in viral infectivity, maintenance of high viral loads, host cell activation, and

regulation of gene transcription 244-257. Even though the proteins were not directly

19

detected in HTLV-1 infected cells, indirect evidences indicate that pX ORFs I and II

mRNAs and proteins are expressed both in vitro and in vivo. mRNAs of pX ORF I and

II were detected by reverse transcription-PCR (RT-PCR) assays or semi-quantitative

RNase protection assays in infected cell lines and freshly isolated cells from HTLV-1-

infected subjects 242,258 . Furthermore, humoral antibody 254,259 and cytotoxic T cell 260

responses against recombinant proteins or peptides of the pX ORF I and II proteins are

detected in HTLV-1 infected patients, and asymptomatic carriers. These findings

strongly indicate that ORF I/II encoding proteins are produced in vivo and play roles in

HTLV-1 infection and leukemogenesis.

Chronically infected HTLV-1 cell lines were found to have 100 to 1,000-fold

less ORF I mRNA expressed compared to ORF III/IV mRNA species, and ORF II

mRNA amounts were 500 to 2,500-fold lower than ORF III/IV mRNA, suggesting that

HTLV-1 proteins expression are differentially regulated by unknown mechanism 261.

However, these data were collected from chronically infected cell lines. Therefore, a

requirement of ORF I and ORF II encoded proteins in HTLV-1 pathogenesis can not be

ruled out, especially in the early phases of infection in vivo. Furthermore, the nucleotide

sequence and genomic region encoding these accessory proteins, particularly p12I, are

highly conserved among different HTLV-1 strains, HTLV-2 and the highly related

simian T lymphotropic virus type 1 (STLV-1) 240,262-264.

In addition to these regulatory and accessory proteins, a novel viral protein

HTLV-1 bZIP factor (HBZ), encoded from antisense RNA was recently identified 110.

20

1.7.1 Minus Strand Encoded HBZ

Several ORFs encoded from the complementary (minus) strand of HTLV-1

RNA genome were reported initially by Larocca et al 265. Through RNA blotting, 2.5

and 2.9 kb minus strand RNAs were detected in HTLV-1 infected T-cells, but not in an

uninfected control. A novel viral protein, HBZ encoded by these complementary strands

of the HTLV-1 RNA genome was identified recently 110. HBZ is a nuclear localizing

protein composed of 209 amino acids that contains an N-terminal transcriptional

activation domain and a leucine zipper motif in its C terminus 110, as well as three

nuclear localizing signals 266.

By interacting with bZIP transcription factor, CREB-2, HBZ abolishs the ability

of CREB-2 to activate Tax mediated viral transcription, when it exogeneously

overexpressed. As a result, HBZ was suggested to repress Tax-mediated viral

transcription from the HTLV-1 LTR 110.

Moreover, HBZ interacts with the activator protein-1 (AP-1) transcription

factors c-Jun , JunB and JunD 267, 268. While the interaction with c-Jun results in down

regulation of AP-1 transcriptional activity, transcription function of JunB and JunD was

enhanced 267, 268. HBZ suppresses c-Jun- mediated AP-1 transcription by preventing c-

Jun DNA-binding 267 and promoting c-Jun degradation through a proteasome-dependent

pathway 269.

Rescently Arnold et al 270 evalutated the role of HBZ in vitro cellular

immortalization, and in vivo viral infectivity and persistence by generating HBZ

21

deletion mutatants. Mutation of HBZ did not affect in vitro viral replication and cellular

immortalization. However, rabbits inoculated with HBZ mutant cells displayed a

decreased antibody response and reduced infectivity when compared to wild type

HTLV-1 infected rabbits. These findings indicated that role of HBZ was not related

with repressive effect on Tax and AP-1 - mediated in vitro immortalization, but was

related with in vivo viral infectivity and persistence.

The role of HBZ in the pathogenesis of HTLV-1 infection is unclear, but

studies to date susggest that HBZ may contribute to the dysregulation of viral or cellular

gene expession to alter viral replication and perhaps disease.

1.7.2 pX ORF II Proteins: p13II and p30II

A 241 amino acid protein, p30II is generated from the doubly spliced mRNA

pX-tax-orf II 95,96 while p13II is translated from the singly spliced message pX-orf II 95-

97. However, an internal start codon in p30II message can be used to produce the smaller

p13II protein, which represents the C-terminal 87 amino acids of p30II.

Earlier studies suggested that ORF II encoding proteins were dispensable for

viral replication or immortalization of primary cells, because a viral strain isolated from

leukemic cells was reported to contain a premature stop condon in pX ORF II 271.

However, this study did not consider the possible role of ORF II encoded proteins in the

early viral infection in vivo. Furthermore, rabbits inoculated with a HTLV-1 proviral

clone with selective mutations in ORF II (ACH.p30/p13) failed to develop productive

infections 248.

22

1.7.2.1 p13II – Role in Viral Replication and Cell Survival

HTLV-1 p13II localizes to both the nucleus 242 and to mitochondria 256,272.

Because p13II has no DNA binding motifs nor transcriptional activity 273, a role of p13II

in mitochondrial function has been emphasized. p13II localizes to inner mitochondrial

membranes 272 using an atypical mitochondrial targeting sequence (MTS) located in the

N terminus (amino acids 22 and 31) 256. Unique features of p13II MTS include: its

short length, lack of positive charge and its unusual property of remaining intact upon

mitochondria import 256,274.

Accumulation of p13II in mitochondria results in disruption of the mitochondrial

inner membrane potential (∆ψ) with altered conductance to Ca2+ and K+. It also induces

mitochondrial swelling and fragmentation, suggesting a possible role of p13II in

apoptosis 256. Although p13II causes these alterations in mitochondrial morphology, the

protein does not change the permeability transition pore (PTP) driven by cation fluxes

or release of cytochrome c, which are two key events in some apoptosis pathway 256.

The potential role of p13II in cell population was implicated by Silic-Benussi et

al. 275. Expression of p13II resulted in the growth suppression in both HeLa and Jurkat

T cells in vitro, and reduced tumorgenicity of Ras and Myc co-transfected rat

embryonal fibroblasts (REFs). Hiraragi et al. 276 recently found that Jurkat T cells

expressing p13II were more sensitive to apoptosis when treated with apoptosis inducing

agents, such as ceramide and Fas ligand (FasL). Furthermore this apoptosis was

inhibited by a farnesyl transferase inhibitor (FTI) treatment, which blocks the

posttranslational modification of Ras to its active form, suggesting a role for p13II in

23

Ras-mediated cell signaling. These data indicate that exogenously expressed p13II

affects HTLV-1-induced lymphocyte proliferation by altering the balance in Ras-

mediated lymphocyte survival and death.

Cellular proteins that are physically associated with p13II have been identified

by yeast two-hybrid screening. p13II associated with the products of two cDNA clones,

C44 and C254, encoding nucleoside monophosphate kinase superfamily and actin-

binding protein 280 (ABP280), respectively 277. The protein product of C44 has

structural similarities to archeal adenylate kinases that is an eukaryotic mitochondrial

protein involved in energy metabolism. Interestingly, this protein is expressed in Jurkat

T cells and proliferating PBMC, but not quiescent PBMC 277. ABP280 is part of the

cytoskeleton and functions in the insertion of adhesion molecules into the cell

membrane 277. Furthermore, similar to an accessory protein of BLV, G4 that localizes to

mitochondria, p13II binds to farnesyl pyrophosphate synthetase (FPPS) 278. FPPS, which

is involved in the mevalonate/squalene pathway to synthesize of FPP, is a substrate

required for prenylation of Ras oncoprotein 279. Interestingly, Hiraragi et al. 276 found

recently that p13II mediated apoptosis sensitization was dependent on Ras mediated

signal pathway.

Although the biological significance of this viral accessory protein in the

pathogenesis of HTLV-1 infection remains still unclear, recent in vivo rabbit inoculation

with selectively p13II expression abrogated ACH cell line (729.ACH.p13) resulted in

significant reduced HTLV-1 infection, indicating the critical requirement of p13II for

establishment of HTLV-1 infection in vivo 280 .

24

1.7.2.2 p30II - A Selective Repressor of Transcription

HTLV-1 p30II mainly localizes to nucleus, specifically the nucleolus through

highly conserved bipartite nuclear localization signals (NLS) 241,281. It also contains

serine/threonine-rich regions that share distant homology to the activation domain of

transcription factors of the POU family, such as POU-2, Oct-1/2 and Pit-1 96,241. In

addition, p30II co-localizes and physically interacts with the cellular transcriptional co-

adaptor p300 in the nucleus 247. Taken together, these findings suggest that p30II has a

role in regulation of viral and cellular gene expression.

Recent studies indicate a role for p30II as a transcriptional regulator or a

negative regulator of mRNA expression. As a transcriptional regulator, p30II alters the

basal level of CRE and TRE - mediated transcription 247,250. Zhang et al.250 reported

that low concentrations of p30II stimulated HTLV-1 TRE-driven reporter gene activity,

whereas higher concentrations repressed LTR (TRE) and CRE-driven reporter gene

activity. Furthermore, p30II activated transcription through its central core region

(amino acid 62 and 220) 250. In a subsequent study, p30II - mediated transcriptional

activity was enhanced by CBP/p300, a critical co-adaptor of cell transcription, via

binding highly conserved KIX region of CBP/p300 247. Through DNA binding assays,

p30II inhibited the CREB-Tax-CBP/p300 complexes on TRE oligonucleotides 247

suggesting that p30II differentially regulates viral transcription via the sequestration of

CBP/p300.

25

Nicot et al. 282 demonstrated that expression of p30II results in selective nuclear

retention of spliced Tax/Rex mRNA resulting in decreased viral gene expression 282.

The mechanism by which p30II regulates the Tax/Rex mRNA retention in the nucleus is

still unclear. Because p30II does not contain a conserved RNA-binding motif, it may

bind directly via unknown motif or indirectly via ribonuceloprotein complexes which

are important for viral nuclear export. Recent studies by Younis et al. 283 support these

findings by showing that p28II, a homologue of p30II, regulates HTLV-2 gene

expression by a similar posttranscriptional manner. This same group reported that p30II

and p28II actually travel with transcription compex using chromatin

immunoprecipitation assay 284.

Microarray gene expression analyses of Jurkat T cells stably expressing p30II

indicated that p30II selectively represses many cellular genes involed in T-cell

activation, adhesion and apoptosis 285. Although the precise mechanism that p30II

affects gene regulation is still unclear, the critical role of p30II in HTLV-1 infection is

implicated from rabbit infection studies 286. Inoculation of the ACH.30 cell line, which

produces a selectively truncated of p30II, resulted in significantly reduced HTLV-1

proviral load in rabbits. Interestingly, sequencing results from ACH.30-inoculated

rabbits revealed a reversion to wild-type sequence, suggesting HTLV-1 requires p30II to

survive in vivo 286 .

26

1.7.3 pX ORF-I Protein: p12I

HTLV-1 p12I can be translated from singly spliced or doubly spliced ORF I

mRNA via initiation at an internal methionine codon. Because full-length p27I cDNA

expression plasmid produced only p12I protein, p27I mRNA was thought to be

preferentially used over the p12I message 242. However, p27I can be generated using in

vitro transcription-translation systems 96 and CTL responses against p27I specific

peptides were demonstrated in asymptomatic and disease subjects 260.

1.7.3.1 Structure of p12I

p12I is a small 99 amino acids hydrophobic protein with a high percent of

leucine (31%) and proline (17%) residues 242. The amino acid sequence of p12I is

highly conserved among viral samples of HTLV-1-infected individuals 47,287 Further

analyses of p12I suggested potential secondary α helix structures at amino acids 12-32

and amino acids 48-67 that function as transmembrane regions, which overlap putative

leucine zipper motifs (Fig. 1.2) 97,288. These distinct secondary structures are thought to

be involved in protein membrane localization or homo-oligomerization of the protein.

Indeed, p12I was demonstrated to form dimers by Trovato et al 288. Furthermore, p12I

contains four proline-rich (PXXP) Src homology 3 (SH3)-binding domains (Fig. 1.2) 47,

a motif implicated in cell signalling, suggesting that p12I plays a role in modulating cell

signal transduction. Among the four PXXP motifs, the first (at amino acids 8-11) and

the third (at amino acids 70-74) motifs are highly conserved among different HTLV-1

27

strains. Interestingly, these PXXP motifs are often preceded by an arginine residue at +2

position 289. In addition, p12I has one dileucine motif (DXXXLL) at amino acid 26-31

(Fig. 1.2), which is known as a sorting motif in HIV Nef 290. Although, its functional

role in p12I has not been determined, the dileucine motif in Nef is required for down

regulation of cell surface molecules, such as CD4 and MHC class I through endocytosis

and trafficking to the ER-Golgi compartments by associating with adapter protein 1

(AP-1), AP-2 and AP-3 290. Furthermore, in studies reported in Chapter 2 in this thesis

we have identified a conserved PSLP(I/L)T sequence in p12I (Fig. 1.2), which has

homology to the calcineurin-binding PxIxIT motif of nuclear factor of activated T cells

(NFAT) 244. In Chapter 2 of this thesis, the role of this motif in modulation of NFAT

activation is addressed.

Sequence analysis of p12I reveals potential post-translational modifications

sites such as ubiquitylation, glycosylation, and phosphorylation. Trovato et al. 288 have

reported that p12I is ubiquitylated on lysine residue at amino acid position 88 (Fig. 1.2).

Substitution of this lysine with an arginine resulted in increased stability of the protein.

Interestingly, they reported that natural alleles of HAM/TSP patients exclusively

contain a lysine residue at amino acid 88, whereas arginine is found in HTLV-1 strains

isolated from most ATLL patients and asymptomatic carriers, suggesting ubiquitylation

of p12I may result in different disease outcomes. However, an analysis by Martins et

al.291 indicted that lysine residue at this position was neither related to disease outcome,

nor is able to be used as a marker of progression to HAM/TSP, because only one out of

37 HAM/TSP patients carried the lysine residue and also one of 40 asymptomatic

HTLV-1 carriers had this rare phenotype. Additionally, HTLV-1 p12I has a potential N-

28

linked glycosylation site at amino acid 51 (asparagine) and multiple potential O-linked

glycosylation sites (serine and threonine). However, a deglycosylation study revealed

that p12I is not a glycoprotein 292. Through sequence analysis of p12I, we have found

several potential phophorylation sites such as protein kinase C (LTMR) site at amino

acid 75. However, p12I does not appear to be phosphorylated as demonstrated by a

phosphate metabolic labeling assay in 293T cells transiently expressing p12I

(unpublished data).

1.7.3.2 p12I Subcellular Localization and Protein Interaction

HTLV-1 p12I localizes in cellular endomembranes 241, predominantly in the ER

and cis-Golgi apparatus which is evidenced by immunofluorescent confocal microscopy,

electron microscopy and subcellular fractionation studies in 293T and Hela-Tat cells 292.

p12I was retained in the ER and cis-Golgi after blocking de novo protein synthesis or

disrupting ER-to-Golgi protein transport 292, suggesting p12I is neither transported to the

cell membrane nor secreted extracellularly. Thus, p12I localization in the ER is required

for its calcium-mediated NFAT activation 293.

Interestingly, two ER localizing proteins, calreticulin and calnexin, appear to

directly bind with exogenously expressed p12I 292. These two proteins are involved in

multiple cellular functions including calcium homeostasis, protein folding, integrin

mediated signaling, and function as molecular chaperones 294,295. The biological

significance of p12I interaction with these proteins in viral pathogenesis remains to be

elucidated.

29

The second putative trans-membrane domain of HTLV-1 p12I shares

approximately 50% amino-acid sequence homology with the bovine papilloma virus

(BPV) E5 protein and Epstein-Barr virus (EBV) LMP-1 protein 296,297. Similar to E5

protein of BPV, p12I binds to the 16 kD subunit of the vacuolar H+-ATPase (16K) 297,298.

When tested in a focus-formation assay, p12I itself did not induce transformation of

mouse C127 fibroblasts, however it enhanced the ability of E5 in inducing

transformation 297, suggesting a role of p12I in the transformation activity by enhancing

Tax oncogenic functions. However, a complex p12I and 16K does not correlate with the

transforming ability of p12I, while the E5 binding to 16 K appears to be important in the

E5-mediated transformation.

p12I was demonstrated to interact with the immature form of the interleukin-2

receptor β and γ chain in a transient over-expression system (Fig. 1.2) 299. Through the

central proline-rich region (amino acid 37-47), p12I interacts with these IL-2 receptors

288. Further studies demonstrated that the p12I-binding on the cytoplasmic domain of the

IL-2 receptor β chain was involved in the recruitment of Janus-associated kinases (Jak)

1 and Jak3 after IL-2 ligand binding 300. As a result of this interaction, the DNA binding

activity and transcriptional activity of signal transducers and activators of transcription-

5 (STAT5) were increased, suggesting p12I expression may decrease the threshold

required for T-cell activation. However, ACH.p12I cells, which are selectively p12I

expression abrogated, have no significant differences in IL-2 receptor chain (α, β, γc)

expression, in IL-2-mediated proliferation, or in IL-2-induced phosphorylated forms of

Stat3, Stat5, Jak1, or Jak3 when compared with wild type cell line ACH 301. These

30

controversial findings can be explained if p12I affects IL-2 receptor -mediated cell

signal during the early stages of HTLV-1 infection, not in the immortalized cell stage.

Furthermore, HTLV-1 p12I was demonstrated to associate with immature forms

of the major histocompatibility complex class I (MHC I) and inhibits interactions of

MHC I with β2-microglobulin, when these proteins were coexpressed in Hela-Tat cells,

which resulted in a decrease of surface expression of MHC I by directing the newly

synthesized MHC-I to the proteasome for degradation 302-304 These data suggest a role

of p12I in escape from immune surveillance. However, cell surface MHC I and MHC II

expression were not significantly different between PBMC immortalized by transfection

of wild type (ACH) and pX-ORF I ablated proviral clones (ACH.p12I) 301, indicating

that effect of p12I on MHC I expression is subtle in immortalized cells. Furthermore,

the possible function of p12I on escaping immune surveillance during the early stage of

viral infection is also unlikely, because p12I message abrogated ACH.p12I cells , which

were inoculated into rabbits, did not elicit strong immune response, while wild type

ACH cells did 253. Therefore, the early loss of viral infectivity of ACH.p12I cells is

unlikely due to immune mediated cell death nor p12I-mediated down regulation of

MHC I.

On work presented herein, we discovered that HTLV-1 p12I binds calcineurin

via a highly conserved PxIxIT binding motif 244 which is found in variety of calcineurin

binding proteins found in yeast, mammalian cells, and viruses 305. The biological

significance of this binding is still undetermined, but it may modulate p12I-mediated

calcium dependent signaling in T cells. In Chapter 2 of this thesis, I discussed the role

of this motif in modulating specifically NFAT activation in T cells.

31

1.7.3.3 Role of p12I in Viral Infectivity

Although pX ORF I of HTLV-1 was initially reported to be dispensable for viral

infectivity and primary lymphocyte transformation in vitro 243,306, studies from our

laboratory indicated that p12I is critically required for establish HTLV-1 infection in

vivo in a rabbit model and in primary T cells 251,253. Rabbits inoculated with ACH.p12I

failed to establish persistent infection, which was indicated by dramatic decreases in

humoral responses against HTLV-1 antigens, absence of viral antigen p19 production

from ex vivo PBMC culture, and transient detection of provirus by PCR 253. These

results suggested that p12I has a role in T cell activation, since most circulating

lymphocytes in vivo are quiescent or non-activated. Because initial studies used

activated PBMC, which were stimulated by IL-2 and phytohemagglutinin (PHA), as

target cells, the function of p12I as a T cell activator might be not required for

productive HTLV-1 infectivity in typical cell culture assays.

Our data supported a hypothesis that p12I is critical for T cell activation during

the early stages of viral infection 251. When ACH.p12I cells were cocultured with naive,

quiescent PBMC in the absence of exogenous stimuli, which more accurately reflect the

virus-cell interactions in vivo, a dramatic reduction in the viral infectivity resulted.

Furthermore, when T cell stimulators were added to the coculture, ACH.p12I cells were

restored in their ability to infect quiescent target T cells 251. These data indicated that

HTLV-1 p12I is required for efficient HTLV-1 infection in quiescent PBMC and

suggest a role for p12I in target T cell activation. The function of p12I is similar to that

32

of HIV-1 Nef, which is required for viral infectivity in quiescent T lymphocytes 307-309.

Interestingly, p12I complemented Nef function for efficient HIV-1 infection of

macrophages 310.

1.7.3.4 Role of p12I in Calcium- Mediated T Cell Signaling.

As implicated from in vivo and in vitro infectivity assays to test p12I function

and its structural properties, the viral protein role in T cell activation has become more

clear 245 . p12I was demonstrated to activate a major T cell transcription factor, NFAT,

in Jurkat T cells, without alteration of AP-1 or NFκB-mediated transcription 245. This

specific NFAT activation was dependent upon the Ras/MAP kinase pathway stimulated

by the phorbol ester, PMA. Further inhibition assays were performed to determine how

p12I activates NFAT. Inhibition of phospholipase C- γ (PLC-γ) and LAT (linker for

activation of T cells), which are upstream of calcium-mediated pathway, did not affect

NFAT activity, whereas inhibition of calcium-dependent signals by calcineurin inhibitor

(cyclosporin A), intracellular calcium chelator (BAPTA-AM), and a dominant negative

mutant of NFAT, abolished NFAT-dependent transcription, indicating direct p12I

function on calcium homeostasis in ER. Furthermore, the functional substitution of p12I

with thapsigargin, which released intracellular calcium from ER stores, supports the

hypothesis that p12I activates NFAT in a calcium-dependent manner (Fig. 1.3).

We have directly tested this hypothesis by measuring cytoplasmic calcium

concentration in Jurkat T cells 246. Expression of HTLV-1 p12I increases the basal

cytoplasmic calcium concentration and concurrently diminishes calcium available from

33

ER stores. Inhibition assays using inhibitors of inositol 1,4,5-triphosphate receptor

(IP3R) or calcium release-activated calcium channels (CRAC) suggested that p12I-

mediated NFAT activation occurs through on IP3R mediated the calcium release from

ER stores and subsequent by extracellular calcium entry via CRAC. Moreover, we

found that p12I stable expression enhanced the production of interleukin-2 (IL-2) which

is a downstream gene of calcium-NFAT mediated pathway, in Jurkat T cells and

primary lymphocytes Fig. 1.3) 293.

In Chapter 2 of this thesis, I provide evidence that p12I can regulate NFAT

activity either positively via elevating cytosolic calcium from ER stores or negatively

by calcineurin binding through PxIxIT motif 244. Although the significance of this

opposed regulatory functions of p12I is still unclear, there are several similarities

between p12I and cellular proteins such as CAML (Ca2+-modulating cyclophilin ligand)

and Bcl-2. CAML which contains two putative transmembrane domains localizes in the

ER and induces calcium release from the ER and leads to NFAT activation like HTLV-

1 p12I. Also CAML indirectly binds with calcineurin through cyclophilin binding. An

anti-apoptotic protein, Bcl-2 also has similar functional properties with p12I: calcium

release from the ER and calcineurin binding 311,312. Bcl-2 localizes to the ER and

mitochondria and maintains calcium homeostasis, suggesting its functions as an ion

channel protein 313 On the other hand, Bcl-2 binds calcineurin via its BH4 domain, and

inhibits NFAT activity by sequestering calcineurin from NFAT binding without

affecting calcineurin catalytic activity 314,315. Therefore, p12I may affect apoptosis in

HTLV-1 infected T cells similar to Bcl-2.

34

Furthermore, in a subsequent study, our laboratory performed gene array

analysis in stably p12I expressing Jurkat T cells to test if p12I regulates the expression

of cellular genes in a calcium-dependent manner 316. Alteration of genes by p12I

involved cell proliferation, apoptosis, cell adhesion, immune response modulation and T

cell signaling predominantly in a calcium-dependent manner. Interestingly, p12I

expression induced increased expression of p300, which is a key cellular transcriptional

co-adaptor 316 and a co-activator for HTLV-1 LTR transcription 172. p300/CBP functions

as a co-adaptor of transcription at the HTLV-1 promoter and plays a critical role in the

regulation of Tax-dependent HTLV-1 transcription in infected T-cells 170. HTLV-1 Tax

trans-activates HTLV-1 and cellular gene transcription through its interaction with the

p300 and CBP co-activators 317. Therefore, p12I may enhance Tax-mediated HTLV-1

viral gene expression and infection by increasing p300 expression. In addition, p12I

may modulate clonal expansion, cell survival and transformation of HTLV-1 infected

cells through p300 and Tax mediated transactivation.

Calcium release from the ER by p12I and subsequent activation of calcium

mediated cell signal transduction including stimulation of NFAT 246, increased IL-2

production 293 and increased p300 expression316, might be critical during the early

stages of HTLV-1 infection. For HIV-1 active replication, activation of NFAT is

required in primary T cells. NFAT activation is sufficient to overcome a blockade at

reverse transcription and induces highly permissive state for HIV-1 replication 318.

Therefore, p12I may induce NFAT activation to induce a permissive state for active

HTLV-1 early replication. Moreover, by increasing production of cytokine IL-2, p12I

35

activates T lymphocytes division, which is a prerequisite for the retroviral provirus to

integrate into the host cell genome and permit the subsequent viral replication and

productive infection. p12I expression in HTLV-1-infected T lymphocytes may lower the

calcium threshold in order to respond to weak stimuli, which would normally not

activate T cells. These stimuli may come from direct contact via T cell receptor or cell

surface receptors or soluble factors such as cytokines and chemokines 319.

Interestingly, many viral proteins modulates the cellular Ca2+ homeostasis

during virus replication or pathogenesis 320. Hepatitis B virus X protein (HBx) 321,

rotavirus NSP4 322, and protein K7 of kaposi's sarcoma-associated herpesvirus (KSHV)

323 increase cytoplasmic calcium for viral replication. Also, HIV-1 Nef, which plays a

role in viral pathognesis, modulates calcium homeostasis through IP3R mediated

pathway to activate NFAT 307. Therefore, modulation of calcium homeostasis by p12I

may play a role in HTLV-1 viral replication and pathogenesis.

In the Chapter 3 of this thesis, I tested the influence of p12I in another calcium -

calmodulin mediated T cell signal pathway. The expression of p12I in Jurkat T cells

induced cell surface clustering of LFA-1, which was inhibited by calcium chelator

BAPTA-AM, calcium channel blocker SK&F 96365, and calpeptin, an inhibitor of the

calcium-dependent protease calpain. These data indicate that HTLV-1 p12I induces

LFA-1 clustering through calcium dependent signaling and may influence cell-to-cell

transmission of virus.

36

1.8 HTLV-1 Cell-to-Cell Transmission

Unlike HIV-1, naturally HTLV-1 infected lymphocytes produce limited cell-free

viral particles, and these are produced only during in vitro conditions such as

continuous culture of transfected or producer cell lines 64,324. Moreover, HTLV-1 is

poorly infectious as cell-free virions 19: only one out of 105 to 106 particles is infectious

325. Cell-to-cell contact between virus infected and target cells is required for efficient

transmission 20,324,326.

The mechanism of cell-to-cell spread of HTLV-1 is still not completely

understood. However, the interaction between HTLV-1 envelope proteins and its

receptor or between cellular adhesion molecules are thought to be required, as cell

fusion or syncytium formation is induced by cell surface expression of envelope protein

327-329, or cellular ICAM-1, ICAM-3, and VCAM adhesion molecules123,330.

A recent confocal observation by Igakura et al 331 indicated that the structure of

spontaneously formed cell conjugate between HTLV-1 infected cell and target

lymphocytes resembles an immunologic synapse, which is a specialized contact

structure made between a lymphocyte and another cell, such as an antigen-presenting

cell. They found that polarization of the adhesion molecule talin and the microtubule

organizing centre (MTOC) to the cell contact areas with accumulation of core protein

Gag and the viral genome at the cell-to-cell junction, followed by transfer of both the

Gag protein and the HTLV-1 genome to the uninfected cell. They termed this structure

the “virological synapse” and highlighted on important difference between the

37

virological and the immunological synapse: the MTOC is oriented toward the synapse

in an immunological synapse, but the MTOC is polarized inside the HTLV-1–infected

cell in a virological synapse. Therefore, they suggested that these events were not

caused by T cell receptor (TCR)-mediated recognition of HTLV-1 antigens, but

triggered by HTLV-1 infection and cell-to-cell contact 64.

In a subsequent study, Barnard et al 332 investigated T-cell surface molecules

involved in triggering the polarization of MTOC in the HTLV-1–associated virological

synapse. They carried out antibody-coated bead-cell conjugate formation assays using

antibodies against T-cell surface molecules such as CD2 and CD3, LFA-1 (CD11a and

CD18), CD28, ICAM-1 and CD25, which are involved in T cell activation or up-

regulated in HTLV-1 infected cells. They found that engagement of ICAM-1 and LFA-

1 molecules caused MTOC polarization in HTLV-1–infected cells. Moreover, they

exhibited that that ICAM-1 was more highly associated with MTOC polarization than

LFA-1 molecules by performing inhibitory assay using soluble inhibitory cyclic

peptides or LFA1 mutated cell lines. These authors strongly suggested that the ICAM-

1/LFA-1 interaction plays an important role in polarization of MTOC and formation of

virological synapse. The increased expression of ICAM-1 in HTLV-1 infection 333 may

facilitate more HTLV-1 viral cell-to-cell spread. Interestingly, in HIV-1 viral

transmission or syncytium formation, ICAM-1/LFA-1 interaction is important 334-336.

In addition to cellular factors, Nejmeddine et al 337 tested if HTLV-1 viral

protein Tax affected MTOC polarization and virological synapse formation. They found

that Tax protein is present in cis-Golgi compartment around the MTOC and in the cell-

38

to-cell contact area, as well as in the nucleus. MTOC and Tax were always polarized in

the same direction and this MTOC/Tax reorientation was dependent on the function of

the Rac family and Cdc42 small GTPase, which are important in cell movement and

polarization 338. Expression of Tax sufficiently induced polarization of the MTOC

toward the target cell, suggesting that Tax promotes cell-to-cell spread of HTLV-1

through virological synapse.

In Chapter 3 of this thesis, I have tested if HTLV-1 nonstructural protein p12I

affects LFA-1 mediated T cell adhesion believed to be important in MTOC

reorientation and virological synapse formation and in facilitating HTLV-1 cell-to-cell

spread. Interaction between LFA-1 and its ligand ICAM-1 is required in the formation

of the immunologic synapse 339 and plays role in reducing the threshold for T-cell

activation 340. Moreover, signal transduction through LFA-1 (outside in signal) which

regulates cytoskeletal rearrangement and cell movement 341, may play a critical role in

MTOC polarization and virological synapse formation. Therefore, HTLV-1 p12I may

enhance HTLV-1 cell-to-cell transfer via promoting LFA-1 mediated adhesion and

virological synapse formation.

1.9 Regulation of Integrin Activity

Integrins are heterodimeric adhesion receptors composed of α and β subunits. T

cells possess at least 12 of the 24 known integrin heterodimers, the pattern of expression

is changed depending on maturation status or subset of T cells. The β1 subunit integrins

(α1-6β), which are commonly found in many cell types, bind extracellular matrix. There

39

are four kinds of leukocyte-specific ß2 integrins αLβ2, αMβ2, αXβ2 and αDβ2) on T

cells, αLβ2 (leukocyte function-associated antigen-1; LFA-1) is most the abundant T

cell integrin 341-343.

Integrins are required for all of T cell function, such as migration to peripheral

lymph nodes and inflammatory sites, and formation of the immunological synapse in

antigen presentation and cytotoxicity. Because the ligands for integrins are widely

expressed in tissue and cells, the activity of integrins are strictly regulated. Integrins are

adhesive to their ligands only when cells were exposed to stimuli such as engagement of

T cell receptors or cytokines and chemokines stimulation. This process termed ‘inside-

out signaling’ starts from intracellular signals that trigger rapid integrin activation. In

addition, ligand binding to the integrin itself transfers intracellular signal transduction

(outside–in signaling) to regulate cytoskeletal polarization and T cell movement and

migration 341,343.

Inside-out signaling causes both clustering of integrins on the cell surface, or an

increase in the affinity for ligand through conformational change. Experimentally

addition of the divalent cations Mn2+ or Mg2+ with EGTA or specific antibodies that

bind to the extracellular portion of LFA-1 cause conformational and affinity alteration

344-346. However, identification of physiological affinity modulation is difficult, because

the increase of LFA-1 affinity is rapid and transient. Therefore, the mechanism of

inside out signaling to activate LFA-1 affinity is not well understood compared to that

of LFA-1 clustering 341. However recent studies exhibited that a few microliters of

chemokine could induce affinity increase, whereas lasting stimulation induced LFA-1

clustering 347. Moreover, talin, Rap1 and RapL proteins are reported to be required for

40

affinity regulation of LFA-1 343,348. The cytoskeletal protein talin is associated with the

integrin cytoplasmic domain, modulating affinity for β1, β2 and β3 integrins 349.

Feigelson et al. 350 showed that functional Src-family kinase p56 (Lck) is required for

maintaining high affinity status of integrin α4β1 on T cells, which leads to rapid binding

of circulating T cell to endothelium.

Release of integrins from the cytoskeletal network is thought to permit increased

lateral movement and clustering of integrins. From TCR to LFA-1 clusteirng, key

signaling proteins such as Vav-1 (guanine exchange factor), ADAP, SKAP-55 (adaptor

protein), Rap1 (small GTPase) and RAPL (Rap1 binding protein) were identified

through studies of transgenic and knockout mice models 341,351-354. For example, LFA-

1clustering is induced by activation of GTPase Rac, which is stimulated by signal

cascade from TCR activaiton, phosphorylation of SKAP-55, a subsequent binding with

ADAP and SLP-76, and complex formation with Vav-1, a GEF for GTPase Rac 349,355.

Furthermore, protein kinase C (PKC) activation by phorbol esters triggers integrin

clustering by interacting with β2 subunit and by phosphorylating residuces in its

cytoplasmic tail 354,356. Moreover, elevation of cytoplasmic calcium by TCR activation

or by treatment with calcium mobilizers such as ionomycin and thapsigargin 345 can

induce clustering of LFA-1. Phorbol esters and calcium mediated LFA-1 activation

also requires the calcium-dependent protease calpain, which is suggested to play a role

in releasing integrin from cytoskeleton 345. Interestingly, the function of HTLV-1 p12I

is similar to that of thapsigargin i.e. inducing calcium release from the ER stores to

activate T cells 246.

41

Signal transduction through activated LFA-1 (outside-in signaling) can cause T

cell attachment and lamellipodial movement at the leading edge by activating myosin

light chain kinase (MLCK), whereas it leads T cell detachment at trailing edge through

RhoA and Rho kinase (ROCK) mediated pathway 341,357,358. Interestingly, LFA-1

interacts with the microtubule network, which are required for HTLV-1 cell-to-cell

transmission 331. LFA-1 engagement triggers translocation of Pyk-2, PKCβI and PKCδ,

which are known to important in T cell migration, to the MTOC 359,360. Therefore,

signaling from activated LFA-1 may enhance translocation of HTLV-1 viral or cellular

proteins, which are required for HTLV-1 cell-to-cell spread, to the MTOC.

Furthermore, these reports suggest that LFA-1 activation by HTLV-1 p12I also may

promote cell-to-cell transmission through LFA-1 mediated microtubular system

activation (MTOC polarization).

1.10 Bystander Cell Activation by HTLV-1

One of immunological hallmarks of HTLV-1 infection is spontaneous

proliferation of PBMC, which are isolated from HAM/TSP patients and asymptomatic

carriers, in vitro without exogenous antigens or stimulants 361-364. The magnitude of

spontaneous proliferation is more pronounced in HAM/TSP patients than in

asymptomatic carriers361-364. It is suggested that spontaneous proliferation of PBMC is

due to Tax mediated up regulation of IL2- IL2 receptor autocrine loop, because Tax

protein, which is not detected in fresh ex vivo PBMC, but can be found after several

hours of culture in vitro 365,366. Moreover, Wucherpfennig et al 367 and Kimata et al 368

42

demonstrated that direct interaction between HTLV-1 infected cells and uninfected

resting T cells induced spontaneous proliferation.

Proliferation or stimulation of uninfected PBMC was induced by direct

interaction with irradiated or fixed HTLV-1 infected cells, but not by isolated virus

particles nor soluble factors produced from infected cells 367,368. This activation process

was blocked by monoclonal antibodies that inhibit interaction of CD2/LFA-3 (CD58),

LFA-1/ICAM-1 (CD54) and IL2/ IL-2 receptor (CD25). The strong inhibition was

found in treatment of antibodies against CD2 and CD58. However, antibodies against

class I or class II MHC complex or against various regions of the HTLV-1 envelope

proteins gp46 and gp21 did not block activation of uninfected cells. These data indicate

that envelope proteins of HTLV-1 viral particles are not mitogenic. Instead, target

uninfected cell activation by HTLV-1 infected cells requires cell-to-cell contact via

interaction of cell surface molecules especially CD2 and LFA-3 (CD58).

Our previous data suggest that the role of p12I is related with T cell activation

during the early HTLV-1 infection, and demonstrated that p12I activates LFA-1

clustering in HTLV-1 infected cells, therefore, interaction between LFA-1 and ICAM-1

may activate target cell T cell activation. I present data in Chapter 4 of this thesis to

address the role of p12I in activation of uninfected target cells.

Overall, data presented in this thesis provided further evidence that a highly

conserved viral accessory protein, p12I is critically required for early HTLV-1 infection.

I demonstrated in this thesis that p12I regulates NFAT activation by interacting with

43

calcineurin, enhances cell-to-cell adhesion by inducing LFA-1 clustering and facilitates

early viral infectivity via regulating viral gene expression during early cell-to-cell

transmission.

1.11 REFERENCES

(1) Poiesz BJ, Ruscetti FW, Gazdar AF et al. 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. 1980;77:7415-7419.

(2) Hinuma Y, Nagata K, Hanaoka M et al. 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. 1981;78:6476-6480.

(3) Hinuma Y, Komoda H, Chosa T et al. 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. 1982;29:631.

(4) Miyoshi I, Kubonishi I, Yoshimoto S et al. Type C virus particles in a cord T-cell line derived by co-cultivating normal human cord leukocytes and human leukaemic T cells. Nature. 1981;294:770-771.

(5) Yoshida M, Miyoshi I, Hinuma Y. Isolation and characterization of retrovirus from cell lines of human T-cell leukemia and its implication in the disease. Proc Natl Acad Sci USA. 1982;79:2031-2035.

(6) Takatsuki K, Uchiyama T, Sagawa K, Yodoi J. Adult T-cell leukemia in Japan. In: Seno S, Takaku F, Irino S, eds. Topics in Hematology. Amsterdam: Excerpta Medica; 1977:73-77.

(7) Uchiyama T, Yodoi J, Sagawa K, Takatsuki K, Uchino H. Adult T-cell leukemia: clinical and hematologic features of 16 cases. Blood. 1977;50:481-492.

(8) Gessain A, Barin F, Vernant J et al. Antibodies to human T lymphotropic virus

type 1 in patients with tropical spastic paresis. Lancet. 1985;2:407-410.

(9) Osame M, Usuku K, Izumo S et al. HTLV-I associated myelopathy, a new clinical entity. Lancet. 1986;i:1031-1032.

(10) Gessain A, Mahieux R. [A virus called HTLV-1. Epidemiological aspects]. Presse Med. 2000;29:2233-2239.

44

(11) Edlich RF, Arnette JA, Williams FM. Global epidemic of human T-cell

lymphotropic virus type-I (HTLV-I). J Emerg Med. 2000;18:109-119.

(12) Blattner W, Kalyanaraman V, Robert-Guroff M et al. The human type-C retrovirus, HTLV, in blacks from the Caribbean region and relationchip to adult T-cell leukemia/lymphoma. Int J Cancer. 1982;30:257.

(13) Saxinger WC, Blattner W, Levine P et al. Human T-cell leukemia virus (HTLV-I) antibodies in Africa. Science. 1984;225:1473.

(14) Reeves W, Saxinger C, Brenes M et al. Human T-cell lymphotropic virus type I (HTLV-I) and risk factors in metropolitan Panama. Am J Epidemiol. 1988;127:539.

(15) Merino F, Robert-Guroff M, Clark J et al. Natural antibodies to human T-cell leukemia/lymphoma virus in healthy Venezuelan populations. Int J Cancer. 1984;34:501.

(16) Asher DM, Goudsmit J, Pomeroy K et al. Antibodies to HTLV-I in populations of the southwestern Pacific. J Med Virol. 1988;26:339.

(17) Blayney DW, Blattner WA, Robert-Guroff M et al. The human T-cell leukemia-lymphoma virus in the southeastern United States. JAMA. 1983;250:1048-1052.

(18) Gessain A, Mahieux R. [A virus called HTLV-1. Epidemiological aspects]

Presse Med. 2000;29:2233-2239. (19) Hollsberg P, Hafler D. Pathogenesis of diseases induced by human lymphotropic

virus type I infection. N Engl J Med. 1993;328:1173-1182.

(20) Yamamoto N, Okada M, Koyanagi Y, Kannagi M, Hinuma Y. Transformation of human Leukocytes by cocultivation with an adult T cell leukemia virus producer cell line. Science. 1982;217:737-739.

(21) Hino S, Yamaguchi K, Katamine S et al. Mother-to-child transmission of human T-cell leukemia virus type-I. Jpn J Cancer Res. 1985;76:474-480.

(22) Hino S. Milk-borne transmission of HTLV-I as a major route in the endemic

cycle. Acta Paediatr Jpn. 1989;31:428-435. (23) Fujino T, Fujiyoshi T, Yashiki S et al. HTLV-I transmission from mother to

fetus via placenta. Lancet. 1992;340:1157. (24) Wiktor SZ, Pate EJ, Murphy EL et al. Mother-to-child transmission of human T-

cell lymphotropic virus type I (HTLV-I) in Jamaica: association with antibodies

45

to envelope glycoprotein (gp46) epitopes. J Acquir Immune Defic Syndr. 1993;6:1162-1167.

(25) Manns A, Wilks RJ, Murphy EL et al. A prospective study of transmission by transfusion of HTLV-I and risk factors associated with seroconversion. Int J Cancer. 1992;51:886-891.

(26) Khabbaz RF, Douglas JM, Judson FN et al. Seroprevalence of human T-lymphotropic virus type I or II in sexually transmitted disease clinic patients in the USA. J Infect Dis. 1990;162:241-244.

