+ All Categories
Home > Documents > ROLE OF HUMAN HERPESVIRUS 8 VIRAL INTERLEUKIN-6 …

ROLE OF HUMAN HERPESVIRUS 8 VIRAL INTERLEUKIN-6 …

Date post: 22-Dec-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
167
ROLE OF HUMAN HERPESVIRUS 8 VIRAL INTERLEUKIN-6 SIGNALING IN VIRUS BIOLOGY by Emily M. Cousins A dissertation submitted to Johns Hopkins University in conformity with the requirements for the degree of Doctor of Philosophy Baltimore, Maryland December 2014 © 2014 Emily M. Cousins All Rights Reserved
Transcript

ROLE OF HUMAN HERPESVIRUS 8 VIRAL INTERLEUKIN-6 SIGNALING IN VIRUS BIOLOGY

by Emily M. Cousins

A dissertation submitted to Johns Hopkins University in conformity with the requirements for the degree of Doctor of Philosophy

Baltimore, Maryland December 2014

© 2014 Emily M. Cousins All Rights Reserved

ii

ABSTRACT

Human herpesvirus 8 (HHV-8) is associated with B cell and endothelial tumors.

HHV-8 encodes a viral homolog of interleukin 6 (vIL-6), which promotes the proliferation

and survival of virally infected and neighboring cells and induces angiogenesis, suggesting

that vIL-6 contributes to HHV-8 malignant pathogenesis. Furthermore, vIL-6 is expressed

during viral latency and lytic replication, allowing vIL-6 to function in viral latency

maintenance and virus productive replication.

Prior to this dissertation research, little was known regarding the role of vIL-6

interactions with its signal transducer (gp130) within the endoplasmic reticulum (ER),

where vIL-6 predominantly localizes. High levels of phosphorylated STAT3, a product of

vIL-6 signaling, had been observed in HHV-8+ primary effusion lymphomas (PEL), but the

mechanism behind STAT3 activation was unclear. The effects of vIL-6 on virus replication

were limited to one published study, which failed to detect any influence of vIL-6 in a model

culture system. Mechanistically, vIL-6 is distinct from its cellular counterparts because it

can signal from the ER, but the biological functions of ER-localized vIL-6/gp130 signaling

were unknown. Therefore, the goals of this dissertation were to further characterize ER-

localized vIL-6/gp130 interactions, investigate the mechanism(s) responsible for

increased STAT3 activation in PEL cells, decipher the roles of gp130-activated STAT and

ERK signaling in PEL cell maintenance, and evaluate the role of vIL-6 and the vIL-6/gp130

interaction in HHV-8 replication.

Studies performed during this dissertation work demonstrated that ER-localized

vIL-6/gp130 signaling in PEL cells contributes to high levels of active STAT3, that vIL-

6/gp130 signaling also activates STAT1 and ERK1/2 in PEL cells, that ER-localized vIL-

6/gp130 signaling can sustain PEL cell growth and viability, and that STAT3 and ERK1/2

are crucial for cell growth and survival. In addition to promoting the growth and survival of

latently infected PEL cells, vIL-6 and gp130 also promote productive virus replication in

iii

PEL and endothelial cells. In PEL cells, vIL-6 pro-replication effects require its interaction

with gp130 in the ER, where vIL-6/gp130-mediated STAT signaling, and not MAPK

signaling, is critical. These studies have expanded our understanding of how vIL-6

functions in normal virus biology and have opened new avenues for targeted HHV-8

therapy.

Readers John Nicholas, PhD (Advisor) Professor of Oncology Johns Hopkins University School of Medicine, Baltimore, MD Gary Ketner, PhD (Committee Member) Professor of Molecular Microbiology and Immunology Johns Hopkins Bloomberg School of Public Health, Baltimore, MD

iv

ACKNOWLEDGEMENTS

Many individuals have supported me during my graduate training at the Johns

Hopkins University School of Medicine. First and foremost, I must acknowledge my

mentor, John Nicholas. My development as a scientist can be attributed to John’s patience

and guidance throughout my time in his laboratory. John taught me how to design

experiments with proper controls, how to write scientific articles and grant applications,

and how to develop a story worthy of publication. Additionally, John fostered a

collaborative environment in the laboratory in which techniques, reagents, and expertise

were shared. I would also like to thank the other members of the Nicholas lab for their

guidance, instruction, and friendship: Daming Chen, Young Choi, Gordon (Sandy)

Sandford, Yang Gao, and Kihyun Yoo. Additionally, much of my dissertation research was

funded by a Ruth L. Kirschstein F31 fellowship (5F31CA171933), and I am grateful for the

support of this program.

Prior to joining the Nicholas laboratory, several other scientists taught and

encouraged me. My first foray into research occurred during my undergraduate studies at

Clemson University. My mentor at Clemson, Michael Sehorn, is passionate about science

(and college football), and his continued support and guidance throughout my scientific

training has been incredibly valuable. Additionally, my research rotations with Joel

Pomerantz and Kathryn Wagner were both fun and instructive; Joel and Kathryn are both

gifted scientists, and they have graciously advised me throughout my time at Hopkins.

Furthermore, I owe my gratitude to members of my thesis committee, whose time

and expertise were crucial to the development of my thesis project. Young Choi, Diane

Hayward, Kuan-Teh Jeang, and Gary Ketner were instrumental in the design and

execution of my research projects. Interactions with my thesis committee encouraged me

to ask the right research questions, to determine the best method for answering my

v

research questions, and to think about my career after graduation. Their help and support

cannot be overstated.

The training environment within the Graduate Training Program in Cellular and

Molecular Medicine encourages collaboration and friendship from the very beginning. I

would like to thank all of my classmates for making graduate school enjoyable. Afternoon

coffee breaks with friends and fellow students were not only stress-relieving, but these

discussions often yielded new approaches for circumventing experimental failures. Thank

you, Adam Moyer, Laurene Cheung, Laura Gottschalk, Sara Sinclair, deMauri Mackie,

Emily Chang, Kate Laws, and Justin Glenn.

I also want to acknowledge and thank my family for their love and support

throughout this journey. I attribute my love of science to my parents, Shawn and Susan

Gaunt, who have spent their careers in the medical field. They have always encouraged

me and my siblings (Patrick, Benjamin, Christine, and Leah) to study hard, help others,

and explore new things.

Lastly, I could not have completed my PhD without the love, encouragement, and

support of my husband, Matthew Cousins. Having a spouse understand not only the time

commitment of research but also one that you can bounce scientific ideas off of is

incredibly helpful. Learning how to be parents to Mark may be our biggest challenge yet,

but I have no doubt that we will succeed together.

vi

TABLE OF CONTENTS

ABSTRACT.................................................................................................................... II

ACKNOWLEDGEMENTS ............................................................................................. IV

TABLE OF CONTENTS ................................................................................................ VI

LIST OF ABBREVIATIONS ........................................................................................ VIII

LIST OF TABLES ....................................................................................................... XIII

LIST OF FIGURES ...................................................................................................... XIV

CHAPTER 1: INTRODUCTION ...................................................................................... 1

CLINICAL SIGNIFICANCE OF HUMAN HERPESVIRUS 8 ................................ 2

HUMAN HERPESVIRUSES ............................................................................... 5

HUMAN HERPESVIRUS GENOME ORGANIZATION ........................................ 7

HISTORY AND UNIQUE FEATURES OF HHV-8 ............................................... 8

HHV-8 LYTIC CYCLE ......................................................................................... 8

HHV-8 LATENCY ...............................................................................................11

VIRAL INTERLEUKIN 6 AND PATHOGENESIS ................................................13

VIRAL INTERLEUKIN 6 CHARACTERISTICS ...................................................16

SIGNIFICANCE OF VIRAL IL-6 AND THIS PROJECT ......................................19

TABLES .............................................................................................................20

FIGURES ...........................................................................................................22

CHAPTER 2: METHODS ..............................................................................................29

INTRODUCTION ...............................................................................................31

RNA EXTRACTION ...........................................................................................31

REVERSE TRANSCRIPTION ............................................................................31

POLYMERASE CHAIN REACTION ...................................................................32

AGAROSE GEL ELECTROPHORESIS .............................................................33

CLONING AND SHORT HAIRPIN RNA DESIGN ..............................................34

MINIPREP DNA EXTRACTIONS .......................................................................36

CELL CULTURE ................................................................................................36

TRANSFECTION OF CULTURED CELLS .........................................................37

WESTERN BLOT ANALYSIS ............................................................................38

IMMUNOFLOURESCENCE STAINING .............................................................39

BRDU INCORPORATION ..................................................................................39

ANNEXIN V-CY3 STAINING ..............................................................................40

LENTIVIRUS PRODUCTION .............................................................................41

LENTIVIRAL TRANSDUCTION OF CULTURED CELLS ...................................41

CELL GROWTH ASSAY ....................................................................................42

HHV-8 R219 GENERATION AND INFECTION OF ENDOTHELIAL CELLS ......42

TITERING OF REACTIVATED HHV-8 ...............................................................43

DEPLETION AND COMPLEMENTATION EXPERIMENTS ...............................44

vii

CHAPTER 3: INVOLVEMENT OF GP130 IN VIL-6-MEDIATED PEL CELL GROWTH

AND SURVIVAL .......................................................................................................46

ROLE OF HUMAN HERPESVIRUS 8 INTERLEUKIN-6-ACTIVATED GP130 SIGNAL TRANSDUCER IN PRIMARY EFFUSION LYMPHOMA CELL GROWTH AND VIABILITY ............................................................................49

INTRODUCTION ....................................................................................49

METHODS .............................................................................................51

RESULTS ...............................................................................................55

DISCUSSION .........................................................................................61

FIGURES ...............................................................................................65

CHAPTER 4: ROLE OF GP130 IN VIL-6-MEDIATED VIRUS REPLICATION ..............80

HUMAN HERPESVIRUS 8 VIRAL INTERLEUKIN-6 SIGNALING THROUGH GP130 PROMOTES VIRUS REPLICATION IN PRIMARY EFFUSION LYMPHOMA AND ENDOTHELIAL CELLS....................................................82

INTRODUCTION ....................................................................................82

METHODS .............................................................................................83

RESULTS ...............................................................................................89

DISCUSSION .........................................................................................91

FIGURES ...............................................................................................93

CHAPTER 5: DISCUSSION AND PERSPECTIVES ................................................... 101

DISCUSSION OF MAJOR FINDINGS .................................................. 102

IMPLICATIONS FOR CLINICAL RESEARCH ...................................... 108

FUTURE DIRECTIONS ........................................................................ 110

FIGURES ............................................................................................. 111

REFERENCES ............................................................................................................ 115

CURRICULUM VITAE ................................................................................................. 151

viii

LIST OF ABBREVIATIONS

AIDS acquired immunodeficiency syndrome

bFGF basic fibroblast growth factor

BrdU 5-bromo-2-deoxyuridine

BSA bovine serum albumin

CatD cathepsin D

cDNA complementary deoxyribonucleic acid

CTL cytotoxic T lymphocyte

DAPI 4',6-diamidino-2-phenylindole

DHFR dihydrofolate reductase

DMEM Dulbecco’s modified Eagle’s medium

DNA deoxyribonucleic acid

dNTP deoxynucleotide triphosphate

Dox doxycycline

DTT dithiothreitol

EBV Epstein-Barr virus (HHV-4)

EBM-2 endothelial growth basal media

EDTA ethylenediaminetetraacetic acid

ER endoplasmic reticulum

FBS fetal bovine serum

FLICE FADD-like interleukin-1 beta-converting enzyme

gB glycoprotein B

GC guanine-cytosine

GFP green fluorescent protein

GSK3β glycogen synthase kinase-3 beta

ix

HAART highly active antiretroviral therapy

HCl hydrochloric acid

HCMV human cytomegalovirus (HHV-5)

HCV hepatitis C virus

MHC major histocompatibility complex

HHV-1 human herpesvirus 1 (HSV-1)

HHV-2 human herpesvirus 2 (HSV-2)

HHV-3 human herpesvirus 3 (VZV)

HHV-4 human herpesvirus 4 (EBV)

HHV-5 human herpesvirus 5 (HCMV)

HHV-6 human herpesvirus 6

HHV-7 human herpesvirus 7

HHV-8 human herpesvirus 8 (KSHV)

hIL-6 human interleukin 6

HIV human immunodeficiency virus type 1 (HIV-1)

HRP horseradish peroxidase

hrs hours

HSV-1 herpes simplex virus 1 (HHV-1)

HSV-2 herpes simplex virus 2 (HHV-2)

IFN interferon

IL-1 interleukin 1

IL-6 interleukin 6

IL6R interleukin 6 receptor (alpha subunit)

JAK Janus kinase

kb kilobase

KICS KSHV inflammatory cytokine syndrome

x

KS Kaposi’s sarcoma

KSHV Kaposi’s sarcoma-associated herpesvirus (HHV-8)

LANA latency-associated nuclear antigen

LB Luria broth

MAPK mitogen-activated protein kinase

MCD multicentric Castleman’s disease

MCMV murine cytomegalovirus

MHC major histocompatibility complex

min minute

MIR modulator of immune recognition

miRNA micro ribonucleic acid

mRNA messenger ribonucleic acid

N asparagine

NaB sodium butyrate

NK natural killer

ORF open reading frame

OX-2 orexin receptor 2 (CD 200)

PAN RNA polyadenylated nuclear ribonucleic acid

PBS phosphate-buffered saline

PCR polymerase chain reaction

PDGF platelet-derived growth factor

PEL primary effusion lymphoma

PM plasma membrane

PPMO peptide-conjugated phosphorodiamidate morpholino oligomer

pRb retinoblastoma protein

PS phosphatidylserine

xi

RNA ribonucleic acid

Rta replication and transcriptional activator

RT-PCR reverse transcription-polymerase chain reaction

ROS reactive oxygen species

SCID severe combined immunodeficiency

SDS-PAGE sodium dodecyl sulfate-polyacrylamide gel electrophoresis

shRNA short hairpin ribonucleic acid

SOCS3 suppressor of cytokine signaling 3

STAT signal transducer and activator of transcription

STAT3D dominant negative signal transducer and activator of transcription

TAE Tris-acetate-EDTA

TIME telomerase-immortalized microvascular endothelial

TK thymidine kinase

TLR Toll-like receptor

TNFα tumor necrosis factor alpha

TPA 12-O-tetradecanoylphorbol-13-acetate

TR terminal repeat

TS thymidylate synthase

U unit

UTR untranslated region

UV ultraviolet

vBcl-2 viral B-cell lymphoma 2

vCCL-1 viral chemokine (C-C motif) ligand 1

vCCL-2 viral chemokine (C-C motif) ligand 2

vFLIP viral FLICE-inhibitory protein

vGPCR viral G-protein-coupled receptor

xii

vIRF viral interferon regulatory factor

VKORC1v1 vitamin K epoxide reductase complex subunit 1 variant 1

VKORC1v2 vitamin K epoxide reductase complex subunit 1 variant 2

VKORC1v3 vitamin K epoxide reductase complex subunit 1 variant 3

VSV/G vesicular stomatitis virus G protein

vIL-6 viral interleukin 6

VZV varicella-zoster virus (HHV-3)

WB western blot

WT wild-type

XBP1 X-box binding protein 1 Y tyrosine

xiii

LIST OF TABLES

Table 1.1. Descriptions of HHV-8-specific open reading frames (“K” ORFs) ..................20

Table 1.2. HHV-8 microRNAs ........................................................................................21

xiv

LIST OF FIGURES

Figure 1.1. Genomic organization of HHV-8 ..................................................................22

Figure 1.2. vIL-6 and hIL-6 three-dimensional structures ...............................................24

Figure 1.3. Human and viral interleukin 6 signaling complexes ......................................25

Figure 1.4. Known and proposed functions of viral interleukin 6 ....................................27

Figure 3.1. Effects of gp130 depletion on PEL cell growth and viability ..........................65

Figure 3.2. Contribution of gp130 signaling to levels of active STATs and ERKs in PEL

cells ..........................................................................................................................67

Figure 3.3. Contributions of gp130-activated STATs to PEL cell growth ........................69

Figure 3.4. Analysis of culture growth, cell proliferation, and apoptosis rates as a function

of ERK depletion .......................................................................................................71

Figure 3.5. ER-localized vIL-6 signal transduction via gp130 supports PEL cell growth .74

Figure 3.6. Inhibitory targeting of vIL-6/gp130 signaling diminishes PEL cell proliferation

and survival ..............................................................................................................77

Figure 4.1. Involvement of gp130 and vIL-6 in promotion of HHV-8 replication ..............93

Figure 4.2. Importance of ER-localized vIL-6/gp130 interactions for HHV-8 replication .95

Figure 4.3. Contributions of gp130-mediated STAT and ERK signaling to HHV-8

productive replication ................................................................................................97

Figure 4.4. Role of STAT3 in HHV-8 replication in PEL and endothelial cells ................99

Figure 5.1. ER-localized vIL-6 and gp130 signaling in PEL cell growth and survival .... 111

Figure 5.2. Role of vIL-6 and gp130 in HHV-8 replication in PEL and endothelial cells 113

1

CHAPTER 1: INTRODUCTION

2

CLINICAL SIGNIFICANCE OF HUMAN HERPESVIRUS 8

Human herpesvirus 8 (HHV-8), also known as Kaposi’s sarcoma-associated

herpesvirus (KSHV), has been linked to three diseases in humans: Kaposi’s sarcoma

(KS), primary effusion lymphoma (PEL), and multicentric Castleman’s disease (MCD).