(27) Khabbaz RF, Onorato IM, Cannon RO et al. Seroprevalence of HTLV-1 and HTLV-2 among intravenous drug users and persons in clinics for sexually transmitted diseases. N Engl J Med. 1992;326:375-380.

(28) Nagai M, Osame M. Human T-cell lymphotropic virus type I and neurological diseases. J Neurovirol. 2003;9:228-235.

(29) Courouce AM, Pillonel J, Lemaire JM, Saura C. HTLV testing in blood transfusion. Vox Sang. 1998;74 Suppl 2:165-169.

(30) Ferreira OC, Planelles V, Rosenblatt JD. Human T-cell leukemia viruses: Epidemiology, biology, and pathogenesis. Blood Rev. 1997;11:91-104.

(31) Murphy EL, Glynn SA, Fridey J et al. Increased incidence of infectious diseases during prospective follow-up of human T-lymphotropic virus type II- and I-infected blood donors. ARCH INTERN MED. 1999;159:1485-1491.

(32) Murphy EL, Busch MP, Tong M, Cornett P, Vyas GN. A prospective study of the risk of transfusion-acquired viral infections. Transfus Med. 1998;8:173-178.

(33) Schreiber GB, Murphy EL, Horton JA et al. 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. 1997;14:263-271.

(34) Stuver SO, Tachibana N, Okayama A et al. 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. 1993;167:57-65.

(35) Anderson DW, Epstein JS, Lee TH et al. Serological confirmation of human T-lymphotropic virus type I infection in healthy blood and plasma donors. Blood. 1989;74:2585-2591.

46

(36) Lairmore MD, Lal R. Other Human Retrovirus Infections: HTLV-I and HTLV-II. In: Schocketman G, George JR, eds. AIDS Testing. Methodology and Management Issues. New York: Springer-Verlag; 1995:168-188.

(37) Toedter GP, Pearlman S, Hofheinz D et al. Development of a monoclonal antibody-based p24 capsid antigen detection assay for HTLV-I, HTLV-II, and STLV-I infection. AIDS Res Hum Retroviruses. 1992;8:527-532.

(38) Bangham CR. HTLV-1 infections. J Clin Pathol. 2000;53:581-586.

(39) Arisawa K, Soda M, Endo S et al. Evaluation of adult T-cell leukemia/lymphoma incidence and its impact on non-Hodgkin lymphoma incidence in southwestern Japan. Int J Cancer. 2000;85:319-324.

(40) Shimoyama M. 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. 1991;79:428-437.

(41) Kamihira S, Sohda H, Atogami S et al. Phenotypic diversity and prognosis of

adult T-cell leukemia. Leuk Res. 1992;16:435-441.

(42) Nicot C. Current views in HTLV-I-associated adult T-cell leukemia/lymphoma. Am J Hematol. 2005;78:232-239.

(43) Kawano F, Yamaguchi K, Nishimura H, Tsuda H, Takatsuki K. Variation in the clinical courses of adult T-cell leukemia. Cancer. 1985;55:851-856.

(44) Bazarbachi A, Hermine O. Treatment of adult T-cell leukaemia/lymphoma: current strategy and future perspectives. Virus Res. 2001;78:79-92.

(45) Yao J, Wigdahl B. Human T cell lymphotropic virus type I genomic expression and impact on intracellular signaling pathways during neurodegenerative disease and leukemia. Front Biosci. 2000;5:D138-D168.

(46) Yoshida M, Seiki M, Yamaguchi K, Takatsuki K. 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. 1984;81:2534-2537.

(47) Franchini G. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood. 1995;86:3619-3639.

(48) Hollsberg P, Wucherpfennig KW, Ausubel LJ et al. Characterization of HTLV-1 In Vivo Infected T Cell Clones IL-2-independent Growth of Nontransformed T Cells. J Immunol. 1992;148:3256-3263.

47

(49) Gatza ML, Watt JC, Marriott SJ. Cellular transformation by the HTLV-I Tax protein, a jack-of-all-trades. Oncogene. 2003;22:5141-5149.

(50) Franchini G, Nicot C, Johnson JM. Seizing of T cells by human T-cell leukemia/lymphoma virus type 1. Adv Cancer Res. 2003;89:69-132.

(51) Osame M, Janssen R, Kubota H et al. Nationwide survey of HTLV-I-associated Myelopathy in Japan: Association with blood transfusion. Ann Neurol. 1990;28:50-56.

(52) Uchiyama T. Human T cell leukemia virus type I (HTLV-I) and human diseases. [Review] [134 refs]. Annual Review of Immunology. 1997;15:15-37:15-37.

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

(54) Kasahata N, Shiota J, Miyazawa Y, Nakano I, Murayama S. Acute human T-lymphotropic virus type 1-associated myelopathy: a clinicopathologic study. Arch Neurol. 2003;60:873-876.

(55) Furukawa Y, Kubota R, Eiraku N et al. 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. 2003;32:328-334.

(56) Kiwaki T, Umehara F, Arimura Y et al. The clinical and pathological features of peripheral neuropathy accompanied with HTLV-I associated myelopathy. J Neurol Sci. 2003;206:17-21.

(57) Molgaard CA, Eisenman PA, Ryden LA, Golbeck AL. Neuroepidemiology of human T-lymphotrophic virus type-I-associated tropical spastic paraparesis. Neuroepidemiology. 1989;8:109-123.

(58) Castillo LC, Gracia F, Roman GC et al. 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. 2000;101:405-412.

(59) Alamy AH, Menezes FB, Leite AC, Nascimento OM, Araujo AQ. Dysautonomia in human T-cell lymphotrophic virus type I-associated myelopathy/tropical spastic paraparesis. Ann Neurol. 2001;50:681-685.

(60) Osame M. Pathological mechanisms of human T-cell lymphotropic virus type I-associated myelopathy (HAM/TSP). J Neurovirol. 2002;8:359-364.

(61) Jacobson S. Immunopathogenesis of human T cell lymphotropic virus type I-associated neurologic disease. J Infect Dis. 2002;186 Suppl 2:S187-S192.

48

(62) Nitta T, Tanaka M, Sun B, Hanai S, Miwa M. The genetic background as a determinant of human T-cell leukemia virus type 1 proviral load. Biochem Biophys Res Commun. 2003;309:161-165.

(63) Bangham CR. Human T-lymphotropic virus type 1 (HTLV-1): persistence and immune control. Int J Hematol. 2003;78:297-303.

(64) Bangham CR. The immune control and cell-to-cell spread of human T-lymphotropic virus type 1. J Gen Virol. 2003;84:3177-3189.

(65) Yamano Y, Nagai M, Brennan M et al. 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. 2002;99:88-94.

(66) Kitze B, Usuku K. HTLV-1-mediated immunopathological CNS disease. Curr Top Microbiol Immunol. 2002;265:197-211.

(67) Osame M, Izumo S, Igata A et al. Blood transfusion and HTLV-I-associated myelopathy. Lancet. 1986;104.

(68) Osame M, Igata A, Usuku K, Rosales RL, Matsumoto M. Mother-to-child transmission in HTLV-I associated myelopathy. Lancet. 1987;1:106.

(69) Nagai M, Usuku K, Matsumoto W et al. 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. 1998;4:586-593.

(70) Daenke S, Nightingale S, Cruickshank JK, Bangham CRM. 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. 1990;64:1278-1282.

(71) Evangelista A, Maroushek S, Minnigan H et al. Nucleotide sequence analysis of a provirus derived from an individual with tropical spastic paraparesis. Microb Pathog. 1990;8:259-278.

(72) Gould KG, Bangham CR. Virus variation, escape from cytotoxic T lymphocytes and human retroviral persistence. Semin Cell Dev Biol. 1998;9:321-328.

(73) Mahieux R, Ibrahim F, Mauclere P et al. 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. 1997;71:1317-1333.

49

(74) Saito M, Furukawa Y, Kubota R et al. 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. 1995;1:286-294.

(75) Kuroda Y, Matsui M. Cerebrospinal Fluid Interferon-Gamma Is Increased in HTLV- I- Associated Myelopathy. J Neuroimmunol. 1993;42:223-226.

(76) Matsui M, Nagumo F, Tadano J, Kuroda Y. Characterization of humoral and cellular immunity in the central nervous system of HAM/TSP. J Neurol Sci. 1995;130:183-189.

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

(78) Kawakami K, Miyazato A, Iwakura Y, Saito A. Induction of lymphocytic inflammatory changes in lung interstitium by human T lymphotropic virus type I. AMER J RESPIR CRIT CARE MED. 1999;160:995-1000.

(79) Levin MC, Jacobson S. HTLV-I associated myelopathy/tropical spastic paraparesis (HAM/TSP): A chronic progressive neurologic disease associated with immunologically mediated damage to the central nervous system. J Neurovirology. 1997;3:126-140.

(80) Levin MC, Lee SM, Kalume F et al. Autoimmunity due to molecular mimicry as a cause of neurological disease. Nat Med. 2002;8:509-513.

(81) Levin MC, Lee SM, Morcos Y, Brady J, Stuart J. Cross-reactivity between immunodominant human T lymphotropic virus type I tax and neurons: implications for molecular mimicry. J Infect Dis. 2002;186:1514-1517.

(82) Mochizuki M, Watanabe T, Yamaguchi K et al. HTLV-I Uveitis: A distinct clinical entity caused by HTLV-I. Jpn J Cancer Res. 1992;83:236-239.

(83) Nishioka K, Maruyama I, Sato K et al. Chronic inflammatory arthropathy associated with HTLV-I [letter]. Lancet. 1989;i:441.

(84) Terada K, Katamine S, Eguchi K et al. Prevalence of serum and salivary antibodies to HTLV-1 in Sjogren's syndrome. Lancet. 1994;344:1116-1119.

(85) Lagrenade L, Hanchard B, Fletcher V, Cranston B, Blattner W. Infective dermatitis of Jamaican children: a marker for HTLV-I infection. Lancet. 1990;336:1345-1347.

50

(86) Leonmonzon M, Illa I, Dalakas MC. 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. 1994;36:643-649.

(87) Ohshima K, Kikuchi M, Masuda Y et al. Human T-cell leukemia virus type I associated lymphadenitis. Cancer. 1992;69:239-248.

(88) Kimura I, Tsubota T, Tada S, Sogawa J. Presence of antibodies against adult T

cell leukemia antigen in the patients with chronic respiratory diseases. Acta Med Okayama. 1986;40:281-284.

(89) Xu RZ, Gao QK, Wang SJ et al. Human acute myeloid leukemias may be

etiologically associated with new human retroviral infection. Leuk Res. 1996;20:449-455.

(90) Yajima A, Kawada A, Aragane Y, Tezuka T. Detection of HTLV-I proviral DNA in sarcoidosis. Dermatology. 2001;203:53-56.

(91) Terashima A, Gotuzzo E, Alvarez H et al. [STRONGYLOIDES STERCORALIS: CLINICAL SEVERE FORMS ASSOCIATED TO HTLV-1 INFECTION]. Rev Gastroenterol Peru. 1999;19:35-40.

(92) Coffin JM. Structure and Classification of Retroviruses. The Retroviridae. 1992;1:19-49.

(93) Coffin JM. Retroviridae: The Viruses and Their Replication. 1767-1835. 1996. Phildaelphia, PA, Lippinocott-Raven Publishers.

(94) Seiki M, Hattori S, Hirayama Y, Yoshida M. Human adult T-cell leukemia

virus: Complete nucleotide sequence of the provirus genome integrated in leukemia cell DNA. Proc Natl Acad Sci. 1983;80:3618-3622.

(95) Berneman ZN, Gartenhaus RB, Reitz MS et al. 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. 1992;89:3005-3009.

(96) Ciminale V, Pavlakis GN, Derse D, Cunningham CP, Felber BK. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J Virol. 1992;66:1737-1745.

(97) Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type 1. Proc Natl Acad Sci USA. 1992;89:8813-8817.

51

(98) Shimotohno K, Takano M, Teruuchi T, Miwa M. 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. 1986;83:8112-8116.

(99) Brady JN, Jeang KT, Duvall J, Khoury G. Identification of p40x-Responsive

Regulatory Sequences within the Human T-Cell Leukemia Virus Type 1 Long Terminal Repeat. J Virol. 1987;61:2175-2181.

(100) Kitado H, Chen IS, Shah NP et al. U3 sequences from HTLV-I and -II LTRs confer pX protein response to a murine leukemia virus LTR. Science. 1987;235:901-904.

(101) Tillmann M, Wessner R, Wigdahl B. Identification of human T-cell

lymphotropic virus type I 21-base- pair repeat-specific and glial cell-specific DNA- protein complexes. J Virol. 1994;68:4597-4608.

(102) Wessner R, Wigdahl B. AP-1 derived from mature monocytes and astrocytes preferentially interacts with the HTLV-I promoter central 21 bp repeat. Leukemia. 1997;11 Suppl 3:21-4.:21-24.

(103) Wessner R, Tillmann-Bogush M, Wigdahl B. Characterization of a glial cell-specific DNA-protein complex formed with the human T cell lymphotropic virus type I (HTLV-I) enhancer. J Neurovirol. 1995;1:62-77.

(104) Grant C, Barmak K, Alefantis T et al. 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. 2002;190:133-159.

(105) Oroszlan S. Chemical Analysis of HTLV-1 Structural proteins. In: Gallo RC, Essex ME, Gross L, eds. Human T-cell Leukemia Lymphoma Viruses. Cold Spring Harbor Laboratory Press; 1984:101-110.

(106) Kiyokawa T, Seiki M, Iwashita S et al. p27x-III and p21x-III, proteins encoded

by the pX sequence of human T-cell leukemia virus type 1. Proc Natl Acad Sci USA. 1985;82:8359-8363.

(107) Seiki M, Hikikoshi A, Taniguchi T, Yoshida M. Expression of the pX Gene of HTLV-1: General Splicing Mechanism in the HTLV Family. Science. 1985;228:1532-1534.

(108) Furukawa K, Furukawa K, Shiku H. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 1991;295:141-145.

52

(109) Cereseto A, Berneman Z, Koralnik I et al. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia. 1997;11:866-870.

(110) Gaudray G, Gachon F, Basbous J et al. The complementary strand of the human T-cell leukemia virus type 1 RNA genome encodes a bZIP transcription factor that down-regulates viral transcription. J Virol. 2002;76:12813-12822.

(111) Ootsuyama Y, Shimotohno K, Miwa M, Oroszlan S, Sugimura T. Myristylation of gag protein in human T-cell leukemia virus type-I and type-II. Jpn J Cancer Res. 1985;76:1132-1135.

(112) Hattori S, Kiyokawa T, Imagawa K et al. Identification of gag and env gene products of human T-cell leukemia virus (HTLV). Virology. 1984;136:338-347.

(113) Craven RC, Parent LJ. Dynamic interactions of the gag polyprotein. In: Krausslich HG, ed. Morphogenesis and Maturation of Retroviruses. Springer-Verlag Berl; 1996:65-94.

(114) Garnier L, Bowzard JB, Wills JW. Recent advances and remaining problems in HIV assembly. AIDS. 1998;12 Suppl A:S5-16.

(115) Le B, I, Prevost MC, Dokhelar MC, Rosenberg AR. The PPPY motif of human T-cell leukemia virus type 1 Gag protein is required early in the budding process. J Virol. 2002;76:10024-10029.

(116) Blot V, Perugi F, Gay B et al. Nedd4.1-mediated ubiquitination and subsequent recruitment of Tsg101 ensure HTLV-1 Gag trafficking towards the multivesicular body pathway prior to virus budding. J Cell Sci. 2004;117:2357-2367.

(117) Green PL, Chen ISY. Molecular features of the human T-cell leukemia virus. Mechanisms of transformation and leukemogenicity. In: Levy JA, ed. The Retroviridae.3. New York: Plenum Press; 1994:277-311.

(118) Nam SH, Kidokoro M, Shida H, Hatanaka M. Processing of gag Precursor Polyprotein of Human T-Cell Leukemia Virus Type 1 by Virus-Encoded Protease. J Virol. 1988;62:3718-3728.

(119) Pique C, Pham D, Tursz T, Dokhelar MC. Human T-Cell Leukemia Virus Type I Envelope Protein Maturation Process: Requirements for Syncytium Formation. J Virol. 1992;66:906-913.

(120) Le B, I, Grange MP, Delamarre L et al. HTLV-1 structural proteins. Virus Res. 2001;78:5-16.

53

(121) Manel N, Kim FJ, Kinet S et al. The ubiquitous glucose transporter GLUT-1 is a receptor for HTLV. Cell. 2003;115:449-459.

(122) Green PL, Chen ISY. Regulation of human T cell leukemia virus expression. FASEB J. 1990;4:169-175.

(123) Hildreth JEK, Subramanium A, Hampton RA. Human T-cell lymphotropic virus type 1 (HTLV-1)-induced syncytium formation mediated by vascular cell adhesion molecule-1: Evidence for involvement of cell adhesion molecules in HTLV-1 biology. J Virol. 1997;71:1173-1180.

(124) Sagara Y, Ishida C, Inoue Y, Shiraki H, Maeda Y. 71-Kilodalton heat shock cognate protein acts as a cellular receptor for syncytium formation induced by human T-cell lymphotropic virus type 1. J.Virol. 72, 535-541. 1998.

(125) Sagara Y, Inoue Y, Kojima E et al. Phosphatidylglycerol Participates in

Syncytium Formation Induced by HTLV Type 1-Bearing Cells. AIDS Res Hum Retroviruses. 2001;17:125-135.

(126) Niyogi K, Hildreth JE. Characterization of new syncytium-inhibiting monoclonal antibodies implicates lipid rafts in human t-cell leukemia virus type 1 syncytium formation. J Virol. 2001;75:7351-7361.

(127) Pinon JD, Klasse PJ, Jassal SR et al. Human T-cell leukemia virus type 1 envelope glycoprotein gp46 interacts with cell surface heparan sulfate proteoglycans. J Virol. 2003;77:9922-9930.

(128) Petrini M, Pelosi-Testa E, Sposi NM et al. Constitutive expression and abnormal glycosylation of transferrin receptor in acute T-cell leukemia. Cancer Res. 1989;49:6989-6996.

(129) Manel N, Kinet S, Battini JL et al. The HTLV receptor is an early T-cell activation marker whose expression requires de novo protein synthesis. Blood. 2003;101:1913-1918.

(130) Nath MD, Ruscetti FW, Petrow-Sadowski C, Jones KS. Regulation of the cell-surface expression of an HTLV-I binding protein in human T cells during immune activation. Blood. 2003;101:3085-3092.

(131) Overbaugh J. HTLV-1 sweet-talks its way into cells. Nat Med. 2004;10:20-21.

(132) Bukrinsky MI, Sharova N, Dempsey MP et al. Active nuclear import of human immunodeficiency virus type 1 preintegration complexes. Proc Natl Acad Sci U S A. 1992;89:6580-6584.

54

(133) Cereseto A, Giacca M. Integration site selection by retroviruses. AIDS Rev. 2004;6:13-21.

(134) Cann AJ, Chen ISY. Human T-Cell Leulemia Virus Types I and II. 1849-1880. 1996. Philadelphia, PA, Lippincott-Raven Publishers.

(135) Cann AJ, Rosenblatt JD, Wachsman W, Shah NP, Chen IS. Identification of the

gene responsible for human T-cell leukaemia virus transcriptional regulation. Nature. 1985;318:571-574.

(136) Felber BK, Paskalis H, Kleinmanewing C. The pX Protein of HTLV-1 is a

Transcriptional Activator of its Long Terminal Repeats. Science. 1985;229:675-679.

(137) Fujisawa J, Seiki M, Kiyokawa T, Yoshida M. 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. 1985;82:2277-2281.

(138) Sodroski J, Rosen C, Goh WC, Haseltine W. A transcriptional activator protein encoded by the x-lor region of the human T-cell leukemia virus. Science. 1985;228:1430-1434.

(139) Sodroski JG, Rosen CA, Haseltine WA. Trans-Acting Transcriptional

Activation of the Long Terminal Repeat of Human T Lymphotropic Viruses in Infected Cells. Science. 1984;223:381-385.

(140) Burton M, Upadhyaya CD, Maier B, Hope TJ, Semmes OJ. Human T-cell leukemia virus type 1 tax shuttles between functionally discrete subcellular targets. J Virol. 2000;74:2351-2364.

(141) Semmes OJ, Jeang KT. Localization of human T-cell leukemia virus type 1 Tax to subnuclear compartments that overlap with interchromatin speckles. J Virol. 1996;70:6347-6357.

(142) Jeang KT, Giam CZ, Majone F, Aboud M. Life, death, and tax: role of HTLV-I oncoprotein in genetic instability and cellular transformation. J Biol Chem. 2004;279:31991-31994.

(143) Franklin AA, Kubik MF, Uittenbogaard MN et al. Transactivation by the human T-cell leukemia virus Tax protein is mediated through enhanced binding of activating transcription factor-2 (ATF-2) ATF-2 response and cAMP element-binding protein (CREB). J Biol Chem. 1993;268:21225-21231.

(144) Pise-Masison CA, Jeong SJ, Brady JN. Human T cell leukemia virus type 1: the role of Tax in leukemogenesis. Arch Immunol Ther Exp (Warsz ). 2005;53:283-296.

55

(145) Portis T, Harding JC, Ratner L. The contribution of NF-kappa B activity to spontaneous proliferation and resistance to apoptosis in human T-cell leukemia virus type 1 Tax- induced tumors. Blood. 2001;98:1200-1208.

(146) Sun SC, Ballard DW. Persistent activation of NF-kappa B by the Tax transforming protein of HTLV-1: hijacking cellular I kappa B kinases. Oncogene. 1999;18:6948-6958.

(147) Suzuki T, Hirai H, Fujisawa J, Fujita T, Yoshida M. 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. Oncogene. 1993;8:2391-2397.

(148) Yoshida M. HTLV-1 Tax: regulation of gene expression and disease. [Review].

Trends in Microbiology. 1993;1(4):131-135.

(149) Suzuki T, Fujisawa J, Toita M, Yoshida M. 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. 1993;90:610-614.

(150) Zhao LJ, Giam CZ. 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. 1992;89:7070-7074.

(151) Hai TW, Liu F, Coukos WJ, Green MR. Transcription factor ATF cDNA clones: an extensive family of leucine zipper proteins able to selectively form DNA-binding heterodimers. Genes Dev. 1989;3:2083-2090.

(152) Beraud C, Lombard-Platet G, Michal Y, Jalinot P. Binding of the HTLV-I Tax1 transactivator to the inducible 21 bp enhancer is mediated by the cellular factor HEB1. EMBO J. 1991;10:3795-3803.

(153) Barnhart MK, Connor LM, Marriott SJ. Function of the human T-cell leukemia

virus type 1 21-base-pair repeats in basal transcription. J Virol. 1997;71:337-344.

(154) Jeang KT, Chiu R, Santos E, Kim SJ. Induction of the HTLV-I LTR by Jun occurs through the Tax-responsive 21- bp elements. Virology. 1991;181:218-227.

(155) Muchardt C, Seeler JS, Nirula A, Gong S, Gaynor R. Transcription factor AP-2

activates gene expression of HTLV-I. EMBO Journal. 1992;11:2573-2581.

(156) Marriott SJ, Lindholm PF, Brown KM et al. A 36-kilodalton cellular transcription factor mediates an indirect interaction of human T-cell

56

leukemia/lymphoma virus type 1 Tax1 with a responsive element in the viral long terminal repeat. Mol Cell Biol. 1990;10:4192-4201.

(157) Tanimura A, Teshima H, Fujisawa JI, Yoshida M. 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. 1993;67:5375-5382.

(158) Bosselut R, Duvall JF, Gegonne A et al. 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. 1990;9:3137-3144.

(159) Bosselut R, Lim F, Romond PC et al. Myb protein binds to multiple sites in the

human T cell lymphotropic virus type 1 long terminal repeat and transactivates LTR-mediated expression. Virology. 1992;186:764-769.

(160) Clark NM, Smith MJ, Hilfinger JM, Markovitz DM. 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. 1993;67:5522-5528.

(161) Colgin MA, Nyborg JK. 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. 1998;72:9396-9399.

(162) Torgeman A, Mor-Vaknin N, Aboud M. 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. 1999;254:279-287.

(163) Kimzey AL, Dynan WS. Specific regions of contact between human T-cell leukemia virus type I Tax protein and DNA identified by photocross- linking. J Biol Chem. 1998;273:13768-13775.

(164) Lenzmeier BA, Baird EE, Dervan PB, Nyborg JK. The Tax protein-DNA interaction is essential for HTLV-I transactivation in vitro. J Mol Biol. 1999;291:731-744.

(165) Lundblad JR, Kwok RP, Laurance ME et al. 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. 1998;273:19251-19259.

(166) Perini G, Wagner S, Green MR. Recognition of bZIP proteins by the human T-cell leukaemia virus transactivator Tax. Nature. 1995;376:602-605.

57

(167) Chrivia JC, Kwok RP, Lamb N et al. Phosphorylated CREB binds specifically to the nuclear protein CBP. Nature. 1993;365(6449):855-859.

(168) Kwok RP, Lundblad JR, Chrivia JC et al. Nuclear protein CBP is a coactivator for the transcription factor. Nature. 1994;370(6486):223-226.

(169) Bex F, Yin MJ, Burny A, Gaynor RB. 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. 1998;18:2392-2405.

(170) Lemasson I, Polakowski NJ, Laybourn PJ, Nyborg JK. Transcription factor binding and histone modifications on the integrated proviral promoter in human T-cell leukemia virus-I-infected T-cells. J Biol Chem. 2002;277:49459-49465.

(171) Harrod R, Kuo YL, Tang Y et al. 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. 2000;275:11852-11857.

(172) Jiang H, Lu H, Schiltz RL et al. 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. 1999;19:8136-8145.

(173) Ng PW, Iha H, Iwanaga Y et al. Genome-wide expression changes induced by HTLV-1 Tax: evidence for MLK- 3 mixed lineage kinase involvement in Tax-mediated NF-kappaB activation. Oncogene. 2001;20:4484-4496.

(174) Maruyama M, Shibuya H, Harada H et al. Evidence for aberrant activation of the interleukin-2 autocrine loop by HTLV-1-encoded p40x and T3/Ti complex triggering. Cell. 1987;48:343-350.

(175) Muraoka O, Kaisho T, Tanabe M, Hirano T. Transcriptional activation of the

interleukin-6 gene by HTLV-1. Immunology Letters. 1993;37(2-3):159-165.

(176) Mori N, Mukaida N, Ballard DW, Matsushima K, Yamamoto N. 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. 1998;58:3993-4000.

(177) Mori N, Prager D. Transactivation of the interleukin-1 alpha promoter by human T- cell leukemia virus type I and type II tax proteins. Blood. 1996;87:3410-3417.

(178) Miyatake S, Seiki M, Malefijt RD et al. Activation of T cell-derived lymphokine genes in T cells and fibroblasts: effects of human T cell leukemia virus type I

58

p40x protein and bovine papilloma virus encoded E2 protein. Nucleic Acids Res. 1988;16:6547-6566.

(179) Cowan EP, Alexander RK, Daniel S, Kashanchi F, Brady JN. Induction of

tumor necrosis factor alpha in human neuronal cells by extracellular human T-cell lymphotropic virus type 1 Tax(1). J Virol. 1997;71:6982-6989.

(180) Lindholm PF, Reid RL, Brady JN. Extracellular Tax Protein Stimulates Tumor Necrosis Factor-B and Immunoglobulin Kappa Light Chain Expression in Lymphoid Cells. J Virol. 1992;66:1294-1302.

(181) Duyao MP, Kessler DJ, Spicer DB et al. Transactivation of the c-myc Promoter by Human T Cell Leukemia virus Type 1 tax is Mediated by NFkB. J Biol Chem. 1992;267:16288-16291.

(182) Fujii M, Sassone-Corsi P, Verma I. c-fos promotor trans-activation of cAMP on IL-2 stimulated gene expression. Proc Natl Acad Sci U S A. 1988;8526-8530.

(183) Ratner L. Regulation of expression of the c-sis proto-oncogene. Nucleic Acids Res. 1989;17:4101-4115.

(184) Fujii M, Tsuchiya H, Chuhjo T, Akizawa T, Seiki M. Interaction of HTLV-1

Tax1 with p67SRF causes the aberrant induction of cellular immediate early genes through CArG boxes. Genes Dev. 1992;6:2066-2076.

(185) Li CC, Ruscetti FW, Rice NR et al. 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. 1993;67:4205-4213.

(186) Nicot C, Mahieux R, Takemoto S, Franchini G. Bcl-X(L) is up-regulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL samples. Blood. 2000;96:275-281.

(187) Tsukahara T, Kannagi M, Ohashi T et al. 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. 1999;73:7981-7987.

(188) Ressler S, Morris GF, Marriott SJ. Human T-cell leukemia virus type 1 Tax transactivates the human proliferating cell nuclear antigen promoter. J Virol. 1997;71:1181-1190.

(189) Mesnard JM, Devaux C. Multiple control levels of cell proliferation by human T- cell leukemia virus type 1 tax protein. Virology. 1999;257:277-284.

59

(190) Jeang KT. Functional activities of the human T-cell leukemia virus type I Tax oncoprotein: cellular signaling through NF-kappa B. Cytokine Growth Factor Rev. 2001;12:207-217.

(191) Teshigawara K, Maeda M, Nishino K et al. Adult T leukemia cells produce a lymphokine that augments interleukin 2 receptor expression. J Mol Cell Immunol. 1985;2:17-26.

(192) Inoue J, Seiki M, Taniguchi T, Tsuru S, Yoshida M. Induction of interleukin 2 receptor gene expression by p40x encoded by human T-cell leukemia virus type 1. EMBO J. 1986;5:2883-2888.

(193) Mori N, Fujii M, Ikeda S et al. Constitutive activation of NF-kappa B in primary adult T- cell leukemia cells. Blood. 1999;93:2360-2368.

(194) Suzuki T, Hirai H, Yoshida M. 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. 1994;9:3099-3105.

(195) Murakami T, Hirai H, Suzuki T, Fujisawa JI, Yoshida M. 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. 1995;206:1066-1074.

(196) Gerritsen ME, Williams AJ, Neish AS et al. CREB-binding protein/p300 are transcriptional coactivators of p65. Proc Natl Acad Sci U S A. 1997;94:2927-2932.

(197) Perkins ND, Felzien LK, Betts JC et al. Regulation of NF-kappaB by cyclin-dependent kinases associated with the p300 coactivator. Science. 1997;275:523-527.

(198) Zhong H, Voll RE, Ghosh S. Phosphorylation of NF-kappa B p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with the coactivator CBP/p300. Mol Cell. 1998;1:661-671.

(199) Baeuerle PA, Baltimore D. I kappa B: a specific inhibitor of the NF-kappa B transcription factor. Science. 1988;242:540-546.

(200) Hirai H, Suzuki T, Fujisawa J, Inoue J, Yoshida M. Tax Protein of Human T-Cell Leukemia Virus Type I Binds to the Ankyrin Motifs of Inhibitory Factor kappa b and Induces Nuclear Translocation of Transcription Factor NF- kappa b Proteins for Transcriptional Activation. Proc Natl Acad Sci USA. 1994;91:3584-3588.

60

(201) Maggirwar SB, Harhaj E, Sun SC. Activation of NF-kappa B/Rel by Tax involves degradation of I kappa B alpha and is blocked by a proteasome inhibitor. Oncogene. 1995;11:993-998.

(202) Suzuki T, Hirai H, Murakami T, Yoshida M. 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. 1995;10:1199-1207.

(203) Li XH, Murphy KM, Palka KT, Surabhi RM, Gaynor RB. The human T-cell leukemia virus type-1 Tax protein regulates the activity of the I kappa B kinase complex. J Biol Chem. 1999;274:34417-34424.

(204) Brauweiler A, Garrus JE, Reed JC, Nyborg JK. Repression of Bax gene expression by the HTLV-I tax protein: Implications for suppression of apoptosis in virally infected cells. Virology. 1997;231:135-140.

(205) Kwok RPS, Laurance ME, Lundblad JR et al. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature. 1996;380:642-646.

(206) Suzuki T, Uchidatoita M, Yoshida M. 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. 1999;18:4137-4143.

(207) Eckner R, Yao TP, Oldread E, Livingston DM. Interaction and functional collaboration of p300/CBP and bHLH proteins in muscle and B-cell differentiation. Genes Dev. 1996;10:2478-2490.

(208) Lundblad JR, Kwok RP, Laurance ME, Harter ML, Goodman RH. Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature. 1995;374:85-88.

(209) Yoshida M. Multiple viral strategies of htlv-1 for dysregulation of cell growth control. Annu Rev Immunol. 2001;19:475-496.

(210) Suzuki T, Kitao S, Matsushime H, Yoshida M. HTLV-1 Tax protein interacts with cyclin-dependent kinase inhibitor p16(INK4A) and counteracts its inhibitory activity towards CDK4. EMBO J. 1996;15:1607-1614.

(211) Kamb A, Gruis NA, Weaver-Feldhaus J et al. A cell cycle regulator potentially involved in genesis of many tumor types. Science. 1994;264:436-440.

(212) Suzuki T, Narita T, Uchidatoita M, Yoshida M. Down-regulation of the INK4 family of cyclin-dependent kinase inhibitors by tax protein of HTLV-1 through two distinct mechanisms. Virology. 1999;259:384-391.

61

(213) Jin DY, Spencer F, Jeang KT. Human T cell leukemia virus type 1 oncoprotein Tax targets the human mitotic checkpoint protein MAD1. Cell. 1998;93:81-91.

(214) Laherty CD, Perkins ND, Dixit VM. Human T-Cell Leukemia Virus Type-I Tax and Phorbol 12- Myristate 13-Acetate Induce Expression of the A20 Zinc Finger Protein by Distinct Mechanisms Involving Nuclear Factor-kappaB. J Biol Chem. 1993;268:5032-5039.

(215) Wu K, Bottazzi ME, de La FC et al. Protein profile of tax-associated complexes. J Biol Chem. 2004;279:495-508.

(216) Inoue J, Seiki M, Yoshida M. The second pX product p27 chi-III of HTLV-1 is required for gag gene expression. FEBS Lett. 1986;209:187-190.

(217) Inoue J, Yoshida M, Seiki M. 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. 1987;84:3653-3657.

(218) Grone M, Koch C, Grassmann R. The HTLV-1 Rex protein induces nuclear accumulation of unspliced viral RNA by avoiding intron excision and degradation. Virology. 1996;218:316-325.

(219) Seiki M, Inoue J, Hidaka M, Yoshida M. 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. 1988;85:7124-7128.

(220) Ballaun C, Farrington GK, Dobrovnik M et al. 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. 1991;65:4408-4413.

(221) Barshira A, Panet A, Honigman A. An RNA Secondary Structure Juxtaposes Two Remote Genetic Signals for Human T-Cell Leukemia Virus Type 1 RNA 3'-End Processing. J Virol. 1991;65:5165-5173.

(222) Black AC, Ruland C, Yip MT et al. 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. 1991;65:6645-6653.

(223) Black AC, Chen ISY, Arrigo S et al. Regulation of HTLV-II gene expretion by rex involves positive and negative cis-Acting elements in the 5 long terminal repeat. Virology. 1991;181:433-444.

(224) Kim JH, Kaufman PA, Hanly SM, Rimsky LT, Greene WC. Rex transregulation of human T-cell leukemia virus type II gene expression. J Virol. 1991;65:405-414.

62

(225) Ahmed YF, Gilmartin GM, Hanly SM, Nevins JR, Greene WC. The HTLV-1 Rex Response Element Mediates a Novel Form of mRNA Polyadenylation. Cell. 1991;64:727-737.

(226) Derse D. Bovine leukemia virus transcription is controlled by a virus-encoded trans-acting factor and by cis-acting response elements. J Virol. 1987;61:2462-2471.

(227) Black AC, Ruland CT, Luo J et al. Binding of Nuclear Proteins to HTLV-II Cis-Acting Repressive Sequence (CRS) RNA Correlates with CRS Function. Virology. 1994;200:29-41.

(228) King JA, Bridger JM, Lochelt M et al. Nucleocytoplasmic transport of HTLV-1 RNA is regulated by two independent LTR encoded nuclear retention elements. Oncogene. 1998;16:3309-3316.

(229) Seiki M, Hikikoshi A, Yoshida M. The U5 sequence is a cis-acting repressive element for genomic RNA expression of human T cell leukemia virus type I. Virology. 1990;176:81-86.

(230) Bogerd HP, Huckaby GL, Ahmed YF, Hanly SM, Greene WC. The type I human T-cell leukemia virus (HTLV-I) Rex trans-activator binds directly to the HTLV-I Rex and the type 1 human immunodeficiency virus Rev RNA response elements. Proc Natl Acad Sci U S A. 1991;88:5704-5708.

(231) Grassmann R, Berchtold S, Aepinus C et al. In Vitro Binding of Human T-Cell Leukemia Virus rex Proteins to the rex-Response Element of Viral Transcripts. J Virol. 1991;65:3721-3727.

(232) Nosaka T, Siomi H, Adachi Y et al. 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. 1989;86:9798-9802.

(233) Siomi H, Shida H, Nam SH et al. Sequence requirements for nucleolar

localization of human T cell leukemia virus type I pX protein, which regulates viral RNA processing. Cell. 1988;55:197-209.

(234) Bogerd HP, Echarri A, Ross TM, Cullen BR. 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. 1998;72:8627-8635.

(235) Hakata Y, Umemoto T, Matsushita S, Shida H. 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. 1998;72:6602-6607.

63

(236) Palmeri D, Malim MH. The human T-cell leukemia virus type 1 posttranscriptional trans- activator Rex contains a nuclear export signal. J Virol. 1996;70:6442-6445.

(237) Adachi Y, Copeland TD, Takahashi C et al. Phosphorylation of the Rex protein of human T-cell leukemia virus type I. Journal of Biological Chemistry. 1992;267:21977-21981.

(238) Orita S, Saiga A, Takagi S et al. 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. 1991;295:127-134.

(239) Kubota S, Hatanaka M, Pomerantz RJ. Nucleo-cytoplasmic redistribution of the

HTLV-I rex protein: Alterations by coexpression of the HTLV-I p21(x) protein. Virology. 1996;220:502-507.

(240) Ciminale V, D'Agostino D, Zotti L et al. Expression and characterization of proteins produced by mRNAs spliced into the X region of the human T-cell leukemia/lymphotropic virus type II. Virology. 1995;209:445-456.

(241) Koralnik IJ, Fullen J, Franchini G. 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. 1993;67:2360-2366.

(242) Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type I. Proc Natl Acad Sci U S A. 1992;89:8813-8817.

(243) Derse D, Mikovits J, Ruscetti F. 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. 1997;237:123-128.

(244) Kim SJ, Ding W, Albrecht B, Green PL, Lairmore MD. 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. 2003;278:15550-15557.

(245) Albrecht B, D'Souza CD, Ding W et al. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J Virol. 2002;76:3493-3501.

(246) Ding W, Albrecht B., Kelley RE et al. 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. 2002;76:10374-10382.

64

(247) Zhang W, Nisbet JW, Albrecht B et al. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J Virol. 2001;75:9885-9895.

(248) Bartoe JT, Albrecht B, Collins ND et al. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J Virol. 2000;74:1094-1100.

(249) Lairmore MD, Albrecht B, D'Souza C et al. 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. 2000;16:1757-1764.

(250) Zhang W, Nisbet JW, Bartoe JT, Ding W, Lairmore MD. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J Virol. 2000;74:11270-11277.

(251) Albrecht B, Collins ND, Burniston MT et al. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J Virol. 2000;74:9828-9835.

(252) Johnson JM, Franchini G. Retroviral proteins that target the major histocompatibility complex class I. Virus Res. 2002;88:119-127.

(253) Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

(254) Dekaban GA, Peters AA, Mulloy JC et al. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology. 2000;274:86-93.

(255) D'Agostino DM, Zotti L, Ferro T et al. The p13II protein of HTLV type 1: comparison with mitochondrial proteins coded by other human viruses. AIDS Res Hum Retroviruses. 2000;16:1765-1770.

(256) Ciminale V, Zotti L, Dagostino DM et al. 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. 1999;18:4505-4514.

(257) Trovato R, Mulloy JC, Johnson JM et al. 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. 1999;73:6460-6467.

65

(258) Cereseto A, Kislyakova T, Washington-Parks R, Franchini G. Differential response to genotoxic stress in immortalized or transformed HTLV-I-infected T-cells. J Gen Virol. 1999;80:1575-1581.

(259) Chen YA, Chen S, Fu C, Chen J, Osame M. 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. 1997;71:196-202.

(260) Pique C, Uretavidal A, Gessain A et al. 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. 2000;191:567-572.

(261) Princler GL, Julias JG, Hughes SH, Derse D. Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology. 2003;317:136-145.

(262) Ciminale V, D'Agostino DM, Zotti L, Chieco-Bianchi L. Coding potential of the X region of human T-cell leukemia/lymphotropic virus type II. J Acquir Immune Defic Syndr Hum Retrovirol. 1996;13 Suppl 1:S220-S227.

(263) Saksena NK, Srinivasan A, Ge YC et al. Simian T cell leukemia virus type I from naturally infected feral monkeys from Central and West Africa encodes a 91-amino acid p12 (ORF-I) protein as opposed to a 99-amino acid protein encoded by HTLV type I from humans. AIDS Res Hum Retroviruses. 1997;13:425-432.

(264) Shaw GM, Gonda MA, Flickinger GH et al. Genomes of evolutionarily divergent members of the human T-cell leukemia virus family (HTLV-I and HTLV-II) are highly conserved, especially in pX. Proc Natl Acad Sci USA. 1984;81:4544-4548.

(265) Larocca D, Chao LA, Seto MH, Brunck TK. Human T-cell leukemia virus minus strand transcription in infected T-cells. Biochem Biophys Res Commun. 1989;163:1006-1013.

(266) Hivin P, Frederic M, rpin-Andre C et al. Nuclear localization of HTLV-I bZIP factor (HBZ) is mediated by three distinct motifs. J Cell Sci. 2005;118:1355-1362.

(267) Thebault S, Basbous J, Hivin P, Devaux C, Mesnard JM. HBZ interacts with JunD and stimulates its transcriptional activity. FEBS Lett. 2004;562:165-170.

(268) Basbous J, Arpin C, Gaudray G et al. The HBZ factor of human T-cell leukemia virus type I dimerizes with transcription factors JunB and c-Jun and modulates their transcriptional activity. J Biol Chem. 2003;278:43620-43627.

66

(269) Matsumoto J, Ohshima T, Isono O, Shimotohno K. HTLV-1 HBZ suppresses AP-1 activity by impairing both the DNA-binding ability and the stability of c-Jun protein. Oncogene. 2005;24:1001-1010.

(270) Arnold J, Yamamoto B, Li M et al. Enhancement of infectivity and persistence in vivo by HBZ, a natural antisense coded protein of HTLV-1. Blood. 2006.

(271) Chou KS, Okayama A, Tachibana N, Lee TH, Essex M. 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. 1995;60:701-706.

(272) D'Agostino DM, Ranzato L, Arrigoni G et al. Mitochondrial alterations induced by the p13II protein of human T-cell leukemia virus type 1. Critical role of arginine residues. J Biol Chem. 2002;277:34424-34433.

(273) Roithmann S, Pique C, Le Cesne A et al. 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. 1994;68:3448-3451.

(274) D'Agostino DM, Zotti L, Ferro T et al. Expression and functional properties of proteins encoded in the x-II ORF of HTLV-I. Virus Res. 2001;78:35-43.

(275) Silic-Benussi M, Cavallari I, Zorzan T et al. Suppression of tumor growth and cell proliferation by p13II, a mitochondrial protein of human T cell leukemia virus type 1. Proc Natl Acad Sci U S A. 2004;101:6629-6634.

(276) Hiraragi H, Michael B, Nair A et al. Human T-lymphotropic virus type 1 mitochondria localizing protein p13II sensitizes Jurkat T cells to Ras-mediated apoptosis. J Virol. 2005;In press.

(277) Hou X, Foley S, Cueto M, Robinson MA. The Human T-Cell Leukemia Virus Type I (HTLV-I) X Region Encoded Protein p13(II) Interacts with Cellular Proteins. Virology. 2000;277:127-135.

(278) Lefebvre L, Vanderplasschen A, Ciminale V et al. Oncoviral Bovine Leukemia Virus G4 and Human T-Cell Leukemia Virus Type 1 p13(II) Accessory Proteins Interact with Farnesyl Pyrophosphate Synthetase. J Virol. 2002;76:1400-1414.

(279) Lefebvre L, Ciminale V, Vanderplasschen A et al. Subcellular localization of the bovine leukemia virus R3 and G4 accessory proteins. J Virol. 2002;76:7843-7854.

(280) Hiraragi H, Kim SJ, Phipps AJ et al. Human T-Lymphotropic Virus Type 1 Mitochondrion-Localizing Protein p13II Is Required for Viral Infectivity In Vivo. J Virol. 2006;80:3469-3476.

67

(281) D'Agostino DM, Ciminale V, Zotti L, Rosato A, Chieco-Bianchi L. 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. 1997;71:75-83.

(282) Nicot C, Dundr M, Johnson JM et al. HTLV-1-encoded p30II is a post-transcriptional negative regulator of viral replication. Nat Med. 2004;10:197-201.

(283) Younis I, Khair L, Dundr M et al. Repression of Human T-Cell Leukemia Virus Type 1 and Type 2 Replication by a Viral mRNA-Encoded Posttranscriptional Regulator. J Virol. 2004;78:11077-11083.

(284) Younis I, Boris-Lawrie K, Green PL. Human T-cell leukemia virus open reading frame II encodes a posttranscriptional repressor that is recruited at the level of transcription. J Virol. 2006;80:181-191.

(285) Michael B, Nair AM, Hiraragi H et al. Human T lymphotropic virus type-1 p30II alters cellular gene expression to selectively enhance signaling pathways that activate T lymphocytes. Retrovirology. 2004;1:39.

(286) Silverman LR, Phipps AJ, Montgomery A, Ratner L, Lairmore MD. 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. 2004;78:3837-3845.

(287) Iniguez AM, Otsuki K, Magalhaes GP, Silva EA, Vicente AC. Genetic markers on the HTLV-1 p12I protein sequences from Brazilian HAM/TSP patients and asymptomatic HTLV-1 carrier isolates. AIDS Res Hum Retroviruses. 2005;21:580-582.

(288) Trovato R, Mulloy JC, Johnson JM et al. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic leukemia virus type 1 p12 protein greatly influences its stability. J Virol. 1999;73:6460-6467.

(289) Pawson T. Protein modules and signalling networks. Nature. 1995;373:573-580. (290) Craig HM, Reddy TR, Riggs NL, Dao PP, Guatelli JC. 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. 2000;271:9-17.

(291) Martins ML, Soares BC, Ribas JG et al. 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. 2002;18:899-902.

68

(292) Ding W, Albrecht B, Luo R et al. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J Virol. 2001;75:7672-7682.

(293) Ding W, Kim SJ, Nair AM et al. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol. 2003;77:11027-11039.

(294) Krause KH, Michalak M. Calreticulin. Cell. 1997;88:439-443.

(295) Michalak M, Corbett EF, Mesaeli N, Nakamura K, Opas M. Calreticulin: one protein, one gene, many functions. Biochem J. 1999;344 Pt 2:281-292.

(296) Garry RF. Sequence similarities between latent membrane protein LMP- 1 of Epstein-Barr virus, integral membrane protein p12(I) of human T cell leukemia lymphotropic virus type 1, E5 transformation protein of bovine papillomavirus, and the transmembrane proteins of slowly transforming retroviruses. AIDS Res Hum Retroviruses. 1995;11:431-432.

(297) Franchini G, Mulloy JC, Koralnik IJ et al. 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. Journal of Virology. 1993;67:7701-7704.

(298) Koralnik I, Mulloy JC, Andresson T, Fullen J, Franchini G. Mapping of the intermolecular association of the human T-cell leukemia/lymphotropic virus type 1 p12I and the vacuolar H+ATPase 16 kDa subunit protein. J Gen Virol. 1995;76:1909-1916.

(299) Trovato R, Mulloy JC, Johnson JM et al. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic leukemia virus type 1 p12 protein greatly influences its stability. J Virol. 1999;73:6460-6467.

(300) Nicot C, Mulloy JC, Ferrari MG et al. HTLV-1 p12(I) protein enhances STAT5 activation and decreases the interleukin-2 requirement for proliferation of primary human peripheral blood mononuclear cells. Blood. 2001;98:823-829.

(301) Collins ND, D'Souza C, Albrecht B et al. 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. 1999;73:9642-9649.

(302) Johnson JM, Mulloy JC, Ciminale V et al. 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. 2000;16:1777-1781.

69

(303) Johnson JM, Fullen J, Casareto L, Mulloy JC. p12(I) May Contribute TO HTLV-1 Escape from Immune Detection. Leukemia. 2000;14:558.

(304) Johnson JM, Nicot C, Fullen J et al. 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. 2001;75:6086-6094.

(305) Boustany LM, Cyert MS. Calcineurin-dependent regulation of Crz1p nuclear export requires Msn5p and a conserved calcineurin docking site. Genes Dev. 2002;16:608-619.

(306) Robek MD, Wong FH, Ratner L. 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. 1998;72:4458-4462.

(307) Manninen A, Saksela K. HIV-1 Nef interacts with inositol trisphosphate receptor to activate calcium signaling in T cells. J Exp Med. 2002;195:1023-1032.

(308) Petit C, Buseyne F, Boccaccio C et al. Nef is required for efficient hiv-1 replication in cocultures of dendritic cells and lymphocytes. Virology. 2001;286:225-236.

(309) Wu Y, Marsh JW. Selective transcription and modulation of resting T cell activity by preintegrated HIV DNA. Science. 2001;293:1503-1506.

(310) Tsukahara T, Ratner L. Substitution of HIV Type 1 Nef with HTLV-1 p12. AIDS Res Hum Retroviruses. 2004;20:938-943.

(311) Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

(312) He H, Lam M, McCormick TS, Distelhorst CW. Maintenance of calcium homeostasis in the endoplasmic reticulum by Bcl-2. J Cell Biol. 1997;138:1219-1228.

(313) Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

(314) Shibasaki F, Kondo E, Akagi T, Mckeon F. Suppression of signalling through transcription factor NF-AT by interactions between calcineurin and Bcl-2. Nature. 1997;386:728-731.

70

(315) Srivastava RK, Sasaki CY, Hardwick JM, Longo DL. Bcl-2-mediated drug resistance: inhibition of apoptosis by blocking nuclear factor of activated T lymphocytes (NFAT)-induced Fas ligand transcription. J Exp Med. 1999;190:253-265.

(316) Nair A, Michael B, Hiraragi H et al. Human T lymphotropic virus type 1 accessory protein p12I modulates calcium-mediated cellular gene expression and enhances p300 expression in T lymphocytes. AIDS Res Hum Retroviruses. 2005;21:273-284.

(317) Bex F, Gaynor RB. Regulation of gene expression by HTLV-I Tax protein. Methods. 1998;16:83-94.

(318) Kinoshita S, Chen BK, Kaneshima H, Nolan GP. Host control of HIV-1 parasitism in T cells by the nuclear factor of activated T cells. Cell. 1998;95:595-604.

(319) Ballard DW. Molecular mechanisms in lymphocyte activation and growth. Immunol Res. 2001;23:157-166.

(320) Ganem D. Virology. The X files--one step closer to closure. Science. 2001;294:2299-2300.

(321) Bouchard MJ, Wang LH, Schneider RJ. Calcium signaling by HBx protein in hepatitis B virus DNA replication. Science. 2001;294:2376-2378.

(322) Tian P, Estes MK, Hu Y et al. The rotavirus nonstructural glycoprotein NSP4 mobilizes Ca2+ from the endoplasmic reticulum. J Virol. 1995;69:5763-5772.

(323) Feng P, Park J, Lee BS et al. Kaposi's sarcoma-associated herpesvirus mitochondrial K7 protein targets a cellular calcium-modulating cyclophilin ligand to modulate intracellular calcium concentration and inhibit apoptosis. J Virol. 2002;76:11491-11504.

(324) Popovic M, Sarin PS, Robert-Gurroff M et al. Isolation and transmission of human retrovirus (human t-cell leukemia virus). Science. 1983;219:856-859.

(325) Fan N, Gavalchin J, Paul B et al. Infection of Peripheral Blood Mononuclear Cells and Cell Lines by Cell-Free Human T-Cell Lymphoma/Leukemia Virus Type 1. J Clin Microbiol. 1992;30:905-910.

(326) Okochi K, Sato H, Hinuma Y. A retrospective study on transmission of adult T cell leukemia virus by blood transfusion: seroconversion in recipients. Vox Sang. 1984;46:245-253.

71

(327) Delamarre L, Rosenberg AR, Pique C, Pham D, Dokhelar MC. A novel human T-Leukemia virus type 1 cell-to-cell transmission assay permits definition of SU glycoprotein amino acids important for infectivity. J Virol. 1997;71:259-266.

(328) Jassal SR, Lairmore MD, Leigh-Brown AJ, Brighty DW. Soluble recombinant HTLV-1 surface glycoprotein competitively inhibits syncytia formation and viral infection of cells. Virus Res. 2001;78:17-34.

(329) Nagy K, Clapham P, Cheingsong-Popov R, Weiss R. Human T cell leukemia virus type 1: Induction of syncytia and inhibition by patient's sera. Int J Cancer. 1983;32:321-328.

(330) Daenke S, Mccracken SA, Booth S. Human T-cell leukaemia lymphoma virus type 1 syncytium formation is regulated in a cell-specific manner by ICAM-1, ICAM-3 and VCAM-1 and can be inhibited by antibodies to integrin beta(2) or beta(7). J Gen Virol. 1999;80:1429-1436.

(331) Igakura T, Stinchcombe JC, Goon PK et al. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science. 2003;299:1713-1716.

(332) Barnard AL, Igakura T, Tanaka Y, Taylor GP, Bangham CR. Engagement of specific T cell surface molecules regulates cytoskeletal polarization in HTLV-1-infected lymphocytes. Blood. 2005.

(333) Fukudome K, Furuse M, Fukuhara N et al. Strong induction of ICAM-1 in human T cells transformed by human T-cell-leukemia virus type 1 and depression of ICAM-1 or LFA-1 in adult T-cell-leukemia-derived cell lines. Int J Cancer. 1992;52:418-427.

(334) Bounou S, Giguere JF, Cantin R et al. The importance of virus-associated host ICAM-1 in human immunodeficiency virus type 1 dissemination depends on the cellular context. FASEB J. 2004;18:1294-1296.

(335) Fortin JF, Barbeau B, Hedman H, Lundgren E, Tremblay MJ. Role of the leukocyte function antigen-1 conformational state in the process of human immunodeficiency virus type 1-mediated syncytium formation and virus infection. Virology. 1999;257:228-238.

(336) Giguere JF, Tremblay MJ. Statin compounds reduce human immunodeficiency virus type 1 replication by preventing the interaction between virion-associated host intercellular adhesion molecule 1 and its natural cell surface ligand LFA-1. J Virol. 2004;78:12062-12065.

72

(337) Nejmeddine M, Barnard AL, Tanaka Y, Taylor GP, Bangham CR. HTLV-1 tax protein triggers microtubule reorientation in the virological synapse. J Biol Chem. 2005.

(338) Allen WE, Zicha D, Ridley AJ, Jones GE. A role for Cdc42 in macrophage chemotaxis. J Cell Biol. 1998;141:1147-1157.

(339) Sims TN, Dustin ML. The immunological synapse: integrins take the stage. Immunol Rev. 2002;186:100-117.

(340) Perez OD, Mitchell D, Jager GC et al. Leukocyte functional antigen 1 lowers T cell activation thresholds and signaling through cytohesin-1 and Jun-activating binding protein 1. Nat Immunol. 2003;4:1083-1092.

(341) Hogg N, Laschinger M, Giles K, Mcdowall A. T-cell integrins: more than just sticking points. J Cell Sci. 2003;116:4695-4705.

(342) von Andrian UH, Mackay CR. T-cell function and migration. Two sides of the same coin. N Engl J Med. 2000;343:1020-1034.

(343) Carman CV, Springer TA. Integrin avidity regulation: are changes in affinity and conformation underemphasized? Curr Opin Cell Biol. 2003;15:547-556.

(344) Stewart M, Hogg N. Regulation of leukocyte integrin function: affinity vs. avidity. J Cell Biochem. 1996;61:554-561.

(345) Stewart MP, Mcdowall A, Hogg N. LFA-1-mediated adhesion is regulated by cytoskeletal restraint and by a Ca2+-dependent protease, calpain. J Cell Biol. 1998;140:699-707.

(346) Kucik DF, Dustin ML, Miller JM, Brown EJ. Adhesion-activating phorbol ester increases the mobility of leukocyte integrin LFA-1 in cultured lymphocytes. J Clin Invest. 1996;97:2139-2144.

(347) Constantin G, Majeed M, Giagulli C et al. Chemokines trigger immediate beta2 integrin affinity and mobility changes: differential regulation and roles in lymphocyte arrest under flow. Immunity. 2000;13:759-769.

(348) Dustin ML, Bivona TG, Philips MR. Membranes as messengers in T cell adhesion signaling. Nat Immunol. 2004;5:363-372.

(349) Kim M, Carman CV, Springer TA. Bidirectional transmembrane signaling by cytoplasmic domain separation in integrins. Science. 2003;301:1720-1725.

(350) Feigelson SW, Grabovsky V, Winter E et al. The Src kinase p56(lck) up-regulates VLA-4 integrin affinity. Implications for rapid spontaneous and

73

chemokine-triggered T cell adhesion to VCAM-1 and fibronectin. J Biol Chem. 2001;276:13891-13901.

(351) Ardouin L, Bracke M, Mathiot A et al. Vav1 transduces TCR signals required for LFA-1 function and cell polarization at the immunological synapse. Eur J Immunol. 2003;33:790-797.

(352) Griffiths EK, Krawczyk C, Kong YY et al. Positive regulation of T cell activation and integrin adhesion by the adapter Fyb/Slap. Science. 2001;293:2260-2263.

(353) Wang H, Moon EY, Azouz A et al. SKAP-55 regulates integrin adhesion and formation of T cell-APC conjugates. Nat Immunol. 2003;4:366-374.

(354) Katagiri K, Hattori M, Minato N et al. Rap1 is a potent activation signal for leukocyte function-associated antigen 1 distinct from protein kinase C and phosphatidylinositol-3-OH kinase. Mol Cell Biol. 2000;20:1956-1969.

(355) D'Souza-Schorey C, Boettner B, Van AL. Rac regulates integrin-mediated spreading and increased adhesion of T lymphocytes. Molecular & Cellular Biology. 1998;18:3936-3946.

(356) Fagerholm S, Morrice N, Gahmberg CG, Cohen P. Phosphorylation of the cytoplasmic domain of the integrin CD18 chain by protein kinase C isoforms in leukocytes. J Biol Chem. 2002;277:1728-1738.

(357) Hogg N, Henderson R, Leitinger B et al. Mechanisms contributing to the activity of integrins on leukocytes. Immunol Rev. 2002;186:164-171.

(358) Smith A, Bracke M, Leitinger B, Porter JC, Hogg N. LFA-1-induced T cell migration on ICAM-1 involves regulation of MLCK-mediated attachment and ROCK-dependent detachment. J Cell Sci. 2003;116:3123-3133.

(359) Rodriguez-Fernandez JL, Gomez M, Luque A et al. The interaction of activated integrin lymphocyte function-associated antigen 1 with ligand intercellular adhesion molecule 1 induces activation and redistribution of focal adhesion kinase and proline-rich tyrosine kinase 2 in T lymphocytes. Mol Biol Cell. 1999;10:1891-1907.

(360) Volkov Y, Long A, Kelleher D. Inside the crawling T cell: leukocyte function-associated antigen-1 cross-linking is associated with microtubule-directed translocation of protein kinase C isoenzymes beta(I) and delta. J Immunol. 1998;161:6487-6495.

74

(361) Itoyama Y, Minato S, Kira J et al. Spontaneous proliferation of peripheral blood lymphocytes increased in patients with HTLV-I-associated myelopathy. Neurology. 1988;38:1302-1307.

(362) Jacobson S, Zaninovic V, Mora C et al. Immunological findings in neurological diseases associated with antibodies to HTLV-I: activated lymphocytes in tropical spastic paraparesis. Ann Neurol. 1988;23 Suppl:S196-S200.

(363) Matsumoto M, Sugimoto M, Nakashima H et al. Spontaneous T cell proliferation and release of soluble interleukin-2 receptors in patients with HTLV-I-associated myelopathy. Am J Trop Med Hyg. 1990;42:365-373.

(364) Usuku K, Sonoda S, Osame M et al. HLA haplotype-linked high immune responsiveness against HTLV-I in HTLV-I-associated myelopathy: comparison with adult T-cell leukemia/lymphoma. Ann Neurol. 1988;23 Suppl:S143-50.:S143-S150.

(365) Hanon E, Hall S, Taylor GP et al. Abundant Tax protein expression in CD4+T cells infected with human T-cell lymphotropic virus type I (HTLV-I) is prevented by cytotoxic T lymphocytes. Blood. 2000;95:1386-1392.

(366) Tendler CL, Greenberg SJ, Blattner WA et al. Transactivation of interleukin 2 and its receptor induces immune activation in human T-cell lymphotropic virus type I-associated myelopathy: pathogenic implications and a rationale for immunotherapy. Proceedings of the National Academy of Sciences of the United States of America. 1990;87:5218-5222.

(367) Wucherpfennig KW, Hollsberg P, Richardson JH, Benjamin D, Hafler DA. T-cell activation by autologous human T-cell leukemia virus type I- infected T-cell clones. Proc Natl Acad Sci U S A. 1992;89:2110-2114.

(368) Kimata JT, Palker TJ, Ratner L. The Mitogenic Activity of Human T-Cell

Leukemia Virus Type- I Is T-Cell Associated and Requires the CD2/LFA-3 Activation Pathway. J Virol. 1993;67:3134-3141.

75

Figure 1.1 Schematic illustration of HTLV-1 proviral genome, mRNA and protein

species Diagram of HTLV-1 proviral genome (top), mRNA (middle), and protein

species (bottom). Numbers correspond to nucleotide positions of each exon splice

acceptor and donor sites with respect to the full length HTLV-1 genome. (Modified

from Wycuff and Marriott, 2005 Frontiers in Biosciences)

76

Figure 1.2. Schematic Illustration of HTLV-1 non-structural protein p12I

Diagram of p12I with putative and identified 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); PxIxIT, coserved calcineurin binding site.

Numbers below bars indicate amino acid position numbers.

77

Figure 1.3. A model of p12I role in calcium-mediated signaling

Abbreviations: TCR, T cell receptor ; PTKs, protein tyrosine kinases ; PLCγ,

phospholipase Cγ ; DAG, diacylglycerolotein ; PKC, protein kinase C ;MEK mitogen-

activated protein (MAP) or extracellular signal-regulated (Erk) kinase ; AP-1, activator

protein-1 ;IP-3, inositol trisphosphate ; IP3R, inositol trisphosphate receptor ; ER,

endoplasmic reticulum; MLCK, myosin light chain kinase ; CaMKII,

Ca2+/Calmodulin-Dependent Protein Kinase II ;NFAT-ppp, phophorylated nuclear

factor of activated T-cell nuclear factor; NFATre, NFAT responsive element ; IL-2,

Interleukin 2

78

CHAPTER 2

A CONSERVED CALCINEURIN-BINDING MOTIF IN HUMAN T-

LYMPHOTROPIC VIRUS TYPE 1 p12I FUNCTIONS TO MODULATE

NFAT ACTIVATION

2.1 INTRODUCTION

As a complex retrovirus, human T lymphotropic virus type 1 (HTLV-1) encodes

common retrovirus structural and enzymatic proteins, as well as regulatory (Tax and

Rex) and accessory (p12I, p27I, p13II and p30II) proteins from four open reading frames

(ORF I-IV) in its unique pX region 1-4. The understanding of the molecular pathogenesis

of HTLV-1 infection, immortalization and transformation has been primarily focused

on the role of Tax and Rex 5,6, however emerging evidence also indicates the important

roles of accessory proteins in establishment of HTLV-1 infection 7.

The ORF I encoded protein HTLV-1 p12I is highly conserved among different

isolates and its messenger RNA can be detected in infected cell lines and directly from

infected patient cells 2-4,8,9. Humoral antibody and cytotoxic T lymphocyte responses

79

against ORF I encoded viral proteins are elicited in infected carriers and diseased

patients indicating the expression of the protein during the natural infection 10,11. p12I

localizes in the endoplasmic reticulum and cis-Golgi compartment 12-14, and associates

with the H+ vacuolar ATPase 15, the interleukin-2 (IL-2) receptor β- and γ-chain 16 and

major histocompatibility complex class I heavy chain 17. A small hydrophobic protein,

p12I contains several potential functional motifs including two transmembrane domains,

leucine zipper motifs, and four SH3 domain-binding motifs (PxxP) 2,15,18. However,

defined motifs of p12I that may alter key T cell signaling pathways that thereby

influence cell activation have not been identified.

We have reported that p12I is required for early establishment of HTLV-1

infection 19,20. When compared with the wild type infectious molecular clone (ACH),

ORF I mutated ACH immortalized T cell lines (ACH·p12I) failed to establish infection

in rabbits 21,22. ACH·p12I also exhibited reduced infectivity in quiescent peripheral

blood mononuclear cells (PBMC), but not in activated PBMC, suggesting a role for

p12I in T cell activation during the early stages of infection 20. Recently we have

demonstrated that p12I localizes to the ER and associates with an ER luminal protein,

calreticulin 12 and selectively activates the nuclear factor of activated T cells (NFAT) in

a calcium- dependent manner 12,23. Importantly, p12I expression in lymphocytes directly

increases cytoplasmic calcium from ER calcium stores 12,24, which serves to explain

how the protein activates the NFAT signaling pathway.

The transcriptional activation of NFAT is regulated by the calcium/calmodulin

dependent serine/threonine phosphatase, calcineurin 25,26. Through protein-protein

interaction, calcineurin triggers the dephosphorylation and subsequent nuclear

80

translocation of NFAT, which results in transactivation of NFAT inducible cytokine

genes including IL-2 25,26. A conserved calcineurin-binding motif (PxIxIT) has been

identified in the N-terminal regulatory domain of NFAT, which serves as a major

binding site with both inactivated and activated calcineurin 27,28. When this motif is

mutated, NFAT is functionally impaired 27,28. In addition, secondary calcineurin binding

sites in NFAT have been identified that contribute to higher affinity binding with

calcineurin 29,30.

The NFAT SPRIEIT calcineurin binding sequence and synthetic peptides

containing the critical binding amino acids (PxIxIT) inhibit binding between calcineurin

and NFAT without affecting calcineurin phosphatase activity 27,28 . A number of

calcineurin binding proteins inhibit either calcineurin phosphatase activity or its

substrate NFAT transcriptional activity. These include the anti-apoptotic protein BCL-

2 31,32, calcineurin B homologous protein (CHP) 33, AKAP79 (A kinase anchoring

protein) 34, and myocyte –enriched calcineurin-interacting protein 1(MCIP1) 35.

Interestingly, the African swine fever virus immunosuppressive protein A238L inhibits

NFAT activation, and has calcineurin binding sequence resembling that of NFAT 36,37.

Herein, we identified a conserved sequence (PSLPI/LT) in p12I, which is highly

homologous to the PxIxIT calcineurin-binding motif of NFAT. Full-length p12I and the

PSLPI/LT motif containing mutants bound calcineurin in both immunoprecipitation and

calmodulin bead pull-down assays. In contrast, serial mutations of p12I that lacked

PSLPI/LT motif or had selective alanine substitutions of the motif (p12I AxAxAA)

exhibited abolished or decreased binding affinity with calcineurin. Furthermore, p12I

competed with NFAT for calcineurin binding in calmodulin bead pull-down

81

experiments. Interestingly, the p12I AxAxAA mutant induced more NFAT nuclear

translocation than wild type and increased NFAT transcriptional activity (~ 2 fold) in a

reporter gene assay when compared to wild type p12I. The existence of a calcineurin

binding motif in p12I suggests that there may be at least two regulatory actions for p12I

that modulate NFAT activation: 1) to cause calcium release from ER stores and 2)

through calcineurin binding. Collectively our data provide new insight into the role of a

retroviral accessory protein in mediating early events in T cell infection and activation.

2.2 MATERIALS AND METHODS

Cell lines. The 293T HEK (American Type Culture Collection, catalog number 1573)

cells and Hela-Tat cells (AIDS Research and Reference Reagent Program, National

Institutes of Health, HLtat, catalog number 1293) were maintained in Dulbecco's

modified Eagle medium (supplemented with 10% fetal bovine serum, 100 µg of

streptomycin plus penicillin/ml, and 2 mM L-glutamine [Life Technologies, Rockville,

MD]). Jurkat T cells (clone E6-1, American Type Culture Collection catalog number

TIB-152) were maintained in RPMI media (Life Technologies, Rockville, MD)

supplemented with 15% fetal bovine serum, 100 µg of streptomycin plus penicillin/ml,

and 2 mM L-glutamine, and 10 mM HEPES (Life Technologies, Rockville,

MD)(complete RPMI).

82

Plasmids and site directed mutagenesis. pME 18S and pMEp12I plasmids were kindly

provided by G. Franchini (National Institute of Health)14. pMEp12I expresses an

influenza hemagglutinin (HA1) tagged HTLV-1 p12I fusion protein. HA-NFAT1c-GFP

expression vector 27 (A. Rao, Harvard Medical School) was used for NFAT expression

in 293T and Hela-Tat cell lines. The serial deletion mutants of p12I (1-47, 48-99, 15-69,

and 15-86) were previously described 12. Point mutants of p12I were generated by PCR

based site directed mutagenesis using a QuickChangeTM site-directed mutagenesis kit

(Stratagene La Jolla, CA). Point mutant L74I was made by substituting the 74th leucine

residue to isoleucine within the 99-amino acid of p12I. Point mutant p12I AxAxAA was

made by multiple alanine substitution of the PxLxLT motif (e.g., 70th proline to alanine,

71st leucine to alanine, 73rd leucine to alanine and 75th threonine to alanine) motif. To

construct L74I mutant, 5'CCTCAGCCCGTCGCTGCCGATCACGATGCGTTTCCCC

3' and 5' GGGGAAACGCATCGTGATCGGCAGCGACGGGCTGAGG 3' primers

were used. To construct p12I AxAxAA mutant, 5'

CCTCTTCTCCTCAGCGCGTCGGCGCCGGCAGCGATGCGTTTCCCC 3' and 5'

GGGGAAACGCATCGCTGCCGGCGCCGACGCGCTGAGGAGAAGAGG 3'

primers were used. Sanger sequencing was used to confirm that the sequences of each

p12I mutant plasmids were correct and in-frame. An NFAT-luciferase construct

pNFAT-luc, which has a trimerized human distal IL-2 NFAT site inserted into the

minimal IL-2 promoter was used for the reporter gene assay 38.

Transient transfection and immunoprecipitation assay. 293T cells were seeded at

50% confluence in 10-cm-diameter tissue culture dishes 24 h before transfection. Cells

83

were transfected with 10 µg of pME-18S, wild type pMEp12I and mutants of p12I using

LipofectamineTM reagent 12 (Invitrogen, Carlsbad, CA). For Jurkat T cell transfection,

107 cells were electroporated in complete RPMI (350 V, 975 µF; Biorad Gene Pulser II)

with 30 µg pME 18S, wild type pMEp12I and mutant expression vectors. To test the

binding between p12I and calcineurin, 293T cells or Jurkat T cells were transfected with

wild type pMEp12I or mutated p12I-expression vectors. Transfected 293T cells (1-2 x

107) or Jurkat T cells (1 x 107) were lysed with Triton X-100 buffer [1% deoxycholic

acid, 0.1% sodium dodecyl sulfate (SDS), 1% Triton X-100, 150 mM NaCL, 50 mM

Tris-HCl (pH 7.5), with Complete® Protease Inhibitor (Roche, Indianapolis, IN)]. CaCl2

(2 mM), EDTA (5 mM) or EGTA (5 mM) were added to test the calcium chelation

effect in p12I and calcineurin binding. Cell lysates were incubated with 1:200 diluted

rabbit polyclonal calcineurin antibody (Chemicon International, Inc., Temecular, CA)

overnight. The immune complex mixture was then incubated with 50 µl of protein A/G

plus agarose beads (Santa Cruz Biotechnology, Santa Cruz, CA) for 4 h. Beads were

washed three times in Triton X-100 buffer, boiled in SDS sample buffer and

supernatants were analyzed by Western immunoblotting. To test pharmacological

inhibitors in p12I and calcineurin binding assay, cyclosporin A (Sigma, St. Louis, MD)

(10 µM) or BAPTA-AM [Glycine, N, N’-1, 2-ethanediylbis (oxy-2, 1-phenylene)-bis-

N-2- (acetyloxy) methoxy-2-oxoethyl]-, [bis(acetyloxy)methyl-ester] (Molecular Probes,

Eugene, OR) (1 µM) were added in culture 30 min before cell lysis.

84

Calmodulin agarose bead pull-down assay. To test p12I binding with calcineurin and

to test for competition between p12I and NFAT for calcineurin binding, a calmodulin

bead assay was performed as previously described 39. Briefly, 293T cells or Jurkat T

cells were transfected with wild type pMEp12I or mutated p12I expression vectors or a

NFAT expression vector. Transfected 293T (2 x 107) or Jurkat T cell (1 x 107) were

lysed in Triton X-100 buffer [1% deoxycholic acid, 0.1% sodium dodecyl sulfate (SDS),

1% Triton X-100, 150 mM NaCL, 50 mM Tris-HCl (pH 7.5), with Complete® protease

inhibitor (Roche, Indianapolis, IN)] with CaCl2 (2 mM). Calmodulin-agarose

(Phosphodiesterase 3’, 5’-Cyclic Nucleotide Activator from bovine brain) beads

(Sigma) were washed three times in 1 ml of preactivation buffer (10 mM Tris, pH 8.0,

150 mM NaCl, and 2 mM CaCl2) to activate calmodulin. Lysates were then incubated

with 60 µl of packed beads for 3 h at 4 ° C. To remove unbound proteins, the beads

were washed three times with washing buffer [20 mM Hepes, pH 7.4, 150 mM NaCl, 5

mM sodium pyrophosphate, 10% glycerol, 1% Triton X-100, and protease inhibitors

supplemented with CaCl2 (2 mM)]. The beads were boiled in SDS sample buffer and

supernatants were analyzed by Western immunoblotting.

Immunoblotting. Western immunoblotting was used to analyze immunoprecipiated

and calmodulin bead pull downed products. Protein concentrations in all lysates were

determined by BCA assay (micro BCA Protein Assay, Pierce, Rockford, IL). Cell

lysates were separated by SDS-polyacrylamide gel electrophoresis (SDS-PAGE)

followed by transfer to nitrocellulose membranes. Membranes were blocked with 5%

non-fat dry milk for 2 h, incubated overnight at 4° C with monoclonal anti-HA antibody

85

(1:1000) (clone 16B-12, Covance Research Products, CA), with polyclonal calcineurin

antibody (1:1000) (Chemicon International Inc., Temecular, CA) or with monoclonal

NFAT1c antibody (1:500)(Santa Cruz Biotechnology, Santa Cruz, CA) for detection of

p12I, calcineurin and NFAT2 23 respectively. The blots were developed using

horseradish peroxidase-labeled secondary antibody and enhanced chemiluminescence

(Cell Signaling Technologies, Beverly, MA). The density of each band was measured

using a commercial software package (Gel-pro Analyzer software, Media Cybernetic,

Inc., Silver Spring, MD).

NFAT Reporter gene assay. For analysis of NFAT transcriptional activity in wild type

pMEp12I and alanine substituted mutant (p12I AxAxAA) transfected Jurkat T cells,

NFAT driven luciferase activity assay was performed as previously described 23. Briefly

Jurkat T cells (107) were electroporated as described above with 30 µg expression

vectors (pME 18S, pMEp12I, or p12I AxAxAA mutant), 10 µg reporter plasmid

(NFAT-luc) and 1 µg pCMV•SPORT-β-gal to normalize for transfection efficiency.

The transfected cells were plated in 6-well plates at a density of 5x105/ml. Cells were

stimulated with 20 ng/ml phorbol myristate acetate (PMA, Sigma, St. Louis, MO) at 6 h

post-transfection. After 18 h stimulation, cells were lysed (Cell Culture Lysis Reagent,

Promega, Madison, WI) and analyzed for luciferase activity according to the

manufacturer’s protocol (Promega, Madison, WI). Values were normalized for

transfection efficiency based on β-gal activity. Data expressed in reporter gene assay

86

were the mean of at least two independent experiments conducted in triplicate.

Statistical analysis was performed using Student’s t-test.

NFAT-GFP localization. HeLa-Tat cells were seeded into chamber slides (Fisher

Scientific, Pittsburg, PA) and were cotransfected with 2 µg of HA-NFAT1c-GFP

expression vector 27 and 4 µg pME 18S, pMEp12I, or p12I AxAxAA mutant expression

vectors. Two days post-transfection cells were fixed for 15 min with 2%

paraformaldehyde, and were visualized by an Olympus BH-2 fluorescence microscope

or Zeiss LSM510 confocal microscope. To quantify the HA-NFAT1c-GFP nuclear

localization, cells that had sole or predominant nuclear fluorescence were counted as

positive cells. Values were obtained from two wells of a single representative

experiment (~600 cells). To measure GFP intensity in cytoplasmic and nuclear regions

MetaMorph® Imaging software (Universal Imaging Corporation, Downingtown PA)

was used. Values represent nuclear/cytoplasmic ratio of GFP intensity of up to 25 cells.

Calcineurin phosphatase activity. Endogenous calcineurin phosphatase activity of

Jurkat cells was measured by using the calcineurin assay kit (Biomol, Plymouth

Meeting, PA) according to the manufacturer's instructions. 107 Jurkat cells were

transfected with 30 µg expression vectors (pME 18S, pMEp12I, or p12I AxAxAA

mutant). At 24 h post-transfection, phosphatase activity was measured using

phosphopeptide substrate (RII peptide) in the presence or absence of EGTA. The

released free PO4 was measured colorimetrically by using the Biomol Green reagent.

87

2.3 RESULTS

The putative calcineurin-binding motif in p12I is highly conserved. The peptide

alignment of p12I sequences with the calcineurin binding peptide sequence from NFAT

(SPRIEIT) was carried out using a commercial software package (Align X software of

Vector NTI suite, InforMax Inc., Bethesda, MD). We compared p12I sequences from

pMEp12I plasmid, ACH molecular proviral clone, and 5 different p12I sequences from

the Entrez protein database (http://www.ncbi.nlm.nih.gov). We identified a highly

conserved 70PSLPI/LT75 sequence in p12I (Fig. 2.1). p12I sequences from various

HTLV-1 strains had nearly identical sequences compared with the consensus

calcineurin binding PxIxIT sequence of NFAT except the 72nd leucine residue. The

pMEp12I plasmid was the only clone that had leucine residue at amino acid 74. This

putative calcineurin-binding motif is also conserved in the closely related simian T cell

leukemia virus type 1 (STLV-1) p12I 40. However, the accessory proteins R3 and G4 of

bovine leukemia virus (BLV), also a member of the Deltaretrovirus group, did not share

these consensus sequences 41.

The C-terminal half (48-99) of p12I containing PSLPI/LT binds calcineurin. To test

if p12I binds calcineurin, we transiently transfected p12I into both 293T and Jurkat T

cells and performed both immunoprecipitation and calmodulin bead pull-down

experiments (Fig. 2.2). In calcium containing buffer, calmodulin beads precipitated both

calcineurin and its binding protein NFAT 39. p12I was consistently pulled down in

calmodulin agarose bead and immunoprecipitation assays (Fig. 2.2B). Thus, p12I, as a

88

full-length protein, had the capacity to bind endogenously expressed calcineurin in both

293T and Jurkat T cells. As expected using the pMEp12I vector, the detected p12I in

higher percentage SDS gels (15 %) showed typical doublet bands as described 18. The

two forms of p12I may be due to post-translational modification of p12I, such as

phosphorylation or ubiquitylation 18.

To define the regions in p12I that are responsible for binding with calcineurin,

we tested serially deleted p12I mutants for calcineurin binding (Fig. 2.2A). The mutants

p12I 48-99, p12I 15-86, and p12I 15-69 bound calcineurin in immunoprecipitation and

calmodulin bead pull-down assays (Fig. 2.2B). However, the mutant p12I 1-47 (the N

terminal half of p12I) failed to bind calcineurin in either assay (Fig. 2.2B). These data

indicated that amino acids 48- 99 of p12I, which contains the putative calcineurin-

binding (70PSLPI/LT75) motif was likely responsible for calcineurin binding. The

mutant p12I 15-69, which did not contain the putative calcineurin binding motif, weakly

bound to calcineurin (Fig. 2.2B). This finding was not unexpected, because secondary

calcineurin binding sites are predicted from the p12I amino acid sequence when

compared to homologous alignment of secondary binding sequences of NFAT (data not

shown).

PSLPI/LT motif is responsible for interaction between p12I and calcineurin. To

more specifically test if the calcineurin-binding region in p12I was due to the PSLPI/LT

motif, we constructed an alanine substitution mutant of p12I (p12I AxAxAA) and

carried out immunoprecipitation and calmodulin bead pull-down assays (Fig. 2.3).

89

These substitutions were designed to replace critical proline, leucine, isoleucine, and

threonine amino residues with alanine residues (Fig. 2.3A). The p12I AxAxAA mutant

had significantly reduced binding affinity for calcineurin (Fig. 2.3B). This result

indicated that the PSLPI/LT motif in p12I is largely responsible for calcineurin binding.

Our data, however, did not exclude the possibility secondary binding sites in p12I as

implicated from our serially deletion mutant results. In addition, we constructed a

mutant L74I that converted a leucine from pMEp12I to an isoleucine more common at

this position in p12I sequences from most HTLV-1 strains including the ACH molecular

clone (PSLPLT vs. PSLPIT) (Fig. 2.1). There was no significant difference in binding

affinity between these two sequences (Fig 2.3). The other single substitution mutants in

the putative calcineurin-binding (70PSLPI/LT75) motif, P70A, S71A, P73A, and T75A

did not result in significant reduction in binding affinity (data not shown).

p12I and calcineurin binding is calcium-dependent but is not affected by

cyclosporin A. Calcineurin is activated via cooperative interactions between calcium

and calmodulin 42. Previous reports indicated that calcium was required for binding of

NFAT to calcineurin as a complex with calmodulin sepharose beads 39. To test the

calcium requirement for the association between p12I and calcineurin, we treated cell

lysates with the calcium chelators, EDTA and EGTA or used BAPTA-AM treated cells

prior to conducting immunoprecipitation experiments. The strong calcium chelator

EGTA eliminated p12I binding to calcineurin (Fig. 2.4, lane 4) whereas the weaker

calcium chelator EDTA and BAPTA-AM reduced the binding affinity in

90

immunoprecipitation assays (Fig. 2.4, lanes, 3, and 5). p12I binding was decreased when

cell lysates were treated with decreased concentration of CaCl2 or when treated with

increasing concentration of EDTA (data not shown). These data indicated that the

binding of p12I to calcineurin required calcium, which is similar to the requirement of

calcium for effective NFAT binding to calcineurin 39. In parallel, because the binding of

NFAT to calcineurin was not affected by calcineurin inhibitor cyclosoporin A 39, we

tested the affect of cyclosporin A in the p12I binding to calcineurin. As expected, p12I

binding was not inhibited by cyclosporin A (Fig. 2.4, lane 6). This result suggested that

the cyclosporin A-cyclophilin complex site is distinct from sites required for the

calcium/calmodulin activation and NFAT and p12I binding 43. Collectively our data

indicated that the calcineurin binding properties of NFAT and p12I were similar in

requiring calcium for calmodulin binding and each did not compete with binding of the

cyclosporin A-cyclophilin complex. The results suggested both proteins might have

similar binding mechanism or share a same binding site in calcineurin.

p12I competes with NFAT for calcineurin binding. Because synthetic peptides or

proteins, which contain the calcineurin-binding motif, compete with NFAT for

calcineurin binding 27,28,37, we tested if the p12I and calcineurin binding inhibits NFAT

and calcineurin binding. We transiently cotransfected 293T cells with a constant amount

of NFAT expression plasmid with increasing amounts of the wild type pMEp12I

expression plasmid (Fig. 2.5A), as well as plasmids expressing p12I mutants (p12I 1-47,

48-99,15-69, or AxAxAA mutants) (Fig 2.5B). We then performed calmodulin bead

pull-down assay to detect the amount of bound NFAT and wild type p12I or its mutant

91

proteins to the calcineurin A subunit. As expected, wild type p12I and the calcineurin

binding mutants, p12I 48-99, p12I 15-69 and p12I AxAxAA were bound in

concentration–dependent manner, but the non-calcineurin binding mutant p12I 1-47

expressed in a parallel manner failed to bind calcineurin (Fig. 2.5B). The amount of

calcineurin (A subunit) from pull-down experiments remained constant in all assays

(Fig. 2.5A). Both wild type p12I (Fig. 2.5A) and p12I 48-99 (Fig. 2.5B) effectively

competed against NFAT for calcineurin binding. In the non-calcineurin binding mutant

p12I 1-47 transfection, the amount of pull-downed NFAT was relatively constant

regardless of the amount of mutant p12I 1-47 transfection (Fig 2.5B). In addition, the

weakly calcineurin binding p12I 15-69 mutant, which did not contain calcineurin

binding sequence, did not compete with NFAT for calcineurin binding (Fig 2.5B). As

expected, mutation in calcineurin binding sequence (p12I AxAxAA mutant) resulted in

reduction of competition between p12I and NFAT (Fig 5B). Together, these results

indicated PSLPI/LT calcineurin binding sequence is critical for p12I competition with

NFAT for calcineurin binding.