HHV-8 infection is required for the development of these neoplasias; however, disease

progression is rare in immunocompetent hosts. These diseases are most prevalent in

populations that are immunocompromised, such as those with acquired immunodeficiency

syndrome (AIDS) and those who are immunosuppressed due to organ transplant or

treatment of unrelated disease. It is believed that immunosuppression allows the virus to

replicate and for latently infected cell populations to expand in the absence of host-

immune pressures, leading to more rapid and aggressive disease progression. While KS,

PEL, and MCD share an etiologic cause in HHV-8, they differ in target cell types, clonality

of lesions, and host cell responses to viral infection.

KS is a distinctive condition that inspired one of the alternative names given to

HHV-8, KSHV. Four forms of KS have been described: classical, AIDS-associated,

endemic, and iatrogenic [1-8]. These classifications are based on geographical

prevalence, disease progression, and other factors. Classical KS was first described by

dermatologist Dr. Moritz Kaposi in 1872 [9]. This form of the disease is characterized by

cutaneous lesions of the skin, which are commonly observed in the extremities [9].

Classical KS is most common in elderly men of eastern European and Mediterranean

descent [10], but with the emergence of HIV/AIDS in Africa beginning in the 1980s, the

AIDS-associated form of KS became more prevalent in this geographical region [11]. The

prevalence of AIDS-associated KS has mirrored the prevalence of HIV infection in Africa,

and the advent of highly active antiretroviral therapy (HAART) has led to a rapid decline

in both HIV incidence and AIDS-associated KS [12]. AIDS-associated KS is the most

aggressive form of KS and has the highest prevalence of the four subtypes [11,13].

3

Endemic KS is most commonly observed in African children and is characterized by high

mortality rates [14]. Finally, iatrogenic KS is observed in transplant recipients that have

undergone immunosuppressive therapy [15].

Nearly all KS lesions contain cells infected with HHV-8, though the fraction of virus-

positive cells within each lesion varies widely (partially dependent on the stage of disease)

[16,17]. HHV-8 infects endothelial cells, which subsequently become spindloid and

proliferate rapidly [17]. Spindle cells have a distinct morphology and a unique gene

expression profile; these cells express a mixture of blood and lymphatic endothelial cell

markers [18-20]. KS lesions are generally polyclonal in nature, though clonal KS has also

been observed [21,22]. Lesions are reddish, brown, or purple in color due to the induction

of angiogenesis and extravasation of erythrocytes from malformed microvasculature.

Cytokine dysregulation has been implicated as a cause of KS. Pro-inflammatory

factors promote the infiltration of immune cells into infected tissues. Cultured KS cells

require basic fibroblast growth factor (bFGF), interleukins 1 and 6 (IL-1, IL-6), and platelet-

derived growth factor (PDGF) for growth, and cells found in KS lesions express IL-6 and

PDGF receptors [23-27]. Furthermore, cells in KS lesions secrete IL-1β, tumor necrosis

factor alpha (TNFα), and IL-6 [23]. Virus infection drives this cytokine-rich environment,

which promotes virus replication and the infection of new host cells that secrete additional

cellular and viral cytokines, further exacerbating the infection.

HHV-8 is also the causative agent of PEL, a B-cell malignancy that involves the

pleural, peritoneal, and pericardial spaces [28,29]. Infected B-cells express both B-cell

and plasma cell markers (CD138, CD38, and CD23), suggesting that the HHV-8-infected

cells are in fact immature B-cells that were in the process of differentiating into plasma

cells when the virus promoted their proliferation and halted their differentiation [30-33].

PEL tumors are monoclonal, and each infected B-cell contains many (approximately 50)

copies of the viral genome maintained as circular episomes [34,35]. PEL is generally

4

diffuse with no detectable mass, but solid tumors have also been reported [36]. Cases of

PEL are most commonly observed in HIV-positive patients or other immunocompromised

individuals; the median survival of patients with this aggressive cancer is only six months

[37]. The initiation of HAART prior to PEL diagnosis is thought to improve clinical prognosis

in patients co-infected with HIV [37].

Finally, HHV-8 is the etiologic agent of MCD, which is characterized by viral

infection of plasmablastic B-cells within the mantle zone of B-cell follicles [38]. Unlike PEL,

MCD is generally polyclonal, though cases of clonal MCD have been reported [39]. MCD

progression in HIV-infected individuals is rapid; the disease is driven by the elevated

expression of pro-inflammatory cytokines, including IL-6, IL-10, and vascular endothelial

growth factor (VEGF) [40-43]. Furthermore, levels of IL-6 in MCD patients were inversely

correlated with patient prognosis [40]. Recently, it was reported that six patients coinfected

with HIV and HHV-8 presented with pathologies and symptoms similar to MCD, but these

patients could not be classified as having MCD and did not develop MCD during follow-up

of 3-60 months [44]. These patients had elevated levels of HHV-8-encoded viral IL-6 (vIL-

6) (similar to levels observed in patients diagnosed with MCD) compared to patients with

mild or severe KS [44]. Furthermore, those with MCD-like inflammatory disease also had

significantly increased levels of serum human IL-6 (hIL-6), serum IL-10, and HHV-8 viral

load compared to control patients with KS [44]. These levels were comparable to serum

hIL-6 levels, IL-10 levels, and HHV-8 viral loads in MCD patients [44]. This MCD-like

disease has since been termed KSHV inflammatory cytokine syndrome (KICS) [45]. In an

AIDS patient presenting with KICS, circulating HHV-8 viral load was calculated as

5,300,000 copies per ml; this HHV-8 viral load is the highest value reported to date [46].

This patient also had high HHV-6A viremia, though the significance of HHV-6A infection

as it relates to KICS is unclear [46]. KICS is thought to arise from the dysregulation of pro-

inflammatory cytokines in the context of increased lytic replication of HHV-8 [45].

5

HHV-8 causes neoplasia in the context of an immunocompromised host. This virus

has been linked to three malignancies that complicate the management of infectious and

induced immunodeficiencies. vIL-6 is believed to contribute to all three diseases via its

pro-proliferative, anti-apoptotic, pro-angiogenic, and pro-inflammatory activities. Though

HHV-8 is unique in its possession of vIL-6 and other virus-specific genes, this virus shares

many characteristics with other human herpesviruses.

HUMAN HERPESVIRUSES

Herpesviruses are ancient (approximately 180-220 million years old) [47].

Members of the family Herpesviridae can infect a variety of hosts, including humans, birds,

pigs, mice, elephants, and oysters [47]. Each of the herpesviruses has a linear, double-

stranded, deoxyribonucleic acid (DNA) genome, a lipid bilayer envelope surrounding the

tegument, and an icosahedral capsid architecture (triangulation T=16) [48]. Herpesvirus

genomes range in size from approximately 125-225 kilobases (kb) and contain 70-200

open reading frames (ORFs) [48]. Mature virions measure approximately 120-260

nanometers in diameter [49]. Of the greater than 100 identified members of the

herpesvirus family, eight infect humans [50]. The eight known human herpesviruses fall

into three subfamilies: Alphaherpesvirinae (α), Betaherpesvirinae (β), and

Gammaherpesvirinae (Ɣ) [51].

Alphaherpesviruses include herpes simplex virus 1 (HSV-1; also called HHV-1),

herpes simplex virus 2 (HSV-2; also called HHV-2), and varicella-zoster virus (VZV; also

called HHV-3). HSV-1 and HSV-2 are responsible for cold sores and genital warts,

respectively. During outbreaks, HSV-1 and HSV-2 are reactivated from host neuronal

cells, and virions are shed from the oral or genital mucosa. VZV is associated with both

chicken pox, a primary infection characterized by a vesicular skin rash, and shingles, a

reactivation of latent VZV from neural ganglia. The alphaherpesviruses are characterized

6

by accelerated replication in culture, efficient cell lysis of infected host cells, and diverse

cell tropism compared to members of Betaherpesvirinae and Gammaherpesvirinae [48].

The betaherpesvirus subfamily includes human cytomegalovirus (HCMV; formally

HHV-5), human herpesvirus 6 (HHV-6), and human herpesvirus 7 (HHV-7). HCMV infects

cells of the salivary glands, causing characteristic cytomegaly [48]. HCMV has the largest

genome of the eight human herpesviruses at 235 kb and encodes more than 200 ORFs

[52]. HCMV infection is generally asymptomatic in immunocompetent individuals;

however, HCMV infection of immunocompromised individuals can result in pneumonia,

dyspepsia, meningitis, and pericardial effusions [53-56]. Though highly prevalent in

children, HHV-6 (95% seropositive by age 3) and HHV-7 (90% seropositive by age 5) have

minor clinical significance [57,58]. HHV-6 causes roseola infantum, and HHV-7 is often

associated with concurrent HHV-6 infection, though HHV-7 infection alone can also cause

roseola infantum [59-61]. Roseola infantum, also known as exanthema subitum, is

characterized by high fevers and a morbilliform rash that first appears on the trunk and

later on the face and extremities [62]. Viruses within the betaherpesvirinae subfamily

undergo slow replication and spread in culture and have asynchronous lytic cycles [48].

The gammaherpesviruses include Epstein-Barr virus (EBV; formally HHV-4) and

HHV-8 (also called KSHV). According to the 2013 release from the International

Committee on the Taxonomy of Viruses, EBV can be further classified as a member of the

lymphocryptovirus (gamma-1) genus while HHV-8 has been placed within the rhadinovirus

(gamma-2) genus. EBV is associated with infectious mononucleosis, Burkitt’s lymphoma,

Hodgkin’s lymphoma, and nasopharyngeal carcinoma. As noted previously, HHV-8 is the

causative agent of KS, MCD, and PEL [28,38,63]. Gammaherpesviruses exhibit latent

tropism for lymphocytes (with EBV and HHV-8 also infecting epithelial and endothelial

cells, respectively), have transforming capabilities, and are more likely to be oncogenic in

7

immunocompromised hosts. HIV-positive individuals and those having received organ

transplants are particularly susceptible to EBV- and HHV-8-associated cancers [64].

HUMAN HERPESVIRUS GENOME ORGANIZATION

Though each of the human herpesviruses has a complement of unique genes,

these viruses also share sets of genes encoded in conserved gene blocks. Though the

relative orientations and genomic locations of the gene blocks differ between the

herpesvirus subfamilies (α, β, and Ɣ), the genes encoded within a specific gene block are

maintained in the same relative arrangement. Many of these conserved genes (>40) are

involved in viral entry, replication, virion packaging, virion maturation, and egress [65].

Enzymes involved in basic metabolism, including thymidine kinase (TK), ribonucleotide

reductase, and uracil-DNA glycosylase, are also encoded within the genomes of all human

herpesviruses. Many viral structural proteins that make up the virion capsid are also

conserved within the human herpesviruses; these capsid proteins include the major capsid

protein, small capsid protein, triplex monomer, portal protein, and portal capping protein

[65].

A number of the human herpesviruses encode “accessory” (non-core) cellular

gene homologs. Frequently, these genes are specific to one or a few members of

Herpesviridae. These genes encode proteins that function as chemokines (CXC

chemokines in HCMV, CC chemokines in HHV-8), cytokines (vIL-10 in EBV and vIL-6 in

HHV-8), and cell cycle regulators (vCyc in HHV-8) [66-73]. Several anti-apoptotic proteins

are also encoded by human herpesviruses (e.g., vBcl-2 in EBV and HHV-8 and viral

FLICE-inhibitory protein [vFLIP] in HHV-8) [74-78]. Additionally, microRNAs (miRNAs) are

encoded by several herpesviruses, including HSV-1, HSV-2, HCMV, HHV-6, EBV, and

HHV-8 [79].

8

HISTORY AND UNIQUE FEATURES OF HHV-8

HHV-8, the eighth human herpesvirus to be identified, was discovered in 1994

using representational difference analysis [63]. HHV-8 DNA sequences were found in KS

lesions from an HIV-positive patient but not in adjacent unaffected tissue [63]. Shortly

thereafter, the virus was also causally linked to PEL and MCD [38,80].

HHV-8 encodes at least 90 ORFs within its genome of approximately 170 kb

(Figure 1.1) [34,63]. Guanine-cytosine (GC)-rich terminal repeat (TR) sequences are

present at both ends of the genome and comprise about 30 kb of genomic sequence [81].

The HHV-8 ORFs that are shared with herpesvirus saimiri (the first-sequenced gamma-2

herpesvirus) are designated ORF1 to ORF75 starting at the left end of the genome, and

the HHV-8-unique ORFs are designated K1-K15 and are located at particular regions of

gammaherpesvirus genome divergence (Table 1.1) [82]. The HHV-8-specific ORFs are

also numbered from left to right within the genome. In addition to these protein-encoding

ORFs, 12 microRNAs (miR-K12-1 to miR-K12-12) have recently been identified within the

HHV-8 genome (Tables 1.2). The majority (10) of these miRNAs are located between

ORF71 and K12, while the other two miRNAs are located within the 3’ untranslated region

(UTR) and coding sequence of K12 [83,84]. These miRNAs facilitate host immune system

evasion, cell cycle regulation, latency maintenance, inhibition of apoptosis, promotion of

angiogenesis, growth signaling, and chromatin modification [83,84].

HHV-8 LYTIC CYCLE

Immediately upon primary infection of a cell by HHV-8, a burst in lytic gene

expression is triggered and quickly aborted in favor of latency. Under normal conditions,

the virus remains latent until stimulated to reactivate. The mechanisms governing this

latent-to-lytic switch are unclear, though latently infected cultured cells can be

experimentally reactivated by the addition of phorbol esters or sodium butyrate (NaB)

9

[35,85-87]. In addition to chemically-induced reactivation, the lytic cycle can be reactivated

by biological stressors, including interferon-Ɣ, toll-like receptor 7 and 8 (TLR7 and TLR8)

agonists, reactive oxygen species (ROS), and the ER stress-activated X-box binding

protein 1 (XBP1) transcription factor [88-91].

Following reactivation or primary infection, lytic genes are transcribed and

expressed. These lytic genes are transcribed in a specific order and are categorized as

immediate-early, delayed-early, and late genes [92]. Immediate-early genes do not require

protein synthesis for their expression, whereas delayed-early genes require protein

synthesis but not DNA replication. Late genes require DNA replication and the expression

of immediate-early and delayed-early gene products for expression. Nine viral genes have

been classified as being expressed very early after butyrate induction, and 27 have been

classified as late genes [93-95].

The major immediate-early gene product, Rta (replication and transcriptional

activator, encoded by ORF50), is required for lytic replication and the induction of

downstream genes (delayed-early and late genes). Rta alone is sufficient to initiate the

lytic cascade, and inhibition of Rta prohibits lytic reactivation [96-99]. Ectopic expression

of Rta in PEL cells leads to the transcription of viral lytic genes [100]. Rta contains both a

DNA-binding domain and a transactivation domain and is detected within 4 hours (hrs) of

n-butyrate induction [92,101]. The DNA-binding domain of Rta allows the viral protein to

bind to viral DNA promoter sequences and to recruit cellular transcription factors, such as

RBP-Jk, via its transactivation domain [101]. Rta can also induce genes lacking Rta-

binding cis sequences via association with promoter-bound transcription factors including

C/EBPα; C/EBPα and Rta cooperatively activate the Rta and PAN promoters [101]. Rta

functions as a tetramer and can be observed as decamers in solution [102]. Genes that

are transcribed soon after Rta expression include ORF45, K3, K5, polyadenylated nuclear

(PAN) ribonucleic acid (RNA), vIL-6, and viral chemokine (C-C motif) ligand 2 (vCCL-

10

2/vMIP II); these genes can be detected within 13 hrs of n-butyrate induction [92,93].

Delayed-early genes include TK synthase, vCCL-1/vMIP I, viral G-protein-coupled

receptor (vGPCR), viral B-cell lymphoma 2 (vBcl-2), K12, and dihydrofolate reductase

(DHFR) [92]. The late genes are maximally expressed following viral DNA replication.

Lytic viral DNA replication occurs from two lytic origins (OriLyt-L and OriLyt-R) and

proceeds via a rolling circle mechanism [103]. Like other herpesviruses, HHV-8 encodes

a viral polymerase, polymerase processivity factor, primase, primase-associated factor,

helicase, and single-strand binding protein [82]. When these six gene products were co-

transfected into Vero cells, globular pseudo-replication complexes were observed in the

nucleus [104]. DNA replication is required for the expression of HHV-8 late genes such as

capsid proteins (e.g. ORF65) and other structural proteins found in the tegument [95].