The PSLPI/LT calcineurin binding sequence in p12I is an inhibitory motif for

NFAT transcriptional activity. p12I activates NFAT transcriptional activity in T cells

in a calcium-dependent manner 25. In contrast, the NFAT SPRIEIT calcineurin binding

sequence and synthetic peptides based on this motif inhibit NFAT transcriptional

activation by competing with NFAT for calcineurin binding 27. To further investigate if

p12I binding to calcineurin affected NFAT transcriptional activity, we performed an

NFAT driven reporter gene assay using p12I alanine substitution mutant, p12I AxAxAA,

92

which exhibited lowered binding affinity for calcineurin in the immunoprecipitation and

calmodulin bead pull-down assay. Jurkat T cells were transiently transfected with wild

type pMEp12I or p12I AxAxAA expression plasmids and then measured NFAT

transcriptional activities. While the wild type pMEp12I showed typical approximate 20

fold induction in the presence of PMA stimulation 23 compared to the pME empty

vector, the p12I AxAxAA mutant induced an approximate 40 fold enhancement of

luciferase activity compared to the pME control vector (Fig. 2.6). These data suggested

that the lowered binding between calcineurin and the p12I AxAxAA mutant allows

more NFAT to be available for the reporter gene assay. Alternatively, the calcineurin-

binding motif in p12I may function as a negative modulator motif for NFAT activation.

In addition, co-transfection of wild type pMEp12I and p12IAxAxAA mutant resulted in

approximate 25 fold induction over the pME control vector, which indicated pME p12I

and p12I AxAxAA did not compete for NFAT activation.

Lowered p12I binding affinity to calcineurin induces more NFAT nuclear

translocation. To test if p12I binding to calcineurin affected NFAT nuclear

translocation, we performed NFAT-GFP fusion protein localization assay when co-

expressed with wild type p12I or p12I AxAxAA mutant in Hela-Tat cells. As expected,

both wild type p12I and p12I AxAxAA mutant expression induced NFAT-GFP nuclear

translocation, while NFAT-GFP was retained in the cytoplasm in Hela-Tat cells

transfected with pME vector control (Fig. 2.7 A, B, and C). Nuclear translocation of

NFAT-GFP was activated by ionomycin and was inhibited by calcineurin inhibitor

cyclosporin A (Fig. 2.7 D). When untreated, cells transfected wild type p12I or p12I

93

AxAxAA mutant had approximate 80 % nuclear NFAT-GFP translocation, which was

5-6 fold greater than in pME control transfected (Fig. 2.7 D). We then measured GFP

intensity in nuclear and cytoplasmic regions of NFAT-GFP expressing cells to quantify

NFAT nuclear translocation. As demonstrated in Fig. 2.7 E, The nuclear/cytoplasmic

ratio of NFAT-GFP was less than 1 in the control pME transfected cells, approximately

2 in pME p12I and 2.5 in p12I AxAxAA mutant transfection. Thus p12I AxAxAA, the

lowered binding mutant to calcineurin, induced more NFAT translocation than wild

type p12I.

p12I binding to calcineurin does not affect phosphatase activity. Because many

calcineurin binding proteins inhibit calcineurin phosphatase activity 42, we tested if p12I

binding to calcineurin affected calcineurin phosphatase activity. We measured

endogenous calcineurin phosphatase activity in Jurkat T cells expressing pME, pME

p12I and the p12I AxAxAA mutant. As expected wild type p12I stimulated calcineurin

phosphatase activity ~ 3 fold over pME empty vector (Fig. 2.8). However, the

difference between phosphatase activity induced by wild type p12I and the p12I

AxAxAA mutant was not statically significant (p > 0.5). Wild type p12I, which binds

calcineurin with higher affinity compared to the p12I AxAxAA mutant, did not show

significant reduction in phosphatase activity (Fig. 2.8). Similar to NFAT, our data

indicated that p12I binding to calcineurin via PSLPI/LT calcineurin binding sequence

did not inhibit calcineurin phosphatase activity. These data indicate that the increased

NFAT transcription and nuclear translocation induced by the p12I AxAxAA mutant was

94

independent of calcineurin binding and was more likely due to the known calcium

mobilization properties of the viral protein.

2.4 DISCUSSION

Herein, we are the first to demonstrate that the HTLV-1 accessory protein p12I

binds to calcineurin and modulates NFAT activation. We identified in p12I, a highly

conserved calcineurin-binding motif, which is homologous to a PxIxIT calcineurin-

binding motif of NFAT.

p12I was consistently co-precipitated with calcineurin in both

immunoprecipitation and calmodulin bead pull-down assays. We found that PSLPI/LT

motif in p12I was largely responsible for calcineurin binding through the use of serial

deletion and alanine substitution mutants. Like NFAT, p12I binding required calcium

but was not inhibited by cyclosporin A. Thus, these two proteins appear to bind

calcineurin via similar motifs.

The PxIxIT calcineurin binding motif or highly related sequences are found in a

variety of calcineurin binding proteins of yeast, mammalian cells, and viruses 44. These

proteins are either calcineurin substrates or inhibitors of calcineurin mediated signaling.

For example, the yeast transcription factor Crz1p has a PIISIQ motif that mediates

calcineurin interaction 35 and its pattern of regulation is similar to that of NFAT.

Therefore, the dephosphorylation, nuclear accumulation and transcriptional activity of

Crz1p depend on this calcineurin-binding motif 44. Like, NFAT or Crz1p, p12I itself

may be a substrate of calcineurin. We are currently investigating if p12I is regulated by

95

calcineurin-mediated dephosphorylation. Also, calcineurin binding motifs are found in

calcineurin inhibitor proteins. Cellular protein MCIP1 and cain/cabin1 (calcineurin

inhibitor) are endogenous calcineurin inhibitors. MCIP1, which contains a PKIIQT

calcineurin-binding motif, is expressed abundantly in striated muscle, and inhibits

calcineurin dependent signaling pathways when over-expressed 35. Cain/cabin1 has a

PEITVT motif in a 38-amino acid putative calcineurin-binding domain 45. A viral

protein A238L of African swine fever virus, which inhibits NFAT transcriptional

activity, interacts with calcineurin via the PKIIIT sequence 37.

Our data indicates that PSLPI/LT calcineurin binding sequence of p12I

modulates NFAT nuclear translocation and transcription activity. This sequence of p12I

is critical for competition with NFAT for calcineurin binding. This result is consistent

with our previous study that tested NFAT driven reporter gene assay using truncation

mutants and alanine substitution mutants (AxxA) of putative PxxP motifs (SH3 domain

binding motifs) 46. There were two positive (aa 33-47, aa 87-99) and two negative

regions (aa 1-14, aa 70-86) for NFAT transcription in p12I and the mutation in the 3rd

PxxP motif (P70xxP73) resulted in increased NFAT activation about 2 fold more than

wild type. The 70th –75th PSLPI/LT calcineurin binding site corresponds to the aa 70-86

negative region and 3rd PxxP motif. This NFAT inhibitory function of the PSLPI/LT

motif was not from the inhibition of calcineurin phosphatase activity. Calcineurin

binding proteins such as AKAP 79, CHP, cain/cabin1, and A238L inhibit calcineurin

phosphatase activity by presumably affecting calcineurin active sites 42. However,

PxIxIT motif-mediated binding itself does not inhibit the catalytic activity of

calcineurin, because NFAT activation requires enzymatic activity of calcineurin, as well

96

as binding via this motif. Synthetic peptides based on PxIxIT motif inhibit NFAT

activation without affecting calcineurin catalytic activity 27,28. Likewise, p12I binding to

calcineurin via a similar motif does not inhibit calcineurin catalytic activity, but instead

influences NFAT and calcineurin interaction by competing for binding with NFAT

similar to artificial peptides representing this motif 27,28.

It is unclear why p12I has two regulatory functions for NFAT transcriptional

activity; positive modulation by increasing cytosolic calcium concentration from ER

stores 12 and negative modulation by calcineurin binding. Interestingly, Bcl-2 has these

similar properties with p12I, and 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 47,48. When Bcl-2 is at the ER membrane, its

function appears to increase ER calcium permeability, like p12I 47. Although its function

at the ER membrane is still poorly understood, it has been suggested that Bcl-2

functions as ion channel protein 47. On the other hand, Bcl-2 binds calcineurin via its

BH4 domain, and inhibits NFAT activity by sequestering calcineurin from NFAT

binding without affecting calcineurin catalytic activity 31,32. By inhibiting NFAT

activation, Bcl-2 may prevent Fas ligand (FasL) expression and further apoptosis. Like

Bcl-2, p12I may affect apoptosis in HTLV-1 infected T cells. Another ER membrane

protein, calcium-modulator and cyclophilin ligand (CAML) also has parallel properties

to p12I. It activates NFAT by increasing calcium flux when overexpressed in Jurkat T

cells 49,50. Also, CAML binds with calcineurin indirectly, through its association with

cyclophilin 51.

97

In summary, we demonstrate that the HTLV-1 accessory protein p12I interacts

with calcineurin, an important regulator of NFAT signaling, via a highly conserved

PSLPI/LT motif. Thus, HTLV-1, a retrovirus associated with lymphoproliferative

disease, expresses a conserved accessory protein to further T cell activation, an

important antecedent to effective viral infection, via a calcium/calcineurin/NFAT

pathway.

2.5 REFERENCES

(1) Seiki M, Hikikoshi A, Taniguchi T, Yoshida M. Expression of the pX Gene of HTLV-1: General Splicing Mechanism in the HTLV Family. Science. 1985;228:1532-1534.

(2) Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type 1. Proc Natl Acad Sci USA. 1992;89:8813-8817.

(3) Furukawa K, Furukawa K, Shiku H. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 1991;295:141-145.

(4) Berneman ZN, Gartenhaus RB, Reitz MS et al. 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. 1992;89:3005-3009.

(5) Mesnard JM, Devaux C. Multiple control levels of cell proliferation by human T- cell leukemia virus type 1 tax protein. Virology. 1999;257:277-284.

(6) Yoshida M. Multiple viral strategies of htlv-1 for dysregulation of cell growth control. Annu Rev Immunol. 2001;19:475-496.

(7) Lairmore MD, Albrecht B, D'Souza C et al. 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. 2000;16:1757-1764.

98

(8) Cereseto A, Berneman Z, Koralnik I et al. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia. 1997;11:866-870.

(9) Ciminale V, Pavlakis GN, Derse D, Cunningham CP, Felber BK. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J Virol. 1992;66:1737-1745.

(10) Dekaban GA, Peters AA, Mulloy JC et al. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology. 2000;274:86-93.

(11) Pique C, Uretavidal A, Gessain A et al. 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. 2000;191:567-572.

(12) Ding W, Albrecht B, Luo R et al. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J Virol. 2001;75:7672-7682.

(13) Koralnik IJ, Mulloy JC, Andresson T, Fullen J, Franchini G. Mapping of the intermolecular association of the human T cell leukaemia/lymphotropic virus type I p12(I) and the vacuolar H+- ATPase 16 kDa subunit protein. J Gen Virol. 1995;76:1909-1916.

(14) Koralnik IJ, Fullen J, Franchini G. 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. 1993;67:2360-2366.

(15) Franchini G. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood. 1995;86:3619-3639.

(16) Mulloy JC, Crowley RW, Fullen J, Leonard WJ, Franchini G. The human T-cell leukemia/lymphotropic virus type 1 p12(I) protein binds the interleukin-2 receptor beta and gamma(c) chains and affects their expression on the cell surface. J Virol. 1996;70:3599-3605.

(17) Johnson JM, Nicot C, Fullen J et al. 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. 2001;75:6086-6094.

99

(18) Trovato R, Mulloy JC, Johnson JM et al. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic leukemia virus type 1 p12 protein greatly influences its stability. J Virol. 1999;73:6460-6467.

(19) Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12(I) reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

(20) Albrecht B, Collins ND, Burniston MT et al. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J Virol. 2000;74:9828-9835.

(21) Collins ND, Newbound GC, Ratner L, Lairmore MD. In vitro CD4(+) lymphocyte transformation and infection in a rabbit model with a molecular clone of human T-cell lymphotropic virus type 1. J Virol. 1996;70:7241-7246.

(22) Kimata JT, Wong F, Wang J, Ratner L. Construction and characterization of infectious human T-cell leukemia virus type 1 molecular clones. Virology. 1994;204:656-664.

(23) Albrecht B, D'Souza CD, Ding W et al. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J Virol. 2002;76:3493-3501.

(24) Ding W, Albrecht B., Kelley RE et al. 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. 2002;76:10374-10382.

(25) Masuda ES, Imamura R, Amasaki Y, Arai K, Arai N. Signalling into the T-cell nucleus: NFAT regulation. Cell Signal. 1998;10:599-611.

(26) Rao A, Luo C, Hogan PG. Transcription factors of the NFAT family: regulation and function. Annu Rev Immunol. 1997;15:707-747.

(27) Aramburu J, Garcia-Cozar F, Raghavan A et al. Selective inhibition of NFAT activation by a peptide spanning the calcineurin targeting site of NFAT. Mol Cell. 1998;1:627-637.

(28) Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A. Science. 1999;285:2129-2133.

(29) Garcia-Cozar FJ, Okamura H, Aramburu JF et al. Two-site interaction of nuclear factor of activated T cells with activated calcineurin. J Biol Chem. 1998;273:23877-23883.

100

(30) Park S, Uesugi M, Verdine GL. A second calcineurin binding site on the NFAT regulatory domain. Proc Natl Acad Sci U S A. 2000;97:7130-7135.

(31) Shibasaki F, Kondo E, Akagi T, Mckeon F. Suppression of signalling through transcription factor NF-AT by interactions between calcineurin and Bcl-2. Nature. 1997;386:728-731.

(32) Srivastava RK, Sasaki CY, Hardwick JM, Longo DL. Bcl-2-mediated drug resistance: inhibition of apoptosis by blocking nuclear factor of activated T lymphocytes (NFAT)-induced Fas ligand transcription. J Exp Med. 1999;190:253-265.

(33) Lin X, Sikkink RA, Rusnak F, Barber DL. Inhibition of calcineurin phosphatase activity by a calcineurin B homologous protein. J Biol Chem. 1999;274:36125-36131.

(34) Kashishian A, Howard M, Loh C et al. AKAP79 inhibits calcineurin through a site distinct from the immunophilin-binding region. J Biol Chem. 1998;273:27412-27419.

(35) Rothermel B, Vega RB, Yang J et al. A protein encoded within the Down syndrome critical region is enriched in striated muscles and inhibits calcineurin signaling. J Biol Chem. 2000;275:8719-8725.

(36) Miskin JE, Abrams CC, Goatley LC, Dixon LK. A viral mechanism for inhibition of the cellular phosphatase calcineurin. Science. 1998;281:562-565.

(37) Miskin JE, Abrams CC, Dixon LK. African swine fever virus protein A238L interacts with the cellular phosphatase calcineurin via a binding domain similar to that of NFAT. J Virol. 2000;74:9412-9420.

(38) Northrop JP, Ullman KS, Crabtree GR. Characterization of the nuclear and cytoplasmic components of the lymphoid-specific nuclear factor of activated T cells (NF-AT) complex. J Biol Chem. 1993;268:2917-2923.

(39) Loh C, Shaw KT, Carew J et al. Calcineurin binds the transcription factor NFAT1 and reversibly regulates its activity. J Biol Chem. 1996;271:10884-10891.

(40) Saksena NK, Srinivasan A, Ge YC et al. Simian T cell leukemia virus type I from naturally infected feral monkeys from Central and West Africa encodes a 91-amino acid p12 (ORF-I) protein as opposed to a 99-amino acid protein encoded by HTLV type I from humans. AIDS Res Hum Retroviruses. 1997;13:425-432.

101

(41) Lefebvre L, Ciminale V, Vanderplasschen A et al. Subcellular localization of the bovine leukemia virus R3 and G4 accessory proteins. J Virol. 2002;76:7843-7854.

(42) Rusnak F, Mertz P. Calcineurin: form and function. Physiol Rev. 2000;80:1483-1521.

(43) Griffith JP, Kim JL, Kim EE et al. X-ray structure of calcineurin inhibited by the immunophilin-immunosuppressant FKBP12-FK506 complex. Cell. 1995;82:507-522.

(44) Boustany LM, Cyert MS. Calcineurin-dependent regulation of Crz1p nuclear export requires Msn5p and a conserved calcineurin docking site. Genes Dev. 2002;16:608-619.

(45) Lai MM, Burnett PE, Wolosker H, Blackshaw S, Snyder SH. Cain, a novel physiologic protein inhibitor of calcineurin. J Biol Chem. 1998;273:18325-18331.

(46) Ding W, Kim SJ, Nair AM et al. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol. 2003;77:11027-11039.

(47) Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

(48) He H, Lam M, McCormick TS, Distelhorst CW. Maintenance of calcium homeostasis in the endoplasmic reticulum by Bcl-2. J Cell Biol. 1997;138:1219-1228.

(49) Holloway MP, Bram RJ. A hydrophobic domain of Ca2+-modulating cyclophilin ligand modulates calcium influx signaling in T lymphocytes. J Biol Chem. 1996;271:8549-8552.

(50) von Bulow GU, Bram RJ. NF-AT activation induced by a CAML-interacting member of the tumor necrosis factor receptor superfamily. Science. 1997;278:138-141.

(51) Holloway MP, Bram RJ. Co-localization of calcium-modulating cyclophilin ligand with intracellular calcium pools. J Biol Chem. 1998;273:16346-16350.

102

Figure 2.1. HTLV-1 p12I contains a highly conserved putative calcineurin-binding

motif in p12I. Sequences from HTLV-1 and STLV-1 were aligned with calcineurin

binding sequence (SPRIEIT) of NFAT using a commercial software program (Align

X,Vector NTI version 7.0). ACH 22 pMEp12I 16, HTLV-1A (Entrez Protein AAA45390),

HTLV-1B (AAA85329), HTLV-1C (AAB23360), HTLV-1D (B46181), HTLV-1E

(C61547), and STLV-1 isolate number 14 40 were compared. The conserved SPSLPI/LT

sequence is indicated in shaded box and black boxes indicate amino acid residues that

have similar properties. Consensus motif is indicated in bottom row.

103

Figure 2.2. The C-terminal half of p12I containing the PSLPI/LT motif is required

for calcineurin binding. (A) Schematic representation of wild type pMEp12I and serial

p12I deletion mutants. Shaded boxes indicate PxxP motifs and bar represents the

location of the PSLPI/LT conserved sequence. (B) Co-precipitation of wild type p12I

and serial deletion p12I mutants with calcineurin in immunoprecipitation (IP) and

calmodulin agarose bead pull-down (CaM bead) assays. Jurkat T or 293T cells

transfected with pME empty vector wild type pMEp12I or mutant p12I plasmids were

lysed and analyzed by Western blot. Polyclonal calcineurin antibody was used for

immunoprecipitation assay. Monoclonal HA-1 antibody was used to detect wild type

and mutant p12I.

104

Figure 2.3. Alanine substitution mutant of p12I (AxAxAA) decreases binding

affinity for calcineurin. (A) Schematic representation of wild type pMEp12I, alanine

substitution mutant p12I AxAxAA, and L74I plasmid. Grey boxes are PxxP motifs and

bar represents the PSLPI/LT conserved sequence. (B) 293T cells were transfected with

wild type pMEp12I, p12I AxAxAA, and L74I plasmid. Binding was tested by

immunoprecipitation (IP) and calmodulin agarose bead pull-down (CaM bead) assays.

105

Figure 2.4. p12I and calcineurin binding is inhibited by calcium chelators but not

inhibited by cyclosporin A. CaCl2 (2 mM), Calcium chelators EDTA (5 mM) (lane 3),

or EGTA (5 mM) (lane 4) were added in Triton X-100 cell lysis buffer and calcium

chelator BAPTA-AM (1 µM) (lane 5), or calcineurin inhibitor cyclosporin A (10 µM)

(lane 6) were added in 293T cells 30 min before cell lysis and following

immunoprecipitation (IP) assay.

106

Figure 2.5. p12I binding to calcineurin decreases the amount of NFAT binding to

calcineurin (A) Constant amount of HA-NFAT1c-GFP (7.5 µg) and increasing amount

of pMEp12I (0-15 µg) were cotransfected into 293T cells. The amount of bound NFAT

and wild type p12I proteins to calcineurin A subunit were detected by calmodulin pull-

down assay. (B) 293T cells were cotransfected with increasing amount (0-15 µg) of

wild type pMEp12I and mutant (p12I 1-47, 48-99,15-69, or AxAxAA mutants) plasmids

and constant amount of HA-NFAT1c-GFP (7.5 µg). Then wild type p12I or p12I

mutants, NFAT and calcineurin A subunit were precipitated by calmodulin bead.

Optical density of NFAT indicating the amount of calcineurin binding was measured.

Optical density of each NFAT bands was normalized by corresponding calcineurin A

bands.

107

Figure 2.6. The substitution mutation in calcineurin binding sequence in p12I

induces increased NFAT transcription activity and did not compete against wild

type p12I for NFAT transcription activity. Jurkat T cells were co-transfected with 30

µg of effector plasmid [pME control vector, pMEp12I, p12I AxAxAA mutant, or both

pMEp12Iand p12I AxAxAA mutant (15 µg each)] and 10 µg of NFAT reporter plasmid

and then were left un-stimulated or stimulated with PMA. Each bar represent the fold

induction of NFAT-luciferase activity in wild type pMEp12I, p12I AxAxAA mutant, or

both plasmids transfected cells over luciferase activity of the pME empty vector

transfected cells. Values were the means of quadruplicate samples in two independent

experiments and were statistically significant (*, P < 0.01, pMEp12I compared to

p12IAxAxAA alone).

108

Figure 2.7. Reduced p12I binding to calcineurin induces increased NFAT nuclear

translocation. Confocal microscopic image of Hela-Tat cell cotransfected with HA-

NFAT1c-GFP expression vector and A) pME empty vector, B) pMEp12I, or C) p12I

AxAxAA mutant expression vectors.

109

D) The percent of cells that had nuclear predominant fluorescence. Cells were untreated,

stimulated with ionomycin (3 µM, 10 min before fixation), or inhibited with

cyclosporin A (1 µM, 16 h before fixation). E) Nuclear/cytoplasmic ratio of NFAT-GFP

intensity of cells transfected with pME, pMEp12I or p12I AxAxAA vector (*, P < 0.05,

pME vector control compared to pMEp12I. **, P < 0.05, p12I AxAxAA compared to

pMEp12I).

110

Figure 2.8. Calcineurin phosphatase activity is not affected by p12I binding to

calcineurin. Endogenous calcineurin activity from Jurkat T cells transfected with pME,

pMEp12I or p12I AxAxAA mutant was tested by measuring free phosphate (nmol)

released from phosphopeptide substrate (RII peptide) (*, P < 0.01 vs. pMEp12I and p12I

AxAxAA).

111

CHAPTER 3

ENHANCEMENT OF LFA-1-MEDIATED T-CELL ADHESION BY

HUMAN T-LYMPHOTROPIC VIRUS TYPE 1 p12I

3.1 INTRODUCTION

Human T-lymphotropic virus type 1 (HTLV-1), the first identified human

retrovirus, infects approximately 15 to 25 million people worldwide 1 and causes adult

T-cell leukemia/lymphoma (ATLL) and other lymphocyte-mediated diseases 2-5. The

primary targets for HTLV-1 infection are CD4+ T-cells, although CD8+ T-cells, B-cells,

and macrophages also can be infected 4,6. HTLV-1 is naturally transmitted in a highly

cell-associated manner via beast milk, semen and blood 2,7,8. Thus, cell-to-cell contact

is required for efficient HTLV-1 transmission both in vivo 9 and in vitro 10,11.

Furthermore, HTLV-1 transmission occurs directly through cell-to-cell junctions

112

(virological synapse) that mimic immunological synapses 12,13. HTLV-1- infected cells

have reorganized microtubule-organizing centers (MTOC) near points of cell contacts 14.

At these sites structural proteins such as Gag, along with the HTLV-1 genome,

aggregate and transfer from infected to target lymphocytes. Interestingly, microtubule

reorganization is triggered by antibodies to ICAM-1 (CD54) or the IL-2 receptor

(CD25) 14,15. While HTLV-1 Tax contributes to microtubule reorganization 12-15, it is

unclear whether other viral proteins, known to be important in the early events of

infection modulate cell-to-cell contact to accomplish viral transmission.

The HTLV-1 genome encodes structural and enzymatic (Gag, Pol, and Env),

regulatory (Tax and Rex) and nonstructural or “accessory” (p12I, p27I, p13II, and p30II)

proteins 16-20. HTLV-1 nonstructural proteins are important in viral replication and viral

spread in vivo 18,20-23. A highly conserved pX ORF I-encoded protein, p12I, is expressed

during the natural viral infection 24,25 and is critical for the establishment of HTLV-1

infection in non-activated T-cells in vitro and in animals models 26,27. p12I localizes to

the endoplasmic reticulum (ER) and Golgi apparatus, releases calcium from ER stores

to trigger nuclear factor of activated T-cells (NFAT)-mediated transcription leading to

IL-2 production and T-cell activation 21,28-30.

One of the key events of T-cell activation is the modulation of the expression of

adhesion molecules such as lymphocyte function-associated antigen 1 (LFA-1), which

mediates lymphocyte adherence to the vascular wall and lymphocyte migration into

tissues, and contributes to the immunological synapse 31. LFA-1 is a heterodimeric

transmembrane molecule composed of a unique α subunit (αL; CD11a) and a β2 subunit

(CD18) 32,33. Activation of LFA-1 is tightly controlled by “inside-out” signaling, which

113

is induced by ligation of the T-cell receptor (TCR) or certain chemokine receptors 32-35.

LFA-1-mediated T-cell adhesion results from either increasing the affinity for its ligand,

intercellular adhesion molecule-1 (ICAM-1), or clustering of LFA-1 (avidity) on the

cell membrane 32-35. Although recent studies suggested that Talin, Rap1 and RapL

proteins are required for regulation of LFA-1 expression 31,36, the detailed mechanism of

signaling leading to increased affinity is unclear 34. Under experimental conditions,

divalent cations such as Mn2+ or Mg2+ and activating antibodies can induce affinity

activation of LFA-1 34. The signaling pathways that result in increased integrin lateral

mobility and clustering include proteins such as Vav-1 (guanine exchange factor),

ADAP, SKAP-55 (adaptor protein), and Rap1 (small GTPase) 34,37-40. Protein kinase C

(PKC) activation by phorbol esters also can trigger integrin clustering by interacting

with the β2 subunit 34,40. Furthermore, integrin clustering can be induced by elevated

cytoplasmic calcium via TCR activation or treatment with calcium mobilizers such as

ionomycin and thapsigargin (TG) 41. Calcium-mediated LFA-1 activation also requires

the calcium-dependent protease calpain 41. Interestingly, the HTLV-1 accessory protein

p12I mimics the function of TG and induces calcium release from the ER stores to

activate T-cells 29.

Herein, we tested if the expression of HTLV-1 p12I could activate LFA-1-

mediated T-cell adhesion in a calcium-dependent manner. LFA-1-mediated adhesion of

wild type HTLV-1 (ACH) T-cells was greater than T-cells expressing a proviral clone

lacking p12I (ACH.p12). Expression of p12I in Jurkat T-cells also enhanced LFA-1-

mediated cell adhesion, which was inhibited by the calcium chelator BAPTA-AM, the

calcium channel blocker SK&F 96365 and calpeptin, an inhibitor of the calcium-

114

dependent protease calpain. Similar to the effects of TG, the expression of p12I in

Jurkat T-cells induced cell surface clustering of LFA-1 without changing LFA-1

expression. Our data indicate that p12I induces LFA-1 clustering on T-cells via

calcium-dependent signaling, which would promote spread of HTLV-1 to target T-cells.

3.2 MATERIALS AND METHODS

Cell lines

ACH.2 and ACH.p12.4 cell lines were generated from the outgrowth of

immortalized PBMC previously transfected with molecular clones of HTLV-1 (ACH

and ACH.p12I) 29,42. The ACH.p12I plasmid has a mutation resulting in complete

ablation of p12I mRNA expression without affecting the expression of other viral genes

26. ACH.2p and ACH.p12.4p cells were infected cell lines created by co-culturing naïve

human PBMC with lethally γ-irradiated (10,000 rads) ACH.2 and ACH.p12.4 cells,

respectively. Target PBMC were pre-activated for four days with human IL-2

(10 U/ml) and phytohemagglutinin (PHA) (2 µg/ml). ACH.2, ACH.p12.4, ACH.2p and

ACH.p12.4p cells were maintained in RPMI 1640 supplemented with 15% fetal bovine

serum, L-glutamine (0.3 mg/ml), penicillin (100 U/ml), streptomycin (100 µg/ml), and

recombinant IL-2 (10 U/ml). Viral p19 antigen production from the ACH.2, ACH.p12.4,

ACH.2p and ACH.p12.4p cell lines was measured in cell culture supernatants by

enzyme-linked immunosorbent assay (ELISA) in quadruplicate samples (Zeptomatrix)

according to the manufacturer's protocol. Jurkat T-cells (clone E6-1; American Type

Culture Collection catalog number TIB-152) were maintained in RPMI 1640

115

(Invitrogen) supplemented with 10 % fetal bovine serum, 100 µg of streptomycin-

penicillin/ml, and 2 mM L-glutamine. For lentivirus production, the 293T cell line

(American Type Culture Collection), which stably expresses the simian virus 40 (SV40)

T antigen, was maintained in Dulbecco’s modified Eagle medium (DMEM, Invitrogen)

supplemented with 10% fetal bovine serum (FBS), 100 µg/ml streptomycin/penicillin

plus 2 mM L-glutamine.

Recombinant lentivirus production and infection of Jurkat T-cells

Recombinant lentiviruses were prepared as previously described 21. pWPT-

IRES-GFP (empty plasmid) was generated by cloning an internal ribosome entry site

(IRES) sequence (pHR’CMV/Tax1/eGFP, G. Feuer, SUNY, Syracuse, NY) into

pWPT-GFP (D. Trono, University of Geneva, Geneva, Switzerland). pWPT-p12IHA-

IRES-GFP was generated by cloning p12IHA from pME-p12I (G. Franchini, National

Cancer Institute, National Institutes of Health) into pWPT-IRES-GFP. To generate

recombinant virus, 293T cells (5 x 106) were transfected with 2 µg of pHCMV-G, 10 µg

of pCMV∆R8.2 and 10 µg of pWPT-p12IHA-IRES-GFP or pWPT-IRES-GFP using

calcium phosphate. Supernatants were collected at 24, 48 and 72 h post transfection,

and filtered through a 0.2 µm filter. To obtain viral pellets, supernatants were

centrifuged at 6,500 xg for 16 h at 4° C, and then pellets were suspended in DMEM

medium overnight at 4° C. The concentrated virus was aliquotted and stored at -80°C.

After determination of viral titer, Jurkat T-cells were infected with recombinant virus at

a multiplicity of infection of 3 in the presence of 8 µg/ml polybrene (Sigma) and spin-

infected at 2700 rpm for 1 h at 25° C. Expression of eGFP and p12I was confirmed by

116

flow cytometry, western blot or reverse transcription- polymerase chain reaction (RT-

PCR) 21. For detection of p12I mRNA, cellular RNA was isolated using RNAqueous

(Ambion). RNA was converted to cDNA using the Reverse Transcription system

(Promega) and was amplified with AmpliTaq DNA polymerase (Perkin Elmer) using

CCTCTTTCTCCCGCTCTTTT (forward) and GGCCAAGCTAGCGTAATCTG

(reverse) primers. Transduced Jurkat T-cells (control and p12I expressing cells) were

used for experiments at 10 to 60 days post infection when expression of eGFP was

greater than 85% as determined by flow cytometric analysis.

Cell adhesion to immobilized ICAM-1

Cell-adhesion assays were performed as described elsewhere 41,43. Briefly,

recombinant human ICAM-1/Fc chimera (R&D system) in phosphate-buffered saline

(PBS) (0.5ug/well) was used for coating 96-well tissue culture plates overnight at 4° C.

Nonspecific binding was blocked with bovine serum albumin (BSA) in PBS for 1 h at

room temperature followed by three washes in PBS and once wash with RPMI 1640 or

HEPES buffer (20 mM HEPES, 140 mM NaCl, 2 mg/ml glucose, pH 7.4) for

MgCl2/EGTA stimulation. Cells (2 x 105 in 50 µl of RPMI 1640 or HEPES buffer)

were added to each well in the presence or absence of 50 µl of integrin stimulatory or

inhibitory reagents (2 x final concentrations) as indicated in results and figure legends.

Ionomycin (Sigma), TG (Calbiochem), phorbol 12-myristate 13-acetate (PMA, Sigma)

and MgCl2/EGTA were used as stimuli to test LFA-1-mediated adhesion. The

intracellular calcium chelator BAPTA-AM (Invitrogen), calcium release-activated

calcium (CRAC) channel blocker SK&F 96365 (Calbiochem) and calpain inhibitor

117

calpeptin (Calbiochem) were used for LFA-1 inhibition. The plates were incubated on

ice for 20 min, were centrifuged for one min at 30 ×g and incubated for 30 min at 37°

C. The plates were washed carefully with pre-warmed RPMI 1640 or HEPES buffer

three times. The percentage of attached cells was calculated by viable cell staining with

tetrazolium dye (CellTiter 96 Cell Proliferation Assay, Promega). Statistical differences

between groups were determined using the Wilcoxon- Mann Whitney test because we

assumed that differences between the two groups were not normally distributed.

P < 0.05 was considered significant.

Flow cytometry and soluble ICAM-1Fc binding assay

Flow cytometric analysis of the expression of cell surface markers CD11a and

CD18 was performed as described 42. Red-phycoerythrin (R-PE)-conjugated anti-

CD11a and FITC-conjugated anti-CD18 antibodies were used according to the

manufacturer's recommendations (Southern Biotechnology). For soluble ICAM-1 Fc

(sICAM-1) binding assay, cells were washed with RPMI 1640 or HEPES buffer and

resuspended into 50 µl of stimulator (2 x final concentrations) containing media or

HEPES buffer. Cells were mixed with 50 µl of recombinant human ICAM-1 Fc

chimera (R&D System) at a final concentration 5 µg/ml and incubated for 30 min at 37°

C. Subsequently, the cells were washed twice in ice-cold 0.2 % BSA and 0.2 % sodium

azide in PBS and incubated with R-PE-conjugated goat anti-human IgG Fc-specific

antibody (Jackson Immunoresearch Labs) for 20 min on ice. Cells were washed twice

in ice-cold 0.2 % BSA and 0.2 % sodium azide in PBS and fixed in 1 %

118

paraformaldehyde. Fluorescence of cells was measured by FACScan (Becton

Dickinson) analysis.

Confocal microscopy to detect LFA-1 distribution

For detection of LFA-1 distribution on the surface of T-cells by confocal

microscopy, Jurkat T-cells were attached to cover slips coated with a 0.01 % solution of

poly-L-lysine (Sigma). Cells were incubated for 30 min at 37° C with or without

stimulation and then were fixed with 0.5 % paraformaldehyde for 10 min. After

washing once with PBS and three times with RPMI 1640, cells were stained with R-PE-

conjugated CD11a antibody (Southern Biotechnology) for 30 min at 37° C, washed

once with RPMI 1640, mounted and analyzed using the Leica TCS SP2 AOBS confocal

system (Leica, Heidelberg, Germany). Leica Confocal Software (Leica Lite) was used

to quantify the fluorescence signal on cell membranes representing LFA-1 clustering as

described previously 41. Student’s t test was used for statistical analysis of differences

between two groups.

3.3 RESULTS

LFA-1-mediated adhesion in HTLV-1 immortalized T-cells lacking p12I expression

To examine the role of p12I in LFA-1-mediated cell adhesion, we first carried

out adhesion assays using wild type HTLV-1 (ACH cell line; ACH.2 and ACH.2p) and

T-cells lacking p12I expression (ACH. p12I; ACH.p12.4 and ACH.p12.4p). We

generated newly infected ACH.2p and ACH.p12.4p cells by co-culturing pre-activated

119

naïve human PBMC with lethally irradiated ACH.2 and ACH.p12.4, respectively, to

rule out the influence of specific clonal effects during in vitro culture. Virus replication

measured by viral p19 antigen ELISA, was similar among ACH.2, ACH.p12.4, ACH.2p

and ACH.p12.4p (Fig. 3.1A). These results are consistent with previous studies that

indicate that p12I is not required for HTLV-1 replication in activated T-cells or

immortalized T-cell lines 26,44.

LFA-1-mediated T-cell binding can result either from increased affinity for

ICAM-1 or clustering of LFA-1 on the cell surface 32-35. LFA-1 clustering can be

activated by TCR signaling, phorbol ester stimulation or increased intracellular calcium

levels 32-35. Thapsigargin and PMA-induced LFA-1-mediated adhesion of the wild type

ACH cell lines, ACH.2 and ACH.2p, was two-fold greater compared to unstimulated

cells (Fig. 3.1B). T-cells lacking p12I expression (ACH.p12.4 and ACH.p12.4p) did not

respond to the LFA-1 clustering reagents thapsigargin or PMA (Fig. 3.1B).

Mg2+/EGTA treatment, which resulted in an increase of LFA-1 affinity in vitro,

increased ACH.2 and ACH.2p cell adhesion up to 50-60%, but did not induce

significant adhesion in ACH.p12.4 and ACH.p12.4p cells (Fig. 3.1B). The failure of

ACH.p12.4 and ACH.p12.4p cells to respond to reagents that cause LFA-1 cluster and

enhance T-cell adhesion could be explained by these cells having either lowered LFA-1

expression or decreased binding affinity.

ACH.p12I cells express LFA-1, but have decreased binding of sICAM-1

To examine if lowered LFA-1 mediated adhesion in ACH.p12 cells was due to

changes in LFA-1 expression, we performed flow cytometric analysis using anti-CD11a

120

and anti-CD18 antibodies. There was no significant difference in integrin expression on

the cell surface among ACH.2, ACH.2p, ACH.p12.4 and ACH.p12.4p (Fig. 3.2A).

T-cell binding to sICAM-1 is increased only when LFA-1 affinity is enhanced

by agonists such as Mg2+-stimulation. In contrast, LFA-1 clustering on T-cells is

stimulated by reagents that stimulate calcium release from cell stores, such as TG or

reagents that induce TCR signaling, such as PMA 41,43. We therefore tested ACH wild

type and ACH.12 mutant T-cells for their sICAM-1 binding properties using flow

cytometry to measure LFA-1 affinity. As expected, none of the clustering stimulants

(TG and PMA) exhibited any induction of sICAM-1 binding in any of the cell lines (Fig.

3.2B). Mg2 -stimulated wild type ACH, ACH.2 and ACH.2p cells had enhanced

sICAM-1 binding, when compared to the ACH.p12I cell lines, ACH.p12.4 and

ACH.p12.4p (Fig. 3.2B). These data were consistent with our Mg2+/EGTA experiments

using immobilized ICAM-1 (Fig. 3.1B).

p12I expression in Jurkat T-cells did not affect LFA-1 affinity

To test whether HTLV-1 p12I expression specifically influenced LFA-1-

mediated T-cell adhesion, we stably expressed p12I in Jurkat T-cells using recombinant

lentiviruses (21). All empty control and p12I-expressing Jurkat T-cells were more than

85% GFP positive (Fig. 3.3A). Expression of HTLV-1 p12I also was confirmed by p12I

mRNA amplification by RT-PCR (Fig. 3.3B) and immunoprecipitation using polyclonal

HA-1 antibody (data not shown).

121

sICAM binding was tested using both empty control and p12I Jurkat T-cells. As

expected, TG and PMA, which activate LFA-1 clustering, did not induce sICAM-1

binding (Fig. 3.3A). Mg2+/EGTA treatment, which stimulates LFA-1 affinity, was used

to compare sICAM-1 binding in both empty control and p12I Jurkat T-cells. Our data

indicated no significant difference in LFA-1 affinity between the two cell types (Fig.

3.3A). These data demonstrated that p12I expression does not affect LFA-1 affinity.

Expression of p12I in Jurkat T cells induced LFA-1 clustering

Calcium mobilizers, such as ionomycin and TG enhance LFA-1-mediated T-cell

adhesion by affecting LFA-1 clustering 34,41. Since HTLV-1 p12I functionally mimics

TG by increasing intracellular calcium release from ER stores, we tested if p12I

expression might affect LFA-1 clustering. Jurkat T-cells that stably expressed p12I

exhibited enhanced cell adhesion to ICAM-1 molecules compared to empty vector-

transduced control Jurkat T-cells in all untreated, TG, ionomycin, PMA and

Mg2+/EGTA treatment groups (Fig. 3.4A). Since p12I did not affect LFA-1 affinity (Fig.

3.3), the enhancement observed in Mg2+/EGTA treated p12I Jurkat T-cell likely was due

to pre-clustered LFA-1 in p12I-expressing cells.

To clarify whether p12I expression affected clustering or conformational

changes of LFA-1, we measured the fold increase of cell adhesion over untreated cells

in the empty control and p12I Jurkat T-cells that were treated with increasing

concentrations of TG, ionomycin, PMA or Mg2+(Fig. 3.4B, C, D and E). As expected,

the calcium mobilizers TG and ionomycin induced increased cell adhesion in a dose-

dependent manner at lower concentrations and led to decreased cell adhesion at higher

122

concentrations in both cell types (Fig. 3.4B and C). The peak adhesion by TG and

ionomycin was observed between 2.5 to 5 µM and 0.5 to 1 µM, respectively, similar to

a previous report 41. Another LFA-1 clustering agent, PMA, also induced a dose-

dependent increase in adhesion to ICAM-1 at lower concentrations, but adhesion was

saturated at higher concentrations in both cell types (Fig. 3.4D). At the lower

concentrations, p12I Jurkat T-cell adhesion was higher than that of empty control Jurkat

T-cells after TG, ionomycin, or PMA stimulation (Fig. 3.4B, C, and D), suggesting that

p12I expression on Jurkat T-cells reduced the threshold required for the activation of

LFA-1 clustering.

As expected, Mg2+ also induced a dose-dependent increase and saturation of cell

adhesion in both cell types, however, the fold increase over untreated cells was not

significantly different between empty control and p12I Jurkat T-cells (Fig. 3.4E). The

affinity of both cell types was increased to a similar level by Mg2+, suggesting that

expression of p12I did not affect LFA-1 affinity. These data indicated that enhanced

LFA-1-mediated cell adhesion in p12I Jurkat T-cells was due to enhanced LFA-1

clustering rather than increased affinity.

HTLV-1 p12I - mediated LFA-1 activation is calcium dependent

LFA-1 clustering requires signaling pathways leading to the elevation of

calcium and activation of calpain, a calcium-dependent protease 41. To determine

whether LFA-1-activation induced by p12I was dependent on calcium-mediated

signaling pathways, we performed cell adhesion assays using inhibitors such as the

intracellular calcium chelator BAPTA-AM, the calcium channel blocker SK&F 96365,

123

and the calpain inhibitor calpeptin (Fig. 3.5). As expected, all inhibitors abolished TG

and PMA-induced cell adhesion in both cell types (Fig. 3.5). These results suggested

that LFA-1 clustering, due to p12I, was dependent on both intracellular and extracellular

calcium, as well as calpain. In contrast to TG and PMA stimulation, cell adhesion

induced by Mg2+/EGTA was not blocked completely by the inhibitors (Fig. 3.5). These

data were consistent with a previous report that Mg2+-induced affinity increases were

not altered by calcium channel blockers or calpain inhibitors 41.

HTLV-1 p12I modulated surface distribution of LFA-1

To confirm if the membrane distribution of LFA-1 was altered in p12I-

expressing Jurkat T-cells, we measured LFA-1 expression by confocal microscopy

using R-PE-conjugated CD11a. Untransduced mock Jurkat T-cells were either

untreated or treated with 1 µM of TG as a control experiment. A higher intensity of

LFA-1 fluorescence was observed frequently in TG-treated cells (Fig. 3.6A). Similar to

TG treatment, p12I expression in Jurkat T-cells had more than a two-fold enhanced

fluorescence indicating increased LFA-1 clustering on the cell membranes (Fig. 3.6B).

Increased LFA-1 fluorescence in p12I expressing Jurkat T cells was not due to increased

cell surface expression of LFA-1. Cell surface expression of CD11a was measured by

flow cytometry and was not significantly different between empty control and p12I

expressing Jurkat T cells (Fig. 3.6C). These data indicated that p12I expression altered

LFA-1 cell surface distribution favoring clustering of the adhesion molecule without

changing total LFA-1 cell surface expression and were consistent with our cell adhesion

and sICAM-1 binding assays.

124

3.4 DISCUSSION

In this study, we found that HTLV-1 p12I expression in T-cells altered LFA-1-

mediated T-cell adhesion to ICAM-1. Abrogation of p12I expression in HTLV-1

infected cells (ACH.p12I) resulted in decreased LFA-1-mediated cell adhesion, which

was due to a reduced affinity of LFA-1 compared wild type HTLV-1-infected cells. To

identify the specific role of p12I in LFA-1-mediated cell adhesion, we stably expressed

p12I in Jurkat T-cells using lentiviral vectors. Expression of p12I increased LFA-1-

mediated cell adhesion to ICAM-1 by inducing clustering of LFA-1 on Jurkat T-cells.

Our data are consistent with previous reports from our group and others 26-28,45 that

demonstrated the important role of HTLV-1 p12I in T-cell activation and cell-to-cell

transmission during the early stages of HTLV-1 infection.

An important role of LFA-1 and ICAM-1 in HTLV-1 cell-to-cell transmission

was implicated in a recent study by Barnard et al 14. By cross-linking with antibodies to

cell surface molecules, they identified that LFA-1 and ICAM-1 were involved in

triggering microtubule-organizing center (MTOC) polarization, which is characteristic

of HTLV-1-induced virological synapses. Our data suggested that p12I may synergize

to modify cell adhesion and contribute to MTOC polarization by increasing cell

adhesion through clustered LFA-1. Interestingly, transmission of other viruses is also

enhanced by the LFA-1/ ICAM-1 interaction. For example, measles virus infection

induced increased LFA-1 expression on monocytes resulting in enhanced viral

transmission to endothelial cells 14,46. Furthermore, LFA-1 activation of T-cells

125

increased their susceptibility to human immunodeficiency virus type 1 (HIV-1)

infection 47,48.

Our results presented herein, support our previous findings, which demonstrated

that T-cells expressing ACH. p12I have reduced viral infectivity when co-cultured with

quiescent T-cells 27. ACH.p12I cells do not exhibit efficient cell-to-cell contact and

therefore are disadvantaged for cell-to-cell transmission. This tenet would explain, in

part, data indicating that HTLV-1 p12I is required for in vivo viral infectivity in rabbits

26. The early block in infection observed in this animal study could be explained by

inadequate LFA-1 activation on HTLV-1-infected cells, which would inhibit the ability

of HTLV-1 to spread to non-activated target cells. Our findings are also consistent

with studies that indicate that HTLV-1 infection is established by early viral spread to

T-cells, but subsequently maintained by clonal expansion of T-cells49-51. If the virus

lacks the ability to rapidly spread from infected cells to its targets early in the infection,

it is likely that HTLV-1 infection is eliminated by a robust immune response52.

The observed affinity reduction in ACH.p12I cells might have resulted from

alteration of inside-out signaling. Although signaling pathways that regulate LFA-1

affinity are incompletely defined compared to that of LFA-1 clustering modulation 34,

transient affinity increases of LFA-1 and α1β4 are induced by chemokine stimulation

53,53,54. In addition, the lymphocyte specific protein tyrosine kinase, Lck, has been

identified as a critical enzyme to maintain the high affinity status of integrin α1β4 55.

Interestingly, expression of Lck gene was down-regulated by the HTLV-1 accessory

protein p30II 56. Thus, a balance of HTLV-1 regulatory and accessory protein

expression is likely required to ensure cell-to-cell transmission of the virus.

126

HTLV-1 p12I expression in Jurkat T-cells induced LFA-1-mediated cell

adhesion in a calcium-dependent manner. p12I mimics TG, which is an LFA-1

clustering agent, by releasing calcium from ER stores and increasing intracellular

calcium concentration without activation of the TCR signaling pathway 29. We cannot

rule out the possibility that p12I may activate LFA-1-mediated adhesion via chemokine

secretion, which can increase affinity and induce clustering. We have reported that p12I

enhanced IL-2 production through NFAT activation 28,30,45. IL-2 produced by HTLV-1

infected cells may activate LFA-1 on both infected and target cells in an autocrine or

paracrine manner in the local tissue microenvironment in vivo 30,45. Our data presented

herein, together with these previous studies suggests that p12I expression in HTLV-1-

infected T-lymphocytes lowers the threshold of calcium signaling events to promote T-

cell activation and cell to cell transmission.

In summary, our data indicate that HTLV-1 p12I expression induced

LFA-1-mediated T-cell adhesion to ICAM-1 by activating LFA-1 clustering in a

calcium-dependent manner. Our data demonstrate that p12I enhances HTLV-1 cell-to-

cell transmission during the early stages of infection by activating LFA-1-mediated

adhesion and by supporting the formation of the HTLV-1 virological synapse.

3.5 REFERENCES

(1) Proietti FA, Carneiro-Proietti AB, Catalan-Soares BC, Murphy EL. Global epidemiology of HTLV-I infection and associated diseases. Oncogene. 2005;24:6058-6068.

(2) Mahieux R, Gessain A. HTLV-1 and associated adult T-cell leukemia/lymphoma. Rev Clin Exp Hematol. 2003;7:336-361.

127

(3) Franchini G, Fukumoto R, Fullen JR. T-cell control by human T-cell leukemia/lymphoma virus type 1. Int J Hematol. 2003;78:280-296.

(4) Kehn K, Berro R, de La FC et al. Mechanisms of HTLV-1 transformation. Front Biosci. 2004;9:2347-2372.

(5) Hall WW, Fujii M. Deregulation of cell-signaling pathways in HTLV-1 infection. Oncogene. 2005;24:5965-5975.

(6) Manns A, Hisada M, Lagrenade L. Human T-lymphotropic virus type I infection. Lancet. 1999;353:1951-1958.

(7) Weiss RA. Milk-borne transmission of HTLV-1. [Review]. Japanese Journal of Cancer Research. 1993;84(3):inside front cover.

(8) Bangham CR. The immune control and cell-to-cell spread of human T-lymphotropic virus type 1. J Gen Virol. 2003;84:3177-3189.

(9) Okochi K, Sato H, Hinuma Y. A retrospective study on transmission of adult T cell leukemia virus by blood transfusion: seroconversion in recipients. Vox Sang. 1984;46:245-253.

(10) Yamamoto N, Okada M, Koyanagi Y, Kannagi M, Hinuma Y. Transformation of human Leukocytes by cocultivation with an adult T cell leukemia virus producer cell line. Science. 1982;217:737-739.

(11) Popovic M, Sarin PS, Robert-Gurroff M et al. Isolation and transmission of human retrovirus (human t-cell leukemia virus). Science. 1983;219:856-859.

(12) Dustin M. Viral spread through protoplasmic kiss. Nat Cell Biol. 2003;5:271-272.

(13) Igakura T, Stinchcombe JC, Goon PK et al. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science. 2003;299:1713-1716.

(14) Barnard AL, Igakura T, Tanaka Y, Taylor GP, Bangham CR. Engagement of specific T cell surface molecules regulates cytoskeletal polarization in HTLV-1-infected lymphocytes. Blood. 2005.

(15) Nejmeddine M, Barnard AL, Tanaka Y, Taylor GP, Bangham CR. Human T-lymphotropic Virus, Type 1, Tax Protein Triggers Microtubule Reorientation in the Virological Synapse. J Biol Chem. 2005;280:29653-29660.

(16) Pise-Masison CA, Jeong SJ, Brady JN. Human T cell leukemia virus type 1: the role of Tax in leukemogenesis. Arch Immunol Ther Exp (Warsz ). 2005;53:283-296.

128

(17) Manel N, Battini JL, Taylor N, Sitbon M. HTLV-1 tropism and envelope receptor. Oncogene. 2005;24:6016-6025.

(18) Bindhu M, Nair A, Lairmore MD. Role of accessory proteins of HTLV-1 in viral replication, T cell activation, and cellular gene expression. Front Biosci. 2005;10:17-37.

(19) Franchini G, Nicot C, Johnson JM. Seizing of T cells by human T-cell leukemia/lymphoma virus type 1. Adv Cancer Res. 2003;89:69-132.

(20) Nicot C, Harrod RL, Ciminale V, Franchini G. Human T-cell leukemia/lymphoma virus type 1 nonstructural genes and their functions. Oncogene. 2005;24:6026-6034.

(21) Nair A, Michael B, Hiraragi H et al. Human T lymphotropic virus type 1 accessory protein p12I modulates calcium-mediated cellular gene expression and enhances p300 expression in T lymphocytes. AIDS Res Hum Retroviruses. 2005;21:273-284.

(22) Hiraragi H, Michael B, Nair A et al. Human T-lymphotropic virus type 1 mitochondrion-localizing protein p13II sensitizes Jurkat T cells to Ras-mediated apoptosis. J Virol. 2005;79:9449-9457.

(23) D'Agostino DM, Silic-Benussi M, Hiraragi H, Lairmore MD, Ciminale V. The human T-cell leukemia virus type 1 p13(II) protein: effects on mitochondrial function and cell growth. Cell Death Differ. 2005.

(24) Dekaban GA, Peters AA, Mulloy JC et al. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology. 2000;274:86-93.

(25) Pique C, Uretavidal A, Gessain A et al. 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. 2000;191:567-572.

(26) Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

(27) Albrecht B, Collins ND, Burniston MT et al. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J Virol. 2000;74:9828-9835.

129

(28) Albrecht B, D'Souza CD, Ding W et al. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J Virol. 2002;76:3493-3501.

(29) Ding W, Albrecht B., Kelley RE et al. 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. 2002;76:10374-10382.

(30) Ding W, Kim SJ, Nair AM et al. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol. 2003;77:11027-11039.

(31) Dustin ML, Bivona TG, Philips MR. Membranes as messengers in T cell adhesion signaling. Nat Immunol. 2004;5:363-372.

(32) Hogg N, Henderson R, Leitinger B et al. Mechanisms contributing to the activity of integrins on leukocytes. Immunol Rev. 2002;186:164-171.

(33) van KY, Figdor CG. Activation and inactivation of adhesion molecules. Curr Top Microbiol Immunol. 1993;184:235-256.

(34) Hogg N, Laschinger M, Giles K, Mcdowall A. T-cell integrins: more than just sticking points. J Cell Sci. 2003;116:4695-4705.

(35) Hogg N, Smith A, Mcdowall A et al. How T cells use LFA-1 to attach and migrate. Immunol Lett. 2004;92:51-54.

(36) Carman CV, Springer TA. Integrin avidity regulation: are changes in affinity and conformation underemphasized? Curr Opin Cell Biol. 2003;15:547-556.

(37) Ardouin L, Bracke M, Mathiot A et al. Vav1 transduces TCR signals required for LFA-1 function and cell polarization at the immunological synapse. Eur J Immunol. 2003;33:790-797.

(38) Griffiths EK, Krawczyk C, Kong YY et al. Positive regulation of T cell activation and integrin adhesion by the adapter Fyb/Slap. Science. 2001;293:2260-2263.

(39) Wang H, Moon EY, Azouz A et al. SKAP-55 regulates integrin adhesion and formation of T cell-APC conjugates. Nat Immunol. 2003;4:366-374.

(40) Katagiri K, Hattori M, Minato N et al. Rap1 is a potent activation signal for leukocyte function-associated antigen 1 distinct from protein kinase C and phosphatidylinositol-3-OH kinase. Mol Cell Biol. 2000;20:1956-1969.

130

(41) Stewart MP, Mcdowall A, Hogg N. LFA-1-mediated adhesion is regulated by cytoskeletal restraint and by a Ca2+-dependent protease, calpain. J Cell Biol. 1998;140:699-707.

(42) Collins ND, D'Souza C, Albrecht B et al. 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. 1999;73:9642-9649.

(43) Stewart MP, Cabanas C, Hogg N. T cell adhesion to intercellular adhesion molecule-1 (ICAM-1) is controlled by cell spreading and the activation of integrin LFA-1. J Immunol. 1996;156:1810-1817.

(44) Robek MD, Wong FH, Ratner L. 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. 1998;72:4458-4462.

(45) Nicot C, Mulloy JC, Ferrari MG et al. HTLV-1 p12(I) protein enhances STAT5 activation and decreases the interleukin-2 requirement for proliferation of primary human peripheral blood mononuclear cells. Blood. 2001;98:823-829.

(46) Hummel KB, Bellini WJ, Offermann MK. Strain-specific differences in LFA-1 induction on measles virus-infected monocytes and adhesion and viral transmission to endothelial cells. J Virol. 1998;72:8403-8407.

(47) Fortin JF, Cantin R, Tremblay MJ. T cells expressing activated LFA-1 are more susceptible to infection with human immunodeficiency virus type 1 particles bearing host-encoded ICAM-1. J Virol. 1998;72:2105-2112.

(48) Arias RA, Munoz LD, Munoz-Fernandez MA. Transmission of HIV-1 infection between trophoblast placental cells and T-cells take place via an LFA-1-mediated cell to cell contact. Virology. 2003;307:266-277.

(49) Wattel E, Vartanian JP, Pannetier C, Wainhobson S. Clonal expansion of human T-Cell leukemia virus type I- infected cells in asymptomatic and symptomatic carriers without malignancy. J Virol. 1995;69:2863-2868.

(50) Cavrois M, Leclercq I, Gout O et al. 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. 1998;17:77-82.

(51) Mortreux F, Kazanji M, Gabet AS, de Thoisy B, Wattel E. Two-step nature of human T-cell leukemia virus type 1 replication in experimentally infected squirrel monkeys (Saimiri sciureus). J Virol. 2001;75:1083-1089.

131

(52) Mosley AJ, Asquith B, Bangham CR. Cell-mediated immune response to human T-lymphotropic virus type I. Viral Immunol. 2005;18:293-305.

(53) Constantin G, Majeed M, Giagulli C et al. Chemokines trigger immediate beta2 integrin affinity and mobility changes: differential regulation and roles in lymphocyte arrest under flow. Immunity. 2000;13:759-769.

(54) Chan JR, Hyduk SJ, Cybulsky MI. Chemoattractants induce a rapid and transient upregulation of monocyte alpha4 integrin affinity for vascular cell adhesion molecule 1 which mediates arrest: an early step in the process of emigration. J Exp Med. 2001;193:1149-1158.

(55) Feigelson SW, Grabovsky V, Winter E et al. The Src kinase p56(lck) up-regulates VLA-4 integrin affinity. Implications for rapid spontaneous and chemokine-triggered T cell adhesion to VCAM-1 and fibronectin. J Biol Chem. 2001;276:13891-13901.

(56) Michael B, Nair AM, Hiraragi H et al. Human T lymphotropic virus type-1 p30II alters cellular gene expression to selectively enhance signaling pathways that activate T lymphocytes. Retrovirology. 2004;1:39.

132

Figure 3.1. LFA-1 mediated adhesion is reduced in ACH.p12I cell lines.

(A) p19 antigen production from ACH.2, ACH.p12.4, ACH.2p and ACH.p12.4p cell

lines was measured using ELISA. Supernatants were collected 24 days post-infection.

Values are the means (± SEM) of triplicate samples and represent two independent

experiments. (B) Cells were treated with TG (5 µM), PMA(50 ng/ml) or

Mg2+/EGTA(5 mM/1 mM) and incubated on ICAM-1Fc-coated 96-well plates for

30 min at 37° C. The proportion of adhesion was measured as a percentage of total cells

added per well. Values are the means (± SEM) of triplicate sample and represent four

independent experiments. Significantly reduced adhesion was observed in TG, PMA

and Mg2+/EGTA treated ACH.p12.4 and ACH.p12.4p cell lines (Wilcoxon- Mann

Whitney test, P < 0.05).

133

Figure 3.2. ACH.p12I cells have decreased binding of sICAM-1 without alteration

of LFA-1 expression.

(A) Flow cytometry specific for LFA-1 molecules CD11a and CD18 was performed on

ACH.2, ACH.p12.4, ACH.2p and ACH.p12.4p (106cells per sample). (B) Soluble

ICAM-1Fc binding assay was performed on ACH.2, ACH.p12.4, ACH.2p and

ACH.p12.4p cell lines. Cells were treated with TG (5 uM), PMA(50 ng/ml) or

Mg2+/EGTA(5 mM/1 mM) and incubated with sICAM-1Fc for 30 min at 37° C. Bound

sICAM-1 was detected with R-PE-conjugated goat anti-human IgG Fc-specific

antibody and analyzed by flow cytometry. Data are expressed as mean fluorescence

intensity. Figure is representative of two experiments.

134

Figure 3.3. p12I stable expression in Jurkat T-cells did not alter LFA-1 affinity.

(A) Flow cytometric analysis for both GFP expression and sICAM-1 binding was

performed in empty control and p12I-expressing Jurkat T-cells. Both samples were

more than 85% GFP positive indicating high transduction efficiency. Soluble ICAM-

1Fc binding was measured concurrently. Cells were treated with TG (5 µM), PMA(50

ng/ml) or Mg2+/EGTA(5 mM/1 mM) and incubated with sICAM-1Fc for 30 min at 37°

C. Bound sICAM-1 was detected with R-PE-conjugated goat anti-human IgG Fc-

specific antibody and analyzed by flow cytometry. (B) RT-PCR was performed using

total cellular RNA isolated from mock, empty control and p12I-expressing Jurkat T-

cells at 30 and 60 days post-infection with lentiviral vectors.

135

Figure 3.4. Expression of p12I in Jurkat T-cells induced LFA-1-mediated cell

adhesion. Adhesion to immobilized ICAM-1 was performed using empty control and

p12I expressing Jurkat T-cells. (A) Cells were treated with TG (5 µM), Ionomycin (0.5

µM) PMA (50 ng/ml) or Mg2+/EGTA (5 mM/1 mM) and incubated on ICAM-1Fc-

coated for 30 min at 37° C. The adhesion was measured as a percentage of total cells

added per well. Values are the means (± SEM) of triplicate samples and represent three

independent experiments. P < 0.05 was by Wilann-Whitney test, (B-E) Cells were

treated with increasing concentrations of (B) TG, (C) Ionomycin, (D) PMA, and (E)

Mg2+ (with constant 1 mM EGTA concentration) and incubated for the adhesion assay.

The fold increase was calculated by dividing the percentages of adhered cells in the

presence of each treatment with that of untreated cell adhesion. Figure is representative

of two experiments.

136

Figure 3.5. HTLV-1 p12I-mediated LFA-1 activation is inhibited by calcium signal

inhibitors. Empty control and p12I-expressing Jurkat T-cells were pretreated with

BAPTA-AM (50 µM), SK&F 96365 (100 µM) or calpeptin (100 µg/ml) for 30 min at

37° C and adhesion assay was performed in presence or absence of Thapsigargin (5

µM), PMA (50 ng/ml) or Mg2+ /EGTA (5 mM/1 mM) stimulation.

137

Figure 3.6. Expression of p12I in Jurkat T-cells modulated surface distribution of

LFA-1 on the cell membrane. (A) Untransduced mock Jurkat T cells or (B)

transduced empty control and p12I expressing Jurkat T cells were stained with R-PE

conjugated LFA-1 mAb and analyzed by confocal microscopy. Fluorescent signal

exceeding 50 unit was expressed as bright blue color. Signal of LFA-1 was quantified

on the membrane of 30 cells. Mean fluorescence units (± SEM) were compared in right

bar graphs (A) between TG untreated and treated groups, or (B) between empty control

and p12I expressing Jurkat T cells. Significant (t test P < 0.01) increased fluorescence in

TG treated and p12I expressing Jurkat T cells was observed. C) Flow cytometric

analysis for LFA-1 cell surface expression in GFP expressing empty control and in

empty control and p12I expressing Jurkat T cells was performed.

138

CHAPTER 4

EXPRESSION OF HTLV-1 PX OPEN READING FRAME I ENHANCES

EARLY VIRAL INFECTIVITY DURING CELL-TO-CELL

TRANSMISSION IN T-LYMPHOCYTES

4.1 INTRODUCTION

Human T-cell lymphotropic virus type 1 (HTLV-1), which infects 15 to 25

million people worldwide 1, is the etiological agent of adult T cell leukemia/lymphoma

(ATLL), and is highly associated with HTLV-1-associated myelopathy/tropical spastic

paraparesis (HAM/TSP) and a variety immune-mediated disorders 2-4. In addition to

retroviral common structural and enzymatic proteins (Gag, Pol, and Env), HTLV-1

genome encodes, regulatory (Tax and Rex) and nonstructural accessory proteins (p12I,

p27I, p30II and p13II) 5-9, which are generated by alternative splicing and internal

initiation, from four open reading frames (ORF) located in pX region.

139

Both p27I (doubly spliced) and p121 (singly spliced) are produced from pX ORF

I. p12I can also be translated from the p27I message via initiation at an internal

methionine codon. It is thought that p12I is preferentially expressed from the p27I

message 10. Importantly, the expression of pX ORF I proteins during natural infection is

evidenced by detection of its mRNA in infected cells and elicitation of humoral and

cytotoxic T cell responses against pX ORF I-encoded proteins in infected carriers and

patients 11,12. Moreover, in chronically HTLV-1 infected cell lines, the expression of pX

ORF I mRNA was measured at 100 to 1,000-fold less than pX ORF III/IV mRNA by

real time reverse transcription-polymerase chain reaction (RT-PCR), suggesting that

expression of HTLV-1 pX genes are differentially regulated 13. However, regulation of

pX mRNA expression was not tested during early phase of HTLV-1 infection. The

expression of mRNA can be modulated by specific stimuli resulted from cell-to-cell

contact between HTLV-1 infected cells and target cells. Furthermore, events during de

novo HTLV-1 infection in target cells may affect expression pattern of pX gene

products.

HTLV-1 infection is naturally acquired following contact between target cells

and virus infected cells contained in breast milk, semen and blood 14,15. CD4 + T cells

are the major target of HTLV-1 infection, immortalization and transformation 16,17.

Mechanisms of HTLV-1 mediated immortalization and transformation have been

extensively investigated, however early events of HTLV-1 transmission and infection

are poorly understood. Activated T cells are more susceptible to HTLV-1 infection and

transformation compared to non-stimulated or quiescent/resting T cells 18, suggesting

140

that T cell activation is required for the virus to efficiently establish infection. HTLV-1

infection can further activate uninfected, resting T cells in an antigen-nonspecific

manner or stimulate antiviral T cells in an antigen-specific manner. Non-specific

proliferation and stimulation of uninfected PBMC can be induced by direct contact with

irradiated or fixed HTLV-1 infected cells through the combined interaction of adhesion

molecules CD2/LFA-3 (CD58), LFA-1/ICAM-1 (CD54), and IL-2/IL-2R autocrine

activity 4,19,20. CD2/CD58 interaction is particularly important for the activation of

uninfected T cells by HTLV-I-infected T cells 11.

Studies from our laboratory have implicated HTLV-1 pX ORF I expression in

HTLV-1 infection in a rabbit model and in T cell activation 21,22. Selective ablation of

pX ORF I mRNA expression from a HTLV-1 proviral clone (ACH.p12) resulted in

dramatic reduction of viral infectivity compared to wild type (ACH) in a rabbit model 23.

Furthermore, when cocultured with naïve, quiescent peripheral blood mononuclear cells

(PBMC), ACH.p12I cells exhibited reduced viral infectivity 24. However, in the

presence of the T cell stimulants, phytohemagglutinins (PHA) and IL-2, ACH.p12I

cells were restored in their ability to infect target T cells 24. These reports indicate that

pX ORF I expression is required for efficient HTLV-1 infection in PBMC and

suggesting a role for p12I in target T cell activation.

Ectopic expression of p12I causes calcium mediated T cell activation 25-27. The

small hydrophobic protein accumulates in the endoplasmic reticulum (ER) 28, where it

associates with calreticulin, calnexin and calcineurin 29, and leading to the release of

calcium from ER stores 26. As a result of HTLV-1 p12I expression, calcium-dependent

141

signaling causes activation and nuclear translocation of the nuclear factor of activated T

cells (NFAT) 25, and subsequent up-regulation of IL-2 production 27 and p300

expression 30. In addition, p12I expression leads to enhanced LFA-1 clustering on Jurkat

T cells, implying a role for pX ORF I in cell-to-cell transmission of HTLV-1 31.

Cell-to-cell HTLV-1 transmission occurs directly through virological synapse by

inducing polarization of microtubule-organizing centers (MTOC) to cell contact

junction 32-35. Engagement of adhesion molecules, ICAM-1, LFA-1, and IL-2 receptor

(CD25) triggered strong polarization of the MTOC, suggesting their critical role in

HTLV-1 transmission. HTLV-1 Tax is suggested as a viral factor contributing viral

cell-to-cell transmission, since Tax expression triggered MTOC polarization directly

and may synergize by up-regulating expression of ICAM-1 32 36.

In this study, we investigated the role of expression of HTLV-1 pX ORF I in

early HTLV-1 transmission by comparing the activation status between wild type

HTLV-1 infected cells (ACH) and T cells lacking pX ORF I expression (ACH.p12). We

then tested if the expression of ORF I is required to induce bystander activation of

uninfected target cells by performing both proliferation assays and flow cytometric

analysis of target cells We found that abrogation of Px ORF I message decreased CD69

expression in HTLV-1 infected cells, but did not affect the HTLV-1- induced bystander

target cell activation.

Finally, we performed real-time RT-PCR to detect pX ORF I and pX ORF III/IV

mRNA after coculture of both wild type ACH and ACH.p12 cells with pre-stimulated

142

or non-stimulated T cells. By measuring Tax/Rex (pX ORF III/IV) mRNA in target

cells, we were able to compare viral infectivity of wild type ACH to ACH.p12

immortalized T cells and viral mRNA expression in newly infected target cells. Our

data indicated that expression of pX ORF I is required for efficient HTLV-1

transmission to target cells and defined for the first time Tax/Rex and pX ORF I mRNA

expression in early cell-to-cell transmission.

4.2 MATERIALS AND METHODS

Cells

HTLV-1 infected cell lines, ACH.2, ACH.2p, ACH.p12.4 and ACH.p12.4p have

been previously described 23,31. ACH.2 and ACH.p12.4 cells were generated following

transfection of molecular clones of HTLV-1 (ACH and ACH.p12I) 23,37. ACH.p12I

plasmid contains a mutation in the splice acceptor of the third exon of the pX ORF I

DNA, resulting in complete ablation of pX ORF I expression without affecting the

expression of other viral genes 23,37. ACH.2p and ACH.p12.4p cells were created from

ACH.2 and ACH.p12.4 cells by coculturing PBMC isolated from same dono 31. ACH.2,

ACH.p12.4, ACH.2p and ACH.p12.4p cells were cultured in RPMI 1640 supplemented

with 15% fetal bovine serum, L-glutamine (0.3 mg/ml), penicillin (100 U/ml),

streptomycin (100 µg/ml), and recombinant IL-2 (20 U/ml). HTLV-1 infected human T

cell line, MT-2 cells were grown in RPMI 1640 supplemented with 10% fetal bovine

serum, L-glutamine (0.3 mg/ml), penicillin (100 U/ml), streptomycin (100 µg/ml).

Normal uninfected human PBMC were obtained by leukophoresis and maintained as

143

previously described 24. For coculture experiment, target PBMC were either

unstimulated or pre-activated for four days with human IL-2 (20 U/ml) and

phytohemagglutinin (PHA) (2 µg/ml).

Coculture of HTLV-1 infected cells and target PBMC

HTLV-1 infected effecter cells (ACH.2, ACH.2p, ACH.p12.4 and ACH.p12.4p)

were cocultivated as live or fixed cells with freshly isolated non-stimulated or pre-

activated PBMC. For fixation, cells were incubated 1% paraformaldehye (PFA) for 30

min and washed twice with media before coculture. Permeable transwell inserts with

diameter of 6.5 mm and pore size of 0.4 µm (Corning Inc.) were used to separately

culture live effecter cells from target PBMC and to test the role of soluble factors on

cell activation. Live, fixed or separated effecter cells were cocultured with non-

stimulated or activated PBMC at 1:2 or 1:1 ratio at 2 million cells/ml prior to

proliferation assay, flow cytometric analysis or RNA extraction for real time RT-PCR.

Proliferation assay

Tetrazolium dye (MTS) proliferation assay (CellTiter 96 Cell Proliferation

Assay, Promega) was used according to the manufacturer's instructions. To measure

proliferation of target PBMC cocultured with fixed HTLV-1 infected effecter cells, the

MTS assay was performed at the time of coculture (time 0) and at 3 days after coculture.

Data was expressed as fold proliferation of target cells, which was measured by optical

density in triplicate samples.

144

Flow cytometric analysis and cell sorting.

Flow cytometric analysis was performed as described 38 to measure expression

of cell surface markers. Red-phycoerythrin (R-PE) conjugated antibodies recognizes

CD25 (Southern Biotechnology), CD58 (Southern Biotechnology), and CD11a

(Southern Biotechnology) and FITC conjugated antibodies recognizes CD69 (BD

Bioscience), CD2 (Southern Biotechnology), and CD54 (Southern Biotechnology)

antibodies were used for cell surface staining according to the manufacturer's

recommendations. Fluorescence of cells was measured by FACScan (Becton

Dickinson) analysis.

For evaluating bystander target PBMC activation, target PBMC were pre-

labeled with 0.5 µM succinimidyl ester of carboxyfluorescein diacetate (CFDA-SE)

(Invitrogen) according to the manufacturer's instructions to differentiate from unlabeled

HTLV-1 infected effecter cells during flow cytometric analysis. After 24 and 36 h

coculture, cells were stained with R-PE conjugated antibodies recognizes CD25

(Southern Biotechnology) and CD69 (BD Bioscience). CD25 and CD69 percentage

positive were measured in the target PBMC (CMFD-SE positive cells).

For further real time RT-PCR analysis, cell sorting using a FACSAria flow

cytometer (Becton Dickinson) was performed to selectively collect target PBMC from

HTLV-1 infected cells (ACH.2p, and ACH.p12.4p) at 1 and 3 days after coculture.

Before cell sorting, we pre-labeled HTLV-1 infected cells with 0.5 µM 5-

chloromethylfluorescein diacetate (CMFDA) (Invitrogen) according to the

manufacturer's recommendation.

145

Real time RT-PCR

Total RNA was extracted from cells using RNAqueous Kit (Ambion) and

complementary DNA (cDNA) was generated from 1 µg of RNA using a Reverse

Transcription system kit (Promega) as described by manufacturer. Taqman real time

PCR was performed using a Prism 7000 sequence detector system (Applied

Biosystems). Primers and probes used for detection of HTLV-1 pX ORF III/IV and pX

ORF I mRNAs were previously described 13. Amplication of Tax/Rex cDNA were

performed using forward splice site-specific primer (ex2/6950) 5’-

ACCAACACCATGGCCCA-3’ and reverse specific primer (3’7170) 5’-

GAGTCGAGGGATAAGGAAC-3’. Taqman probe for Tax/Rex cDNA was (TMP-2)

5’FAM-ATC ACC TGG GAC CCC ATC- TAMRA-3’. For amplication of pX ORF I

cDNA (p12I and p27I messages), same reverse primer (3’6552) 5’-

GGAGAAAGCAGGAAGAGC-3’ and Taqman probe (TMP-1), 5’FAM-

TTCGCCTTCTCAGCCCCTTGTCT-TAMRA-3’ were used. The p27I forward splice

site-specific primer (ex2/6383) 5’-ACCAACACCATGGCAACT-3’ and p12I forward

specific primer (ex1/6383) 5’-GTCCGCCGTCTAGCAAC- 3’ were used. For

normalization purpose, we measured human GAPDH copy number at the same cDNA

sample. Forward primer 5’- CATCAATGACCCCTTCATTGAC- 3’, reverse primer 5’-

CGCCCCACTTGATTTTGGA-3’, and Taqman probe 5’FAM-

TGGCAAATTCCATGGCACCGTC-TAMRA-3’ were used to detect human GAPDH

cDNA. Standard curves for each real time PCR reaction was generated from cDNA

plasmids: SE356 for Tax/Rex cDNA (P. Green, Ohio State University); pMS0-1.8 for

146

p27 I (D. Derse, National Cancer Institute, Frederick, MD); pMT2-2.3 for p12I (D.

Derse, National Cancer Institute) and pBluescript hGAPDH (K. Boris-Lawrie, Ohio

State University). These cDNAs were diluted from 1 X 101 to 1 X 107 copies to create

standard curves. Twenty five µl volumes containing 12.5 µl of universal PCR master

mix (Applied Biosystems), 200 nM of specific probe, 600nM of each specific primers

and 2.5 µl of cDNA template were used for each reaction. Each assays were performed

in 96-well optical plates (Applied Biosystems) containing duplicate wells for standard,

non-template control and samples. Real time PCR reaction was 50 ° C for 2 min, 95 ° C

for 10 min for activation of DNA polymerase, and 40 cycles of 15 s at 94.5 ° C and 1

min at 60 ° C. The threshold cycle (Ct) values were used to plot standard curves. The

correlation values of standard curves for each assay were always more than 97 %. The

copy numbers of each cDNA samples were calculated from their measured Ct value

base on standard curves and normalized to 1 x 106 copies of human GAPDH.

4.3 RESULTS

Abrogation of pX ORF I expression influences activation status of HTLV-1

infected cells

To determine if the expression of HTLV-1 pX ORF I could activate HTLV-1

infected cells, we performed flow cytometric analysis of T-cell activation markers (CD

25/IL-2 receptor α and CD69) on wild type HTLV-1 infected cells (ACH cell line;

ACH.2 and ACH.2p) and T cells lacking pX ORF I expression (ACH.p12 cell line;

147

ACH.p12.4 and ACH.p12.4p). ACH.2p and ACH.p12.4p cells were newly generated

from PBMC donated from same donor by cocultivating with ACH.2 and ACH.p12.4

cells respectively. Expression of IL-2 receptor α (CD25) was approximately 99% in

both wild type ACH cells and ACH.p12 cells (Fig. 4.1A) and was not significantly

different between each cell line. On the other hand, the expression of CD69 was

consistently higher in wild type ACH cells compare to ACH.p12 cells. The enhanced

CD69 expression was evident comparing expression between ACH.2 (68.6 % positive)

and ACH.p12.4 (42.4 %) and between newly generated cell lines ACH.2p (70.9 %) and

ACH.p12.4p (49.6 %), These data suggest that abrogation of pX ORF I mRNA

expression may influence CD69 cell surface expression. These data are also consistent

with the known calcium regulatory property of pX ORF I encoded p12I, the known

responsiveness of CD69 expression to calcium signaling 39, and the activation status of

HTLV-1 infected cells.

Abrogation of pX ORF I expression in HTLV-1 infected cells did not affect

bystander proliferation of uninfected target PBMC.

Because bystander activation of uninfected target cells can be induced by

interaction between CD2/CD58, and LFA-1/ICAM-1 interaction 19,20, we measured

expression of CD2, CD58 (LFA-3), CD11a (LFA-1), and CD54 (ICAM-1) in ACH and

ACH.p12 cell lines to test whether the deletion of pX ORF I protein expression could

modulate cell surface expression of these adhesion molecules. Approximately 99 % of

all cell types (ACH.2, ACH.2p, ACH.p12.4 and ACH.p12.4p) were positive for CD2,

CD58, CD11a and CD54 (Fig. 4.1B and C). These results indicate that abrogation of

148

p12I did not affect cell surface expression of adhesion molecules, which are critical for

the cell contact-dependent bystander target cell activation.

We then tested whether the expression of pX ORF I in HTLV-1 infected cells

could induce proliferation of bystander uninfected target cells. As previous studies

indicated 19,20, fixed or irradiated HTLV-1 infected cells were able to induce non-

specific activation of uninfected target cells when they were cocultured. We measured

proliferation of target quiescent PBMC by cocultivating fixed effecter cells (ACH.2,

ACH.2p, ACH.p12.4 and ACH.p12.4p) using MTS colorimetric assays in the presence

or absence of IL-2 and PHA stimulation (Fig. 4.2). As expected, fixed HTLV-1 infected

cells induced significant proliferation of target PBMC after 3 days of coculture in the

absence of stimulation. However, the fold proliferation of target PBMC induced by wild

type ACH and ACH.p12 cells was not significantly different, indicating abrogation of

p12I in HTLV-1 infected cells did not affect the HTLV-1–mediated non-specific

bystander target cell proliferation.

Abrogation of pX ORF I expression in HTLV-1 infected cells did not affect

bystander activation of target PBMC.

To further test if pX ORF I expression could influence the activation of target

PBMC, we measured expression of activation markers (CD25 and CD69) on target

PBMC by flow cytometric analysis. HTLV-1 infected effecter cells (ACH.2, ACH.2p,

ACH.p12.4 and ACH.p12.4p) were cocultivated as live or fixed cells with freshly

prepared quiescent PBMC. Live effecter cells were also cocultured separately in

permeable transwell inserts to test the role of soluble factors in cell activation. Target

149

PBMC were pre-labeled with CFDA-SE to differentiate target PBMC from HTLV-1

infected effecter cells.

By flow cytometric analysis, we were able to differentiate CFDA-SE labeled

target PBMC from unlabeled HTLV-1 infected effecter cells (Fig. 4.3 A). CD25 and

CD69 percentage positive were measured in the target PBMC (CMFD-SE positive

cells) after 24 h (Fig. 4.3 B-D) or 36 h (Fig. 4.3 E-G) coculture. As expected, when

PBMC were stimulated with IL-2 and PHA, the expression of activation marker CD69

and CD25 was up-regulated at 24 h and 36 h after stimulation, with CD69 earlier

induction than CD25 (Fig. 4.3 B-G). The expression of CD69 and CD25 in target

PBMC, which were cocultured with live or fixed effecter cells (ACH.2, ACH.2p,

ACH.p12.4 and ACH.p12.4p), was significantly increased when compared that of

unstimulated PBMC control (Fig.4.3 B, C, E and F). However, effecter cells cultured in

transwell insert did not up-regulate CD69 and CD25 expression in target PBMC, which

was not significantly greater than that of unstimulated PBMC control (Fig.4.3 D and G).

These results are consistent with previous studies 19,20 that indicated that non-specific

stimulation of uninfected PBMC was induced by direct contact, but not by soluble

factors from infected cells.

However, the induction ability of activation markers up-regulation on target

PBMC was not significantly different between wild type ACH (ACH.2 and ACH.2p)

and ACH.p12 cell lines (ACH.p12.4 and ACH.p12.4p) (Fig.4.3 B-G). As consistent

with proliferation assay result, the data indicated abrogation of p12I in HTLV-1 infected

cells did not affect the non-specific target cell activation. These results suggest that

150

bystander target cell activation can be sufficiently induced without expression of pX

ORF I by other mechanism, such as Tax- mediated T cell activation 40.

HTLV-1 pX ORF I and pX ORF III/IV mRNAs were differentially expressed

during the early HTLV-1 transmission.

To examine role of pX ORF I messages during early HTLV-1 infection and

transmission, we carried out real time-RT-PCR. Before coculture assay, we measured

relative mRNA copy number (per one million human GAPDH mRNA) of Tax/Rex,

p27I, and p12I messages in HTLV-1 infected cells (MT-2, ACH.2p, and ACH.p12.4p),

as well as in uninfected PBMC. As consistent with previous report 13, Tax/Rex mRNA

was the most abundant, and p27I and p12I mRNA were detected at 1000-2000 fold less

than Tax/Rex mRNA in MT-2 cells (Fig. 4.4 A). In wild type ACH cell line, ACH.2p

cells, p27I and p12I mRNAs were detected approximately at 400-1000 fold less than

Tax/Rex mRNA (Fig. 4.4 A). As expected, both p27I and p12I mRNA were not detected

in ACH.p12 cells (Fig. 4.4 A), which are lacking pX ORF I mRNA expression due to an

introduction of a mutation in splice accepter site of the third exon 23,41.