Following expression of late genes, which occurs more than 30 hrs after n-butyrate

induction, viral capsid proteins are assembled, and viral DNA is loaded through the viral

portal protein [92]. Capsid assembly and loading occurs in the host cell nucleus,

necessitating a nuclear escape mechanism for efficient virion release. The nuclear egress

complex, which assists with virion egress from the nucleus, is comprised of the ORF67

and ORF69 gene products [105,106]. The process of virion envelopment and maturation

is complex, and HHV-8 encoded viral glycoprotein B (gB) plays an important role in this

process [107-109]. The virion gains a primary envelope and some tegument proteins as it

traverses the inner nuclear membrane [109]. The viral envelope then fuses with the outer

nuclear membrane as the virus moves across the perinuclear space; the virus particle

then deenvelopes, and the particle traverses the outer nuclear leaflet [109].

Reenvelopment and the acquisition of tegument proteins occur as the virus particle

traverses membranes of the trans-Golgi network [109]. Vesicles containing enveloped and

tegumented virus particles then fuse with the plasma membrane, releasing mature virions

[109]. Studies conducted using wild-type and gB-deleted HHV-8 bacmids demonstrated

11

that gB is required for the complete maturation of HHV-8 virions; cells infected with gB-

deleted HHV-8 bacmid did not produce completely enveloped viral particles [107]. After

escaping the cell, progeny virus can infect naïve cells, and the original host cell undergoes

apoptosis.

HHV-8 LATENCY

Viral productive replication is required for the infection of naïve cells and the

expansion of the viral reservoir. However, most infected cells remain latently infected

following primary HHV-8 infection. In B-cells, the virus is nearly completely latent within

hrs of de novo infection, as only a small subset of cells continues to produce virions. In a

study of KS, approximately 3% of cells were found to be lytic, while the remaining infected

cells were latent [92]. The latent virus is maintained as a circular episome within the host

cell nucleus [34,80,110,111]. During latency, a small number of viral genes are expressed,

allowing the virus to evade host immune responses that target viral gene products. Latent

gene products include latency-associated nuclear antigen (LANA), vFLIP, vCyclin, viral

microRNAs, and kaposins A, B, and C [112]. These latency products are encoded in a

cluster within the viral genome known as the latency locus. Two promoters are responsible

for the expression of these latency genes, the LANA promoter (drives expression of LANA,

vCyclin, and vFLIP) and the kaposin promoter (regulates expression of the kaposins and

viral microRNAs) [113,114]. Recently, studies have demonstrated that vIL-6 is also

expressed at low levels by latently infected PEL cells; expression of viral interferon

regulatory factors (vIRFs) 1, 2, and 3 have also been reported in latently infected PEL

cells [115-119].

LANA is a crucial protein of latency that functions to tether the viral episome to the

host chromosome through its interactions with TR regions of the HHV-8 genome and host

cell histone H2A/H2B proteins [120-123]. This tethering allows the HHV-8 genome to be

12

replicated during host cell DNA replication and ensures appropriate segregation of viral

genomes to daughter cell nuclei [120,124,125]. Plasmids containing TR regions can be

replicated in the presence of LANA but are not replicated in the absence of LANA,

indicating that LANA is required for episomal replication and maintenance

[123,124,126,127].

LANA expression regulates both host and viral gene transcription. Overexpression

of LANA results in the upregulation of DNA methyltransferase Dnmt3a, which is

responsible for methylating specific promoters, repressing the transcription of these

promoter-driven genes (e.g. cadherin 13) [128]. One study found 186 cellular genes that

were subject to greater than 2-fold positive or negative regulation by LANA; several of

these LANA-responsive genes are involved in the p53 and retinoblastoma protein (pRB)

pathways [129]. Moreover, LANA binds to and inactivates both the p53 and pRb tumor

suppressors, repressing apoptosis and driving E2F-dependent gene transcription

[130,131]. LANA also autoregulates its own transcription and the expression of Rta;

silencing of Rta by LANA inhibits lytic reactivation and promotes latency [132-134].

LANA achieves its pro-growth and survival functions by targeting multiple cellular

pathways in addition to p53 and pRb. LANA enhances levels of β-catenin via sequestration

of glycogen synthase kinase 3β (GSK-3β), ultimately leading to increased expression of

cyclin D and c-Myc [135]. Cellular Myc is an oncogene and a transcription factor

associated with growth, proliferation, and apoptosis [136]. LANA stabilizes c-Myc

expression by repressing the phosphorylation and ubiquitination of the transcription factor

[137,138]. Telomere length is also maintained by LANA, increasing the lifespan of virally-

infected cells [139]. Furthermore, LANA appears to be involved in the initial stages of

lymphogenesis. A transgenic mouse model expressing LANA driven by the LANA

promoter showed increased germinal center formation and an increased incidence of

13

lymphoma [140]. Thus, LANA is critical for the maintenance and replication of HHV-8

episomes, the promotion of infected cell growth, and likely contributes to lymphogenesis.

VIRAL INTERLEUKIN 6 AND PATHOGENESIS

Prior to the discovery of HHV-8, elevated levels of IL-6 were observed in KS and

MCD patients, and disease severity was correlated with IL-6 levels [25,40]. Moreover, the

growth of cells derived from an HIV-positive KS patient could be suppressed (reduced by

~67%) by the inhibition of IL-6 via antisense oligomers, indicating potential IL-6

involvement in the pathogenesis of KS [25]. Soon after the discovery of HHV-8, one of the

15 unique ORFs of the virus was found to encode a homolog of interleukin 6 [67,69,141].

Cellular IL-6 is involved in many normal physiologic processes, including hematopoiesis,

inflammation, cell growth, and cell survival. Early studies demonstrated that the vIL-6

homolog could substitute for hIL-6 in some experimental systems. vIL-6 could prevent

murine plasmacytoma B9 cell apoptosis and promote B9 cell proliferation; B9 cells are IL-

6-dependent [67,69]. Furthermore, antibodies specific to the IL-6 receptor inhibited hIL-6-

and vIL-6-mediated growth of B9 cells in culture while control antibodies specific for IL-2

had no effect on cell growth [69]. hIL-6 is also known to induce the expression of acute-

phase proteins in hepatocytes [142]. Thus, cell culture-based methods were employed to

determine whether vIL-6 could similarly promote acute-phase protein production [69]. The

Hep3B hepatocyte cell line was transfected with plasmids encoding vIL-6, hIL-6, or control

sequences, and Northern blots were conducted using harvested RNA from the transfected

cells [69]. Both human and viral cytokines could induce the messenger RNA (mRNA)

expression of α1 acid glycoprotein, a prototypical acute-phase protein [69]. Thus, at the

time of its discovery, many surmised that vIL-6 might play a role in the development and

progression of HHV-8-associated disease [67-69].

14

Following the discovery of vIL-6, clinical and other data were reported that

implicated vIL-6 as an important factor in the progression of KS, PEL, and MCD. Elevated

levels of serum vIL-6 were correlated with episodes of acute inflammation in MCD patients

[143]. Additionally, high levels of both hIL-6 and vIL-6 were observed in PEL tumors, and

these cytokines were required for PEL cell growth [115,144]. Depletion of vIL-6 via a short

hairpin RNA (shRNA)-encoding lentiviral vector and inhibition of vIL-6 using neutralizing

antibodies led to decreased PEL growth in culture [115,144]. The engraftment of PEL

tumors in severe combined immunodeficiency (SCID) mice could be inhibited by the

inhibition of vIL-6 via morpholino oligomers specific to vIL-6 [145]. In KS tumors, vIL-6 can

induce the expression of angiogenic factors (including VEGF), which are critical in the

development of KS [146]. In these studies, vIL-6-expressing NIH3T3 fibroblasts were

injected into nude mice, and these mice were compared to control mice that received

injections of NIH3T3 cells expressing control plasmid; vIL-6-expressing mice exhibited

increased hematopoiesis in specific lineages (myeloid, erythroid, and megakaryocytic),

greater vascularization, and polyclonal hypergammaglobulinemia [147]. In NIH3T3 cells

expressing vIL-6, VEGF expression could be detected by immunofluorescence staining;

in control cells, VEGF was not detected [147]. In a mouse model of peritoneal

inflammation, vIL-6 was found in enhance CCL-2 secretion and inhibit CXCL-8 expression,

ultimately blocking the recruitment of neutrophils to the site of inflammation [148]. Finally,

vIL-6 can also enhance the secretion of hIL-6 [149]. Plasmids encoding vIL-6 or the

reverse vIL-6 sequence were transfected into cells, and culture medium was collected for

use in experimental assays [149]. Collected culture medium was then added to several B

and T cell lines, and hIL-6 levels were assessed; the introduction of vIL-6 led to significant

increases (up to 30-fold) in hIL-6 expression in five cell lines (MT-4, THP-1, U937, Raji,

and CESS) [149]. Additionally, vIL-6-containing culture medium was capable of inducing

a 2-fold increase in hIL-6 secretion in cell lines derived from MCD patients [149].

15

Other studies have further characterized the role of vIL-6 in HHV-8 pathogenesis.

vIL-6 is expressed maximally during lytic replication, though it is also expressed at low,

though functional, levels during PEL latency [115,116]. vIL-6 transcripts were identified via

limiting dilution reverse transcription-polymerase chain reaction (RT-PCR) and array-

based methods in HHV-8 latently infected SLK cells (KS-derived, HIV-) [116].

Immunofluorescence staining of HHV-8-positive PEL cells for vIL-6 indicates that the viral

cytokine is expressed during latency; vIL-6-specific immunofluorescence staining in PEL

cells was detected above background staining in co-mixed HHV-8-negative Akata cells

[115]. Depletion of vIL-6 via shRNA-mediated lentiviral transduction led to decreased

growth and survival of PEL cells in culture [115]. Following vIL-6 depletion, PEL cell growth

could be rescued by the addition of fully ER-retained, shRNA-resistant vIL-6 [115]. Similar

results were observed when vIL-6 was neutralized using an ER-directed vIL-6-specific

antibody (MAV) or peptide-conjugated phosphorodiamidate morpholino oligomers

(PPMO) directed to vIL-6 [150,151]. These studies are significant because they indicate

that vIL-6 plays an important role in cell growth and viability during latency when the

protein is expressed at much lower levels than during lytic reactivation.

vIL-6 is known to activate both mitogen-activated protein kinase (MAPK) and signal

transducer and activator of transcription (STAT) signaling (discussed in more detail

below), and high levels of active (phosphorylated) STAT3 have been observed in PEL

cells [152]. Several PEL cells lines, including BC-1, BCBL-1, and VG-1, express

constitutively phosphorylated STAT3 [152]. Furthermore, inactivation of STAT3 (via

treatment with JAK2 pharmacological inhibitor AG490 or the overexpression of dominant-

negative STAT3 [STAT3D]) led to decreased PEL cell growth, increased apoptosis, and

decreased expression of survivin, a protein that is required for PEL cell survival [152].

STAT3D contains a mutation within the DNA binding domain of STAT3; thus, STAT3D

can dimerize with wild-type STAT3 but cannot bind STAT3-responsive DNA elements to

16

initiate transcription, effectively acting as a competitive inhibitor of wild-type STAT3

[152,153]. Survivin is an inhibitor of apoptosis (IAP) family member, and its inhibition has

been implicated in several cancers [154]. Forced overexpression of survivin in VG-1 PEL

cells was capable of rescuing AG490-mediated apoptosis and STAT3 inhibition,

confirming the importance of STAT3 and survivin for PEL cell viability [152]. Importantly,

STAT3 can also be activated by VEGF in bovine aortic endothelial cells, and VEGF is

upregulated by vIL-6 in other cell culture models [147,155]. Thus, vIL-6 can be linked to

pro-growth and survival as well as pro-angiogenic pathways, further implicating the viral

cytokine in HHV-8 malignant pathogenesis.

VIRAL INTERLEUKIN 6 CHARACTERISTICS

hIL-6 and vIL-6 share approximately 25% amino acid sequence identity and 63%

similarity [67-69]. While the primary sequences of the homologs are quite divergent, both

cytokines fold into similar three-dimensional structures consisting of 4-alpha helical

bundles (Figure 1.2) [156]. This structural similarity allows both cytokines to engage the

IL-6 α receptor (IL6R or gp80) and gp130 (β subunit) and initiate signaling, leading to the

induction of MAPK and STAT signaling (Figure 1.3) [156-160]. Phosphorylation of tyrosine

(Y) 759 (Y759) of gp130 results in MAPK signaling, while the phosphorylation of Y767,

Y804, Y905, or Y915 leads to the initiation of STAT signaling [156,161-163]. Ultimately,

hIL-6 and vIL-6 signaling lead to cell growth and survival, and both cytokines can sustain

IL-6-dependent B9 cells in culture [67,141]. While both proteins can initiate signaling via

interactions with gp80 and gp130, hIL-6 and vIL-6 have different requirements for signaling

induction. hIL-6-containing signaling complexes comprise two molecules each of hIL-6,

gp80, and the gp130 signal transducer (hIL-62-gp802-gp1302) [156-160]. vIL-6 can also

form identical hexameric complexes but is unique among IL-6 proteins in that it can signal

in the absence of gp80 via tetrameric complexes (vIL-62-gp1302) [158,164-166]. While

17

gp80 is not required for vIL-6-initiated signaling, the presence of gp80 facilitates a more

sustained signaling event [159,167,168]. Hexameric (gp80-containing) signaling

complexes are found at the plasma membrane (PM), while (unique) vIL-6 signaling from

the ER occurs exclusively through tetrameric (gp80-devoid) complexes [115,169]. vIL-6-

containing tetrameric complexes may also form at the cell surface. Thus, hIL-6 can signal

at the cell surface, while vIL-6 can signal from the cell surface and intracellularly within the

ER. These distinct signaling mechanisms allow vIL-6 to signal in both a paracrine and

strictly autocrine (intracellular) manner.

In addition to differences in signaling complex composition, vIL-6 is mostly retained

within the ER compartment, while hIL-6, like other cellular cytokines, is efficiently secreted

from the cell [115,170,171]. Secretion kinetics indicate that vIL-6 is secreted approximately

8-fold slower than hIL-6 [170]. Within the ER, vIL-6 interacts with calnexin, which is

involved in proper folding of the cytokine [171]. Depletion of calnexin led to reduced levels

of intracellular vIL-6 but did not alter levels of secreted vIL-6; vIL-6 levels could be rescued

by overexpression of gp130 in calnexin-depleted cells, but the underlying mechanism is

unclear [171]. Furthermore, vIL-6 contains two potential asparagine glycosylation sites

(asparagines [N] 78 and 89), and secreted vIL-6 is completely glycosylated [170,172].

Glycosylation of vIL-6 N89 is required for the induction of vIL-6-initiated signaling but not

for its interaction with calnexin, for which N78 glycosylation is sufficient [171,172]. Thus,

posttranslational modification differences between the human and viral cytokines may

partially explain differences in secretion kinetics and interactions with ER-resident

proteins.

vIL-6, but not hIL-6, was found to bind directly to a previously uncharacterized

protein called vitamin K epoxide reductase complex subunit 1 variant 2 (VKORC1v2) in

addition to calnexin [173]. The major isoform, VKORC1v1 (vitamin K epoxide reductase

complex subunit 1 variant 1), plays a role in the vitamin K cycle and is the target of the

18

anticoagulant warfarin [174]. Evidence demonstrating that VKORC1v2 interacts with vIL-

6 was the first reported function of VKORC1v2. VKORC1 variants 1 and 2 share the first

58 amino acids, and splicing gives rise to variant 2 [173]. Only VKORC1v2 interacts with

vIL-6, though the interaction domain (vIL-6 binding domain [vBD], residues 31-39) of

VKORC1v2 is shared with VKORC1v1 [173]. The difference in vIL-6-binding between the

two variants can be explained by the opposing orientations of the two proteins in the ER

membrane. Both VKORC1 variants contain a transmembrane domain between amino

acids 11 to 29 [173]. For VKORC1v1, the N-terminus of the protein is located in the ER

lumen, and amino acids 31-39 (vBD) are located in the cytoplasm [173]. Conversely, the

N-terminus of VKORC1v2 is located in the cytoplasm, and amino acids 31-39 are located

within the ER lumen and are accessible to vIL-6 [173]. A third VKORC1 variant,

VKORC1v3, is similar to VKORC1v1 but has a deletion of amino acids 58-73 due to

alternative splicing, and VKORC1v3 is also incapable of binding to vIL-6 [175]. Studies

have shown that depletion of VKORC1v2, but not VKORC1v1, leads to reduced rates of

PEL cell growth and increased rates of apoptosis in these cells, identifying the importance

of VKORC1v2 for PEL cell viability [173]. Growth of VKORC1v2-depleted cells could be

rescued by the addition of shRNA-resistant wild-type but not vBD-mutated VKORC1v2

[173]. Similarly, overexpression of vBD led to reduced growth and increased apoptosis in

PEL cultures, presumably due to disruption of the vIL-6/VKORC1v2 interaction [173].

These findings indicate the importance of the vIL-6/VKORC1v2 interaction for PEL cell

viability.