We then profiled pX ORF I mRNA (p12I and p27I messages) and pX ORF

III/IV mRNA (Tax/Rex message) from 3 hours to 24 hours after cocultivating HTLV-1

infected cells (ACH.2p or ACH.p12.4p) with unstimulated quiescent or activated target

PBMC. Tax/Rex, p12I and p27I mRNAs from unsorted coculture samples of HTLV-1

infected and target cells were differentially expressed through out coculture (Fig. 4.4 B-

D). Up-regulation of Tax/Rex mRNA was induced from 3 hours and peaked 6 hours

151

after coculture when wild type ACH.2p cells were coculture with quiescent PBMC.

This up-regulation of Tax/Rex message was greater and faster than that observed in

coculture of ACH.p12.4p cells and quiescent PBMC (Fig. 4.4 B). Interestingly, when

cocultured with activated PBMC, expression of Tax/Rex messages were not highly up-

regulated compared to cocultured with quiescent PBMC. These results indicated that

expression pattern of Tax/Rex mRNA during the HTLV-1 early transmission is affected

by the pX ORF I proteins (p12I and p27I) expression and by activation status of target

cells. Also, these results suggest that stimuli induced by pX ORF I proteins or target

cells may modulate Tax/Rex mRNA expression during HTLV-1 transmission.

Furthermore, p27I and p12I mRNA were also differentially expressed during

early HTLV-1 transmission (Fig. 4.4 C and D). As expected, p12I and p27I messages

were not detected throughout the all time points, when cocultured with pX ORF I

expression abrogated ACH.p12.4p cells. The peak expression of both p27I and p12I

mRNA was detected at 12 hours after coculture. Expression of p12I mRNA was not

affected by activation status of target PBMC (Fig. 4.4 D). These results indicate that

expression of pX ORF I mRNA expressed early in cell-to-cell transmission of HTLV-1

and is tightly regulated during HTLV-1 transmission, suggesting a role of pX ORF I

proteins in HTLV-1 early infection.

152

Expression of pX ORF I proteins increased HTLV-1 viral transmission to target

PBMC.

To examine de novo expression of pX ORF I and pX ORF III/IV mRNA in

target PBMC, we selectively collected target PBMC by sorting out HTLV-1 infected

cells (ACH.2p, and ACH.p12.4p) at 1 and 3 days after coculture. Before coculture, we

pre-labeled HTLV-1 infected cells with CMFDA, which freely diffuses into cells and

reacts with intracellular thiols and generates green florescent by being cleaved by

cytoplasmic easterase. This long term cell tracer is retained in live cells through several

generations, without being transferred to adjacent cells 42. As expected, we could

differentiate CMFDA unlabeled target PBMC from stained ACH.2p and ACH.p12.4p

cells by flow cytometric analysis at 1 or 3 day after coculture (Fig. 4.5 A). Collected

target PBMC were consistently more than 95 % CMFDA negative (Fig. 4.5 B) and

labeled HTLV-1 cells were more than 90 % CMFDA positive (Fig. 4.5 C) after cell

sorting.

To test role of pX ORF I protein expression in HTLV-1 viral infectivity on

target PBMC, we then measured Tax/Rex mRNA both in CMFDA positive HTLV-1

producer cells and CMFDA negative target PBMC after cell sorting of 1 and 3 day

coculture. Because expression of Tax/Rex mRNA was significantly correlated with

HTLV-1 proviral DNA load 43, viral infectivity was evaluated by measuring percent of

Tax/Rex mRNA expressed in target PBMC at the point of cell sorting (Fig. 4.6). As

expected, a longer 3 day-coculture increased viral infectivity of both ACH.2p and

ACH.p12.4p cells. At 1 day-coculture, viral infectivity of ACH.p12.4p cells to the

153

quiescent PBMC was significantly decreased compared to that of wild type ACH.2p

cells. However, to the activated PBMC, viral infectivity between wild type ACH.2p and

ACH.p12.4p cells was not significantly different. Whereas, after 3 days coculture, viral

infectivity of wild type ACH.2p cells was consistently greater than that of pX ORF I

expression abrogated ACH.p12.4p cells both in quiescent and activated PBMC (Fig.

4.6). These results were consistent with previous in vitro infectivity data measured

HTLV-1 p19 antigen, suggesting the expression of pX ORF I mRNA is required for

efficient viral transmission to target cells.

Expression of pX ORF I proteins enhanced early establishment HTLV-1 de novo

infection in target PBMC.

To profile expression of Tax/Rex, p27I and p12I mRNA in newly infected target

PBMC, we selectively cultured target PBMC (CMFDA negative), which were sorted

from the 3 day-coculture cells and carried out real time RT-PCR. The relative copy

number of Tax/Rex mRNA was measured in quiescent or activated target PBMC, which

cocultivated with wild type ACH.2p or ACH.p12.4p cells (Fig. 4.7 A). The fold

increase of Tax/Rex mRNA expression was measured on day 5 (2 days after sorting)

and on day 7 (4 day after sorting) (Fig. 4.7 B). From day 3 to day 5, the fold increase of

Tax/Rex mRNA expression was greater in activated PBMC (8 to 10 fold) than in

quiescent PBMC (1 to 3 fold), and the expression Tax/Rex mRNA was not significantly

increased from day 5 to day 7, suggesting establishment of HTLV-1 infection in target

PBMC occurs in 5 days after coculture. Furthermore, the fold increase of Tax/Rex

mRNA expression was consistently higher in target PBMC when they were cocultivated

154

with wild type ACH.2p cells than when they were cultured with ACH.p12.4p cells (Fig.

4.7 B). These results indicate that expression of pX ORF I protein in newly infected

cells facilitated establishment of HTLV-1 infection.

Furthermore, we profiled pX ORF I mRNA expression in sorted target PBMC to

examine how expression of pX ORF I mRNAs are regulated in newly infected cells.

The expression of p27I and p12I mRNA was consistently detected in activated target

PBMC that were cocultivated with wild type ACH.2p cells (Fig. 4.7 C). Relative copy

number of pX ORF I mRNA was consistently increased after cell sorting. However,

ratio to Tax/Rex mRNA was decreased on day 5 and was slightly increased on day 7

(Fig. 4.7 D), indicating that expression of pX ORF I mRNAs were differentially

regulated during early establishment of HTLV-1 infection.

4.4 DISCUSSION

In this study, we demonstrated a role of expression of HTLV-1 pX ORF I

during early HTLV-1 transmission and infection. The abrogation of pX ORF I message

influenced the activation status of HTLV-1 infected cells, but did not effect HTLV-1-

induced bystander target cell activation. Our data indicates that the infectivity of T cell

lines expressing the full length infectious clone of HTLV-1 (wild type ACH) was

greater than that of ACH.p12 immortalized T cells as measured by Tax/Rex mRNA in

target cells. We also demonstrated that pX ORF I is an early transcript expressed cell-

to-cell transmission of HTLV-1. Our results were consistent with previous studies that

155

implicated the expression of pX ORF I plays a role in activation of HTLV-1 infected T

cell and is required for efficient HTLV-1 cell-to-cell transmission.

Expression of IL-2 receptor α (CD25) was consistently enhanced in both wild

type ACH cells and ACH.p12 cells lacking pX ORF I expression, indicating CD25

expression is not influenced by abrogation of pX ORF I expression. However, the early

activation marker CD69 expression was consistently higher in wild type ACH cells.

Interestingly, expression of CD69 is dependent on calcium mediated signaling 39.

Previous results from our laboratory indicated that pX ORF I encoded p12I is a

regulatory of calcium signaling leading the downstream gene expression including

NFAT activation 25-27. Our data links these two observations and offers a mechanism for

enhancement of CD69 expression on HTLV-1 infected cells. CD69 expression was up-

regulated in HTLV infected cells from HAM patients or asymptomatic carriers after 2

days ex vivo culture, but not was not significantly increased when they were freshly

isolated compared to uninfected control cells 44, indicating stimulation was generated

during ex vivo culture. Although CD69 is a well established and an earliest activation

marker of lymphocytes, the precise role in immunity is still elusive due to lack of

known ligands and a suitable in vivo model 45. Recent findings suggested that CD69

acts as immunoregulatory function by producing transforming growth factor-β (TGF-β)

45,46, which down-regulates antigen-presenting cell (APC) function 47, suggesting a

possible role of CD69 expression on HTLV-1 infected cell in escaping immune

surveillance. However, our previous rabbit model results do not support this hypothesis

because ACH.p12 cells did not elicit strong immune response compare to wild type

ACH cells, which have higher CD69 expression, when they were inoculated into rabbits

156

23. Further studies will be required to understand the role of CD69 in HTLV-1 cell-to-

cell transmission and the role of viral gene products in its expression.

Activation of target PBMC is advantageous for retroviral integration, subsequent

viral replication and productive infection. Bystander non-specific activation of PBMC

can be induced by direct contact with HTLV-1 infected cells via interaction of adhesion

molecule such as CD2 / LFA-3 (CD58) and LFA-1(CD11a) / ICAM (CD54)19,20.

Furthermore, signaling via CD2 and ICAM-1 receptor enhanced HTLV-1 replication

48,49, suggesting that the receptor- mediated signaling generated during HTLV-1 cell-to-

cell transmission is important for early viral infection. Because both expressions of

LFA-3 (CD58) and ICAM-1 (CD54) were constitutively enhanced in HTLV-1 infected

cells 50,51, target cells may be activated via CD2 or LFA-1 mediated signaling.

Although the cell surface expression of CD2, LFA-3 (CD58), LFA-1 (CD11a),

ICAM (CD54), and IL-2 receptor (CD25) was not significantly different between wild

type ACH and ACH.p12 cells in this study, LFA-1-mediated adhesion due to LFA-1

clustering in ACH cells was significantly enhanced compared to ACH.p12 cells in

previous study 31. Together our studies indicate that pX ORF I expression influences

LFA-1 clustering on infected cells to enhance cell-to-cell adhesion and stimuli derived

from cell contact with wild ACH cells may be greater than that from ACH.p12 cell

contact. We have previously demonstrated that ACH cells have increased IL-2

production in presence of T cell stimulation compared to ACH.2 cells 27. However,

bystander activation of target PBMC induced by these two cell types was not

significantly different, indicating that target cell activation can be sufficiently induced

in the absence of pX ORF I mRNA. Interestingly, HTLV-1 Tax induced transcriptional

157

activation ICAM-1 gene 40, suggesting that the ICAM-1 expression may be up-regulated

by Tax in HTLV-1 infected cells and may resulting in sufficient activation of target

PBMC by cell-to-cell contact. Furthermore, antibodies against LFA-1 and ICAM-1

triggered re-polarization of MTOC, which is characteristic of HTLV-1-induced

virological synapses 33, suggesting that pX ORF I protein may synergize to MTOC

polarization and contribute to HTLV-1 cell-to-cell transmission.

Accumulation of Gag and Env proteins was observed in cell contact junction

(virological synapse) with in 40 min after coculture of HTLV-1 infected cells with

target cells, and the transfer of Gag protein in target cells was detected within 2 hours 35.

Furthermore, newly synthesized unintegrated HTLV-1 DNA (de novo reverse

transcription) was detected within 4-8 hours after coculture 52, indicating that HTLV-1

cell-to-cell transmission occur in a short time frame. Early events during cell-to-cell

transmission may determine HTLV-1 vial infectivity to target cells by modulating

expression of HTLV-1 gene expression. Therefore, we profiled expression of pX ORF I

mRNA(p12I and p27I) and pX ORF III/IV (Tax/Rex) mRNA from 3 hours to 7 days

after cocultivating HTLV-1 infected cells. We were particularly interested in the role of

pX ORF I protein during early HTLV-1 transmission and therefore utilized ACH.p12

cells, which are lacking pX ORF I mRNA expression.

In this study, we demonstrated that expression of Tax/Rex mRNA was

differentially regulated during early HTLV-1 transmission. Furthermore, Tax/Rex

expression was enhanced by pX ORF I expression and activation status of target cells.

Our data suggested that stimuli and their corresponding signaling pathways, which were

generated from pX ORF I proteins or cell contacts with target cells, may modulate

158

Tax/Rex mRNA expression and subsequent Tax-mediated viral production during

HTLV-1 transmission. Interestingly, elevation of intracellular calcium resulted in

enhancement of structural protein synthesis and activation of long terminal repeat

(LTR) - mediated transcription in cells infected with bovine leukemia virus (BLV),

another member of the deltaretrovirus genus 53 A number of studies have been reported

a role of Tax in transcriptional regulation of expression of viral and cellular genes 54,

but the influence of other viral ORFs in Tax expression is unclear. Signaling via CD2

and ICAM-148,49, which enhanced viral replication and production, may affect

expression of Tax/Rex in HTLV-1 cells.

Our data indicate that the expression of pX ORF I (p12I and p27I) mRNA while

during early HTLV-1 transmission was always lower than that of Tax/Rex mRNA. Our

result combined with our previous studies of HTLV-1 infectivity in target PBMC verify

a critical role of pX ORF I expression in viral transmission. Our data indicate that an

early increase of Tax/Rex mRNA expression in sorted target cells, which reflects de

novo HTLV-1 gene expression. Tax/Rex mRNA expression was increased in cells that

also expressed pX ORF I, suggesting a another role of pX ORF I protein in establishing

HTLV-1 infection in target cells. While the mechanism of pX ORF I expression in viral

replication is beyond the scope of our current study, based on previous studies of pX

ORF I encoded p12I, the viral protein may influence the viral life cycle (e.g., reverse

transcription) directly or indirectly through its unknown cell activation properties and

calcium signaling properties. If pX ORF I encoded p12I enhances early events

following virus entry such as reverse transcription and integration, the protein would

need to be present in viral pre-integration complexes or expressed early following initial

159

integration events. Our data is the first to indicate that indeed pX ORF I is expressed

early (within 12 h) of cell-to-cell contact. Analogously, the human immunodeficiency

virus (HIV) accessory protein, Nef, which functions in a manner similar to HTLV-1

p12I, is found as a protein in HIV virions 55 and is selectively expressed before viral

integration 56. Similarly, our data would support a role for HTLV-1 pX ORF I protein in

the formation of the viral synapse or selectively expressed before or immediately

following integration to enhance de novo HTLV-1 infection in target cells.

In summary, our study further defines a role of pX ORF I expression in

activation of HTLV-1 infected cells. Our data indicated that expression of pX ORF I

enhanced early establishment HTLV-1 infection in target cells during cell-to-cell

transmission and that abrogation of pX ORF I expression did not influence HTLV-1

mediated bystander cell activation. Collectively, we have clarified the role of pX ORF I

in HTLV-1 replication and viral transmission during the earliest event of viral infection

in T lymphocytes.

4.5 References

(1) Gessain A, Mahieux R. [A virus called HTLV-1. Epidemiological aspects. Presse Med. 2000;29:2233-2239.

(2) Bangham CR. HTLV-1 infections. J Clin Pathol. 2000;53:581-586.

(3) Franchini G. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood. 1995;86:3619-3639.

(4) Hollsberg P. Mechanisms of T-cell activation by human T-cell lymphotropic virus type I. Microbiol Mol Biol Rev. 1999;63:308-333.

(5) Berneman ZN, Gartenhaus RB, Reitz MS et al. Expression of alternatively spliced human T-lymphotropic virus type 1 pX mRNA in infected cell lines and

160

in primary uncultured cells from patients with adult T-cell leukemia/lymphoma and healthy carriers. Proc Natl Acad Sci USA. 1992;89:3005-3009.

(6) Ciminale V, Pavlakis GN, Derse D, Cunningham CP, Felber BK. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J Virol. 1992;66:1737-1745.

(7) Furukawa K, Furukawa K, Shiku H. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 1991;295:141-145.

(8) Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type 1. Proc Natl Acad Sci USA. 1992;89:8813-8817.

(9) Cereseto A, Berneman Z, Koralnik I et al. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia. 1997;11:866-870.

(10) Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type I. Proc Natl Acad Sci U S A. 1992;89:8813-8817.

(11) Dekaban GA, Peters AA, Mulloy JC et al. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology. 2000;274:86-93.

(12) Pique C, Uretavidal A, Gessain A et al. 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. 2000;191:567-572.

(13) Princler GL, Julias JG, Hughes SH, Derse D. Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology. 2003;317:136-145.

(14) Mueller N.E., Blattner W.A. Retroviruses-Human T-Cell Lymphotropic Virus. Fourth Edition, 785-814. 1997. New York, N.Y., Plenum Publishing Corp.

(15) Bangham CR. The immune control and cell-to-cell spread of human T-lymphotropic virus type 1. J Gen Virol. 2003;84:3177-3189.

(16) Manns A, Hisada M, Lagrenade L. Human T-lymphotropic virus type I infection. Lancet. 1999;353:1951-1958.

161

(17) Kehn K, Berro R, de La FC et al. Mechanisms of HTLV-1 transformation. Front Biosci. 2004;9:2347-2372.

(18) Merl S, Kloster B, Moore J et al. Efficient Transformation of Previously Activated and Dividing T Lymphocytes by Human T Cell Leukemia-Lymphoma Virus. Blood. 1984;64:967-974.

(19) Kimata JT, Palker TJ, Ratner L. The Mitogenic Activity of Human T-Cell Leukemia Virus Type- I Is T-Cell Associated and Requires the CD2/LFA-3 Activation Pathway. J Virol. 1993;67:3134-3141.

(20) Wucherpfennig KW, Hollsberg P, Richardson JH, Benjamin D, Hafler DA. T-cell activation by autologous human T-cell leukemia virus type I- infected T-cell clones. Proc Natl Acad Sci U S A. 1992;89:2110-2114.

(21) Albrecht B, Lairmore MD. Critical role of human T-lymphotropic virus type 1

accessory proteins in viral replication and pathogenesis. Microbiol Mol Biol Rev. 2002;66:396-406.

(22) Michael B, Nair A, Lairmore MD. Role of accessory proteins of HTLV-1 in viral replication, T cell activation, and cellular gene expression. Front Biosci. 2004;9:2556-2576.

(23) Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

(24) Albrecht B, Collins ND, Burniston MT et al. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J Virol. 2000;74:9828-9835.

(25) Albrecht B, D'Souza CD, Ding W et al. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J Virol. 2002;76:3493-3501.

(26) Ding W, Albrecht B., Kelley RE et al. 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. 2002;76:10374-10382.

(27) Ding W, Kim SJ, Nair AM et al. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol. 2003;77:11027-11039.

(28) Ding W, Albrecht B, Luo R et al. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J Virol. 2001;75:7672-7682.

162

(29) Kim SJ, Ding W, Albrecht B, Green PL, Lairmore MD. 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. 2003;278:15550-15557.

(30) Nair A, Michael B, Hiraragi H et al. Human T lymphotropic virus type 1 accessory protein p12I modulates calcium-mediated cellular gene expression and enhances p300 expression in T lymphocytes. AIDS Res Hum Retroviruses. 2005;21:273-284.

(31) Kim SJ, Nair AM, Fernandez S, Mathes L, Lairmore MD. Enhancement of LFA-1-mediated T cell adhesion by human T lymphotropic virus type 1 p12I1. J Immunol. 2006;176:5463-5470.

(32) Nejmeddine M, Barnard AL, Tanaka Y, Taylor GP, Bangham CR. Human T-lymphotropic Virus, Type 1, Tax Protein Triggers Microtubule Reorientation in the Virological Synapse. J Biol Chem. 2005;280:29653-29660.

(33) Barnard AL, Igakura T, Tanaka Y, Taylor GP, Bangham CR. Engagement of specific T cell surface molecules regulates cytoskeletal polarization in HTLV-1-infected lymphocytes. Blood. 2005.

(34) Dustin M. Viral spread through protoplasmic kiss. Nat Cell Biol. 2003;5:271-272.

(35) Igakura T, Stinchcombe JC, Goon PK et al. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science. 2003;299:1713-1716.

(36) Mori N, Murakami S, Oda S, Eto S. Human T-cell leukemia virus type I tax induces intracellular adhesion molecule-1 expression in T cells. Blood. 1994;84:350-351.

(37) Robek MD, Wong FH, Ratner L. 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. 1998;72:4458-4462.

(38) Collins ND, D'Souza C, Albrecht B et al. 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. 1999;73:9642-9649.

(39) Rodrigues MS, Echevarria-Lima J, Fernandes dos SN, Rumjanek VM. CD69 expression induced by thapsigargin, phorbol ester and ouabain on thymocytes is dependent on external Ca2+ entry. Life Sci. 2003;73:1037-1051.

163

(40) Owen SM, Rudolph DL, Dezzutti CS et al. Transcriptional activation of the intercellular adhesion molecule 1 (CD54) gene by human T lymphotropic virus types I and II Tax is mediated through a palindromic response element. AIDS Res Hum Retroviruses. 1997;13:1429-1437.

(41) Kimata JT, Wong F, Wang J, Ratner L. Construction and characterization of infectious human T-cell leukemia virus type 1 molecular clones. Virology. 1994;204:656-664.

(42) Lantz RC, Lemus R, Lange RW, Karol MH. Rapid reduction of intracellular glutathione in human bronchial epithelial cells exposed to occupational levels of toluene diisocyanate. Toxicol Sci. 2001;60:348-355.

(43) Yamano Y, Nagai M, Brennan M et al. 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. 2002;99:88-94.

(44) Alfahim A, Cabre P, Kastrukoff L, Dorovinizis K, Oger J. Blood mononuclear cells in patients with HTLV-I-associated myelopathy: Lymphocytes are highly activated and adhesion to endothelial cells is increased. Cell Immunol. 1999;198:1-10.

(45) Sancho D, Gomez M, Sanchez-Madrid F. CD69 is an immunoregulatory molecule induced following activation. Trends Immunol. 2005;26:136-140.

(46) Sancho D, Gomez M, Viedma F et al. CD69 downregulates autoimmune reactivity through active transforming growth factor-beta production in collagen-induced arthritis. J Clin Invest. 2003;112:872-882.

(47) Kulkarni AB, Huh CG, Becker D et al. Transforming growth factor beta 1 null mutation in mice causes excessive inflammatory response and early death. Proc Natl Acad Sci U S A. 1993;90:770-774.

(48) Yamamoto A, Hara H, Kobayashi T. Induction of the expression of gag protein in HTLV-I infected lymphocytes by anti-ICAM 1 antibody in vitro. J Neurol Sci. 1997;151:121-126.

(49) Guyot DJ, Newbound GC, Lairmore MD. Signaling via the CD2 receptor enhances HTLV-1 replication in T lymphocytes. Virology. 1997;234:123-129.

(50) Fukudome K, Furuse M, Fukuhara N et al. Strong induction of ICAM-1 in human T cells transformed by human T-cell-leukemia virus type 1 and depression of ICAM-1 or LFA-1 in adult T-cell-leukemia-derived cell lines. Int J Cancer. 1992;52:418-427.

164

(51) Imai T, Tanaka Y, Fukudome K et al. Enhanced expression of LFA-3 on human T-cell lines and leukemic. International Journal of Cancer. 1993;55(5):811-816.

(52) Benovic S, Kok T, Stephenson A et al. De novo reverse transcription of HTLV-1 following cell-to- cell transmission of infection. Virology. 1998;244:294-301.

(53) Bondzio A, Abraham-Podgornik A, Blankenstein P, Risse S. Involvement of intracellular Ca2+ in the regulation of bovine leukemia virus expression. Biol Chem. 2001;382:407-416.

(54) Kashanchi F, Brady JN. Transcriptional and post-transcriptional gene regulation of HTLV-1. Oncogene. 2005;24:5938-5951.

(55) Zhou J, Aiken C. Nef enhances human immunodeficiency virus type 1 infectivity resulting from intervirion fusion: evidence supporting a role for Nef at the virion envelope. J Virol. 2001;75:5851-5859.

(56) Wu Y, Marsh JW. Selective transcription and modulation of resting T cell activity by preintegrated HIV DNA. Science. 2001;293:1503-1506.

165

Figure 4.1.Cell surface expression of early activation markers and adhesion

molecules on wild type HTLV-1 infected cells and T cells lacking pX ORF I

expression. Flow cytometric analysis specific for activation markers (A) CD25 and

CD69, adhesion molecules (B) CD58 and CD2, and (C) CD11a and CD54 was

performed on ACH cell line (ACH.2 and ACH.2p) and ACH.p12 cell line (ACH.p12.4

and ACH.p12.4p) (106 cells/sample). The results were representative of at least 2

independent experiments.

166

167

Figure 4.2. Deletion of pX ORF I expression in HTLV-1 infected cells did not affect

bystander proliferation of uninfected target PBMC.

ACH cell line (ACH.2 and ACH.2p) and ACH.p12 cell line (ACH.p12.4 and

ACH.p12.4p) were fixed with 1% PFA and cocultured target PBMC in the presence or

absence of human IL-2 (20 U/ml) and PHA (2 µg/ml) stimulation. Fixed effecter cells

were cocultured with target PBMC at 1:2 ratios. Target PBMC were cultured without

effecter cells for control. The MTS colorimetric assay was performed at 0 days and at 3

days after coculture by measuring optical density (OD) in triplicate samples. Graph

depicts the fold proliferation of target cell OD at day 3 over day 0. Values are mean ±

SEM of triplicate samples of 3 independent experiments. Statistical significance was

analyzed by Student’s t test.

168

Figure 4.3. Abrogation of pX ORF I expression in HTLV-1 infected cells did not

affect bystander activation of target PBMC.

Target PBMC were pre-labeled with 0.5µM CFDA-SE and were cocultured with live (B

and E) or fixed (C and F) effecter ACH cell line (ACH.2 and ACH.2p) or ACH.p12 cell

line (ACH.p12.4 and ACH.p12.4p). Live effecter cells were also cocultured separately

in permeable transwell inserts (D and G). HTLV-1 infected effecter cells were

cocultured with target PBMC at 1:2 ratios. After 24 h (B-D) and 36 h (E-G) coculture,

cells were stained with R-PE conjugated antibodies against CD25 and CD69. (A) Flow

cytometric analysis was performed at 36 h after coculture. CMFD-SE labeled target

PBMC and live ACH.2p and ACH.p12.4p effecter cells were cocultured and stained

with RPE- conjugated CD25 antibody. (B-G) CD25 and CD69 percentage positive

were measured in the target PBMC (CMFD-SE positive cells) by flow cytometric

analysis. Target PBMC were cultured without effecter cells in the presence or absence

of human IL-2 (20 U/ml) and PHA (2 µg/ml) stimulation for activation control. Values

are mean ± SEM of duplicate samples and representative of two independent

experiments. Statistical significance was analyzed by Student’s t test.

169

170

171

Figure 4.4. HTLV-1 pX ORF I and pX ORF III/IV mRNAs were differentially

expressed during the early HTLV-1 transmission.

Tax/Rex, p27I, and p12I mRNA copy numbers were determined as described under

Material and Methods and normalized relative to 1.0 x 106 copies of human GAPDH

mRNA. (A) Relative mRNA copy number of Tax/Rex, p27I, and p12I messages in

HTLV-1 infected cells (MT-2, ACH.2p, and ACH.p12.4p), as well as in uninfected

PBMC. Values are mean ± SEM of duplicate samples and representative at least 3

independent experiments. Relative copy number of (B) Tax/Rex, (C) p27I and (D)12I

mRNA were measured from 3 hours to 24 hours after coculture in unsorted samples.

HTLV-1 infected cells (ACH.2p or ACH.p12.4p) were cocultured with unstimulated

quiescent or pre- activated target PBMC at 1:2 ratios. Values are mean ± SEM of

duplicate samples and representative at least 2 independent experiments.

172

173

Figure 4.5. Target PBMC were selectively sorted from HTLV-1 infected cells.

HTLV-1 infected cells (ACH.2p or ACH.p12.4p) were pre-labeled with CMFDA and

cocultured with unstimulated quiescent or pre-activated target PBMC at 1:2 ratios. (A)

CMFDA negative unstimulated target PBMC and CMFDA positive ACH.2p cells were

analyzed at 3 day after coculture by flow cytometry. (B) Collected target PBMC were

99 % CMFDA negative and (C) labeled ACH.2p cells were 95 % CMFDA positive

after cell sorting.

174

Figure 4.6. Expression of pX ORF I enhanced HTLV-1 viral transmission to target

PBMC.

HTLV-1 infected cells (ACH.2p or ACH.p12.4p) were cocultured with unstimulated

quiescent or pre-activated target PBMC at 1:2 ratios. Relative copy number of Tax/Rex

mRNA was measured in CMFDA positive HTLV-1 producer cells and CMFDA

negative target PBMC after cell sorting of 1 and 3 day coculture. Percent of Tax/Rex

mRNA in target PBMC was measured by dividing relative copy number of target

PBMC with that of total cells (i.e., Copy number of target PBMC / {copy number of

infected cells + target PBMC} x 100) at the time point of cell sorting. Values are mean

± SEM of duplicate samples and representative at least 2 independent experiments.

175

Figure 4.7. Expression of HTLV-1 pX ORF I and pX ORF III/IV mRNAs in de

novo HTLV-1 infected target PBMC.

(A, B) HTLV-1 infected cells (ACH.2p or ACH.p12.4p) were cocultured with

unstimulated quiescent or pre-activated target PBMC at 1:2 ratios. (A) Relative copy

number of Tax/Rex mRNA was measured in CMFDA negative target PBMC after cell

sorting of 3 day coculture. (B) Fold increase of Tax/Rex mRNA in sorted PBMC from

day 3 to day5 and from day 5 to day 7 after coculture was measured. Values are mean ±

SEM of duplicate samples and representative 2 independent experiments.

176

(C, D) HTLV-1 infected ACH.2p cells were cocultured with pre-activated target PBMC

at 1:1 ratios. (C) Relative copy number of p27I and p12 I mRNA measured in CMFDA

negative target PBMC after cell sorting of 3 day coculture. (D) p27I and p12 I mRNA

ratio to Tax/Rex mRNA at each day point was measured Values are mean ± SEM of

duplicate samples.

177

CHAPTER 5

SYNOPSIS AND FUTURE DIRECTIONS

Since the initial discovery of HTLV-1, extensive studies have been performed to

understand mechanisms of HTLV-1-mediated lymphoproliferative and

neurodegenerative diseases. Most of these studies have been focused on the role of the

transcriptional activator Tax and its role in immortalization and transformation of T

cells. However, the early events of viral infection, such as cell-to-cell transmission,

regulation of reverse transcription, integration and regulation of viral and cellular gene

transcription are incompletely understood. Recent studies have produced an

accumulating body of data that a critical role for pX ORF I and II encoded proteins in

viral infectivity, maintenance of viral loads, host cell activation, and regulation of gene

transcription 1-5. Particularly, the role of an HTLV-1 p12I encoded by pX ORF I was

discovered to be required for productive infection in a rabbit model and viral infectivity

178

in non-stimulated PBMC 4,5. Moreover, our laboratories data demonstrated the role of

p12I in calcium-dependent signaling, activation of NFAT 1, and enhanced production of

IL-2 6 and p300 expression 7. Data presented in this thesis provided further evidence

that the highly conserved viral accessory protein, p12I is required for early HTLV-1

infection. I demonstrated in this thesis that HTLV-1 p12I modulates NFAT activation

by interacting with calcineurin (Chapter 2), enhances cell-to-cell adhesion by inducing

LFA-1 clustering on the surface of T cells (Chapter 3) and facilitates early viral

infectivity during cell-to-cell transmission (Chapter 4).

5.1 Studies to test the role of p12I binding to calcineurin on HTLV-1-mediated T

cell activation and in vivo and in vitro viral infectivity

We identified a PSLP(I/L)T motif in p12I, which is highly homologous with the

PxIxIT calcineurin-binding motif of NFAT. Through this motif, p12I modulates NFAT

nuclear translocation and transcription activity 8. However it is unclear why p12I has

two regulatory functions for NFAT transcriptional activity; positive modulation by

increasing cytosolic calcium concentration from ER stores 2 and negative modulation by

calcineurin binding. Therefore, further analysis using an infectious HTLV-1 proviral

clone such as ACH that contains mutations in PSLP(I/L)T motif, would be required to

understand the role of p12I binding to calcineurin in context to HTLV-1- mediated cell

activation and during in vivo and in vitro viral infectivity. An alanine substitution in this

motif within an infectious HTLV-1 clone will be helpful to define the role of p12I in

binding to calcineurin in infected cells. The p12I mutant (p12I AxAxAA) exhibited a

179

greatly reduced binding affinity for calcineurin, but retained the ability to enhance

NFAT nuclear translocation and transcriptional activity 8. These properties of p12I,

however, were not tested in context to an HTLV-1 provirus and would be an important

extension of the findings in the thesis. By measuring IL-2 production 6, or detecting

activation markers (e.g., CD25 and CD69), the activation status of HTLV-1 infected

cells expressing an HTLV-1 pX ORF I virus containing AxAxAA mutations could be

evaluated and compared to wild type HTLV-1 infected cells (ACH). Furthermore, in

vitro and in vivo viral infectivity of HTLV-1 infected cells containing this mutant

provirus can be directly compared to wild type HTLV-1 infected cells (ACH) using in

vitro p19 detection 4 or approaches used in Chapter 4 of this thesis such as coculture and

real time RT-PCR methods followed by cell sorting to differentiate target and producer

cells, as well as rabbit inoculation to direct test infectivity in vivo 5.

5.2 Further investigation of the role of p12I in MTOC polarization and virological

synapse formation

In Chapter 3 of this thesis, our data indicated that HTLV-1 p12I promotes cell-

to-cell spread by inducing LFA-1 clustering on T-cells. Furthermore, a recent cross-

linking study indicated that LFA-1 and ICAM-1 interaction is important in HTLV-1

cell-to-cell transmission 9. Antibodies against LFA-1 and ICAM-1 triggered

microtubule-organizing center (MTOC) polarization, which is characteristic of HTLV-

1-induced virological synapses. Taken together my data presented in Chapter 3

suggests that p12I may contribute to MTOC polarization and enhance formation of the

virological synapse by increasing cell adhesion through clustered LFA-1. However, to

180

date, there is no direct published reports that p12I can modulate MTOC polarization and

virological synapse formation. Moreover, p12I may enhance MTOC re-polarization

directly through calcium mediated signaling without upstream activation of signaling

pathway, because calcium-mediated events are required for TCR-mediated MTOC

polarization 10. The role of p12I in MTOC polarization and formation of virological

synapse can be tested by comparing wild type HTLV-1 infected cells (ACH cell line)

and T cells lacking p12I expression (ACH.p12 cell line). The frequency of MTOC

polarization can be measured under the influence of variety of stimulation protocols,

such as engagement of TCR, LFA-1, ICAM-1, CD2 and LFA-3, which are important in

HTLV-1 cell-to-cell mediated bystander target cell activation 11,12. Because LFA-1

clustering was enhanced in wild type ACH cells 13, engagement of LFA-1 may result in

higher MTOC frequency in wild type when compared to ACH.p12 cell lines. By

measuring conjugation rates between HTLV-1 infected cells and target cells, the role of

p12I in formation of the virological synapse could be also tested. Furthermore, studies

designed with exogenously expressed p12I in Jurkat T cell could be expanded to

understand the role of p12I in MTOC polarization and virological synapse formation.

5.3 Studies to explore the mechanisms of LFA-1 affinity regulation influenced by

p12I expression

Our data in Chapter 3 demonstrated that abrogation of ORF I expression in

HTLV-1 infected cells resulted in decreased LFA-1 affinity. However, expression of

p12I in Jurkat T cells did not affect LFA-1 affinity, indicating that the role of p12I is not

directly related to the regulation of LFA-1 affinity control mechanisms. Also these data

181

suggested that in context to HTLV-1 infection the balance of HTLV-1 regulatory and

accessory protein expression is required for LFA-1 affinity control and efficinet cell-to-

cell HTLV-1 transmission. Due to incompletely defined mechanisms of LFA-1 affinity

control in the literature compared to reports of regulation of LFA-1 clustering 14,

investigating mechanism of LFA-1 affinity reduction in p12I expression abrogated cells

(ACH.p12 cell line) may be challenging. In this regard, studies to test the role of other

regulatory or accessory protein in LFA-1 affinity control would be useful to further

understanding the role of this p12I in context to HTLV-1 infection. Interestingly, the

expression of the Lck gene, a critical plasma membrane associated kinase in CD4+ T

cells to maintain the high affinity status of integrin α1β4 15, was down-regulated by the

HTLV-1 accessory protein p30II 16.

5.4 Studies required to test the role of p12I in HTLV-1 envelope glycoprotein

folding and cell surface expression

HTLV-1 p12I was reported to be required for efficient in vivo and in vitro rabbit

viral infectivity 4,5. Because interactions between the HTLV-1 envelope protein and its

receptor are critical for cell-to-cell transmission of the virus, the requirement of p12I on

efficient viral transmission may be due to viral protein’s role or pX ORF I mRNA in

regulating envelope expression. Our previous data demonstrated that p12I interacts with

both calreticulin and calnexin 2, which are molecular chaperones to newly synthesized

and unfolded glycoproteins 17. Furthermore, the folding of most viral envelope proteins

including HIV-1 envelope are associated with calreticulin and calnexin in the lumen of

the endoplasmic reticulum of infected cells 18. Thus the binding of p12I to both

182

calreticulin and calnexin may help the proper folding of HTLV-1 envelope

glycoproteins or increase their cell surface expression to enhance receptor interactions.

5.5 Possible role of p12I in regulating CD69 expression during HTLV-1 infection

The expression of the early activation marker CD69 was consistently higher in

wild type HTLV-1 infected cells compared to ACH.p12 cells. The expression of CD69

is dependent on calcium mediated signaling 19. Thus, higher CD69 expression in wild

type ACH cells may have resulted from activation of p12I-mediated calcium signaling.

The role of CD69 in immune function is still largely unknown, due to the lack of

known ligands and a suitable in vivo model 20. However, recent findings suggested that

CD69 plays a immunoregulatory function by increasing the production of transforming

growth factor-β (TGF-β) 20,21. Interestingly, TGF-β down-regulates antigen-presenting

cell (APC) function 22, suggesting a potential role of CD69 expression in escaping

immune surveillance during HTLV-1 pathogenesis. Further studies will be required to

understand the role of CD69 in HTLV-1 pathogenesis, particularly in escaping immune

surveillance as well as cell-to-cell transmission.

5.6 The potential role of p12I in regulation of viral gene expression during HTLV-1

cell-to-cell transmission

In Chapter 4, my data demonstrated that expression of HTLV-1 pX ORF

messages including Tax/Rex, p27I and p12I mRNA were differentially regulated during

early HTLV-1 transmission. Furthermore, Tax/Rex expression was enhanced by pX

ORF I expression and was modulated by the activation status of target cells. These

183

results suggest that Tax/Rex mRNA expression and subsequent Tax-mediated viral

production may be influenced by pX ORF I mRNA or encoded proteins during cell-to-

cell contact required for efficient HTLV-1 transmission. Therefore, studies designed to

identify stimuli and their corresponding signaling pathways modulating viral gene

expression during early HTLV-1 cell-to-cell transmission will be required to more

completely understand molecular mechanisms of HTLV-1 early cell-to-cell

transmission. Moreover, a number of reports have been identified a role of Tax in

transcriptional regulation of expression of viral and cellular genes 23, however, the

effect of other viral ORF proteins in Tax expression has not been reported.

HTLV-1 pX ORF I encoded p12I is a modulator of calcium signal transduction

1,3-5, suggesting a role of viral gene regulation during early viral transmission.

Interestingly, calcium-mediated signaling induced activation of LTR-mediated

transcription and viral structural protein synthesis in cells infected with bovine

leukemia virus (BLV), another deltaretrovirus 24. Thus further studies are required to

directly test a role of calcium-mediated signaling in the regulation of HTLV-1 gene

expression in particularly during HTLV-1 cell-to-cell transmission. Moreover, other

signaling pathways, particularly, signaling via CD2 and ICAM-125,26, which enhanced

viral replication, may directly enhance Tax/Rex expression during early viral

transmission. Furthermore, a more complete understanding of cell signaling pathways

that modulate viral gene expression during viral transmission would be useful for

development of drugs that are designed to block early viral transmission events.

184

5.7 Studies to test the influence of pX OFR I or p12I in early expression of HTLV-1

regulatory proteins

Our data in Chapter 4 indicated that de novo Tax/Rex mRNA expression was

enhanced in cells expressing pX ORF I, suggesting a role of pX ORF I mRNA or its

encoded proteins in establishing HTLV-1 infection in target cells. Furthermore, our data

is the first to indicate that pX ORF I is expressed early (within 12 h) of cell-to-cell

contact. Therefore, pX ORF I encoded p12I may influence the earliest events of the viral

life cycle directly or indirectly through its known cell activation properties and calcium

signaling properties. In this regard the early events of the viral life cycle such as virus

entry, uncoating, reverse transcription and provirus integration may be modulated by

p12I. However, if p12I influences these early events, the protein would need to be

present in viral pre-integration complexes or expressed early following initial

integration events. Similarly, the HIV accessory protein, Nef, is found as a protein in

HIV virions 27,28 and is selectively expressed before viral integration 29 . Because, some

functions of Nef are similar to that of p12I in T cell activation, it would be interesting to

test if p12I is present in HTLV-1 viral particles or is transferred during the virological

synapse formation. Moreover, tests to determine if p12I is expressed selectively before

viral integration or expressed early following initial integration, would be required for

understanding the role of p12I in early establishment of HTLV-1 infection. Further

studies using single-round infection assays, reverse transcriptase inhibitor or integrase

deficient proviral clones, would be helpful in identifying the function of p12I in the

early viral life cycle. Finally, direct tests to determine if calcium-mediated signaling

could affect HTLV-1 reverse transcription, integration and early viral gene expression

185

would be helpful to indirectly implicate p12I in de novo HTLV-1 infection in target

cells.

Overall, the data from this thesis have contributed to our understanding of the

role of pX ORF I and its encoded p12I in HTLV-1 infection. It appears that pX ORF I

mRNA or proteins are indeed not “accessory”, but essential in the control of the early

events of HTLV-1 infection in T cells.

5.8 REFERENCES

Reference List

(1) Albrecht B, D'Souza CD, Ding W et al. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J Virol. 2002;76:3493-3501.