The interaction between vIL-6 and VKORC1v2 can enhance VKORC1v2

interactions with the proenzyme form of cathepsin D (CatD) [175]. CatD, in its mature,

lysosomally localized form, is an aspartate protease, which can be released into the

cytoplasm in response to certain stress signals [176]. Depletion of vIL-6 results in

increased levels of ER-localized/transiting pro-CatD and mature CatD, leading to

19

decreased PEL cell viability in culture [175]. Introduction of vIL-6/VKORC1v2-disrupting

vBD also led to increased levels of CatD and reduced PEL cell viability [175]. PEL cell

apoptosis could be induced with the overexpression of CatD, suggesting that vIL-6 may

suppress CatD function to maintain cell viability [175]. PEL cell viability is critical for

efficient virus replication, and depletion of vIL-6 or introduction of vIL-6/VKORC1v2-

disrupting vBD in lytically reactivated BCBL-1 and JSC-1 PEL cells led to increased levels

of CatD and decreased HHV-8 viral titers [175]. In contrast, depletion of CatD in these

PEL cell lines led to increased viral titers [175]. These results suggest a role for vIL-6 in

HHV-8 replication via its interactions with VKORC1v2 and regulation of CatD.

SIGNIFICANCE OF VIRAL IL-6 AND THIS PROJECT

In order to develop therapeutic agents for the treatment of HHV-8-associated

disease, it will be crucial to develop a better understanding of the mechanisms governing

vIL-6-mediated pathogenesis. Several research questions must be addressed before

beginning the process of selecting therapeutic targets. First, the role of ER-localized vIL-

6/gp130 interactions must be investigated. vIL-6/gp130-mediated signaling likely activates

STAT and MAPK pathways, though this has not been shown in the context of PEL.

Furthermore, the mechanism(s) responsible for elevated levels of active STAT3 in PEL

tumors should be elucidated. In addition to its proposed roles in HHV-8 latency, functions

of the vIL-6/gp130 interaction in the context of virus replication are unclear. The focus of

this dissertation project was to functionally characterize ER-localized vIL-6/gp130

interactions in the context of both viral latency and virus productive replication.

20

TABLES

Table 1.1. Descriptions of HHV-8-specific open reading frames (“K” ORFs).

ORF Encoded Protein Protein Function Reference

K1 VIPa activation of NFĸB; pro-angiogenic; downregulation of B-cell

receptor; cytokine induction

[177-182]

K2 vIL-6 pro-growth; pro-survival; pro-angiogenic; pro-virus replication [67-69,115,152,169,175,183]

K3 K3 protein downregulation of MHCb class I; NKc cell and CTLd evasion [94,184-186]

K4 vCCL-2 pro-angiogenic; pro-survival; pro-virus replication; Th2

polarization

[187-193]

K4.1 vCCL-3 pro-angiogenic; Th2 polarization [193,194]

K5 K5 protein downregulation of MHC class I; NK and CTL evasion [94,184-186,195]

K6 vCCL-1 pro-angiogenic; pro-survival; pro-virus replication; Th2

polarization

[187,196]

K7 vIAPe pro-survival [197-199]

K8 K8.1A/K8.1B target cell recognition; heparin binding [200,201]

K9 vIRF1 pro-survival; inhibition of interferon signaling [202-211]

K10 vIRF4 p53 destabilization [202,212,213]

K10.5 vIRF3 pro-angiogenic; pro-survival; inhibition of IFN signaling [117,202,214-217]

K11 vIRF2 inhibition of interferon signaling [119,202,218]

K12 kaposin A, B, & C prevention of cytokine mRNA decay; activation of MAPK

signaling; PROX1 stabilization

[219-223]

K13 vFLIPf anti-apoptotic via inhibiting caspase-8 activation; pro-survival

via NFĸB activation

[78,224-226]

K14 K14 cell surface protein

interacts with CD200 receptor; downregulation of TNFα and myeloid cell production

[227]

K15 K15 membrane protein

Pro-growth; pro-survival; pro-angiogenic

[228-231]

a Variable immunoreceptor tyrosine-activation motif (ITAM) protein; b major histocompatibility complex; c natural killer; d cytotoxic T lymphocyte; e viral inhibitor of apoptosis; f viral FADD-like interleukin-1 beta-converting enzyme (FLICE) inhibitory protein

21

Table 1.2. HHV-8 microRNAs.

microRNA Targets Function Reference

miR-K12-1 p21, caspase 3, THBS1, IkBα

inhibition of cell-cycle arrest; anti-apoptotic; pro-angiogenic; suppression of viral replication; induction of MAPK signaling [232-235]

miR-K12-2 unknown upregulated in KS lesions [236]

miR-K12-3 caspase 3, NFIB, THBS1 anti-apoptotic; pro-latency; pro-angiogenic [234,235,237]

miR-K12-4 caspase 3, Rbl2 anti-apoptotic; epigenetic modifications [234,238]

miR-K12-5 BCLAF1, Rta, MyD88 anti-apoptotic; pro-latency; suppression of inflammatory cytokine production [238-240]

miR-K12-6 MAF, THBS1 cell fate reprogramming; pro-angiogenic [20,235]

miR-K12-7 Rta, MICB pro-latency; NK evasion [241,242]

miR-K12-8 unknown upregulated in KS lesions [236]

miR-K12-9 BCLAF1, Rta, IRAK1 anti-apoptotic; pro-latency; decreased inflammatory cytokine production [239,240,243]

miR-K12-10 TWEAKR, BCLAF1 repression of caspase activation; anti-apoptotic [239,244]

miR-K12-11 BACH1, SMAD5, MAF, THBS1

pro-survival; suppression of TGF-β signaling; cell fate reprogramming; pro-angiogenic; induction of MAPK signaling

[20,235,245-248]

miR-K12-12 unknown upregulated during lytic reactivation [249,250]

22

FIGURES

Figure 1.1. Genomic organization of HHV-8. Gammaherpesvirus-conserved “core” gene

blocks are shown in blue. Divergent loci (DL-A through DL-D and IRF locus) outside of the

conserved blocks are labelled. Signaling membrane receptors (pink), cytokines (blue),

interferon regulatory factors (orange), and latency genes (green) are highlighted. The K1,

vOX2/vCD200 (K14), and K15 receptors, the viral cytokines, and the viral IRFs have thus

far been identified only in HHV-8 and closely-related macaque rhadinoviruses [251-253].

K7 (survivin) and K12 (kaposin) are unique to HHV-8. Latency genes encoding LANA,

vCyclin, and vFLIP are present in and characteristic of other gamma-2 herpesviruses. .

Abbreviations: see Table 1.1; modulator of immune recognition 2 (MIR2), thymidylate

synthase (TS), modulator of immune recognition 1 (MIR1), viral B-cell lymphoma 2 (vBcl-

2), viral chemokine ligand (vCCL), orexin receptor 2 (OX-2).

23

Figure 1.1

24

Figure 1.2. vIL-6 and hIL-6 three-dimensional structures. hIL-6 and vIL-6 are composed

of four alpha helices (A-D) and fold into similar three-dimensional structures. This

structural similarity allows both cytokines to initiate signaling through complexes

containing the IL-6α receptor and gp130 signal transducer.

Figure 1.2

The above figure has been published previously and is reproduced here with permission from the American Society for Microbiology (Copyright © American Society for Microbiology, [Journal of Virology, 2012, 86(3):1577-1588, doi: 10.1128/JVI.05782-11]). Chen D, Cousins E, Sandford G, and Nicholas J (2012). Human herpesvirus 8 viral interleukin-6 interacts with splice variant 2 of vitamin K epoxide reductase complex subunit 1. 86(3):1577-1588.

25

Figure 1.3. Human and viral interleukin 6 signaling complexes. Human IL-6 (hIL-6) initiates

signaling through hexameric signaling complexes containing two molecules each of hIL-

6, gp80, and gp130; these complexes form exclusively at the plasma membrane. Viral IL-

6 (vIL-6) is capable of signaling via identical hexameric complexes, though vIL-6 is unique

in its ability to signal via tetrameric complexes (vIL-62-gp1302) within the endoplasmic

reticulum (ER). Following complex formation, tyrosine (Y) residues within the cytoplasmic

tail of gp130 are phosphorylated. Depending on the specific tyrosine phosphorylation

event, signal transducer and activator of transcription (STAT) 1 or 3 or mitogen-activated

protein kinase (MAPK) signaling cascades are activated downstream of gp130.

26

Figure 1.3

27

Figure 1.4. Known and proposed functions of viral interleukin 6. Viral interleukin 6 (vIL-6)

is a multifunctional viral protein and is active during both the latent and lytic phases of the

virus life cycle. During latency in PEL cells, vIL-6 acts as an autocrine factor to promote

cell growth and survival; it may perform a similar role in other cell types. During the lytic

phase, vIL-6 is thought to promote pathogenically significant angiogenesis and

inflammation through induction of cellular factors, including VEGF and hIL-6. vIL-6 inhibits

neutrophil recruitment via downregulation of CCL-2. These activities are believed to

promote virus replication in vivo. The hypotheses that vIL-6 autocrine signaling via gp130

can contribute to latently infected cell viability and to productive replication form the basis

of the thesis project.

28

Figure 1.4

29

CHAPTER 2: METHODS

30

Summary

Many laboratory techniques were utilized to conduct the studies described in this

dissertation. Detailed methods are provided in the chapters describing the studies in

which they were used.

31

INTRODUCTION

In order to answer the research questions posed in this body of work, many

techniques were employed. Several of the techniques have been well-described by others,

but other protocols were specifically developed in the course of this project. Classic and

novel techniques are described in general terms below; more detailed methods are

provided in the relevant chapters that follow.

RNA EXTRACTION

HHV-8 and cellular mRNA were extracted from cell cultures in order to quantify

gene expression levels. To extract RNA, cells were washed with phosphate-buffered

saline (PBS) and suspended in TRizol (Thermo Fisher Scientific, Inc., Waltham, MA).

Following a 3 minute (min) incubation in TRizol, chloroform was added to the mixture, and

the samples were vortexed briefly and incubated for 10 min at room temperature. Samples

were then centrifuged in a microcentrifuge (10 min at maximum speed at room

temperature), and the top layer was removed and placed in a clean Eppendorf tube.

Isopropanol was added to precipitate RNA, and samples were centrifuged in a

microcentrifuge (10 min at maximum speed at 4°C) to pellet the RNA. Finally, RNA pellets

were washed with 70% ethanol and resuspended in RNase- and DNase-free water.

Concentrations of RNA samples were determined via spectrophotometer.

REVERSE TRANSCRIPTION

After viral and cellular RNA were extracted from cultured cells, the RNA was

reverse-transcribed to complementary deoxyribonucleic acid (cDNA) to allow amplification

of a gene of interest. The reverse-transcription reaction requires a pure RNA template

devoid of contaminating DNA molecules. As a result, the extracted RNA was treated with

DNase I (45 min incubation at room temperature) to destroy contaminating DNA

32

molecules; ethylenediaminetetraacetic acid (EDTA) was subsequently used to inactivate

the DNase I enzyme by chelating magnesium and calcium required for enzyme activity.

Random hexamer primers and deoxynucleotide triphosphates (dNTPs) were added to

DNase-treated samples. Samples were then heated (65°C for 5 min) and cooled on ice (5

min); this process facilitated hexamer primer annealing. Finally, dithiothreitol (DTT;

Invitrogen, Carlsbad, CA), reverse transcriptase buffer (5X First Strand Buffer; Invitrogen),

RNase inhibitor (RNaseOUT; Invitrogen), and reverse transcriptase (SuperScript II RT;

Invitrogen) were added to the samples. RNA samples were then reverse-transcribed into

cDNA by incubating at 25°C for 10 min, 42°C for 50 min, and 70°C for 15 min. cDNA

samples were stored at -20°C until further processing.

POLYMERASE CHAIN REACTION

Polymerase chain reaction (PCR) was used to amplify a DNA template in order to

create vast numbers of copies of a given amplicon. PCR products are more stable than

cDNA derived from reverse-transcription and can be further analyzed or manipulated.

PCR reactions included a DNA template, dNTPs, forward and reverse primers,

polymerase buffer, and DNA polymerase. PCR amplification times depended on the length

of the PCR product and the polymerase used. Generally, PCR reactions included

approximately 50 nanograms of DNA template, 4 μl of dNTPs (2.5 mM), 1 μl each of

forward and reverse primers (10 μM each), 1X polymerase buffer, 1.5 μl of MgCl2 (50 mM),

and 1 unit (U) of Taq polymerase isolated from Thermus aquaticus (Invitrogen) in 1X

polymerase buffer. For cloning purposes, 2.5 U of PFU Ultra High-Fidelity polymerase

(Stratagene, La Jolla, CA) was used, and MgCl2 was not added. PFU Ultra High-Fidelity

polymerase is less error-prone than Taq polymerase and is therefore more suitable for

cloning.

33

Amplification using Taq polymerase was conducted as follows: 94°C for 2 min

(denaturing); 25-35 cycles of 94°C for 30 sec (denaturing), 55°C for 30 sec (annealing),

and 72°C for 1 min per kb of DNA template (extension); and a final extension of 72°C for

10 min. Samples were then cooled to 4°C and stored at -20°C until they were used for

further analysis. A similar protocol was employed for reactions containing PFU Ultra High-

Fidelity enzyme: 95°C for 2 min; 25-35 cycles of 95°C for 30 sec, 50-55°C for 30 sec, 72°C

for 1 min per kb of template DNA (for vector templates <10 kb); followed by a final

extension of 72°C for 10 min. For vector templates greater than 10 kb, an extension time

of 2 min per kb and an extension temperature of 68°C were used. PCR amplification was

confirmed and visualized by agarose gel electrophoresis (described below).

AGAROSE GEL ELECTROPHORESIS

PCR amplification was verified by running PCR products on an agarose gel with

an appropriate DNA size marker. PCR samples were mixed with DNA loading buffer (40%

sucrose and 0.25% bromophenol blue) and loaded into agarose gel lanes. DNA size

marker was loaded into a gel lane as a reference to determine the size of PCR products.

As DNA is negatively charged, PCR products migrated from the negatively charged pole

toward the positively charged pole through the gel matrix. Ethidium bromide was included

in the agarose gel and intercalated into the DNA duplexes, allowing for the detection of

labeled DNA fragments by ultraviolet (UV) light. Smaller DNA fragments migrated through

the gel more rapidly than larger DNA fragments; thus, DNA was separated based on size.

While agarose gel electrophoresis was used to determine PCR product size,

relative abundance was also assessed in some instances. For example, RT-PCR was

often used to determine whether a specific shRNA adequately repressed its target gene

expression. In this case, RT-PCR-derived DNA samples from control and shRNA-treated

cells were compared on an agarose gel. Amplified target gene DNA should be less

34

abundant than amplified target gene DNA from control cells if the shRNA used was

effective. Thus, target gene expression was semi-quantified by agarose gel

electrophoresis.

CLONING AND SHORT HAIRPIN RNA DESIGN

Several plasmid constructions were prepared in order to address questions related

to vIL-6 functions in cell growth, apoptosis, and virus replication. Some genes were cloned

with tags (e.g. FLAG) to facilitate detection, while others were cloned with localization

sequences (e.g. KDEL motif) to restrict the gene product to specific cellular locations.

Gene segments were also cloned (e.g. soluble gp130 [sgp130]) and introduced into cells

to inhibit the interaction between vIL-6 and gp130. Additionally, shRNAs were cloned into

lentiviral vectors to deplete endogenous cellular or viral proteins in cultured cells. Finally,

shRNA-resistant genes were cloned into lentiviral vectors for use in depletion-

complementation experiments. Specific details regarding the primers, restriction enzyme

sites, and vectors utilized for these constructions are provided in the relevant chapters.

In general, genes were amplified from various vector templates using PFU Ultra

High-Fidelity polymerase and then purified. In brief, PCR products were precipitated from

PCR reaction mixtures with sodium acetate and ethanol to remove primers, excess

dNTPs, polymerase, and reaction buffer components. Purified PCR products were then

digested with restriction enzymes to create appropriate 5’ and 3’ overhangs (or left

undigested in the event of a blunt-ended ligation). Restriction enzyme-digested PCR

products and vectors were then separated on an agarose gel, and the appropriate DNA

bands were excised from the gel under the guidance of UV light. DNA was purified from

the agarose gel using a DNA gel extraction kit (QIAquick Gel Extraction Kit; Qiagen,

Valencia, CA). Gel-purified DNA vector and PCR product insert were then ligated using

35

T4 DNA ligase (New England Biolabs; Ipswitch, MA) for 1 hr at room temperature or

overnight at 16°C.

Ligated vector-insert constructions were transformed into competent DH5α E. coli

cells via a chemical transformation technique. Briefly, ligated DNA constructions were

incubated with competent bacterial cells on ice for 30 min prior to heat shocking (42°C)

for 30-60 seconds. Ligation product-bacterial mixtures were then placed on ice briefly and

plated on Luria broth (LB) agar plates with appropriate antibiotics for selection of

transformed bacteria. Depending on the plasmid used in the transformation, agar plates

were incubated at 30°C or 37°C overnight to allow for the growth of transformed bacteria.

Individual bacterial colonies were picked from plates, amplified in liquid LB, and

miniprepped to extract bacterial DNA. Colony-derived DNA was used for PCR-based or

restriction enzyme-based techniques to validate cloning constructions. Bacterial-derived

DNA was PCR amplified using construct-specific primers to verify the presence of the

DNA insert, DNA signal sequence, or sequence tag. Restriction enzyme digestion was

also used to verify the insertion of restriction enzyme-containing DNA inserts, signal

sequences, or tags. Prior to use in experiments, all constructions were commercially

sequenced to verify the modified DNA sequence.

Several experiments utilized shRNA-encoded lentiviral vectors to deplete

endogenous viral or host proteins. Proteins were targeted by 19 nucleotide sequences

embedded in a characteristic hairpin structure. Targeting sequences were designed with

the aid of iRNAi (version 2.1; Nucleobytes, The Netherlands). Complementary

oligonucleotides encoding candidate target sequences were annealed, phosphorylated,

and ligated into the pYNC352 lentiviral vector between the BamHI and MluI restriction

sites. shRNA-encoded lentiviral vectors were then sequenced. To assess depletion

efficiency of specific shRNA-encoded lentiviral vectors, vectors were co-transfected into

293T cultured cells with overexpressed target protein. Twenty-four hrs post-transfection,

36

cell lysates were harvested and subjected to western blot analysis to determine protein

levels in control and shRNA-overexpressing cells (described below). In some cases, target

mRNA depletion was assessed via RT-PCR following shRNA-encoding lentiviral

overexpression (described above).

MINIPREP DNA EXTRACTIONS

Transformed bacteria were amplified in liquid LB overnight at 30°C or 37°C

depending on the transformed plasmid and harvested for the extraction of plasmid DNA.

Bacterial cells were pelleted via centrifugation, and cell pellets were resuspended in STET

buffer (8% sucrose, 5% Triton X-100, and 50 mM EDTA, Tris [pH 8.0]). Samples were

vortexed prior to the addition of lysozyme (100 μg/ml) and then boiled for 1 min and

centrifuged at maximum speed for 10 min at room temperature. Supernatants were

transferred to clean tubes, and an equal volume of isopropanol was added to each tube

to precipitate DNA. Tubes containing sample mixtures were chilled at -20°C for 10 min

and then centrifuged at maximum speed for 10 min at 4°C to pellet DNA. DNA pellets were

washed in 70% ethanol and finally resuspended in DNase-free water containing RNase A

to eliminate contaminating RNA.

CELL CULTURE

HEK293T cells were maintained in Dulbecco’s Modified Eagle’s medium (DMEM)

supplemented with 10% heat-inactivated fetal bovine serum (FBS). BCBL-1,

TRExBCBL1-RTA, JSC-1, BJAB, BC-1, and BC-3 cells were maintained in RPMI 1640

media supplemented with 10% FBS. Telomerase-immortalized microvascular endothelial

(TIME) cells and TIME-TRE/RTA cells were grown in endothelial growth basal media

(EBM-2) containing FBS and supplements (Lonza; Walkersville, MD). All cells were grown

in a 5% CO2 humidified chamber maintained at 37°C.

37

TRANSFECTION OF CULTURED CELLS

Transfection of cultured HEK293T cells was utilized for functional experiments

assessing viral and cellular protein-mediated signaling, shRNA depletion efficiency,

protein-protein interactions, and lentivirus production. Transfections of HEK293T cells

were conducted using standard calcium-phosphate protocols. Briefly, HEK293T cells were

plated such that they would achieve 50% confluence at the time of transfection. DNA

plasmids were then mixed with calcium phosphate (2.5 M) and 2X HEPES-buffered saline

(HBS; 12 mM dextrose, 50 mM HEPES, 10 mM KCl, 280 mM NaCl, and 1.5 mM

Na2HPO4·2H2O, pH 7.05), vortexed, incubated at room temperature for 5 min, and added

to HEK293T cell culture media. Following 8 hr of incubation, culture medium was replaced

with fresh DMEM containing 10% FBS. Transfection efficiency was determined for

plasmids encoding green fluorescent protein (GFP) by visualizing GFP fluorescence. To

assess the transfection efficiency of plasmids not encoding fluorescent markers, a subset

of transfected cells were fixed and stained for transfected proteins via

immunohistochemical methods (described below).

Transfected cells were harvested 24-48 hr post-transfection for either RNA or cell

lysates. RNA was harvested using the TRizol method described above. For preparation

of cell lysates, cells were washed with PBS, scraped from the surface of the well or flask,

and incubated in lysis buffer (20 mM Tris [pH 8.0], 5 mM EDTA, 50 mM NaCl, 0.2% NP40,

and 1X protease and phosphatase inhibitors [Sigma; St. Louis, MO]) for 30 min on ice.

Samples were then centrifuged (maximum speed for 5 min at 4°C in a microcentrifuge) to

pellet cellular debris, and supernatants (lysates) were collected and stored at -20°C until

western blot analysis.

38

WESTERN BLOT ANALYSIS

Western blot analysis was conducted to observe signaling events in transfected

HEK293T cells or transduced PEL cell cultures. Western blot analysis was also performed

to determine the efficiency of shRNA-mediated depletion in HEK293T, PEL, and TIME cell

cultures. Preparation of cell lysates used for western blot analysis was described above.

Specific antibodies and product numbers are provided in the methods section of the

relevant chapters.

Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) was

performed to separate lysate proteins via electrophoresis. Similar to agarose gel

electrophoresis, proteins were separated based on size. Smaller proteins migrated

through the gel matrix more rapidly than larger proteins. Depending on the size of the

target protein, 8-12% polyacrylamide gels were used. Smaller target proteins were more

efficiently separated on higher percentage (e.g. 12%) polyacrylamide gels, whereas larger

proteins were better separated on lower percentage (e.g. 8%) gels. SDS loading buffer

(188 mM Tris-HCl [pH 6.8], 3% SDS, 30% glycerol, 0.01% bromophenol blue, and 15% ß-

mercaptoethanol) was added to each protein sample prior to denaturation via boiling (5

min), and a protein size marker was loaded onto the gel as a reference. Following gel

electrophoresis, proteins were electrophoretically transferred onto nitrocellulose

membranes using standard protocols. Membranes were then washed in TBS-T (50 mM

Tris-HCl [pH 7.4], 150 mM NaCl, and 0.1% Tween 20) and blocked with 5% nonfat milk in

TBS-T for 1 hr at room temperature. After briefly washing blocked membranes in TBS-T,

primary antibodies (diluted in SuperBlock Blocking Buffer in TBS; Thermo Scientific,

Rockford, IL) were incubated with membranes for 1 hr at room temperature or overnight

at 4°C. After primary antibody incubation, membranes were washed three times with TBS-

T and then incubated with horseradish peroxidase (HRP)-conjugated secondary antibody

(diluted in 5% non-fat milk in TBS-T) for 1 hr at room temperature. Membranes were

39

continuously rocked during both antibody incubations. Again, membranes were washed

three times in TBS-T and then developed using standard chemiluminescence procedures

with ECL reagents (GE Healthcare Bio-Sciences; Pittsburgh, PA). Images were captured

digitally on a GeneGnomeXRQ machine (Syngene; Frederick, MD).

IMMUNOFLOURESCENCE STAINING

Immunofluorescence staining was performed to determine transfected or

transduced protein expression, to visualize cellular localization of target proteins, and to

calculate the percentage of apoptotic cells, proliferating cells, and LANA-positive cells. To

prepare cells for immunofluorescence staining, cells were washed in PBS, plated on glass

slides, and allowed to air dry. Once dry, slides were fixed in ice cold methanol for 10 min.

Slides were then allowed to air dry completely.

Following fixation, primary antibody (diluted 1:100 in SuperBlock Blocking Buffer

in PBS; Thermo Scientific) was added to slides and incubated in a humidified chamber at

37°C for 1 hr. Primary antibody was then removed, and the slide was washed three times

with PBS. Secondary antibody (diluted 1:200 in SuperBlock Blocking Buffer in PBS) was

then applied to the slide, and the slide was incubated for 1 hr as before. After washing

three times with PBS, nuclei were stained with 4’,6’-diamidino-2-phenylindole (DAPI).

Finally, coverslips were mounted on slides prior to visualization on a UV microscope.

BRDU INCORPORATION

To determine the percentage of dividing (S-phase) cells, 5-bromo-2-deoxyuridine

(BrdU) incorporation was assessed. BrdU is a thymidine analog and is incorporated into

replicating DNA; thus, BrdU incorporation occurs exclusively in actively replicating cells.

BrdU (0.01 M diluted in culture media, 1000X) was added to cell cultures, and the

cultures were incubated for 2 hr at 37°C; cells were then washed in PBS, spotted on glass

40

slides, and allowed to air dry. Slides were fixed in methanol as described above and

washed in PBS. Cells were then permeabilized with 2M hydrochloric acid (HCl) for 20 min

at room temperature, and sodium borate (0.1 M) was then added to the slides for 2 min to

neutralize HCl. Permeabilized cells were then washed with PBS containing 3% bovine

serum albumin (BSA) and incubated with an Alexa Fluor 488-conjugated anti-BrdU

antibody for 1 hr at 37°C. After washing the slide in PBS, nuclei were stained with DAPI

prior to mounting of cover slips. Finally, slides were visualized on a fluorescence

microscope. The percentage of BrdU-positive cells was calculated as the number of BrdU-

positive cells in the DAPI-positive population. Multiple random fields were assessed to

generate an average number of BrdU-positive cells in the sample.

ANNEXIN V-CY3 STAINING

Cells undergoing apoptosis express phosphatidyl serine (PS) on the outer leaflet

of the plasma membrane lipid bilayer. In healthy cells, PS is localized to the inner leaflet

of the lipid bilayer. Annexin A5 (or annexin V) binds with high affinity to PS, which is

accessible only with PS is localized to the outer leaflet. Thus, annexin V can be used as

a marker of apoptotic cells.

To calculate the percentage of cells undergoing apoptosis, cells were first washed

with PBS and resuspended in annexin V binding buffer (10 mM HEPES [pH 7.4], 140 mM

NaCl, 2.5 mM CaCl2) containing stock Annexin V-Cy3 (BioVision; Mountain View, CA).

Following a 5 min incubation at room temperature in the dark, cells were pelleted and

washed with fresh annexin V binding buffer. Washed cell pellets were resuspended in 4%

paraformaldehyde and incubated at room temperature for 20 min. Paraformaldehyde-fixed

cells were spotted on glass slides, permeabilized with 0.25% Triton-X 100 in PBS, and

stained with DAPI prior to visualization by fluorescence microscopy. The percentage of

annexin V-Cy3-positive cells was calculated as the number of annexin V-Cy3-positive cells

41

in the total DAPI-positive population. Multiple random fields were assessed to obtain an

average value of the percentage of annexin V-positive cells.

LENTIVIRUS PRODUCTION

After lentiviral vectors were cloned and assessed for efficiency via HEK293T

transfection and western blot or RT-PCR analysis, lentivirus was produced for use in cell-

based experiments. To produce lentivirus, HEK293T cells were plated in T75 flasks such

that they achieved 50% confluence at the time of transfection. For each T75 flask, 12 μg of

lentiviral vector, 9 μg of Pax2 gag/pol, and 3 μg of vesicular stomatitis virus G protein

(VSV/G) were mixed with 1 ml CaCl2 (2.5 M) and 1 ml 2X HBS. Transfections were

conducted as described above. Two days post-transfection, culture medium was collected

and passed through a 0.45 μm filter. Filtered medium was then centrifuged at 49,000 x g

in a SW28 rotor (Beckman Coulter, Inc; Palo Alto, CA) for 2 hr at 4°C in an ultracentrifuge.

Lentivirus pellets were resuspended in RPMI 1640 medium, aliquoted, and stored at

-80°C. Titering of lentivirus was assessed by GFP fluorescence (for pYNC352-derived

lentiviral vectors) or immunofluorescent staining of cells transduced with limiting dilutions

of aliquoted lentivirus.

LENTIVIRAL TRANSDUCTION OF CULTURED CELLS

Transduction experiments were conducted following lentivirus titration. For

adherent cells (e.g. HEK293T or TIME cells), culture medium was removed. Then, RPMI

1640 medium containing lentivirus and polybrene (1 μg/ml) was added to cells. After 8 hr

of incubation at 37°C, medium containing lentivirus was removed and replaced with fresh

cell-appropriate culture medium. For suspension cells (e.g. PEL cells), cells were pelleted,

and culture medium was removed. Cell pellets were resuspended in lentivirus- and

polybrene-containing RPMI 1640 medium and incubated for 8 hr as described above.

42

Following incubation with lentivirus, PEL cells were pelleted and resuspended in fresh

RPMI 1640 medium containing 10% FBS and allowed to recover for 48 hr prior to initiation

of cell growth assays or HHV-8 reactivation.

CELL GROWTH ASSAY

To determine the role of vIL-6 and cellular proteins on the growth of PEL cells in

culture, PEL cells were first transduced with vIL-6- or cellular protein-specific shRNA-

encoding lentivirus. Two days post-transduction, trypan blue-excluding (viable) cells were

counted using a hemocytometer. Equivalent numbers (1 x 105 cells/well) of control (non-

silencing shRNA-encoding lentivirus) and experimental (e.g. vIL-6-specific shRNA-

encoding lentivirus) cells were plated in duplicate wells of a 24-well cell culture plate.

Plated cells were then counted daily for three additional days to generate growth curves.

At the end of the growth experiment, remaining cells were harvested for preparation of cell

lysates for western blot analysis.

HHV-8 R219 GENERATION AND INFECTION OF ENDOTHELIAL CELLS

HHV-8 r219 virus was designed such that GFP expression is driven by the human

elongation factor 1 alpha (EF-1α) promoter, and RFP expression is driven by the viral PAN

RNA promoter [254]. GFP is expressed constitutively, while RFP is expressed exclusively

in lytic cells as PAN RNA is expressed only during the viral lytic cycle. Therefore, GFP

and RFP fluorescence can be used as markers to determine whether cells have been

infected with HHV-8 r218 (GFP+) or have been reactivated (RFP+, lytic).

HHV-8 r219 virus was generated in Vero-r219 cells infected with baculovirus-

encoded Rta (a gift from Prashant Dasai). Briefly, Vero-r219 cells were grown in DMEM

supplemented with 10% FBS and puromycin (5 μg/ml). Medium was removed from Vero-

r219 cells, and baculovirus-encoded Rta was added to the cell culture medium. The cells

43

and virus were incubated at 37°C for 2 hr with continuous rocking. Cells were then washed

with fresh DMEM and incubated with DMEM containing 10% FBS and NaB (1 mM) for 24

hr. Medium was then replaced with DMEM supplemented with 10% FBS, and cells were

incubated for an additional four days. Finally, culture medium was harvested and

centrifuged briefly at 400 x g for 5 min to remove cellular debris. The supernatant was then

centrifuged at 16,000 rpm (SW41 rotor, Beckman Coulter, Inc) for 2 hr at 4°C in an

ultracentrifuge to pellet the virus. Viral pellets were resuspended in EBM-2 without serum

or supplements, aliquoted, titered on naïve TIME-TRE/RTA cells, and frozen at -80°C until

needed.

TIME-TRE/RTA endothelial cells were infected with HHV-8 r219 virus for use in

virus reactivation experiments. Prior to infection, TIME-TRE/RTA cells were plated such

that they would achieve approximately 90% confluence at the time of infection. Culture

medium was removed, and HHV-8 r219 virus was added to cells. Cell cultures containing

virus were centrifuged clockwise at 400 x g for 30 min at 37°C and then counterclockwise

for an additional 30 min at the same speed and temperature. Following an additional 1 hr

incubation at 37°C at 5% CO2, medium containing virus was removed and replaced with

fresh EBM-2 containing supplements and FBS. Two days post-infection, cells were

visualized for GFP expression via fluorescence microscopy to determine infectious titers.

For reactivation experiments in endothelial cells, HHV-8 r219-infected TIME-TRE/RTA

cells were then transduced with lentivirus targeting viral or host proteins and reactivated

as described below.

TITERING OF REACTIVATED HHV-8

To assess the role of viral and host proteins in HHV-8 productive replication,

protein-depleted TIME and PEL cells were reactivated, and viral titers were calculated.

Transduced JSC-1 PEL cells were reactivated with NaB (0.5 μM) and 12-O-

44

tetradecanoylphorbol-13-acetate (TPA, 20 nM), and medium was harvested four days

post-induction. Transduced TRExBCBL1-RTA and HHV-8 r219-infected TIME-TRE/RTA

cells were induced with doxycycline (Dox, 1 μg/ml) and harvested at four (TRExBCBL1-

RTA) or six (TIME-TRE/RTA) days post-induction. Harvested culture medium containing

virus was passed through a 0.45 μm filter and centrifuged at 16,000 rpm (SW41 rotor,

Beckman Coulter, Inc) in an ultracentrifuge for 2 hr at 4°C to pellet virus. Viral pellets were

resuspended in EBM-2, and aliquots of harvested virus were applied to naïve TIME-

TRE/RTA cells for 2 hrs. Cells and virus-containing medium were centrifuged clockwise

at 400 x g for 30 min at 37°C and then counterclockwise using the same conditions.