(2) Ding W, Albrecht B, Luo R et al. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J Virol. 2001;75:7672-7682.

(3) Ding W, Albrecht B., Kelley RE et al. 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. 2002;76:10374-10382.

(4) Albrecht B, Collins ND, Burniston MT et al. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J Virol. 2000;74:9828-9835.

(5) Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

(6) Ding W, Kim SJ, Nair AM et al. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol. 2003;77:11027-11039.

186

(7) Nair A, Michael B, Hiraragi H et al. Human T lymphotropic virus type 1 accessory protein p12I modulates calcium-mediated cellular gene expression and enhances p300 expression in T lymphocytes. AIDS Res Hum Retroviruses. 2005;21:273-284.

(8) Kim SJ, Ding W, Albrecht B, Green PL, Lairmore MD. 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. 2003;278:15550-15557.

(9) Barnard AL, Igakura T, Tanaka Y, Taylor GP, Bangham CR. Engagement of specific T cell surface molecules regulates cytoskeletal polarization in HTLV-1-infected lymphocytes. Blood. 2005.

(10) Kuhne MR, Lin J, Yablonski D et al. Linker for activation of T cells, zeta-associated protein-70, and Src homology 2 domain-containing leukocyte protein-76 are required for TCR-induced microtubule-organizing center polarization. J Immunol. 2003;171:860-866.

(11) Kimata JT, Palker TJ, Ratner L. The Mitogenic Activity of Human T-Cell Leukemia Virus Type- I Is T-Cell Associated and Requires the CD2/LFA-3 Activation Pathway. J Virol. 1993;67:3134-3141.

(12) Wucherpfennig KW, Hollsberg P, Richardson JH, Benjamin D, Hafler DA. T-cell activation by autologous human T-cell leukemia virus type I- infected T-cell clones. Proc Natl Acad Sci U S A. 1992;89:2110-2114.

(13) Kim SJ, Nair AM, Fernandez S, Mathes L, Lairmore MD. Enhancement of

LFA-1-mediated T cell adhesion by human T lymphotropic virus type 1 p12I1. J Immunol. 2006;176:5463-5470.

(14) Hogg N, Laschinger M, Giles K, Mcdowall A. T-cell integrins: more than just sticking points. J Cell Sci. 2003;116:4695-4705.

(15) Feigelson SW, Grabovsky V, Winter E et al. The Src kinase p56(lck) up-regulates VLA-4 integrin affinity. Implications for rapid spontaneous and chemokine-triggered T cell adhesion to VCAM-1 and fibronectin. J Biol Chem. 2001;276:13891-13901.

(16) Michael B, Nair AM, Hiraragi H et al. Human T lymphotropic virus type-1 p30II alters cellular gene expression to selectively enhance signaling pathways that activate T lymphocytes. Retrovirology. 2004;1:39.

(17) Johnson S, Michalak M, Opas M, Eggleton P. The ins and outs of calreticulin: from the ER lumen to the extracellular space. Trends Cell Biol. 2001;11:122-129.

187

(18) Braakman I, van Anken E. Folding of viral envelope glycoproteins in the endoplasmic reticulum. Traffic. 2000;1:533-539.

(19) Rodrigues MS, Echevarria-Lima J, Fernandes dos SN, Rumjanek VM. CD69 expression induced by thapsigargin, phorbol ester and ouabain on thymocytes is dependent on external Ca2+ entry. Life Sci. 2003;73:1037-1051.

(20) Sancho D, Gomez M, Sanchez-Madrid F. CD69 is an immunoregulatory molecule induced following activation. Trends Immunol. 2005;26:136-140.

(21) Sancho D, Gomez M, Viedma F et al. CD69 downregulates autoimmune reactivity through active transforming growth factor-beta production in collagen-induced arthritis. J Clin Invest. 2003;112:872-882.

(22) Kulkarni AB, Huh CG, Becker D et al. Transforming growth factor beta 1 null mutation in mice causes excessive inflammatory response and early death. Proc Natl Acad Sci U S A. 1993;90:770-774.

(23) Kashanchi F, Brady JN. Transcriptional and post-transcriptional gene regulation of HTLV-1. Oncogene. 2005;24:5938-5951.

(24) Bondzio A, Abraham-Podgornik A, Blankenstein P, Risse S. Involvement of intracellular Ca2+ in the regulation of bovine leukemia virus expression. Biol Chem. 2001;382:407-416.

(25) Yamamoto A, Hara H, Kobayashi T. Induction of the expression of gag protein in HTLV-I infected lymphocytes by anti-ICAM 1 antibody in vitro. J Neurol Sci. 1997;151:121-126.

(26) Guyot DJ, Newbound GC, Lairmore MD. Signaling via the CD2 receptor enhances HTLV-1 replication in T lymphocytes. Virology. 1997;234:123-129.

(27) Kotov A, Zhou J, Flicker P, Aiken C. Association of Nef with the human immunodeficiency virus type 1 core. J Virol. 1999;73:8824-8830.

(28) Zhou J, Aiken C. Nef enhances human immunodeficiency virus type 1 infectivity resulting from intervirion fusion: evidence supporting a role for Nef at the virion envelope. J Virol. 2001;75:5851-5859.

(29) Wu Y, Marsh JW. Selective transcription and modulation of resting T cell activity by preintegrated HIV DNA. Science. 2001;293:1503-1506.

188

BIBLIOGRAPHY

Adachi Y, Copeland TD, Takahashi C et al. Phosphorylation of the Rex protein of human T-cell leukemia virus type I. Journal of Biological Chemistry. 1992;267:21977-21981.

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

Ahmed YF, Gilmartin GM, Hanly SM, Nevins JR, Greene WC. The HTLV-1 Rex Response Element Mediates a Novel Form of mRNA Polyadenylation. Cell. 1991;64:727-737.

Alamy AH, Menezes FB, Leite AC, Nascimento OM, Araujo AQ. Dysautonomia in human T-cell lymphotrophic virus type I-associated myelopathy/tropical spastic paraparesis. Ann Neurol. 2001;50:681-685.

Albrecht B, Collins ND, Burniston MT et al. Human T-lymphotropic virus type 1 open reading frame I p12(I) is required for efficient viral infectivity in primary lymphocytes. J Virol. 2000;74:9828-9835.

Albrecht B, D'Souza CD, Ding W et al. Activation of nuclear factor of activated T cells by human T- lymphotropic virus type 1 accessory protein p12(I). J Virol. 2002;76:3493-3501.

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

Alfahim A, Cabre P, Kastrukoff L, Dorovinizis K, Oger J. Blood mononuclear cells in patients with HTLV-I-associated myelopathy: Lymphocytes are highly activated and adhesion to endothelial cells is increased. Cell Immunol. 1999;198:1-10.

189

Allen WE, Zicha D, Ridley AJ, Jones GE. A role for Cdc42 in macrophage chemotaxis. J Cell Biol. 1998;141:1147-1157.

Anderson DW, Epstein JS, Lee TH et al. Serological confirmation of human T-lymphotropic virus type I infection in healthy blood and plasma donors 5242. Blood. 1989;74:2585-2591.

Aramburu J, Garcia-Cozar F, Raghavan A et al. Selective inhibition of NFAT activation by a peptide spanning the calcineurin targeting site of NFAT. Mol Cell. 1998;1:627-637.

Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A 5399. Science. 1999;285:2129-2133.

Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A 5399. Science. 1999;285:2129-2133.

Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A 5399. Science. 1999;285:2129-2133.

Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A 5399. Science. 1999;285:2129-2133.

Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A 5399. Science. 1999;285:2129-2133.

Aramburu J, Yaffe MB, Lopez-Rodriguez C et al. Affinity-driven peptide selection of an NFAT inhibitor more selective than cyclosporin A 5399. Science. 1999;285:2129-2133.

Ardouin L, Bracke M, Mathiot A et al. Vav1 transduces TCR signals required for LFA-1 function and cell polarization at the immunological synapse. Eur J Immunol. 2003;33:790-797.

Arias RA, Munoz LD, Munoz-Fernandez MA. Transmission of HIV-1 infection between trophoblast placental cells and T-cells take place via an LFA-1-mediated cell to cell contact. Virology. 2003;307:266-277.

Arisawa K, Soda M, Endo S et al. Evaluation of adult T-cell leukemia/lymphoma incidence and its impact on non-Hodgkin lymphoma incidence in southwestern Japan. Int J Cancer. 2000;85:319-324.

190

Arnold J, Yamamoto B, Li M et al. Enhancement of infectivity and persistence in vivo by HBZ, a natural antisense coded protein of HTLV-1. Blood. 2006.

Asher DM, Goudsmit J, Pomeroy K et al. Antibodies to HTLV-I in populations of the southwestern Pacific. J Med Virol. 1988;26:339.

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

Baeuerle PA, Baltimore D. I kappa B: a specific inhibitor of the NF-kappa B transcription factor. Science. 1988;242:540-546.

Ballard DW. Molecular mechanisms in lymphocyte activation and growth. Immunol Res. 2001;23:157-166.

Ballaun C, Farrington GK, Dobrovnik M et al. 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. 1991;65:4408-4413.

Bangham CR. HTLV-1 infections. J Clin Pathol. 2000;53:581-586.

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

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

Barnard AL, Igakura T, Tanaka Y, Taylor GP, Bangham CR. Engagement of specific T cell surface molecules regulates cytoskeletal polarization in HTLV-1-infected lymphocytes. Blood. 2005.

Barnhart MK, Connor LM, Marriott SJ. Function of the human T-cell leukemia virus type 1 21-base-pair repeats in basal transcription. J Virol. 1997;71:337-344.

Barshira A, Panet A, Honigman A. An RNA Secondary Structure Juxtaposes Two Remote Genetic Signals for Human T-Cell Leukemia Virus Type 1 RNA 3'-End Processing. J Virol. 1991;65:5165-5173.

Bartoe JT, Albrecht B, Collins ND et al. Functional role of pX open reading frame II of human T- lymphotropic virus type 1 in maintenance of viral loads in vivo. J Virol. 2000;74:1094-1100.

Basbous J, Arpin C, Gaudray G et al. The HBZ factor of human T-cell leukemia virus type I dimerizes with transcription factors JunB and c-Jun and modulates their transcriptional activity. J Biol Chem. 2003;278:43620-43627.

191

Bazarbachi A, Hermine O. Treatment of adult T-cell leukaemia/lymphoma: current strategy and future perspectives. Virus Res. 2001;78:79-92.

Benovic S, Kok T, Stephenson A et al. De novo reverse transcription of HTLV-1 following cell-to- cell transmission of infection. Virology. 1998;244:294-301.

Beraud C, Lombard-Platet G, Michal Y, Jalinot P. Binding of the HTLV-I Tax1 transactivator to the inducible 21 bp enhancer is mediated by the cellular factor HEB1 5256. EMBO J. 1991;10:3795-3803.

Berneman ZN, Gartenhaus RB, Reitz MS et al. 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. 1992;89:3005-3009.

Bex F, Gaynor RB. Regulation of gene expression by HTLV-I Tax protein. Methods. 1998;16:83-94.

Bex F, Yin MJ, Burny A, Gaynor RB. 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. 1998;18:2392-2405.

Bindhu M, Nair A, Lairmore MD. Role of accessory proteins of HTLV-1 in viral replication, T cell activation, and cellular gene expression. Front Biosci. 2005;10:17-37.

Black AC, Chen ISY, Arrigo S et al. Regulation of HTLV-II gene expretion by rex involves positive and negative cis-Acting elements in the 5 long terminal repeat. Virology. 1991;181:433-444.

Black AC, Ruland C, Yip MT et al. 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. 1991;65:6645-6653.

Black AC, Ruland CT, Luo J et al. Binding of Nuclear Proteins to HTLV-II Cis-Acting Repressive Sequence (CRS) RNA Correlates with CRS Function. Virology. 1994;200:29-41.

Blattner W, Kalyanaraman V, Robert-Guroff M et al. The human type-C retrovirus, HTLV, in blacks from the Caribbean region and relationchip to adult T-cell leukemia/lymphoma. Int J Cancer. 1982;30:257.

Blayney DW, Blattner WA, Robert-Guroff M et al. The human T-cell leukemia-lymphoma virus in the southeastern United States 5419. JAMA. 1983;250:1048-1052.

192

Blot V, Perugi F, Gay B et al. Nedd4.1-mediated ubiquitination and subsequent recruitment of Tsg101 ensure HTLV-1 Gag trafficking towards the multivesicular body pathway prior to virus budding. J Cell Sci. 2004;117:2357-2367.

Bogerd HP, Echarri A, Ross TM, Cullen BR. 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. 1998;72:8627-8635.

Bogerd HP, Huckaby GL, Ahmed YF, Hanly SM, Greene WC. The type I human T-cell leukemia virus (HTLV-I) Rex trans-activator binds directly to the HTLV-I Rex and the type 1 human immunodeficiency virus Rev RNA response elements. Proc Natl Acad Sci U S A. 1991;88:5704-5708.

Bondzio A, Abraham-Podgornik A, Blankenstein P, Risse S. Involvement of intracellular Ca2+ in the regulation of bovine leukemia virus expression. Biol Chem. 2001;382:407-416.

Bosselut R, Duvall JF, Gegonne A et al. 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 5254. EMBO J. 1990;9:3137-3144.

Bosselut R, Lim F, Romond PC et al. 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. 1992;186:764-769.

Bouchard MJ, Wang LH, Schneider RJ. Calcium signaling by HBx protein in hepatitis B virus DNA replication. Science. 2001;294:2376-2378.

Bounou S, Giguere JF, Cantin R et al. The importance of virus-associated host ICAM-1 in human immunodeficiency virus type 1 dissemination depends on the cellular context. FASEB J. 2004;18:1294-1296.

Boustany LM, Cyert MS. Calcineurin-dependent regulation of Crz1p nuclear export requires Msn5p and a conserved calcineurin docking site. Genes Dev. 2002;16:608-619.

Boustany LM, Cyert MS. Calcineurin-dependent regulation of Crz1p nuclear export requires Msn5p and a conserved calcineurin docking site. Genes Dev. 2002;16:608-619.

Braakman I, van Anken E. Folding of viral envelope glycoproteins in the endoplasmic reticulum. Traffic. 2000;1:533-539.

193

Brady JN, Jeang KT, Duvall J, Khoury G. Identification of p40x-Responsive Regulatory Sequences within the Human T-Cell Leukemia Virus Type 1 Long Terminal Repeat. J Virol. 1987;61:2175-2181.

Brauweiler A, Garrus JE, Reed JC, Nyborg JK. Repression of Bax gene expression by the HTLV-I tax protein: Implications for suppression of apoptosis in virally infected cells. Virology. 1997;231:135-140.

Bukrinsky MI, Sharova N, Dempsey MP et al. Active nuclear import of human immunodeficiency virus type 1 preintegration complexes. Proc Natl Acad Sci U S A. 1992;89:6580-6584.

Burton M, Upadhyaya CD, Maier B, Hope TJ, Semmes OJ. Human T-cell leukemia virus type 1 tax shuttles between functionally discrete subcellular targets. J Virol. 2000;74:2351-2364.

Cann AJ, Chen ISY. Human T-Cell Leulemia Virus Types I and II. 1849-1880. 1996. Philadelphia, PA, Lippincott-Raven Publishers. Ref Type: Generic

Cann AJ, Rosenblatt JD, Wachsman W, Shah NP, Chen IS. Identification of the gene responsible for human T-cell leukaemia virus transcriptional regulation 5264. Nature. 1985;318:571-574.

Carman CV, Springer TA. Integrin avidity regulation: are changes in affinity and conformation underemphasized? Curr Opin Cell Biol. 2003;15:547-556.

Castillo LC, Gracia F, Roman GC et al. 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. 2000;101:405-412.

Cavrois M, Leclercq I, Gout O et al. 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. 1998;17:77-82.

Cereseto A, Berneman Z, Koralnik I et al. Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia. 1997;11:866-870.

Cereseto A, Giacca M. Integration site selection by retroviruses. AIDS Rev. 2004;6:13-21.

Cereseto A, Kislyakova T, Washington-Parks R, Franchini G. Differential response to genotoxic stress in immortalized or transformed HTLV-I-infected T-cells. J Gen Virol. 1999;80:1575-1581.

194

Chan JR, Hyduk SJ, Cybulsky MI. Chemoattractants induce a rapid and transient upregulation of monocyte alpha4 integrin affinity for vascular cell adhesion molecule 1 which mediates arrest: an early step in the process of emigration. J Exp Med. 2001;193:1149-1158.

Chen YA, Chen S, Fu C, Chen J, Osame M. 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. 1997;71:196-202.

Chou KS, Okayama A, Tachibana N, Lee TH, Essex M. 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. 1995;60:701-706.

Chrivia JC, Kwok RP, Lamb N et al. Phosphorylated CREB binds specifically to the nuclear protein CBP. Nature. 1993;365(6449):855-859.

Ciminale V, D'Agostino D, Zotti L et al. Expression and characterization of proteins produced by mRNAs spliced into the X region of the human T-cell leukemia/lymphotropic virus type II. Virology. 1995;209:445-456.

Ciminale V, D'Agostino DM, Zotti L, Chieco-Bianchi L. Coding potential of the X region of human T-cell leukemia/lymphotropic virus type II. J Acquir Immune Defic Syndr Hum Retrovirol. 1996;13 Suppl 1:S220-S227.

Ciminale V, Pavlakis GN, Derse D, Cunningham CP, Felber BK. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J Virol. 1992;66:1737-1745.

Ciminale V, Zotti L, Dagostino DM et al. 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. 1999;18:4505-4514.

Clark NM, Smith MJ, Hilfinger JM, Markovitz DM. 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. 1993;67:5522-5528.

Coffin JM. Structure and Classification of Retroviruses. The Retroviridae. 1992;1:19-49.

Coffin JM. Retroviridae: The Viruses and Their Replication. 1767-1835. 1996. Phildaelphia, PA, Lippinocott-Raven Publishers. Ref Type: Generic

Colgin MA, Nyborg JK. 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. 1998;72:9396-9399.

195

Collins ND, D'Souza C, Albrecht B et al. 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. 1999;73:9642-9649.

Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12(I) reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

Collins ND, Newbound GC, Albrecht B et al. Selective ablation of human T-cell lymphotropic virus type 1 p12I reduces viral infectivity in vivo. Blood. 1998;91:4701-4707.

Collins ND, Newbound GC, Ratner L, Lairmore MD. In vitro CD4(+) lymphocyte transformation and infection in a rabbit model with a molecular clone of human T-cell lymphotropic virus type 1. J Virol. 1996;70:7241-7246.

Constantin G, Majeed M, Giagulli C et al. Chemokines trigger immediate beta2 integrin affinity and mobility changes: differential regulation and roles in lymphocyte arrest under flow. Immunity. 2000;13:759-769.

Courouce AM, Pillonel J, Lemaire JM, Saura C. HTLV testing in blood transfusion. Vox Sang. 1998;74 Suppl 2:165-169.

Cowan EP, Alexander RK, Daniel S, Kashanchi F, Brady JN. Induction of tumor necrosis factor alpha in human neuronal cells by extracellular human T-cell lymphotropic virus type 1 Tax(1). J Virol. 1997;71:6982-6989.

Craig HM, Reddy TR, Riggs NL, Dao PP, Guatelli JC. 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. 2000;271:9-17.

Craven RC, Parent LJ. Dynamic interactions of the gag polyprotein. In: Krausslich HG, ed. Morphogenesis and Maturation of Retroviruses. Springer-Verlag Berl; 1996:65-94.

D'Agostino DM, Ciminale V, Zotti L, Rosato A, Chieco-Bianchi L. 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. 1997;71:75-83.

D'Agostino DM, Ranzato L, Arrigoni G et al. Mitochondrial alterations induced by the p13II protein of human T-cell leukemia virus type 1. Critical role of arginine residues. J Biol Chem. 2002;277:34424-34433.

D'Agostino DM, Silic-Benussi M, Hiraragi H, Lairmore MD, Ciminale V. The human T-cell leukemia virus type 1 p13(II) protein: effects on mitochondrial function and cell growth. Cell Death Differ. 2005.

196

D'Agostino DM, Zotti L, Ferro T et al. Expression and functional properties of proteins encoded in the x-II ORF of HTLV-I. Virus Res. 2001;78:35-43.

D'Agostino DM, Zotti L, Ferro T et al. The p13II protein of HTLV type 1: comparison with mitochondrial proteins coded by other human viruses. AIDS Res Hum Retroviruses. 2000;16:1765-1770.

D'Souza-Schorey C, Boettner B, Van AL. Rac regulates integrin-mediated spreading and increased adhesion of T lymphocytes. Molecular & Cellular Biology. 1998;18:3936-3946.

Daenke S, Mccracken SA, Booth S. Human T-cell leukaemia lymphoma virus type 1 syncytium formation is regulated in a cell-specific manner by ICAM-1, ICAM-3 and VCAM-1 and can be inhibited by antibodies to integrin beta(2) or beta(7). J Gen Virol. 1999;80:1429-1436.

Daenke S, Nightingale S, Cruickshank JK, Bangham CRM. 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. 1990;64:1278-1282.

Dekaban GA, Peters AA, Mulloy JC et al. The HTLV-I orfI protein is recognized by serum antibodies from naturally infected humans and experimentally infected rabbits. Virology. 2000;274:86-93.

Delamarre L, Rosenberg AR, Pique C, Pham D, Dokhelar MC. A novel human T-Leukemia virus type 1 cell-to-cell transmission assay permits definition of SU glycoprotein amino acids important for infectivity. J Virol. 1997;71:259-266.

Derse D. Bovine leukemia virus transcription is controlled by a virus-encoded trans-acting factor and by cis-acting response elements. J Virol. 1987;61:2462-2471.

Derse D, Mikovits J, Ruscetti F. 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. 1997;237:123-128.

Ding W, Albrecht B., Kelley RE et al. 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. 2002;76:10374-10382.

Ding W, Albrecht B, Luo R et al. Endoplasmic Reticulum and cis-Golgi Localization of Human T- Lymphotropic Virus Type 1 p12(I): Association with Calreticulin and Calnexin. J Virol. 2001;75:7672-7682.

197

Ding W, Kim SJ, Nair AM et al. Human T-Cell Lymphotropic Virus Type 1 p12(I) Enhances Interleukin-2 Production during T-Cell Activation. J Virol. 2003;77:11027-11039.

Dustin M. Viral spread through protoplasmic kiss. Nat Cell Biol. 2003;5:271-272.

Dustin ML, Bivona TG, Philips MR. Membranes as messengers in T cell adhesion signaling. Nat Immunol. 2004;5:363-372.

Duyao MP, Kessler DJ, Spicer DB et al. Transactivation of the c-myc Promoter by Human T Cell Leukemia virus Type 1 tax is Mediated by NFkB. J Biol Chem. 1992;267:16288-16291.

Eckner R, Yao TP, Oldread E, Livingston DM. Interaction and functional collaboration of p300/CBP and bHLH proteins in muscle and B-cell differentiation. Genes Dev. 1996;10:2478-2490.

Edlich RF, Arnette JA, Williams FM. Global epidemic of human T-cell lymphotropic virus type-I (HTLV-I). J Emerg Med. 2000;18:109-119.

Evangelista A, Maroushek S, Minnigan H et al. Nucleotide sequence analysis of a provirus derived from an individual with tropical spastic paraparesis. Microb Pathog. 1990;8:259-278.

Fagerholm S, Morrice N, Gahmberg CG, Cohen P. Phosphorylation of the cytoplasmic domain of the integrin CD18 chain by protein kinase C isoforms in leukocytes. J Biol Chem. 2002;277:1728-1738.

Fan N, Gavalchin J, Paul B et al. Infection of Peripheral Blood Mononuclear Cells and Cell Lines by Cell-Free Human T-Cell Lymphoma/Leukemia Virus Type 1. J Clin Microbiol. 1992;30:905-910.

Feigelson SW, Grabovsky V, Winter E et al. The Src kinase p56(lck) up-regulates VLA-4 integrin affinity. Implications for rapid spontaneous and chemokine-triggered T cell adhesion to VCAM-1 and fibronectin. J Biol Chem. 2001;276:13891-13901.

Felber BK, Paskalis H, Kleinmanewing C. The pX Protein of HTLV-1 is a Transcriptional Activator of its Long Terminal Repeats. Science. 1985;229:675-679.

Feng P, Park J, Lee BS et al. Kaposi's sarcoma-associated herpesvirus mitochondrial K7 protein targets a cellular calcium-modulating cyclophilin ligand to modulate intracellular calcium concentration and inhibit apoptosis. J Virol. 2002;76:11491-11504.

Ferreira OC, Planelles V, Rosenblatt JD. Human T-cell leukemia viruses: Epidemiology, biology, and pathogenesis. Blood Rev. 1997;11:91-104.

198

Fortin JF, Barbeau B, Hedman H, Lundgren E, Tremblay MJ. Role of the leukocyte function antigen-1 conformational state in the process of human immunodeficiency virus type 1-mediated syncytium formation and virus infection. Virology. 1999;257:228-238.

Fortin JF, Cantin R, Tremblay MJ. T cells expressing activated LFA-1 are more susceptible to infection with human immunodeficiency virus type 1 particles bearing host-encoded ICAM-1. J Virol. 1998;72:2105-2112.

Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

Foyouzi-Youssefi R, Arnaudeau S, Borner C et al. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2000;97:5723-5728.

Franchini G. Molecular mechanisms of human T-cell leukemia/lymphotropic virus type I infection. Blood. 1995;86:3619-3639.

Franchini G, Fukumoto R, Fullen JR. T-cell control by human T-cell leukemia/lymphoma virus type 1. Int J Hematol. 2003;78:280-296.

Franchini G, Mulloy JC, Koralnik IJ et al. 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. Journal of Virology. 1993;67:7701-7704.

Franchini G, Nicot C, Johnson JM. Seizing of T cells by human T-cell leukemia/lymphoma virus type 1. Adv Cancer Res. 2003;89:69-132.

Franklin AA, Kubik MF, Uittenbogaard MN et al. Transactivation by the human T-cell leukemia virus Tax protein is mediated through enhanced binding of activating

199

transcription factor-2 (ATF-2) ATF-2 response and cAMP element-binding protein (CREB). J Biol Chem. 1993;268:21225-21231.

Fujii M, Sassone-Corsi P, Verma I. c-fos promotor trans-activation of cAMP on IL-2 stimulated gene expression. Proc Natl Acad Sci U S A. 1988;8526-8530.

Fujii M, Tsuchiya H, Chuhjo T, Akizawa T, Seiki M. Interaction of HTLV-1 Tax1 with p67SRF causes the aberrant induction of cellular immediate early genes through CArG boxes 5270. Genes Dev. 1992;6:2066-2076.

Fujino T, Fujiyoshi T, Yashiki S et al. HTLV-I transmission from mother to fetus via placenta. Lancet. 1992;340:1157. Fujisawa J, Seiki M, Kiyokawa T, Yoshida M. 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. 1985;82:2277-2281.

Fukudome K, Furuse M, Fukuhara N et al. Strong induction of ICAM-1 in human T cells transformed by human T-cell-leukemia virus type 1 and depression of ICAM-1 or LFA-1 in adult T-cell-leukemia-derived cell lines. Int J Cancer. 1992;52:418-427.

Furukawa K, Furukawa K, Shiku H. Alternatively spliced mRNA of the pX region of human T lymphotropic virus type I proviral genome. FEBS Lett. 1991;295:141-145.

Furukawa Y, Kubota R, Eiraku N et al. 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. 2003;32:328-334.

Ganem D. Virology. The X files--one step closer to closure. Science. 2001;294:2299-2300.

Garcia-Cozar FJ, Okamura H, Aramburu JF et al. Two-site interaction of nuclear factor of activated T cells with activated calcineurin. J Biol Chem. 1998;273:23877-23883.

Garnier L, Bowzard JB, Wills JW. Recent advances and remaining problems in HIV assembly. AIDS. 1998;12 Suppl A:S5-16.

Garry RF. Sequence similarities between latent membrane protein LMP- 1 of Epstein-Barr virus, integral membrane protein p12(I) of human T cell leukemia lymphotropic virus type 1, E5 transformation protein of bovine papillomavirus, and the transmembrane proteins of slowly transforming retroviruses. AIDS Res Hum Retroviruses. 1995;11:431-432.

Gatza ML, Watt JC, Marriott SJ. Cellular transformation by the HTLV-I Tax protein, a jack-of-all-trades. Oncogene. 2003;22:5141-5149.

200

Gaudray G, Gachon F, Basbous J et al. The complementary strand of the human T-cell leukemia virus type 1 RNA genome encodes a bZIP transcription factor that down-regulates viral transcription. J Virol. 2002;76:12813-12822.

Gerritsen ME, Williams AJ, Neish AS et al. CREB-binding protein/p300 are transcriptional coactivators of p65. Proc Natl Acad Sci U S A. 1997;94:2927-2932.

Gessain A, Barin F, Vernant J et al. Antibodies to human T lymphotropic virus type 1 in patients with tropical spastic paresis. Lancet. 1985;2:407-410.

Gessain A, Mahieux R. [A virus called HTLV-1. Epidemiological aspects] Presse Med. 2000;29:2233-2239.

Giguere JF, Tremblay MJ. Statin compounds reduce human immunodeficiency virus type 1 replication by preventing the interaction between virion-associated host intercellular adhesion molecule 1 and its natural cell surface ligand LFA-1. J Virol. 2004;78:12062-12065.

Gould KG, Bangham CR. Virus variation, escape from cytotoxic T lymphocytes and human retroviral persistence. Semin Cell Dev Biol. 1998;9:321-328.

Grant C, Barmak K, Alefantis T et al. 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. 2002;190:133-159.

Grassmann R, Berchtold S, Aepinus C et al. In Vitro Binding of Human T-Cell Leukemia Virus rex Proteins to the rex-Response Element of Viral Transcripts. J Virol. 1991;65:3721-3727.

Green PL, Chen ISY. Regulation of human T cell leukemia virus expression. FASEB J. 1990;4:169-175.

Green PL, Chen ISY. Molecular features of the human T-cell leukemia virus. Mechanisms of transformation and leukemogenicity. In: Levy JA, ed. The Retroviridae.3. New York: Plenum Press; 1994:277-311.

Griffith JP, Kim JL, Kim EE et al. X-ray structure of calcineurin inhibited by the immunophilin-immunosuppressant FKBP12-FK506 complex. Cell. 1995;82:507-522.

Griffiths EK, Krawczyk C, Kong YY et al. Positive regulation of T cell activation and integrin adhesion by the adapter Fyb/Slap. Science. 2001;293:2260-2263.

Grone M, Koch C, Grassmann R. The HTLV-1 Rex protein induces nuclear accumulation of unspliced viral RNA by avoiding intron excision and degradation. Virology. 1996;218:316-325.

201

Guyot DJ, Newbound GC, Lairmore MD. Signaling via the CD2 receptor enhances HTLV-1 replication in T lymphocytes. Virology. 1997;234:123-129.

Hai TW, Liu F, Coukos WJ, Green MR. Transcription factor ATF cDNA clones: an extensive family of leucine zipper proteins able to selectively form DNA-binding heterodimers. Genes Dev. 1989;3:2083-2090.

Hakata Y, Umemoto T, Matsushita S, Shida H. 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. 1998;72:6602-6607.

Hall WW, Fujii M. Deregulation of cell-signaling pathways in HTLV-1 infection. Oncogene. 2005;24:5965-5975.

Hanon E, Hall S, Taylor GP et al. Abundant Tax protein expression in CD4+T cells infected with human T-cell lymphotropic virus type I (HTLV-I) is prevented by cytotoxic T lymphocytes. Blood. 2000;95:1386-1392.

Harrod R, Kuo YL, Tang Y et al. 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. 2000;275:11852-11857.

Hattori S, Kiyokawa T, Imagawa K et al. Identification of gag and env gene products of human T-cell leukemia virus (HTLV). Virology. 1984;136:338-347.

He H, Lam M, McCormick TS, Distelhorst CW. Maintenance of calcium homeostasis in the endoplasmic reticulum by Bcl-2. J Cell Biol. 1997;138:1219-1228.

Hildreth JEK, Subramanium A, Hampton RA. Human T-cell lymphotropic virus type 1 (HTLV-1)-induced syncytium formation mediated by vascular cell adhesion molecule-1: Evidence for involvement of cell adhesion molecules in HTLV-1 biology. J Virol. 1997;71:1173-1180.

Hino S. Milk-borne transmission of HTLV-I as a major route in the endemic cycle. Acta Paediatr Jpn. 1989;31:428-435. Hino S, Yamaguchi K, Katamine S et al. Mother-to-child transmission of human T-cell leukemia virus type-I. Jpn J Cancer Res. 1985;76:474-480. Hinuma Y, Komoda H, Chosa T et al. 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. 1982;29:631.

202

Hinuma Y, Nagata K, Hanaoka M et al. 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. 1981;78:6476-6480.

Hirai H, Suzuki T, Fujisawa J, Inoue J, Yoshida M. Tax Protein of Human T-Cell Leukemia Virus Type I Binds to the Ankyrin Motifs of Inhibitory Factor kappa b and Induces Nuclear Translocation of Transcription Factor NF- kappa b Proteins for Transcriptional Activation. Proc Natl Acad Sci USA. 1994;91:3584-3588.

Hiraragi H, Kim SJ, Phipps AJ et al. Human T-Lymphotropic Virus Type 1 Mitochondrion-Localizing Protein p13II Is Required for Viral Infectivity In Vivo. J Virol. 2006;80:3469-3476.

Hiraragi H, Michael B, Nair A et al. Human T-lymphotropic virus type 1 mitochondrion-localizing protein p13II sensitizes Jurkat T cells to Ras-mediated apoptosis. J Virol. 2005;79:9449-9457.

Hivin P, Frederic M, rpin-Andre C et al. Nuclear localization of HTLV-I bZIP factor (HBZ) is mediated by three distinct motifs. J Cell Sci. 2005;118:1355-1362.

Hogg N, Henderson R, Leitinger B et al. Mechanisms contributing to the activity of integrins on leukocytes. Immunol Rev. 2002;186:164-171.

Hogg N, Laschinger M, Giles K, Mcdowall A. T-cell integrins: more than just sticking points. J Cell Sci. 2003;116:4695-4705.

Hogg N, Smith A, Mcdowall A et al. How T cells use LFA-1 to attach and migrate. Immunol Lett. 2004;92:51-54.

Holloway MP, Bram RJ. A hydrophobic domain of Ca2+-modulating cyclophilin ligand modulates calcium influx signaling in T lymphocytes. J Biol Chem. 1996;271:8549-8552.

Holloway MP, Bram RJ. Co-localization of calcium-modulating cyclophilin ligand with intracellular calcium pools. J Biol Chem. 1998;273:16346-16350.

Hollsberg P. Mechanisms of T-cell activation by human T-cell lymphotropic virus type I. Microbiol Mol Biol Rev. 1999;63:308-333.

Hollsberg P, Hafler D. Pathogenesis of diseases induced by human lymphotropic virus type I infection. N Engl J Med. 1993;328:1173-1182.

Hollsberg P, Wucherpfennig KW, Ausubel LJ et al. Characterization of HTLV-1 In Vivo Infected T Cell Clones IL-2-independent Growth of Nontransformed T Cells. J Immunol. 1992;148:3256-3263.

203

Hou X, Foley S, Cueto M, Robinson MA. The Human T-Cell Leukemia Virus Type I (HTLV-I) X Region Encoded Protein p13(II) Interacts with Cellular Proteins. Virology. 2000;277:127-135.

Hummel KB, Bellini WJ, Offermann MK. Strain-specific differences in LFA-1 induction on measles virus-infected monocytes and adhesion and viral transmission to endothelial cells. J Virol. 1998;72:8403-8407.

Igakura T, Stinchcombe JC, Goon PK et al. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science. 2003;299:1713-1716.

Imai T, Tanaka Y, Fukudome K et al. Enhanced expression of LFA-3 on human T-cell lines and leukemic. International Journal of Cancer. 1993;55(5):811-816.

Iniguez AM, Otsuki K, Magalhaes GP, Silva EA, Vicente AC. Genetic markers on the HTLV-1 p12I protein sequences from Brazilian HAM/TSP patients and asymptomatic HTLV-1 carrier isolates. AIDS Res Hum Retroviruses. 2005;21:580-582.

Inoue J, Seiki M, Taniguchi T, Tsuru S, Yoshida M. Induction of interleukin 2 receptor gene expression by p40x encoded by human T-cell leukemia virus type 1. EMBO J. 1986;5:2883-2888.

Inoue J, Seiki M, Yoshida M. The second pX product p27 chi-III of HTLV-1 is required for gag gene expression 5294. FEBS Lett. 1986;209:187-190.

Inoue J, Yoshida M, Seiki M. 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. 1987;84:3653-3657.

Itoyama Y, Minato S, Kira J et al. Spontaneous proliferation of peripheral blood lymphocytes increased in patients with HTLV-I-associated myelopathy. Neurology. 1988;38:1302-1307.

Jacobson S. Immunopathogenesis of human T cell lymphotropic virus type I-associated neurologic disease. J Infect Dis. 2002;186 Suppl 2:S187-S192.

Jacobson S, Zaninovic V, Mora C et al. Immunological findings in neurological diseases associated with antibodies to HTLV-I: activated lymphocytes in tropical spastic paraparesis. Ann Neurol. 1988;23 Suppl:S196-S200.

Jassal SR, Lairmore MD, Leigh-Brown AJ, Brighty DW. Soluble recombinant HTLV-1 surface glycoprotein competitively inhibits syncytia formation and viral infection of cells. Virus Res. 2001;78:17-34.

204

Jeang KT. Functional activities of the human T-cell leukemia virus type I Tax oncoprotein: cellular signaling through NF-kappa B. Cytokine Growth Factor Rev. 2001;12:207-217.

Jeang KT, Chiu R, Santos E, Kim SJ. Induction of the HTLV-I LTR by Jun occurs through the Tax-responsive 21- bp elements 5255. Virology. 1991;181:218-227.

Jeang KT, Giam CZ, Majone F, Aboud M. Life, death, and tax: role of HTLV-I oncoprotein in genetic instability and cellular transformation. J Biol Chem. 2004;279:31991-31994.

Jiang H, Lu H, Schiltz RL et al. 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. 1999;19:8136-8145.

Jin DY, Spencer F, Jeang KT. Human T cell leukemia virus type 1 oncoprotein Tax targets the human mitotic checkpoint protein MAD1. Cell. 1998;93:81-91.

Johnson JM, Franchini G. Retroviral proteins that target the major histocompatibility complex class I. Virus Res. 2002;88:119-127.

Johnson JM, Fullen J, Casareto L, Mulloy JC. p12(I) May Contribute TO HTLV-1 Escape from Immune Detection 5313. Leukemia. 2000;14:558.

Johnson JM, Mulloy JC, Ciminale V et al. 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. 2000;16:1777-1781.

Johnson JM, Nicot C, Fullen J et al. 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. 2001;75:6086-6094.

Johnson S, Michalak M, Opas M, Eggleton P. The ins and outs of calreticulin: from the ER lumen to the extracellular space. Trends Cell Biol. 2001;11:122-129.

Kamb A, Gruis NA, Weaver-Feldhaus J et al. A cell cycle regulator potentially involved in genesis of many tumor types. Science. 1994;264:436-440.

Kamihira S, Sohda H, Atogami S et al. Phenotypic diversity and prognosis of adult T-cell leukemia. Leuk Res. 1992;16:435-441.

Kasahata N, Shiota J, Miyazawa Y, Nakano I, Murayama S. Acute human T-lymphotropic virus type 1-associated myelopathy: a clinicopathologic study. Arch Neurol. 2003;60:873-876.

205

Kashanchi F, Brady JN. Transcriptional and post-transcriptional gene regulation of HTLV-1. Oncogene. 2005;24:5938-5951.

Kashishian A, Howard M, Loh C et al. AKAP79 inhibits calcineurin through a site distinct from the immunophilin-binding region. J Biol Chem. 1998;273:27412-27419.

Katagiri K, Hattori M, Minato N et al. Rap1 is a potent activation signal for leukocyte function-associated antigen 1 distinct from protein kinase C and phosphatidylinositol-3-OH kinase. Mol Cell Biol. 2000;20:1956-1969.

Kawakami K, Miyazato A, Iwakura Y, Saito A. Induction of lymphocytic inflammatory changes in lung interstitium by human T lymphotropic virus type I. AMER J RESPIR CRIT CARE MED. 1999;160:995-1000.

Kawano F, Yamaguchi K, Nishimura H, Tsuda H, Takatsuki K. Variation in the clinical courses of adult T-cell leukemia. Cancer. 1985;55:851-856.

Kehn K, Berro R, de La FC et al. Mechanisms of HTLV-1 transformation. Front Biosci. 2004;9:2347-2372.

Khabbaz RF, Douglas JM, Judson FN et al. Seroprevalence of human T-lymphotropic virus type I or II in sexually transmitted disease clinic patients in the USA. J Infect Dis. 1990;162:241-244.

Khabbaz RF, Onorato IM, Cannon RO et al. Seroprevalence of HTLV-1 and HTLV-2 among intravenous drug users and persons in clinics for sexually transmitted diseases. N Engl J Med. 1992;326:375-380.

Kim JH, Kaufman PA, Hanly SM, Rimsky LT, Greene WC. Rex transregulation of human T-cell leukemia virus type II gene expression. J Virol. 1991;65:405-414.

Kim M, Carman CV, Springer TA. Bidirectional transmembrane signaling by cytoplasmic domain separation in integrins. Science. 2003;301:1720-1725.

Kim SJ, Ding W, Albrecht B, Green PL, Lairmore MD. 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. 2003;278:15550-15557.

Kim SJ, Nair AM, Fernandez S, Mathes L, Lairmore MD. Enhancement of LFA-1-mediated T cell adhesion by human T lymphotropic virus type 1 p12I1. J Immunol. 2006;176:5463-5470.

Kimata JT, Palker TJ, Ratner L. The Mitogenic Activity of Human T-Cell Leukemia Virus Type- I Is T-Cell Associated and Requires the CD2/LFA-3 Activation Pathway. J Virol. 1993;67:3134-3141.

206

Kimata JT, Wong F, Wang J, Ratner L. Construction and characterization of infectious human T-cell leukemia virus type 1 molecular clones. Virology. 1994;204:656-664.

Kimura I, Tsubota T, Tada S, Sogawa J. Presence of antibodies against adult T cell leukemia antigen in the patients with chronic respiratory diseases 5241. Acta Med Okayama. 1986;40:281-284.