Following a 1 hr incubation at 37°C in a 5% CO2 chamber, medium containing virus was

removed, and fresh EBM-2 containing FBS and supplements was added. The following

day, cells were trypsinized, washed with PBS, and plated on glass chamber slides. One

day after the cells were plated, slides were fixed in methanol as described above and

stained using an anti-LANA primary antibody and a Cy3-conjugated fluorescent secondary

antibody. Nuclei were stained with DAPI, and slides were visualized via fluorescence

microscopy. The percentage of LANA-positive cells was calculated as the number of

LANA-positive cells in the total DAPI-positive cell population. Multiple random fields were

assessed.

DEPLETION AND COMPLEMENTATION EXPERIMENTS

Depletion and complementation experiments were utilized to address several

research questions, including whether an ER-localized vIL-6 signaling mutant could

rescue PEL cell growth in vIL-6 depleted cells and whether ER-localized vIL-6/gp130-

mediated STAT and MAPK signaling played a role in virus replication. These experiments

required that cells be transduced with multiple lentiviruses.

45

For experiments assessing the rescue of vIL-6 depletion with an ER-localized

vIL-6 signaling mutant, cells were first transduced with lentivirus encoding an ER-localized

and shRNA-resistant vIL-6 signaling mutant. After allowing time for the transduced cells

to recover, cultures were selected in puromycin (1 μg/ml) to eliminate untransduced cells.

Then, puromycin-selected cells were transduced with lentivirus encoding vIL-6-specific or

non-silencing shRNAs. Cells were allowed to recover for two days following the second

lentiviral transduction. Then, cells were normalized and plated for cell growth assays as

described above.

For depletion-complementation assays addressing virus replication, TRExBCBL1-

RTA PEL cells were transduced with the first lentivirus (overexpressed vIL-6 variants or

gp130-signaling mutants) and allowed to recover for two days prior to being transduced

with the second (shRNA-encoding) lentivirus. GFP fluorescence was used as a marker of

transduction efficiency of the shRNA-encoded lentivirus. A subset of doubly transduced

TRExBCBL1-RTA cells were harvested in TRIzol for RT-PCR analysis to verify the

expression of vIL-6 and gp130 signaling mutants. As before, cells were allowed to recover

for two days prior to cell count normalization and reactivation with Dox. Virus was

harvested from culture media and applied to naïve TIME-TRE/RTA cells for LANA staining

as described above. Specific procedures are detailed in the methods sections of the

appropriate chapters.

46

CHAPTER 3: INVOLVEMENT OF GP130 IN VIL-6-MEDIATED PEL CELL GROWTH

AND SURVIVAL

47

Summary

Prior to initiating this dissertation work, vIL-6 had been implicated in the

pathogenesis of HHV-8-associated diseases via pro-proliferative, anti-apoptotic, and pro-

angiogenic activities. In PEL cells, vIL-6 is produced in functional amounts during viral

latency and promotes the growth of these cells, mediating its activity from the ER, where

it is predominantly localized. This vIL-6 activity is dependent, in part, on its interaction with

a splice variant of VKORC1, termed VKORC1v2. However, the roles of gp130 and ER-

localized vIL-6/gp130 interactions in PEL cells were unknown. Thus, this study sought to

determine the influence of gp130 on PEL cell viability and growth, the role of ER-localized

vIL-6/gp130 signaling in PEL cell maintenance, and the contributions of gp130-mediated

MAPK and STAT signaling to PEL cell proliferation and survival.

Targeted depletion of gp130, vIL-6, ERKs 1 and 2, STAT1, and STAT3 was

conducted to determine the role of these proteins in PEL cell growth and survival.

Additionally, gp130-activated MAPK and STAT signaling was assessed following gp130

depletion to determine the contribution of gp130 to activation of these signaling pathways

in PEL cells. Finally, two approaches were utilized to establish the significance of ER-

localized vIL-6/gp130 signaling on the proliferation and viability of PEL cells.

This study found that the IL-6 signal transducer, gp130, is required for optimal PEL

cell growth and viability. Levels of activated ERKs 1 and 2 and STATs 1 and 3,

phosphorylated following gp130 stimulation, were reduced in gp130-depleted BCBL-1 and

BC-1 PEL cells. Diminished STAT activation was also detected in two other PEL cell lines,

JSC-1 and BC3. Growth effects of gp130 depletion could be mimicked by shRNA targeting

of ERKs 1 and 2 or by depletion of STAT3. Finally, inhibition of vIL-6/gp130 interactions

specifically within the ER compartment suppressed cell proliferation and viability, mirroring

the effects of gp130 depletion. Combined, these data demonstrate that gp130, in addition

to VKORC1v2, is essential for normal PEL cell growth and survival and that ER-localized

48

vIL-6/gp130 interactions are critical for these activities. Targeting of intracellular vIL-

6/gp130 interactions could potentially provide a means of PEL therapy.

49

ROLE OF HUMAN HERPESVIRUS 8 INTERLEUKIN-6-ACTIVATED GP130 SIGNAL

TRANSDUCER IN PRIMARY EFFUSION LYMPHOMA CELL GROWTH AND

VIABILITY

This work has been published and is reproduced here (with modifications) with permission from the American Society for Microbiology (Copyright © American Society for Microbiology, [Journal of Virology, 2013, 87(19):10816-10827, doi: 10.1128/JVI.02047-13]). Cousins E, Nicholas J (2013) Role of human herpesvirus 8 interleukin-6-activated gp130 signal transducer in primary effusion lymphoma cell growth and viability. 87(19):10816-10827.

INTRODUCTION

HHV-8 encodes several proteins that are believed to contribute to the onset and/or

progression of endothelial KS and B-cell malignancies PEL and MCD [255-258]. vIL-6,

like its cellular counterparts, is a growth factor for B-cells and other cell types and promotes

inflammatory and angiogenic responses. These activities implicate the viral cytokine as a

contributory factor in HHV-8 associated neoplasias [144,259]. In PEL cells, true latent

expression of vIL-6 suggests that the viral protein can contribute to this disease in a direct,

autocrine fashion by promoting PEL cell proliferation and survival, in addition to possibly

maintaining viral latent reservoirs during normal (disease-free) infection [115,116].

While the three-dimensional structures of vIL-6 and hIL-6 are similar and both can

bind to and induce dimerization of the gp130 signal transducer, vIL-6 is unique in that it is

“pre-conformed” to mediate gp130 dimerization without first binding the non-signaling

gp80 IL-6 receptor subunit [165,260,261]. However, vIL-6 can bind gp80 and form

hexameric complexes (vIL-62:gp1302:gp802) in addition to tetrameric (gp80-devoid)

complexes [159,165]. Hexameric and tetrameric complexes have distinguishable

signaling properties [262], likely mediated in part by gp80 stabilization of vIL-6-induced

gp130 dimerization [159,165]. Within the ER, vIL-6 induces the formation of tetrameric

complexes exclusively [115,170]. ER-directed hIL-6 is unable to induce gp130 complexing

50

and signal transduction. vIL-6, hIL-6, and other cellular IL-6 proteins activate STATs 1 and

3 via gp130-associated Janus kinase (JAK)-mediated tyrosine phosphorylation of the

transcription factors [156]. MAPK signaling is activated following SHP2 recruitment to

gp130 and phosphorylation by JAK, which leads to downstream phosphorylation and

activation of ERKs 1 and 2 [156]. In addition to differences in gp80-dependency of ligand-

induced gp130 dimerization and the ability of vIL-6 to signal from the ER, inefficient

secretion of vIL-6 distinguishes it from its cellular counterparts [170]. Thus, vIL-6 is found

largely intracellularly, specifically within the ER, and its ability to signal from this

compartment suggests that this may be functionally important for both virus biology and

viral pathogenesis. Indeed, vIL-6 depletion-mediated inhibition of PEL cell growth in

culture can be reversed by transduction of ER-retained (KDEL motif-tagged) vIL-6 [115].

Also, vIL-6 support of PEL cell growth can be inhibited by an ER-localized single-chain

antibody specific to vIL-6 [150]. It is reasonable to hypothesize that vIL-6 may contribute

to PEL pathogenesis via gp130 signaling. STAT3, a major target of such signaling and a

transcription factor implicated in many human cancers [263-265], is activated in PEL cells

and appears to be important for their viability, in part via the STAT3-induced pro-survival

protein survivin [152].

However, demonstration of vIL-6-mediated signal transduction via gp130 in PEL

cells and the role of gp130 in PEL cell biology have not been reported. Recently, vIL-6

was found to interact with the ER membrane protein VKORC1v2, a splice variant of

warfarin target VKORC1 (variant 1) [174], and this interaction was shown to be important

for the pro-growth and anti-apoptotic activities of vIL-6 in PEL cells [173]. Interaction of

vIL-6 with VKORC1v2 occurs via a transmembrane-proximal region of the ER-lumenal

domain; its precise mapping enabled the development of a vIL-6 refractory variant and an

interaction-inhibitory peptide that helped reveal the functional relevance of the vIL-6-

51

VKORC1v2 interaction [173]. The mechanism of vIL-6 activity via VKORC1v2 appears to

be independent of gp130.

In the present study, we have investigated directly the influence of gp130 on PEL

cell proliferation and viability, the role of the vIL-6-gp130 interaction in gp130-mediated

activities, and the contributions of gp130-activated MAPK and STAT signaling to PEL cell

growth. Our data indicate that gp130 is essential for normal proliferation and viability of

PEL cells and that ER-localized vIL-6/gp130-activated ERK and STAT3 signaling are

critically important. These findings suggest that inhibitory targeting of vIL-6 interactions

with the signal transducer within the ER may provide a useful strategy for PEL therapy or

for targeting HHV-8 latency in normal B-cells should vIL-6 be expressed in this latent

reservoir.

METHODS

Cell culture, transfections, and lentivirus production

Cell cultures were maintained as described in Chapter 2. HEK293T cells were

transfected with plasmids using standard calcium-phosphate-DNA co-precipitation as

explained in Chapter 2. Cells were harvested 24 to 48 hrs post-transfection. Cell lysates

were prepared as outlined in Chapter 2, and lysates were analyzed directly or stored at

-20°C prior to Western blot analysis. Lentivirus was produced by transfecting HEK293T

cells with shRNA-, vIL-6-, or sgp130-encoding or empty lentivirus plasmids (pYNC352 and

pDUET001) together with gag/pol and VSV/G expression vectors, as previously described

and in Chapter 2 [173]. Infectious titers were established by using GFP-based or

immunofluorescence microscopy to identify vector-transduced cells in PEL and other

cultures prior to experimental utilization of lentiviral vectors (see Chapter 2). For vIL-6

“rescue” experiments, PEL cells were first transduced with pYNC419-vIL-6.KDEL or

pYNC419-vIL-6.W167G.KDEL, allowed to recover for 5 days, selected in puromycin (1.0

52

μg/ml) for 2 weeks, and then transduced with either NS or vIL-6-specific shRNA encoded

by pYNC352-derived lentiviruses. For shRNA and sgp130-transduced cultures, cells were

allowed to recover for 2 days prior to normalizing cell numbers for cell growth assays.

Oligonucleotides and plasmids

shRNAs were chosen or designed to selectively deplete gp130, STAT1, STAT3,

ERK1, or ERK2. Complementary oligonucleotides encoding 19-nucleotide short hairpin

regions were annealed, phosphorylated, and cloned into the BamHI and MluI sites of the

pYNC352 lentivirus vector. The following, pre-validated shRNA sequences were used:

STAT1 (5’-GCGTAATCTTCAGGATAAT-3’ [266]), STAT3 (CATCTGCCTAGATCGGCTA

[266,267] used independently or together with AGTCAGGTTGCTGGTCAAA [266]), ERK1

(GCCATGAGAGATGTCTACA [268]), and ERK2 (GAGGATTGAAGTAGAACAG [268]).

Three shRNA sequences were designed and tested for gp130 depletion:

GTGGGATCACCTATGAAGA (sh1), CTGTCCAAGACCTTAAACC (sh2), and

TGCCCTTGGGAAGGTTACA (sh3); sh1 was used in subsequent functional analyses,

alone or together with sh3. vIL-6 shRNA oligonucleotides and non-silencing (NS) control

shRNAs were described previously [115]. For generation of a lentivirus vector expressing

KDEL motif-tagged vIL-6 variant W167G [166], the corresponding vIL-6 coding sequences

were first cloned between the BamHI and SmaI sites of pSG5.KDEL [166] to generate

contiguous vIL-6.W167G-KDEL coding sequences, and these sequences were then PCR

amplified and inserted between the NotI and XhoI sites of pTYB6-derived lentivirus vector

pYNC419 [166,269]. Lentiviral vector-cloned vIL-6 and vIL-6-KDEL and pSG5-cloned vIL-

6, vIL-6-KDEL, vIL-6.W167G, hIL-6, hIL-6-KDEL, gp130-Fc, and chitin binding domain

(CBD)-fused VKORC1v2 have been described previously [69,115,165,166,173]. The

soluble gp130 (sgp130) expression vector was made by amplifying the first 972

nucleotides of gp130 (encoding domains 1-3) and inserting this sequence between the

53

EcoRI and BamHI sites within the pSG5.KDEL vector, generating pSG5-sgp130.KDEL. A

sequence encoding the FLAG epitope was then inserted at the BamHI site using

appropriate complementary oligonucleotides, generating pSG5-sgp130.FLAG.KDEL. The

sgp130.FLAG.KDEL coding region was then amplified and ligated into the pDUET001

lentiviral vector [270] between the BamHI and XhoI sites. The integrities of all

constructions were verified by sequencing.

Immunoblotting and co-precipitation assays

For western blotting, heat-denatured and reduced proteins in cell lysate samples

were size fractionated by SDS-PAGE and electrophoretically transferred to nitrocellulose

membranes using standard procedures (see Chapter 2). Membranes were blocked,

probed with primary antibodies, washed, incubated with appropriate HRP-conjugated

secondary antibodies, and exposed as described in Chapter 2. For co-precipitation

assays, cell lysates (prepared as described above) were added to either chitin beads (for

precipitating VKORC1v2-CBD) or protein-A agarose beads (for precipitating gp130-Fc)

and incubated with continuous rotation overnight at 4°C. Beads were then washed 3 times

with lysis buffer (50 mM Tris-HCl [pH 7.4], 150 mM NaCl, and 0.1% Tween 20) prior to

boiling and loading onto SDS-PAGE gels. Untreated lysate samples were also separated

on SDS-PAGE gels to provide controls for protein expression, extraction, and transfer to

immunoblotting membranes.

Antibodies

The following commercially available antibodies were used for western blotting:

BrdU (BD Biosciences, Rockville, MD; catalog number 558599); pSTAT3 and pSTAT1

(Cell Signaling Technology, Beverly, MA; catalog numbers 9131 and 9171); β-actin

(Sigma, St. Louis, MO; catalog number A2228); pERK, ERK1, STAT3, p-tyrosine, and

54

gp130 (Santa Cruz Biotechnology, Santa Cruz, CA; catalog numbers sc-7383, sc-94, sc-

482, sc-508, and sc-655); hIL-6 (R&D Systems, Minneapolis, MN; catalog number MAB-

206); and CBD (New England Biolabs, Ipswich, MA; catalog number E8034S). Rabbit

antiserum to vIL-6 was generated in our laboratory and reported previously [166]. The

gp130 (BD Biosciences, Rockville, MD; catalog number 555756) and FLAG (Sigma;

catalog number F1804) primary antibodies and the Alexa Fluor 594 donkey anti-mouse

IgG (H+L) secondary antibody (Invitrogen, Carlsbad, CA; catalog number A21203) were

used for immunofluorescence staining.

Apoptosis, proliferation, and cell growth assays

Apoptosis was assessed via annexin V-Cy3 staining of cells as outlined in Chapter

2. Several fields were used to calculate the percentage of annexin V-Cy3-positive cells in

the total DAPI-positive cell population. Cell proliferation rates were determined by BrdU-

incorporation assay, which was detailed in Chapter 2. Slides were visualized under UV

microscopy to assess the percentage of BrdU-positive cells in the total cell population

(DAPI+). Multiple fields were assessed to determine an average percentage of BrdU-

positive cells. For cell growth assays, appropriate, high-efficiency infection with the GFP-

positive lentiviral vectors was verified by >90% GFP fluorescence in the cultures, which

was sustained over the course of the 3-day growth experiment. Cell growth assays were

initiated following cell density normalization and seeding of cells (105) in fresh media 48

hrs after lentiviral infection. Growth of the cultures was quantified by daily counting of

trypan blue-excluding (viable) cells per ml using a hemocytometer. Cultures transduced

with equivalent titers of lentiviral vector expressing NS shRNA were used as controls.