Kimzey AL, Dynan WS. Specific regions of contact between human T-cell leukemia virus type I Tax protein and DNA identified by photocross- linking. J Biol Chem. 1998;273:13768-13775.

King JA, Bridger JM, Lochelt M et al. Nucleocytoplasmic transport of HTLV-1 RNA is regulated by two independent LTR encoded nuclear retention elements. Oncogene. 1998;16:3309-3316.

Kinoshita S, Chen BK, Kaneshima H, Nolan GP. Host control of HIV-1 parasitism in T cells by the nuclear factor of activated T cells. Cell. 1998;95:595-604.

Kitado H, Chen IS, Shah NP et al. U3 sequences from HTLV-I and -II LTRs confer pX protein response to a murine leukemia virus LTR 5250. Science. 1987;235:901-904.

Kitze B, Usuku K. HTLV-1-mediated immunopathological CNS disease. Curr Top Microbiol Immunol. 2002;265:197-211.

Kiwaki T, Umehara F, Arimura Y et al. The clinical and pathological features of peripheral neuropathy accompanied with HTLV-I associated myelopathy. J Neurol Sci. 2003;206:17-21.

Kiyokawa T, Seiki M, Iwashita S et al. p27x-III and p21x-III, proteins encoded by the pX sequence of human T-cell leukemia virus type 1. Proc Natl Acad Sci USA. 1985;82:8359-8363.

Koralnik I, Mulloy JC, Andresson T, Fullen J, Franchini G. Mapping of the intermolecular association of the human T-cell leukemia/lymphotropic virus type 1 p12I and the vacuolar H+ATPase 16 kDa subunit protein. J Gen Virol. 1995;76:1909-1916.

Koralnik IJ, Fullen J, Franchini G. 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. 1993;67:2360-2366.

Koralnik IJ, Fullen J, Franchini G. 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. 1993;67:2360-2366.

207

Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type 1. Proc Natl Acad Sci USA. 1992;89:8813-8817.

Koralnik IJ, Gessain A, Klotman ME et al. Protein isoforms encoded by the pX region of human T-cell leukemia/lymphotropic virus type I. Proc Natl Acad Sci U S A. 1992;89:8813-8817.

Koralnik IJ, Mulloy JC, Andresson T, Fullen J, Franchini G. Mapping of the intermolecular association of the human T cell leukaemia/lymphotropic virus type I p12(I) and the vacuolar H+- ATPase 16 kDa subunit protein. J Gen Virol. 1995;76:1909-1916.

Kotov A, Zhou J, Flicker P, Aiken C. Association of Nef with the human immunodeficiency virus type 1 core. J Virol. 1999;73:8824-8830.

Krause KH, Michalak M. Calreticulin. Cell. 1997;88:439-443.

Kubota S, Hatanaka M, Pomerantz RJ. Nucleo-cytoplasmic redistribution of the HTLV-I rex protein: Alterations by coexpression of the HTLV-I p21(x) protein. Virology. 1996;220:502-507.

Kucik DF, Dustin ML, Miller JM, Brown EJ. Adhesion-activating phorbol ester increases the mobility of leukocyte integrin LFA-1 in cultured lymphocytes. J Clin Invest. 1996;97:2139-2144.

Kuhne MR, Lin J, Yablonski D et al. Linker for activation of T cells, zeta-associated protein-70, and Src homology 2 domain-containing leukocyte protein-76 are required for TCR-induced microtubule-organizing center polarization. J Immunol. 2003;171:860-866.

Kulkarni AB, Huh CG, Becker D et al. Transforming growth factor beta 1 null mutation in mice causes excessive inflammatory response and early death. Proc Natl Acad Sci U S A. 1993;90:770-774.

Kuroda Y, Matsui M. Cerebrospinal Fluid Interferon-Gamma Is Increased in HTLV- I- Associated Myelopathy. J Neuroimmunol. 1993;42:223-226.

Kwok RP, Lundblad JR, Chrivia JC et al. Nuclear protein CBP is a coactivator for the transcription factor. Nature. 1994;370(6486):223-226.

Kwok RPS, Laurance ME, Lundblad JR et al. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature. 1996;380:642-646.

208

Lagrenade L, Hanchard B, Fletcher V, Cranston B, Blattner W. Infective dermatitis of Jamaican children: a marker for HTLV-I infection 5233. Lancet. 1990;336:1345-1347.

Laherty CD, Perkins ND, Dixit VM. Human T-Cell Leukemia Virus Type-I Tax and Phorbol 12- Myristate 13-Acetate Induce Expression of the A20 Zinc Finger Protein by Distinct Mechanisms Involving Nuclear Factor-kappaB. J Biol Chem. 1993;268:5032-5039.

Lai MM, Burnett PE, Wolosker H, Blackshaw S, Snyder SH. Cain, a novel physiologic protein inhibitor of calcineurin. J Biol Chem. 1998;273:18325-18331.

Lairmore MD, Albrecht B, D'Souza C et al. 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. 2000;16:1757-1764.

Lairmore MD, Lal R. Other Human Retrovirus Infections: HTLV-I and HTLV-II. In: Schocketman G, George JR, eds. AIDS Testing. Methodology and Management Issues. New York: Springer-Verlag; 1995:168-188.

Lantz RC, Lemus R, Lange RW, Karol MH. Rapid reduction of intracellular glutathione in human bronchial epithelial cells exposed to occupational levels of toluene diisocyanate. Toxicol Sci. 2001;60:348-355.

Larocca D, Chao LA, Seto MH, Brunck TK. Human T-cell leukemia virus minus strand transcription in infected T-cells. Biochem Biophys Res Commun. 1989;163:1006-1013.

Le B, I, Grange MP, Delamarre L et al. HTLV-1 structural proteins. Virus Res. 2001;78:5-16.

Le B, I, Prevost MC, Dokhelar MC, Rosenberg AR. The PPPY motif of human T-cell leukemia virus type 1 Gag protein is required early in the budding process. J Virol. 2002;76:10024-10029.

Lefebvre L, Ciminale V, Vanderplasschen A et al. Subcellular localization of the bovine leukemia virus R3 and G4 accessory proteins. J Virol. 2002;76:7843-7854.

Lefebvre L, Ciminale V, Vanderplasschen A et al. Subcellular localization of the bovine leukemia virus R3 and G4 accessory proteins. J Virol. 2002;76:7843-7854.

Lefebvre L, Vanderplasschen A, Ciminale V et al. Oncoviral Bovine Leukemia Virus G4 and Human T-Cell Leukemia Virus Type 1 p13(II) Accessory Proteins Interact with Farnesyl Pyrophosphate Synthetase. J Virol. 2002;76:1400-1414.

209

Lemasson I, Polakowski NJ, Laybourn PJ, Nyborg JK. Transcription factor binding and histone modifications on the integrated proviral promoter in human T-cell leukemia virus-I-infected T-cells. J Biol Chem. 2002;277:49459-49465.

Lenzmeier BA, Baird EE, Dervan PB, Nyborg JK. The Tax protein-DNA interaction is essential for HTLV-I transactivation in vitro. J Mol Biol. 1999;291:731-744.

Leonmonzon M, Illa I, Dalakas MC. 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. 1994;36:643-649.

Levin MC, Jacobson S. HTLV-I associated myelopathy/tropical spastic paraparesis (HAM/TSP): A chronic progressive neurologic disease associated with immunologically mediated damage to the central nervous system. J Neurovirology. 1997;3:126-140.

Levin MC, Lee SM, Kalume F et al. Autoimmunity due to molecular mimicry as a cause of neurological disease. Nat Med. 2002;8:509-513.

Levin MC, Lee SM, Morcos Y, Brady J, Stuart J. Cross-reactivity between immunodominant human T lymphotropic virus type I tax and neurons: implications for molecular mimicry. J Infect Dis. 2002;186:1514-1517.

Li CC, Ruscetti FW, Rice NR et al. 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. 1993;67:4205-4213.

Li XH, Murphy KM, Palka KT, Surabhi RM, Gaynor RB. The human T-cell leukemia virus type-1 Tax protein regulates the activity of the I kappa B kinase complex. J Biol Chem. 1999;274:34417-34424.

Lin X, Sikkink RA, Rusnak F, Barber DL. Inhibition of calcineurin phosphatase activity by a calcineurin B homologous protein. J Biol Chem. 1999;274:36125-36131.

Lindholm PF, Reid RL, Brady JN. Extracellular Tax Protein Stimulates Tumor Necrosis Factor-B and Immunoglobulin Kappa Light Chain Expression in Lymphoid Cells. J Virol. 1992;66:1294-1302.

Loh C, Shaw KT, Carew J et al. Calcineurin binds the transcription factor NFAT1 and reversibly regulates its activity. J Biol Chem. 1996;271:10884-10891.

Lundblad JR, Kwok RP, Laurance ME, Harter ML, Goodman RH. Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature. 1995;374:85-88.

210

Lundblad JR, Kwok RP, Laurance ME et al. 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. 1998;273:19251-19259.

Maggirwar SB, Harhaj E, Sun SC. Activation of NF-kappa B/Rel by Tax involves degradation of I kappa B alpha and is blocked by a proteasome inhibitor. Oncogene. 1995;11:993-998.

Mahieux R, Gessain A. HTLV-1 and associated adult T-cell leukemia/lymphoma. Rev Clin Exp Hematol. 2003;7:336-361.

Mahieux R, Ibrahim F, Mauclere P et al. 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. 1997;71:1317-1333.

Manel N, Battini JL, Taylor N, Sitbon M. HTLV-1 tropism and envelope receptor. Oncogene. 2005;24:6016-6025.

Manel N, Kim FJ, Kinet S et al. The ubiquitous glucose transporter GLUT-1 is a receptor for HTLV. Cell. 2003;115:449-459.

Manel N, Kinet S, Battini JL et al. The HTLV receptor is an early T-cell activation marker whose expression requires de novo protein synthesis. Blood. 2003;101:1913-1918.

Manninen A, Saksela K. HIV-1 Nef interacts with inositol trisphosphate receptor to activate calcium signaling in T cells. J Exp Med. 2002;195:1023-1032.

Manns A, Hisada M, Lagrenade L. Human T-lymphotropic virus type I infection. Lancet. 1999;353:1951-1958.

Manns A, Wilks RJ, Murphy EL et al. A prospective study of transmission by transfusion of HTLV-I and risk factors associated with seroconversion. Int J Cancer. 1992;51:886-891.

Marriott SJ, Lindholm PF, Brown KM et al. 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. 1990;10:4192-4201.

Martins ML, Soares BC, Ribas JG et al. 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. 2002;18:899-902.

211

Maruyama M, Shibuya H, Harada H et al. Evidence for aberrant activation of the interleukin-2 autocrine loop by HTLV-1-encoded p40x and T3/Ti complex triggering 5267. Cell. 1987;48:343-350.

Masuda ES, Imamura R, Amasaki Y, Arai K, Arai N. Signalling into the T-cell nucleus: NFAT regulation. Cell Signal. 1998;10:599-611.

Matsui M, Nagumo F, Tadano J, Kuroda Y. Characterization of humoral and cellular immunity in the central nervous system of HAM/TSP. J Neurol Sci. 1995;130:183-189.

Matsumoto J, Ohshima T, Isono O, Shimotohno K. HTLV-1 HBZ suppresses AP-1 activity by impairing both the DNA-binding ability and the stability of c-Jun protein. Oncogene. 2005;24:1001-1010.

Matsumoto M, Sugimoto M, Nakashima H et al. Spontaneous T cell proliferation and release of soluble interleukin-2 receptors in patients with HTLV-I-associated myelopathy. Am J Trop Med Hyg. 1990;42:365-373.

Merino F, Robert-Guroff M, Clark J et al. Natural antibodies to human T-cell leukemia/lymphoma virus in healthy Venezuelan populations. Int J Cancer. 1984;34:501.

Merl S, Kloster B, Moore J et al. Efficient Transformation of Previously Activated and Dividing T Lymphocytes by Human T Cell Leukemia-Lymphoma Virus. Blood. 1984;64:967-974.

Mesnard JM, Devaux C. Multiple control levels of cell proliferation by human T- cell leukemia virus type 1 tax protein. Virology. 1999;257:277-284.

Michael B, Nair A, Lairmore MD. Role of accessory proteins of HTLV-1 in viral replication, T cell activation, and cellular gene expression. Front Biosci. 2004;9:2556-2576.

Michael B, Nair AM, Hiraragi H et al. Human T lymphotropic virus type-1 p30II alters cellular gene expression to selectively enhance signaling pathways that activate T lymphocytes. Retrovirology. 2004;1:39.

Michalak M, Corbett EF, Mesaeli N, Nakamura K, Opas M. Calreticulin: one protein, one gene, many functions. Biochem J. 1999;344 Pt 2:281-292.

Miskin JE, Abrams CC, Dixon LK. African swine fever virus protein A238L interacts with the cellular phosphatase calcineurin via a binding domain similar to that of NFAT. J Virol. 2000;74:9412-9420.

Miskin JE, Abrams CC, Goatley LC, Dixon LK. A viral mechanism for inhibition of the cellular phosphatase calcineurin. Science. 1998;281:562-565.

212

Miyatake S, Seiki M, Malefijt RD et al. 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. 1988;16:6547-6566.

Miyoshi I, Kubonishi I, Yoshimoto S et al. Type C virus particles in a cord T-cell line derived by co-cultivating normal human cord leukocytes and human leukaemic T cells. Nature. 1981;294:770-771.

Mochizuki M, Watanabe T, Yamaguchi K et al. HTLV-I Uveitis: A distinct clinical entity caused by HTLV-I. Jpn J Cancer Res. 1992;83:236-239.

Molgaard CA, Eisenman PA, Ryden LA, Golbeck AL. Neuroepidemiology of human T-lymphotrophic virus type-I-associated tropical spastic paraparesis. Neuroepidemiology. 1989;8:109-123.

Mori N, Fujii M, Ikeda S et al. Constitutive activation of NF-kappa B in primary adult T- cell leukemia cells. Blood. 1999;93:2360-2368.

Mori N, Mukaida N, Ballard DW, Matsushima K, Yamamoto N. 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. 1998;58:3993-4000.

Mori N, Murakami S, Oda S, Eto S. Human T-cell leukemia virus type I tax induces intracellular adhesion molecule-1 expression in T cells. Blood. 1994;84:350-351.

Mori N, Prager D. Transactivation of the interleukin-1 alpha promoter by human T- cell leukemia virus type I and type II tax proteins. Blood. 1996;87:3410-3417.

Mortreux F, Kazanji M, Gabet AS, de Thoisy B, Wattel E. Two-step nature of human T-cell leukemia virus type 1 replication in experimentally infected squirrel monkeys (Saimiri sciureus). J Virol. 2001;75:1083-1089.

Mosley AJ, Asquith B, Bangham CR. Cell-mediated immune response to human T-lymphotropic virus type I. Viral Immunol. 2005;18:293-305.

Muchardt C, Seeler JS, Nirula A, Gong S, Gaynor R. Transcription factor AP-2 activates gene expression of HTLV-I. EMBO Journal. 1992;11:2573-2581.

Mueller N.E., Blattner W.A. Retroviruses-Human T-Cell Lymphotropic Virus. Fourth Edition, 785-814. 1997. New York, N.Y., Plenum Publishing Corp. Ref Type: Generic

Mulloy JC, Crowley RW, Fullen J, Leonard WJ, Franchini G. The human T-cell leukemia/lymphotropic virus type 1 p12(I) protein binds the interleukin-2 receptor beta

213

and gamma(c) chains and affects their expression on the cell surface. J Virol. 1996;70:3599-3605.

Murakami T, Hirai H, Suzuki T, Fujisawa JI, Yoshida M. 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. 1995;206:1066-1074.

Muraoka O, Kaisho T, Tanabe M, Hirano T. Transcriptional activation of the interleukin-6 gene by HTLV-1. Immunology Letters. 1993;37(2-3):159-165.

Murphy EL, Busch MP, Tong M, Cornett P, Vyas GN. A prospective study of the risk of transfusion-acquired viral infections. Transfus Med. 1998;8:173-178.

Murphy EL, Glynn SA, Fridey J et al. Increased incidence of infectious diseases during prospective follow-up of human T-lymphotropic virus type II- and I-infected blood donors. ARCH INTERN MED. 1999;159:1485-1491.

Nagai M, Osame M. Human T-cell lymphotropic virus type I and neurological diseases. J Neurovirol. 2003;9:228-235.

Nagai M, Usuku K, Matsumoto W et al. 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. 1998;4:586-593.

Nagy K, Clapham P, Cheingsong-Popov R, Weiss R. Human T cell leukemia virus type 1: Induction of syncytia and inhibition by patient's sera. Int J Cancer. 1983;32:321-328.

Nair A, Michael B, Hiraragi H et al. Human T lymphotropic virus type 1 accessory protein p12I modulates calcium-mediated cellular gene expression and enhances p300 expression in T lymphocytes. AIDS Res Hum Retroviruses. 2005;21:273-284.

Nam SH, Kidokoro M, Shida H, Hatanaka M. Processing of gag Precursor Polyprotein of Human T-Cell Leukemia Virus Type 1 by Virus-Encoded Protease. J Virol. 1988;62:3718-3728.

Nath MD, Ruscetti FW, Petrow-Sadowski C, Jones KS. Regulation of the cell-surface expression of an HTLV-I binding protein in human T cells during immune activation. Blood. 2003;101:3085-3092.

Nejmeddine M, Barnard AL, Tanaka Y, Taylor GP, Bangham CR. HTLV-1 tax protein triggers microtubule reorientation in the virological synapse. J Biol Chem. 2005.

Nejmeddine M, Barnard AL, Tanaka Y, Taylor GP, Bangham CR. Human T-lymphotropic Virus, Type 1, Tax Protein Triggers Microtubule Reorientation in the Virological Synapse. J Biol Chem. 2005;280:29653-29660.

214

Ng PW, Iha H, Iwanaga Y et al. Genome-wide expression changes induced by HTLV-1 Tax: evidence for MLK- 3 mixed lineage kinase involvement in Tax-mediated NF-kappaB activation. Oncogene. 2001;20:4484-4496.

Nicot C. Current views in HTLV-I-associated adult T-cell leukemia/lymphoma. Am J Hematol. 2005;78:232-239.

Nicot C, Dundr M, Johnson JM et al. HTLV-1-encoded p30II is a post-transcriptional negative regulator of viral replication. Nat Med. 2004;10:197-201.

Nicot C, Harrod RL, Ciminale V, Franchini G. Human T-cell leukemia/lymphoma virus type 1 nonstructural genes and their functions. Oncogene. 2005;24:6026-6034.

Nicot C, Mahieux R, Takemoto S, Franchini G. Bcl-X(L) is up-regulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL samples. Blood. 2000;96:275-281.

Nicot C, Mulloy JC, Ferrari MG et al. HTLV-1 p12(I) protein enhances STAT5 activation and decreases the interleukin-2 requirement for proliferation of primary human peripheral blood mononuclear cells. Blood. 2001;98:823-829.

Nishioka K, Maruyama I, Sato K et al. Chronic inflammatory arthropathy associated with HTLV-I [letter]. Lancet. 1989;i:441.

Nitta T, Tanaka M, Sun B, Hanai S, Miwa M. The genetic background as a determinant of human T-cell leukemia virus type 1 proviral load. Biochem Biophys Res Commun. 2003;309:161-165.

Niyogi K, Hildreth JE. Characterization of new syncytium-inhibiting monoclonal antibodies implicates lipid rafts in human t-cell leukemia virus type 1 syncytium formation. J Virol. 2001;75:7351-7361.

Northrop JP, Ullman KS, Crabtree GR. Characterization of the nuclear and cytoplasmic components of the lymphoid-specific nuclear factor of activated T cells (NF-AT) complex. J Biol Chem. 1993;268:2917-2923.

Nosaka T, Siomi H, Adachi Y et al. 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. 1989;86:9798-9802.

Ohshima K, Kikuchi M, Masuda Y et al. Human T-cell leukemia virus type I associated lymphadenitis 5240. Cancer. 1992;69:239-248.

215

Okochi K, Sato H, Hinuma Y. A retrospective study on transmission of adult T cell leukemia virus by blood transfusion: seroconversion in recipients. Vox Sang. 1984;46:245-253.

Ootsuyama Y, Shimotohno K, Miwa M, Oroszlan S, Sugimura T. Myristylation of gag protein in human T-cell leukemia virus type-I and type-II. Jpn J Cancer Res. 1985;76:1132-1135.

Orita S, Saiga A, Takagi S et al. 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 5300. FEBS Lett. 1991;295:127-134.

Oroszlan S. Chemical Analysis of HTLV-1 Structural proteins 5260. In: Gallo RC, Essex ME, Gross L, eds. Human T-cell Leukemia Lymphoma Viruses. Cold Spring Harbor Laboratory Press; 1984:101-110.

Osame M. Pathological mechanisms of human T-cell lymphotropic virus type I-associated myelopathy (HAM/TSP). J Neurovirol. 2002;8:359-364.

Osame M, Igata A, Usuku K, Rosales RL, Matsumoto M. Mother-to-child transmission in HTLV-I associated myelopathy. Lancet. 1987;1:106.

Osame M, Izumo S, Igata A et al. Blood transfusion and HTLV-I-associated myelopathy. Lancet. 1986;104.

Osame M, Janssen R, Kubota H et al. Nationwide survey of HTLV-I-associated Myelopathy in Japan: Association with blood transfusion. Ann Neurol. 1990;28:50-56.

Osame M, Usuku K, Izumo S et al. HTLV-I associated myelopathy, a new clinical entity. Lancet. 1986;i:1031-1032.

Overbaugh J. HTLV-1 sweet-talks its way into cells. Nat Med. 2004;10:20-21.

Owen SM, Rudolph DL, Dezzutti CS et al. Transcriptional activation of the intercellular adhesion molecule 1 (CD54) gene by human T lymphotropic virus types I and II Tax is mediated through a palindromic response element. AIDS Res Hum Retroviruses. 1997;13:1429-1437.

Palmeri D, Malim MH. The human T-cell leukemia virus type 1 posttranscriptional trans- activator Rex contains a nuclear export signal. J Virol. 1996;70:6442-6445.

Park S, Uesugi M, Verdine GL. A second calcineurin binding site on the NFAT regulatory domain. Proc Natl Acad Sci U S A. 2000;97:7130-7135.

Pawson T. Protein modules and signalling networks 5308. Nature. 1995;373:573-580.

216

Perez OD, Mitchell D, Jager GC et al. Leukocyte functional antigen 1 lowers T cell activation thresholds and signaling through cytohesin-1 and Jun-activating binding protein 1. Nat Immunol. 2003;4:1083-1092.

Perini G, Wagner S, Green MR. Recognition of bZIP proteins by the human T-cell leukaemia virus transactivator Tax. Nature. 1995;376:602-605.

Perkins ND, Felzien LK, Betts JC et al. Regulation of NF-kappaB by cyclin-dependent kinases associated with the p300 coactivator. Science. 1997;275:523-527.

Petit C, Buseyne F, Boccaccio C et al. Nef is required for efficient hiv-1 replication in cocultures of dendritic cells and lymphocytes. Virology. 2001;286:225-236.

Petrini M, Pelosi-Testa E, Sposi NM et al. Constitutive expression and abnormal glycosylation of transferrin receptor in acute T-cell leukemia. Cancer Res. 1989;49:6989-6996.

Pinon JD, Klasse PJ, Jassal SR et al. Human T-cell leukemia virus type 1 envelope glycoprotein gp46 interacts with cell surface heparan sulfate proteoglycans. J Virol. 2003;77:9922-9930.

Pique C, Pham D, Tursz T, Dokhelar MC. Human T-Cell Leukemia Virus Type I Envelope Protein Maturation Process: Requirements for Syncytium Formation. J Virol. 1992;66:906-913.

Pique C, Uretavidal A, Gessain A et al. 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. 2000;191:567-572.

Pise-Masison CA, Jeong SJ, Brady JN. Human T cell leukemia virus type 1: the role of Tax in leukemogenesis. Arch Immunol Ther Exp (Warsz ). 2005;53:283-296.

Poiesz BJ, Ruscetti FW, Gazdar AF et al. 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. 1980;77:7415-7419.

Popovic M, Sarin PS, Robert-Gurroff M et al. Isolation and transmission of human retrovirus (human t-cell leukemia virus). Science. 1983;219:856-859.

Portis T, Harding JC, Ratner L. The contribution of NF-kappa B activity to spontaneous proliferation and resistance to apoptosis in human T-cell leukemia virus type 1 Tax- induced tumors. Blood. 2001;98:1200-1208.

Princler GL, Julias JG, Hughes SH, Derse D. Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology. 2003;317:136-145.

217

Proietti FA, Carneiro-Proietti AB, Catalan-Soares BC, Murphy EL. Global epidemiology of HTLV-I infection and associated diseases. Oncogene. 2005;24:6058-6068.

Rao A, Luo C, Hogan PG. Transcription factors of the NFAT family: regulation and function. Annu Rev Immunol. 1997;15:707-747.

Ratner L. Regulation of expression of the c-sis proto-oncogene 5269. Nucleic Acids Res. 1989;17:4101-4115.

Reeves W, Saxinger C, Brenes M et al. Human T-cell lymphotropic virus type I (HTLV-I) and risk factors in metropolitan Panama. Am J Epidemiol. 1988;127:539.

Ressler S, Morris GF, Marriott SJ. Human T-cell leukemia virus type 1 Tax transactivates the human proliferating cell nuclear antigen promoter. J Virol. 1997;71:1181-1190.

Robek MD, Wong FH, Ratner L. 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. 1998;72:4458-4462.

Rodrigues MS, Echevarria-Lima J, Fernandes dos SN, Rumjanek VM. CD69 expression induced by thapsigargin, phorbol ester and ouabain on thymocytes is dependent on external Ca2+ entry. Life Sci. 2003;73:1037-1051.

Rodriguez-Fernandez JL, Gomez M, Luque A et al. The interaction of activated integrin lymphocyte function-associated antigen 1 with ligand intercellular adhesion molecule 1 induces activation and redistribution of focal adhesion kinase and proline-rich tyrosine kinase 2 in T lymphocytes. Mol Biol Cell. 1999;10:1891-1907.

Roithmann S, Pique C, Le Cesne A et al. 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. 1994;68:3448-3451.

Rothermel B, Vega RB, Yang J et al. A protein encoded within the Down syndrome critical region is enriched in striated muscles and inhibits calcineurin signaling. J Biol Chem. 2000;275:8719-8725.

Rusnak F, Mertz P. Calcineurin: form and function. Physiol Rev. 2000;80:1483-1521.

Sagara Y, Inoue Y, Kojima E et al. Phosphatidylglycerol Participates in Syncytium Formation Induced by HTLV Type 1-Bearing Cells. AIDS Res Hum Retroviruses. 2001;17:125-135.

218

Sagara Y, Ishida C, Inoue Y, Shiraki H, Maeda Y. 71-Kilodalton heat shock cognate protein acts as a cellular receptor for syncytium formation induced by human T-cell lymphotropic virus type 1. J.Virol. 72, 535-541. 1998. Ref Type: Generic

Saito M, Furukawa Y, Kubota R et al. 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. 1995;1:286-294.

Saksena NK, Srinivasan A, Ge YC et al. Simian T cell leukemia virus type I from naturally infected feral monkeys from Central and West Africa encodes a 91-amino acid p12 (ORF-I) protein as opposed to a 99-amino acid protein encoded by HTLV type I from humans. AIDS Res Hum Retroviruses. 1997;13:425-432.

Sancho D, Gomez M, Sanchez-Madrid F. CD69 is an immunoregulatory molecule induced following activation. Trends Immunol. 2005;26:136-140.

Sancho D, Gomez M, Viedma F et al. CD69 downregulates autoimmune reactivity through active transforming growth factor-beta production in collagen-induced arthritis. J Clin Invest. 2003;112:872-882.

Saxinger WC, Blattner W, Levine P et al. Human T-cell leukemia virus (HTLV-I) antibodies in Africa. Science. 1984;225:1473.

Schreiber GB, Murphy EL, Horton JA et al. 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. 1997;14:263-271.

Seiki M, Hattori S, Hirayama Y, Yoshida M. Human adult T-cell leukemia virus: Complete nucleotide sequence of the provirus genome integrated in leukemia cell DNA. Proc Natl Acad Sci. 1983;80:3618-3622.

Seiki M, Hikikoshi A, Taniguchi T, Yoshida M. Expression of the pX Gene of HTLV-1: General Splicing Mechanism in the HTLV Family. Science. 1985;228:1532-1534.

Seiki M, Hikikoshi A, Yoshida M. The U5 sequence is a cis-acting repressive element for genomic RNA expression of human T cell leukemia virus type I. Virology. 1990;176:81-86.

Seiki M, Inoue J, Hidaka M, Yoshida M. 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. 1988;85:7124-7128.

219

Semmes OJ, Jeang KT. Localization of human T-cell leukemia virus type 1 Tax to subnuclear compartments that overlap with interchromatin speckles. J Virol. 1996;70:6347-6357.

Shaw GM, Gonda MA, Flickinger GH et al. Genomes of evolutionarily divergent members of the human T-cell leukemia virus family (HTLV-I and HTLV-II) are highly conserved, especially in pX. Proc Natl Acad Sci USA. 1984;81:4544-4548.

Shibasaki F, Kondo E, Akagi T, Mckeon F. Suppression of signalling through transcription factor NF-AT by interactions between calcineurin and Bcl-2. Nature. 1997;386:728-731.

Shimotohno K, Takano M, Teruuchi T, Miwa M. 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. 1986;83:8112-8116.

Shimoyama M. 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. 1991;79:428-437.

Silic-Benussi M, Cavallari I, Zorzan T et al. Suppression of tumor growth and cell proliferation by p13II, a mitochondrial protein of human T cell leukemia virus type 1. Proc Natl Acad Sci U S A. 2004;101:6629-6634.

Silverman LR, Phipps AJ, Montgomery A, Ratner L, Lairmore MD. 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. 2004;78:3837-3845.

Sims TN, Dustin ML. The immunological synapse: integrins take the stage. Immunol Rev. 2002;186:100-117.

Siomi H, Shida H, Nam SH et al. Sequence requirements for nucleolar localization of human T cell leukemia virus type I pX protein, which regulates viral RNA processing 5296. Cell. 1988;55:197-209.

Smith A, Bracke M, Leitinger B, Porter JC, Hogg N. LFA-1-induced T cell migration on ICAM-1 involves regulation of MLCK-mediated attachment and ROCK-dependent detachment. J Cell Sci. 2003;116:3123-3133.

Sodroski J, Rosen C, Goh WC, Haseltine W. A transcriptional activator protein encoded by the x-lor region of the human T-cell leukemia virus 5017. Science. 1985;228:1430-1434.

220

Sodroski JG, Rosen CA, Haseltine WA. Trans-Acting Transcriptional Activation of the Long Terminal Repeat of Human T Lymphotropic Viruses in Infected Cells. Science. 1984;223:381-385.

Srivastava RK, Sasaki CY, Hardwick JM, Longo DL. Bcl-2-mediated drug resistance: inhibition of apoptosis by blocking nuclear factor of activated T lymphocytes (NFAT)-induced Fas ligand transcription. J Exp Med. 1999;190:253-265.

Srivastava RK, Sasaki CY, Hardwick JM, Longo DL. Bcl-2-mediated drug resistance: inhibition of apoptosis by blocking nuclear factor of activated T lymphocytes (NFAT)-induced Fas ligand transcription. J Exp Med. 1999;190:253-265.

Srivastava RK, Sasaki CY, Hardwick JM, Longo DL. Bcl-2-mediated drug resistance: inhibition of apoptosis by blocking nuclear factor of activated T lymphocytes (NFAT)-induced Fas ligand transcription. J Exp Med. 1999;190:253-265.

Stewart M, Hogg N. Regulation of leukocyte integrin function: affinity vs. avidity. J Cell Biochem. 1996;61:554-561.

Stewart MP, Cabanas C, Hogg N. T cell adhesion to intercellular adhesion molecule-1 (ICAM-1) is controlled by cell spreading and the activation of integrin LFA-1. J Immunol. 1996;156:1810-1817.

Stewart MP, Mcdowall A, Hogg N. LFA-1-mediated adhesion is regulated by cytoskeletal restraint and by a Ca2+-dependent protease, calpain. J Cell Biol. 1998;140:699-707.

Stuver SO, Tachibana N, Okayama A et al. 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. 1993;167:57-65.

Sun SC, Ballard DW. Persistent activation of NF-kappa B by the Tax transforming protein of HTLV-1: hijacking cellular I kappa B kinases. Oncogene. 1999;18:6948-6958.

Suzuki T, Fujisawa J, Toita M, Yoshida M. 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. 1993;90:610-614.

Suzuki T, Hirai H, Fujisawa J, Fujita T, Yoshida M. 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. 1993;8:2391-2397.

221

Suzuki T, Hirai H, Murakami T, Yoshida M. 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. 1995;10:1199-1207.

Suzuki T, Hirai H, Yoshida M. 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. 1994;9:3099-3105.

Suzuki T, Kitao S, Matsushime H, Yoshida M. HTLV-1 Tax protein interacts with cyclin-dependent kinase inhibitor p16(INK4A) and counteracts its inhibitory activity towards CDK4. EMBO J. 1996;15:1607-1614.

Suzuki T, Narita T, Uchidatoita M, Yoshida M. Down-regulation of the INK4 family of cyclin-dependent kinase inhibitors by tax protein of HTLV-1 through two distinct mechanisms. Virology. 1999;259:384-391.

Suzuki T, Uchidatoita M, Yoshida M. 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. 1999;18:4137-4143.

Takatsuki K, Uchiyama T, Sagawa K, Yodoi J. Adult T-cell leukemia in Japan. In: Seno S, Takaku F, Irino S, eds. Topics in Hematology. Amsterdam: Excerpta Medica; 1977:73-77.

Tanimura A, Teshima H, Fujisawa JI, Yoshida M. 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. 1993;67:5375-5382.

Tendler CL, Greenberg SJ, Blattner WA et al. Transactivation of interleukin 2 and its receptor induces immune activation in human T-cell lymphotropic virus type I-associated myelopathy: pathogenic implications and a rationale for immunotherapy. Proceedings of the National Academy of Sciences of the United States of America. 1990;87:5218-5222.

Terada K, Katamine S, Eguchi K et al. Prevalence of serum and salivary antibodies to HTLV-1 in Sjogren's syndrome. Lancet. 1994;344:1116-1119.

Terashima A, Gotuzzo E, Alvarez H et al. [STRONGYLOIDES STERCORALIS: CLINICAL SEVERE FORMS ASSOCIATED TO HTLV-1 INFECTION]. Rev Gastroenterol Peru. 1999;19:35-40.

Teshigawara K, Maeda M, Nishino K et al. Adult T leukemia cells produce a lymphokine that augments interleukin 2 receptor expression. J Mol Cell Immunol. 1985;2:17-26.

222

Thebault S, Basbous J, Hivin P, Devaux C, Mesnard JM. HBZ interacts with JunD and stimulates its transcriptional activity. FEBS Lett. 2004;562:165-170.

Tian P, Estes MK, Hu Y et al. The rotavirus nonstructural glycoprotein NSP4 mobilizes Ca2+ from the endoplasmic reticulum. J Virol. 1995;69:5763-5772.

Tillmann M, Wessner R, Wigdahl B. Identification of human T-cell lymphotropic virus type I 21-base- pair repeat-specific and glial cell-specific DNA- protein complexes. J Virol. 1994;68:4597-4608.

Toedter GP, Pearlman S, Hofheinz D et al. Development of a monoclonal antibody-based p24 capsid antigen detection assay for HTLV-I, HTLV-II, and STLV-I infection. AIDS Res Hum Retroviruses. 1992;8:527-532.

Torgeman A, Mor-Vaknin N, Aboud M. 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. 1999;254:279-287.

Trovato R, Mulloy JC, Johnson JM et al. 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. 1999;73:6460-6467.

Trovato R, Mulloy JC, Johnson JM et al. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic leukemia virus type 1 p12 protein greatly influences its stability. J Virol. 1999;73:6460-6467.

Trovato R, Mulloy JC, Johnson JM et al. A lysine-to-arginine change found in natural alleles of the human T-cell lymphotropic leukemia virus type 1 p12 protein greatly influences its stability. J Virol. 1999;73:6460-6467.

Tsukahara T, Kannagi M, Ohashi T et al. 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. 1999;73:7981-7987.

Tsukahara T, Ratner L. Substitution of HIV Type 1 Nef with HTLV-1 p12. AIDS Res Hum Retroviruses. 2004;20:938-943.

Uchiyama T. Human T cell leukemia virus type I (HTLV-I) and human diseases. [Review] [134 refs]. Annual Review of Immunology. 1997;15:15-37:15-37.

Uchiyama T, Yodoi J, Sagawa K, Takatsuki K, Uchino H. Adult T-cell leukemia: clinical and hematologic features of 16 cases 5209. Blood. 1977;50:481-492.

223

Usuku K, Sonoda S, Osame M et al. HLA haplotype-linked high immune responsiveness against HTLV-I in HTLV-I-associated myelopathy: comparison with adult T-cell leukemia/lymphoma. Ann Neurol. 1988;23 Suppl:S143-50.:S143-S150.

van KY, Figdor CG. Activation and inactivation of adhesion molecules. Curr Top Microbiol Immunol. 1993;184:235-256.

Volkov Y, Long A, Kelleher D. Inside the crawling T cell: leukocyte function-associated antigen-1 cross-linking is associated with microtubule-directed translocation of protein kinase C isoenzymes beta(I) and delta. J Immunol. 1998;161:6487-6495.

von Andrian UH, Mackay CR. T-cell function and migration. Two sides of the same coin. N Engl J Med. 2000;343:1020-1034.

von Bulow GU, Bram RJ. NF-AT activation induced by a CAML-interacting member of the tumor necrosis factor receptor superfamily. Science. 1997;278:138-141.

Wang H, Moon EY, Azouz A et al. SKAP-55 regulates integrin adhesion and formation of T cell-APC conjugates. Nat Immunol. 2003;4:366-374.

Wattel E, Vartanian JP, Pannetier C, Wainhobson S. Clonal expansion of human T-Cell leukemia virus type I- infected cells in asymptomatic and symptomatic carriers without malignancy. J Virol. 1995;69:2863-2868.

Weiss RA. Milk-borne transmission of HTLV-1. [Review]. Japanese Journal of Cancer Research. 1993;84(3):inside front cover.

Wessner R, Tillmann-Bogush M, Wigdahl B. Characterization of a glial cell-specific DNA-protein complex formed with the human T cell lymphotropic virus type I (HTLV-I) enhancer. J Neurovirol. 1995;1:62-77.

Wessner R, Wigdahl B. AP-1 derived from mature monocytes and astrocytes preferentially interacts with the HTLV-I promoter central 21 bp repeat. Leukemia. 1997;11 Suppl 3:21-4.:21-24.

Wiktor SZ, Pate EJ, Murphy EL et al. 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. 1993;6:1162-1167.

Wu K, Bottazzi ME, de La FC et al. Protein profile of tax-associated complexes. J Biol Chem. 2004;279:495-508.

Wu Y, Marsh JW. Selective transcription and modulation of resting T cell activity by preintegrated HIV DNA. Science. 2001;293:1503-1506.

Wucherpfennig KW, Hollsberg P, Richardson JH, Benjamin D, Hafler DA. T-cell activation by autologous human T-cell leukemia virus type I- infected T-cell clones

224

5375. Proc Natl Acad Sci U S A. 1992;89:2110-2114.

Xu RZ, Gao QK, Wang SJ et al. Human acute myeloid leukemias may be etiologically associated with new human retroviral infection. Leuk Res. 1996;20:449-455.

Yajima A, Kawada A, Aragane Y, Tezuka T. Detection of HTLV-I proviral DNA in sarcoidosis. Dermatology. 2001;203:53-56.

Yamamoto A, Hara H, Kobayashi T. Induction of the expression of gag protein in HTLV-I infected lymphocytes by anti-ICAM 1 antibody in vitro. J Neurol Sci. 1997;151:121-126.

Yamamoto N, Okada M, Koyanagi Y, Kannagi M, Hinuma Y. Transformation of human Leukocytes by cocultivation with an adult T cell leukemia virus producer cell line. Science. 1982;217:737-739.

Yamano Y, Nagai M, Brennan M et al. 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. 2002;99:88-94.

Yao J, Wigdahl B. Human T cell lymphotropic virus type I genomic expression and impact on intracellular signaling pathways during neurodegenerative disease and leukemia. Front Biosci. 2000;5:D138-D168.

Yoshida M. HTLV-1 Tax: regulation of gene expression and disease. [Review]. Trends in Microbiology. 1993;1(4):131-135.

Yoshida M. Multiple viral strategies of htlv-1 for dysregulation of cell growth control. Annu Rev Immunol. 2001;19:475-496.

Yoshida M, Miyoshi I, Hinuma Y. Isolation and characterization of retrovirus from cell lines of human T-cell leukemia and its implication in the disease. Proc Natl Acad Sci USA. 1982;79:2031-2035.

Yoshida M, Seiki M, Yamaguchi K, Takatsuki K. 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. 1984;81:2534-2537.

Younis I, Boris-Lawrie K, Green PL. Human T-cell leukemia virus open reading frame II encodes a posttranscriptional repressor that is recruited at the level of transcription. J Virol. 2006;80:181-191.

Younis I, Khair L, Dundr M et al. Repression of Human T-Cell Leukemia Virus Type 1 and Type 2 Replication by a Viral mRNA-Encoded Posttranscriptional Regulator. J Virol. 2004;78:11077-11083.

225

Zhang W, Nisbet JW, Albrecht B et al. Human T-lymphotropic virus type 1 p30(II) regulates gene transcription by binding CREB binding protein/p300. J Virol. 2001;75:9885-9895.

Zhang W, Nisbet JW, Bartoe JT, Ding W, Lairmore MD. Human T-lymphotropic virus type 1 p30(II) functions as a transcription factor and differentially modulates CREB-responsive promoters. J Virol. 2000;74:11270-11277.

Zhao LJ, Giam CZ. 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. 1992;89:7070-7074.

Zhong H, Voll RE, Ghosh S. Phosphorylation of NF-kappa B p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with the coactivator CBP/p300. Mol Cell. 1998;1:661-671.

Zhou J, Aiken C. Nef enhances human immunodeficiency virus type 1 infectivity resulting from intervirion fusion: evidence supporting a role for Nef at the virion envelope. J Virol. 2001;75:5851-5859.


Recommended