55

RESULTS

Contributions of gp130 to PEL cell growth. In previous studies, we found that

depletion of vIL-6 in BCBL-1 and JSC-1 PEL cells led to autocrine-mediated reductions in

cell proliferation and survival and that ER-localized vIL-6 activity was sufficient for these

functions [115]. Subsequently, we determined that expression of VKORC1v2 and its

interaction with vIL-6 were important gp130-independent contributors to the pro-growth

and anti-apoptotic activities of vIL-6 in these cells [173]. However, the possibility remained

that vIL-6 signaling through gp130, in the ER and/or at the cell surface, also contributes

to vIL-6 effects on PEL cells. The potential contribution of gp130 to PEL cell growth was

tested using lentiviral-delivered gp130-specific shRNA to deplete endogenous expression

of the signal transducer. Three shRNAs for gp130 depletion were designed (Methods) and

tested in lentivirus-shRNA vector-transfected HEK293T cells, and each significantly

depleted overexpressed gp130 levels relative to vector-expressed gp130 observed in NS

shRNA-transfected control cultures (data not shown). Two of these gp130 shRNA-

expressing lentivirus vectors were used for single or dual transduction (see Methods) for

subsequent experiments in PEL cells. The control NS shRNA-encoding lentiviral vector

was used for comparison. Equivalent inoculating doses of pre-titered lentiviruses were

used, sufficient to provide ~90% transduction efficiency, as determined by lentivirus-

expressed GFP fluorescence. After a post-infection incubation period of two days, growth

of cell density-normalized PEL cultures was monitored for an additional three days. For all

PEL cell lines tested (BCBL-1, JSC-1, BC-1, and BC-3), gp130-targeted shRNAs led to

significantly reduced cell growth relative to NS shRNA-transduced controls (Figure 3.1A).

Potential off-target, generally cytotoxic effects of gp130-directed shRNA were controlled

for by equivalent experiments in HHV-8-negative BJAB cells, which were essentially

unaffected by gp130 shRNA transduction relative to NS shRNA-transduced controls.

The observed reductions in PEL cell growth following gp130 depletion could be

56

due to decreased cell survival and/or decreased rates of proliferation. This was

investigated utilizing annexin V-Cy3 staining assays to detect apoptotic cells and BrdU

incorporation analysis to identify S-phase cells in BCBL-1 and JSC-1 cultures transduced

with either NS or gp130-specific shRNA. The data from these experiments revealed a

correlation between gp130 depletion and both increased rates of apoptosis (Figure 3.1B)

and decreased rates of DNA synthesis (Figure 3.1C). The combined results indicate that

gp130 signaling supports PEL cell growth by promoting both cell proliferation and cell

survival.

Influence of gp130 on STAT and ERK signaling in PEL cells. Signal

transduction by gp130 involves activation of STAT and MAPK signaling. This proceeds

via JAK-mediated tyrosine phosphorylation of gp130-recruited STATs 1 and 3 and tyrosine

and threonine phosphorylation of ERKs 1 and 2 following the activation of the MAPK

signaling cascade, which is triggered by JAK phosphorylation of gp130-associating SHP2

[156]. Active STAT3 has been reported previously to be expressed at constitutively high

levels in PEL cells and to be important for PEL cell viability in culture [152]. However, the

mechanism of constitutive STAT3 activation in PEL cells and the biological significance of

ERK signaling in these cells are unknown. To address the role of gp130 in the activation

of STAT3 and ERK in the context of PEL cells, we again utilized gp130 depletion via

lentiviral vector-mediated shRNA transduction. Targeting of gp130 resulted in detectably

reduced levels of phospho-STAT3 (pSTAT3, active) and pSTAT1 in JSC-1, BCBL-1,

BC-1, and BC-3 cells (Figure 3.2A,B) and a marked decrease in phospho-ERK (pERK)

levels in BCBL-1 and BC-1 cells but not in JSC-1 and BC-3 cells (Figure 3.2C). The data

from these experiments identify, for the first time, the contribution of gp130 to constitutively

high pSTAT3 and pSTAT1 levels in PEL cells and provide evidence that gp130 can

significantly influence ERK signaling in at least some PEL cell types (shown here for

BCBL-1 and BC-1).

57

Contributions of STATs to PEL cell growth. It has been reported by others that

STAT3 is essential for the viability of PEL cells (BC-1, BCBL-1, and VG-1) and that this

activity involves STAT3-mediated induction of caspase-inhibitory and anti-apoptotic

protein survivin [152]. These studies utilized transduced dominant-negative STAT3 and

the JAK pharmacological inhibitor AG490 to suppress levels of pSTAT3. Others have

documented the importance of JAK-mediated, STAT-activating signaling in several PEL

cell lines via the experimental use of JAK1-inhibitory curcumin; this was found to inhibit

cell proliferation, promote apoptosis, and suppress IAP family proteins, including survivin

[271]. To independently assess the role of STAT3 in PEL cell growth, the transcription

factor was targeted for depletion by shRNA-encoding lentiviruses. Two previously reported

STAT3-targeted shRNAs [266,267] were first tested in transfected HEK293T cells and

verified as effective (data not shown). These shRNAs were each cloned into a lentiviral

vector for introduction into PEL cells. Initial experiments indicated that single shRNA

transduction was insufficient to achieve strong STAT3 depletion in PEL cultures, whereas

co-infection of PEL cells with both lentiviral vectors was effective. For growth assays,

BCBL-1 and JSC-1 cells were simultaneously infected with both STAT3-targeting lentiviral

vectors or with an equivalent infectious dose of NS shRNA-encoding control lentivirus.

Cell growth of post-transduction cultures was monitored by daily counting of trypan blue-

excluding (viable) cells to determine cell densities (Figure 3.3A). Suppression of STAT3

relative to NS shRNA-transduced cultures was verified in terminal (day 3) cultures by

western blotting of cell extracts (Figure 3.3A, right panels). The data revealed that STAT3

depletion substantially diminished growth of these PEL cells, consistent with previously

published studies [152], and identified JSC-1 PEL cells (not previously examined) as

dependent on STAT3 for normal growth.

Further analyses were undertaken to test the contributions of STAT3 to

proliferation and survival of BCBL-1 and JSC-1 cells. These experiments involved

58

immunofluorescence detection of BrdU incorporation and annexin V-Cy3 staining,

respectively, as outlined above and for the equivalent studies following gp130 depletion.

Cells were harvested 5 days following lentiviral transduction (3 days post-normalization

and culture) for BrdU and annexin V analyses. The results mirrored the effects of gp130

depletion, demonstrating that STAT3 depletion inhibited cell proliferation (proportion of

cells in S phase) (Figure 3.3B) and increased the rate of apoptosis (Figure 3.3C). Thus,

both of these components of cell biology are influenced by STAT3, which in turn is

activated by gp130 in these PEL cell lines.

As gp130 signaling involves activation of STAT1 in addition to STAT3, we also

explored the potential contribution of STAT1 activity to BCBL-1 and JSC-1 cell growth.

Lentiviral-delivered shRNA [266] was used to deplete STAT1 in the respective cultures,

and cell growth was compared to that of control (NS) shRNA-transduced cultures. The

data revealed no significant effects of STAT1 depletion on BCBL-1 or JSC-1 cell growth,

despite achieving robust depletion of pSTAT1 in BCBL-1 cells and a nearly 2-fold

reduction of pSTAT1 in JSC-1 cells, as demonstrated by western blotting of terminal

culture extracts (Figure 3.3D). Although STAT1 depletion was somewhat limited in JSC-1

cells, these data suggest that STAT1 activity is not critical for PEL cell growth.

Contributions of ERKs 1 and 2 to PEL cell growth. As gp130 signaling involves

the activation (phosphorylation) of ERKs 1 and 2 in addition to STATs, we undertook

experiments to investigate the influence of ERK signaling on PEL cell growth. ERKs 1 and

2 were simultaneously targeted for depletion in BCBL-1 and JSC-1 cells using techniques

analogous to those employed for STAT depletion. The NS- or ERK-shRNA lentiviral

vector-transduced cells were monitored for growth over a three-day period by daily

calculation of viable (trypan blue-excluding) cell densities from triplicate cultures for each

cell type. For both PEL cell lines, ERK depletion led to dramatic reductions in cell growth

relative to control (NS) shRNA-transduced cultures (Figure 3.4A). Therefore, ERKs 1 and

59

2, in addition to STAT3, are essential for normal growth of these cells.

The shRNA-transduced JSC-1 and BCBL-1 cultures were further analyzed to

assess the effects of ERK depletion on cell proliferation and survival, specifically. Rates

of proliferation and apoptosis in these cells following ERK1 and ERK2 depletion were

examined by BrdU incorporation and annexin V-Cy3 staining, respectively, as outlined

above. These analyses revealed that both proliferation and cell survival were reduced

appreciably (between two- and three-fold) by ERK depletion (Figure 3.4B,C).

Analysis of the effects of ERK depletion in two additional PEL cell lines, BC-1 and

BC-3, indicated the general importance of ERKs 1 and 2 for PEL cell growth (Figure 3.4D).

Effects were not as marked in these cell lines as they were for BCBL-1 and JSC-1 cells,

but it is possible that this was due to reduced efficacy of ERK depletion (Figure 3.4D,

bottom panels). In contrast, ERK depletion in the HHV-8-negative BJAB B-cell line,

reported previously to be insensitive to inhibition of ERK kinase (MEK) by PD98059 [272],

were unaffected by ERK shRNA transduction, verifying lack of general cytotoxic effects of

lentiviral infection and ERK shRNA transduction (Figure 3.4D, right). Combined, the data

from our ERK depletion studies reveal the importance of ERK signaling, sustained in part

via gp130 signal transduction in at least some PEL cell lines (BCBL-1 and BC-1, shown

here), for PEL cell growth and viability.

ER-localized gp130/vIL-6 interaction contributes to vIL-6 and gp130

activities. We have previously reported that ER-localized vIL-6 and its interaction with the

ER protein VKORC1v2 are important for PEL cell proliferation and viability [173]. A

previously characterized vIL-6 variant, vIL-6.W167G, was used to determine the

contribution of ER-localized vIL-6/gp130 interactions to PEL cell growth and survival. This

altered vIL-6 protein is mutated in a region (“site III”) that interacts with gp130 domain 1

and is unable to induce dimerization of gp130 absent gp80 co-expression and hexameric

complexing [159,166]. As ER-localized vIL-6 signaling occurs exclusively through

60

tetrameric (vIL-62:gp1302) complexes [171], vIL-6.W167G should be unable to induce

gp130 dimerization and signaling from the ER compartment. We first verified that ER-

retained, KDEL motif-tagged vIL-6.W167G was unable to induce gp130-mediated signaling

in transfected HEK293T cells. Unlike wild-type vIL-6-KDEL, vIL-6.W167G-KDEL was

unable to induce tyrosine phosphorylation of gp130 or STAT3 (Figure 3.5A). Co-

precipitation assays from transfected cell lysates confirmed that the W167G mutation did

not affect the interaction between vIL-6 and VKORC1v2, which was tagged with CBD to

facilitate precipitation and enable immune-detection with a CBD-specific antibody (Figure

3.5B). KDEL-tagged, shRNA-resistant versions of wild-type and W167G-mutated vIL-6

were then used to supplement vIL-6-depleted PEL cells to determine if the gp130

dimerization-defective but VKORC1v2 binding-competent variant could rescue cell

growth. These experiments were analogous to those undertaken previously with vIL-6-

KDEL, which demonstrated sufficiency of ER-localized vIL-6 for maintenance of BCBL-1

and JSC-1 cell growth and viability [171]. In contrast to vIL-6-KDEL, vIL-6.W167G-KDEL

was unable to complement endogenous vIL-6 depletion (Figure 3.5C), indicating that vIL-

6-induced dimerization of gp130 and its downstream signaling is critically important for

pro-proliferative activity of ER-localized vIL-6 in PEL cells.

The importance of intracellular gp130 signaling was further tested by utilizing

lentiviral-transduced soluble gp130 (sgp130), which was KDEL motif-fused for ER

retention. Initial experiments in transfected HEK293T cells were undertaken to verify the

ability of this protein to inhibit ER-localized (vIL-6.KDEL), but not extracellular (hIL-6),

signal transduction via endogenously expressed gp130 (Figure 3.6A). Additionally, the

intracellular localization of sgp130.KDEL (Flag-tagged) was tested and compared to the

distribution of full-length gp130 by α-Flag and α-gp130 antibody-based

immunofluorescence assays in transfected cells; cells were co-transfected with GFP-

KDEL, which provided an ER marker [115]. In contrast to native gp130, sgp130.KDEL was

61

present at high levels intracellularly with little, if any, surface fluorescence detectable

(Figure 3.6B). Having confirmed the expression and appropriate ER localization of

lentiviral vector-expressed sgp130.KDEL, this vector, or an empty vector control, was then

used to transduce PEL cells to determine the effects of ER-localized gp130 inhibition on

cell growth and apoptosis. The data revealed significant effects of sgp130.KDEL, relative

to the empty vector, on the growth of both BCBL-1 and JSC-1 PEL cells (Figure 3.6C).

BrdU incorporation assays using cells harvested at day 2 from the growth experiment

confirmed inhibition of proliferation (i.e., percentage of cells in S-phase) by introduced

sgp130.KDEL (Figure 3.6D). In a separate experiment, transduction of sgp130.KDEL into

BCBL-1 and JSC-1 cells led to substantial increases in rates of apoptosis, as determined

by annexin V-Cy3 staining (Figure 3.6E). As vIL-6 is the only known cytokine to signal

through gp130 from the ER compartment (and ER-retained hIL-6 has been demonstrated

to be unable to do so [115]), these data suggest strongly that vIL-6/gp130 signaling is of

biological significance and contributes, wholly or in part, to the pro-growth and pro-survival

effects of gp130 in these cells.

DISCUSSION

The work reported here addresses the contribution of gp130 signaling to the

constitutively high levels of active STAT3 reported in PEL cells, the influence of gp130

activity on STAT1 and ERK activation in these cells, and the functional relevance of vIL-

6-mediated gp130 activity to PEL cell growth. To our knowledge, our data are the first to

document a mechanism for the constitutively high levels of STAT3 in PEL cells, to

establish a critical role of the gp130 signal transducer in supporting PEL cell proliferation

and viability, and to identify the relevance of gp130 targeting by vIL-6 to these activities.

Thus, in addition to the important association of vIL-6 with VKORC1v2 for PEL cell growth

and survival [115], activity of the viral cytokine through its well-established signal

62

transducer is also critical. This activity is mediated from the ER, where vIL-6 is

predominantly localized. A VKORC1v2 binding-competent, gp130 dimerization-defective

vIL-6 variant (W167G) that was restricted to the ER was unable to rescue the effects of vIL-

6 depletion, in contrast to KDEL motif-tagged wild-type vIL-6. Furthermore, ER-localized

sgp130, used as a competitive inhibitor of gp130/vIL-6 interactions in the ER

compartment, enabled identification of the specific role of vIL-6 as a mediator of gp130

function in the context of PEL. Thus, gp130 and ER-localized vIL-6 signaling via gp130

have been shown for the first time to support PEL cell proliferation and viability. Therefore,

gp130 and its interaction with vIL-6 have been identified as potential targets for therapeutic

intervention.

Another important finding from this study is that ERK activation contributes

significantly to the growth and survival of PEL cells. Depletion of ERK led to severe

abrogation of cell proliferation (3-fold reductions in BrdU incorporation) and to greatly

increased rates of apoptosis (2- to 3-fold) for BCBL-1 and JSC-1 cultures, effectively

halting culture growth. Significant effects of ERK depletion on the growth of BC-1 and BC-

3 cells were also identified. At least for BCBL-1 and BC-1 cells (included in this study), the

high levels of active, tyrosine-phosphorylated ERK (pERK) are dependent on gp130.

However, gp130 depletion did not diminish the levels of pERK in JSC-1 and BC-3 cells;

pERK was actually increased in both cell lines. These patterns were observed in multiple

repeat experiments in these PEL cell lines. In contrast to the effects of ERK depletion on

PEL cell growth, ERK depletion in HHV-8-negative BJAB cells did not affect cell growth at

all. The role of ERK signaling in HHV-8 biology has previously been noted only in

connection with de novo infection, establishment of latency, and lytic reactivation. Thus,

pharmacological inhibition of ERK activation suppresses TPA-induced lytic gene

expression and replication in PEL cultures [273-276]. Additionally, activation of ERK-

targeting signaling pathways and their requirement for de novo infection, post-entry lytic

63

gene expression, and latency establishment have been reported in endothelial cell

systems [275,277,278]. Our results now add to this body of knowledge regarding the role

of ERK signaling in HHV-8 biology by identifying ERK as an important factor for

maintenance of PEL cell viability and by establishing gp130 signaling as a mechanism

contributing to high levels of active ERK in at least some PEL cell types.

Previous studies by Aoki and colleagues [152] demonstrated that the JAK inhibitor

AG490 and ectopic expression of a dominant-negative STAT3 protein (STAT3D, DNA

binding-refractory) led to dramatically reduced PEL cell viability. The PEL cell lines

investigated were BCBL-1, BC-1 and VG-1. Our experiments investigating the role of

STAT3 utilized BCBL-1 and JSC-1 cells, characterized in depth previously with respect to

latent expression and pro-growth and pro-survival activities of ER-localized vIL-6 [115]. A

caveat of the experiments utilizing AG490 is that this pharmacological inhibitor can affect

JAK-activated receptors other than gp130 and receptor-mediated activation of STATs

other than STAT3; AG490 may also have non-specific effects on other pathways. Of

additional note is that experimental use of the STAT3D protein, while very effective at

blocking endogenous STAT3 activity, also has the potential to inhibit STAT1 and other

STAT3-hetrodimerizing proteins; thus it may affect STAT3-independent mechanisms.

However, the previously reported data from Aoki et al. [152] combined with our

independently-derived and concordant results from STAT3 depletion experiments provide

very strong evidence that STAT3 signaling is critical to PEL cell biology. STAT3 signaling

was shown to induce the pro-survival protein survivin, which was found to be an important

factor in STAT3-mediated pro-survival effects in PEL cells [152]. However, it is noteworthy

that the JAK-targeting STAT signaling inhibitor curcumin, while demonstrated in an

independent study to induce cell death with concomitant suppression of survivin in most

of the PEL cell lines tested, did not lead to a significant loss of survivin expression in

BCBL-1 cells (used in the present study and that of Aoki et al. [152]) [271]. Thus, survivin

64

may not be the only STAT3-regulated factor critical to PEL cell viability. Nonetheless, our

data reveal that gp130 and ER-localized vIL-6 activity through the signal transducer can

contribute to high constitutive levels of active pSTAT3, a factor that has been linked to cell

growth and survival in PEL and other neoplasias as well as under normal physiological

conditions [264,279,280], and to promotion of cell proliferation and viability. It is

noteworthy that gp130-dependent constitutively high levels of active STAT3 have also

been detected in endothelial cells latently infected with HHV-8, although via (paracrine)

mechanisms independent of vIL-6 [281]. Furthermore, de novo infection of blood

endothelial cells leads to STAT3 induction and gp130-dependent, but vIL-6-independent,

lymphatic endothelial cell reprogramming [18]. Thus, STAT3 signaling, in part mediated

via gp130, may be generally important for maintenance of HHV-8 latency and viability of

latently infected cells and may also contribute to virus-associated neoplasia.

In conclusion, the data presented here document for the first time the significance

of gp130-mediated signaling for proliferation and survival of PEL cells, the importance of

gp130 as a contributor to constitutively high levels of functionally significant STAT and/or

ERK signaling in at least some PEL cell lines, and the contribution of ER-localized vIL-

6/gp130 interactions for PEL cell growth and viability. These data broaden our

understanding of the role and mechanisms of action of latently expressed vIL-6 in PEL

cell biology and provide further support for the idea that drug targeting of the viral cytokine

and its signal transducer, within the ER compartment, may represent a useful therapeutic

strategy.

65

FIGURES

Figure 3.1. Effects of gp130 depletion on PEL cell growth and viability. (A) Triplicate

cultures of BCBL-1, JSC-1, BC-1, and BC-3 PEL cells and HHV-8-negative BJAB B-cells

were transduced either singly (BCBL-1, JSC-1, and BJAB) or dually (BC-1 and BC-3) with

gp130-directed shRNA(s) by lentiviral infection, and cells were counted for three days

following normalization of cell numbers (see Chapter 2 and Methods above). Error bars

represent divergence from mean values determined for the triplicate cultures. Western

blotting of cell lysates derived from pooled cultures at the end of the growth experiment

(on day 3) was carried out to verify gp130 depletion (shown below the associated growth

curves). For BC-3 cells, complete cytostatic effects of gp130 shRNA transduction resulted

in insufficient cell numbers to process for immunoblotting. Levels of gp130 in BJAB cell

lysates were below the limit of detection; the growth data show lack of general cytostatic

effects of lentiviral-mediated gp130 shRNA transduction. (B) Samples of gp130 and NS

shRNA-transduced BCBL-1 and JSC-1 cultures harvested on day 2 from the growth curve

experiments shown in panel A were analyzed for proliferation. Rates of proliferation were

determined by detection of BrdU incorporation using Alexa Fluor 488-conjugated BrdU-

specific antibody for immunofluorescence visualization of S-phase cells and

counterstaining with DAPI to detect all nuclei (see Chapter 2 and Methods above).

Relative proliferation rates are shown as the percentage of cells in S-phase (BrdU-

positive) for the respective cultures. Data were derived from two random fields (>100

cells/field) for each slide-spotted culture sample. Examples of BrdU and DAPI co-staining

(sections of a single field, BCBL-1 cells) are shown below the quantified data. (C) Rates

of apoptosis in these same PEL cultures were determined by annexin V-Cy3 binding. DAPI

staining was again used to visualize nuclei, enabling calculation of the percentage of cells

undergoing apoptosis. Examples of annexin V-Cy3 and DAPI co-staining (of JSC-1 cells)

are shown below the graphs. Cells were harvested on day 3 for the apoptosis assays.

66

Figure 3.1

67

Figure 3.2. Contribution of gp130 signaling to levels of active STATs and ERKs in PEL

cells. Phosphotyrosine-specific antibodies were utilized in immunoblotting to detect and

determine the levels of phosphorylated (active) gp130-targeted STAT3 (panel A), STAT1

(panel B), and ERKs 1 and 2 (panel C) (pSTAT3, pSTAT1, and pERK) present in lysates

of PEL cells depleted of gp130 relative to levels of these proteins in NS shRNA-transduced

controls. Antibodies detecting total (phosphorylated and unphosphorylated) STAT3 and

ERK1/2 and β-actin were used to verify equivalent protein loading on SDS-PAGE gels and

transfer to blotting membranes. For each set of lysates (used for single or multiple

phospho-protein detection), effective depletion of endogenous gp130 in the lentiviral-

shRNA infected PEL cultures was verified by immunoblotting for gp130. Monitoring of

growth of NS-shRNA versus gp130-shRNA cultures for BCBL-1, JSC-1, and BC-1 cells

over three days (as in Figure 3.1) confirmed the growth-inhibitory effects of gp130

depletion (data not shown); cells were harvested on the third day for western analyses.

For BC-3 cultures, all cells were harvested 48 hrs after lentiviral infection (“day 0”) to

ensure sufficient material for immunoblotting.

68

Figure 3.2

69

Figure 3.3. Contributions of gp130-activated STATs to PEL cell growth. (A) Appropriate

lentiviral-shRNA vectors (see Methods above) were used in combination to transduce

BCBL-1 and JSC-1 PEL cells to deplete STAT3. Growth curves were derived from

triplicate cultures transduced separately with the NS and STAT3-targeting lentiviral-

shRNA vectors and monitored in parallel (see Chapter 2). Error bars represent deviations

from mean cell density values at each 24-hr timepoint over the three-day experiment.

Pooled cultures were harvested at the end of the experiment (day 3) for western analysis

of cell lysates to check for depletion of STAT3. (B) Proliferation of STAT3-depleted versus

NS shRNA-transduced BCBL-1 and JSC-1 cells were examined by BrdU incorporation

assay, as outlined in Figure 3.1, applied to cells harvested on day 3 (5 days after shRNA

transduction). (C) Rates of apoptosis in these same culture samples were determined by

annexin V-Cy3 binding assays and co-staining of cell nuclei with DAPI (see above, Figure

3.1 legend, Chapter 2). (D) STAT1 depletion experiments were performed to examine the

potential contributions of STAT1 to BCBL-1 and JSC-1 cell growth. STAT1 depletion was

mediated by transduction of PEL cultures with a lentiviral vector encoding a STAT1

mRNA-targeting shRNA (see Methods above), and NS shRNA-encoding lentivirus was

used as a control as before. While substantial depletion of STAT1 was apparent for each

cell line (immunoblots, right), growth of neither was altered significantly as a result.

Chemiluminescence signals from western blots were digitally captured, quantified using

imager-associated software, and normalized to β-actin signal. Levels of pSTAT1 in JSC-

1 cultures was 1.9-fold reduced relative to STAT1 levels in the corresponding NS shRNA-

transduced controls; no pSTAT1-specific signal was detected in BCBL-1 cell lysates from

STAT1 shRNA-transduced cultures.

70

Figure 3.3

71

Figure 3.4. Analysis of culture growth, cell proliferation, and apoptosis rates as a function

of ERK depletion. (A) Potential contributions of ERK to the growth of BCBL-1 and JSC-1

cultures were examined by monitoring growth of cells transduced by lentiviral vectors

encoding either ERK-targeting or NS (control) shRNAs. Cell growth assays were

conducted as described in Chapter 2 and the Methods above. Error bars indicate

deviations from average values derived from triplicate cultures. Effective ERK depletion

was verified by immunoblotting of lysates from BCBL-1 and JSC-1 cells pooled from the

respective triplicate cultures and harvested at the end of the experiment. β-actin was used

as a control for equivalent protein loading and membrane transfer. (B) Relative

proliferation rates of BCBL-1 and JSC-1 cultures transduced with NS or ERK1/2-directed

shRNAs were determined by BrdU incorporation assay (see Chapter 2 and Methods

above). Cultures were harvested 4 days after shRNA transduction and 2 days post-

normalization prior to growth monitoring (panel A). The proportions of BrdU-positive, S-

phase cells are expressed as a percentage of the total cell population (DAPI+) for BCBL-

1 and JSC-1 cultures (dark and light bars, respectively). Data, expressed as averages,

were derived from multiple random microscopic fields (>100 cells/field) for each. (C)

Apoptosis in the same cultures (harvested 5 days post-transduction and on day 3 of

counting) was identified by annexin V-Cy3 staining; data are expressed as a percentage

of the total DAPI+ population. Average values from multiple random microscopic fields are

shown. (D) Growth experiments equivalent to those shown in panel A were undertaken to

determine the effects of ERK depletion in two additional PEL cell lines, BC-1 and BC-3, in

addition to HHV-8-negative BJAB cells, previously shown to be refractory to ERK kinase

(MEK) inhibition [272] and used here as a control for potential non-specific cytotoxic

effects of ERK shRNA transduction. The PEL cell lines, but not BJAB cultures, were

growth-inhibited by ERK shRNA transduction relative to NS shRNA-expressing controls.

Effective ERK depletion in each of the three cell lines was verified by immunoblotting for

72

ERKs 1 and 2; β-actin antibody was used to check for equal protein loading and

membrane transfer. Blots for each of the cell lines are shown below the respective growth

curves.

73

Figure 3.4

74

Figure 3.5. ER-localized vIL-6 signal transduction via gp130 supports PEL cell growth. (A)

Previously described vIL-6 variant W167G, which cannot form tetrameric (vIL-62:gp1302,

ER-localized) complexes [159,166], was tested in transfected HEK293T cells to verify its

predicted inability to mediate signal transduction when restricted to the ER compartment.

KDEL-tagged, ER-directed forms of vIL-6 and vIL-6.W167G were expressed together with

Fc-tagged gp130 from appropriate expression plasmids. After 24 hrs, cell lysates were

analyzed by immunoblotting for pSTAT3, an indicator of gp130 signaling, and protein A-

agarose-precipitated gp130-Fc was probed with phosphotyrosine-specific antibody for

detection of activated gp130. Only wild-type vIL-6 (vIL-6.K) was able to signal via gp130

in the ER compartment, inducing gp130 and STAT3 phosphorylation, in contrast to vIL-

6.W167G-KDEL (W167G.K) and empty vector (vec). (B) Testing of VKORC1v2 binding by

vIL-6.W167G was undertaken by co-immunoprecipitation assay, which was used previously

for the detection of the vIL-6/VKORC1v2 interaction [173]. CBD-tagged VKORC1v2 was

precipitated with chitin beads from lysates of transfected cells, and precipitated material

was analyzed by immunoblotting for the presence of vIL-6.W167G, wild-type vIL-6 (positive

control), or KDEL motif-tagged hIL-6 (hIL-6.K, negative control), expressed in separately

transfected cultures. Both vIL-6 proteins were precipitated and therefore competent for

interacting with VKORC1v2; binding of VKORC1v2 by vIL-6.W167G verified that the “site

III” mutation affected only gp130 binding. (C) Experiments utilizing shRNA-mediated vIL-

6 depletion and complementation with KDEL-tagged, ER-restricted vIL-6 (vIL-6.K) and vIL-

6.W167G (W167G.K) were undertaken to determine whether vIL-6 signaling through ER-

localized gp130 was important for PEL cell growth. BCBL-1 and JSC-1 cells were first

transduced with lentiviral vectors expressing either vIL-6.K or vIL-6.W167G.K (designed to

be refractory to vIL-6 shRNA) and then transduced with lentiviral vectors expressing non-

silencing (NS) or vIL-6-directed shRNAs. Triplicate cultures were monitored for growth as

outlined in Chapter 2 and in the Methods above; error bars indicate standard deviations

75

from mean values of cell density. Below the plots, ethidium bromide-stained agarose gels

are shown containing RT-PCR-amplified vIL-6 mRNA sequences extracted from lysates

of PEL cell samples derived from the growth assay cultures. The data provide evidence

of appropriate and equivalent expression of each of the vIL-6 constructions. RT-PCR

products amplified with PCR primer pairs for vIL-6.K (v) and β-actin (a), used as a

normalization control, are indicated. Omission of reverse transcriptase (-RT) provided a

control for RNA-derived vIL-6.K PCR products (i.e., lack of DNA contamination).

76

Figure 3.5

77

Figure 3.6. Inhibitory targeting of vIL-6/gp130 signaling diminishes PEL cell proliferation

and survival. (A) Functional testing of KDEL motif-fused (ER-directed and retained) and

Flag epitope-tagged soluble gp130 (sgp130.K) as an inhibitor of ER-localized vIL-6

signaling in transfected HEK293T cells. Cultures were co-transfected with expression

vectors for sgp130.K (or empty vector control, vec) and either vIL-6, vIL-6-KDEL (vIL-6.K),

or hIL-6. Cell lysates were harvested 48 hrs post-transfection for preparation of cell

lysates. These were analyzed by western blotting for detection of pSTAT3, an indicator of

gp130 signaling, and β-actin (protein loading control). Both vIL-6 (mainly ER-localized but

also secreted) and KDEL-tagged vIL-6 (completely ER-retained [115]) signaling were

inhibited by sgp130.K, whereas signaling by hIL-6 (efficiently secreted) was unaffected.

The data verified the ER-restricted, vIL-6/gp130-inhibitory activity of sgp130.K. (B)

Confirmation of intracellular localization of KDEL motif-fused and Flag-tagged sgp130 and

comparison to (surface-expressed) full-length gp130 via Flag- and gp130-based

immunofluorescence staining of transfected cells. Alexa Fluor 594-conjugated secondary

antibody was used for detection of (red) fluorescence signal. Each protein was expressed

together with KDEL-tagged GFP (GFP.K, used previously as an ER marker [115]) in

vector-co-transfected HEK293T cells. As expected, Flag-specific staining for

sgp130.Flag.KDEL [sgp130(Flag).K] was evident intracellularly and was coincident with

ER-localized GFP signal, whereas gp130 staining was localized mainly to the cell surface.

Stacked (“total”) and representative individual Z-sections (“slice”) are shown for two

examples (1 and 2) taken from each of the transfected cultures. (C) BCBL-1 and JSC-1

cells were transduced with equivalent titers of “empty” (vec) or sgp130-KDEL (sgp130.K)

expressing lentiviral vectors, and growth of cell density-normalized cultures was monitored

after a two-day recovery period as described in Chapter 2 and in the Methods above. Error

bars represent standard deviations from derived average values. (D) Relative proliferation

rates of PEL cells at “day 2” (from growth assays) were determined by BrdU incorporation

78

assays (see Methods and Chapter 2). (E) In a separate experiment, transduction of

sgp130.K into BCBL-1 and JSC-1 cultures led to induction of apoptosis, as determined by

annexin V-Cy3 staining. Cells were harvested and analyzed 3 days post-transduction with

test (sgp130.K) or control (vec) lentiviral vectors.

79

Figure 3.6

80

CHAPTER 4: ROLE OF GP130 IN VIL-6-MEDIATED VIRUS REPLICATION

81

Summary

HHV-8 vIL-6 has been implicated in the pathogenesis of HHV-8-associated

disease and plays a role in maintaining PEL cell growth and viability in addition to

contributing to elevated levels of phosphorylated STAT3 via its interaction with gp130.

Furthermore, STAT3 and ERK1/2 are required for optimal PEL cell growth and survival.

Additionally, vIL-6 interactions with VKORC1v2 contribute to the pro-growth and survival

activities of vIL-6. While the previous study (Chapter 3) investigated the role of vIL-6 in

PEL cell latency, the effects of vIL-6/gp130 signaling on virus replication are unknown.

Thus, the goal of this study was to determine the role of vIL-6 and gp130 in HHV-8

productive replication in both PEL and endothelial cells.

To assess the effects of vIL-6 and gp130 signaling on HHV-8 replication, vIL-6,

gp130, and gp130-induced STAT3 were depleted in both PEL and endothelial cultures.

Furthermore, depletion-complementation experiments were utilized to determine the

relevance of the ER-localized vIL-6/gp130 interaction as it relates to virus replication.

Finally, ER-localized gp130 signaling mutants were employed to evaluate the specific

contributions of vIL-6/gp130-mediated STAT and ERK signaling to virus replication.

Depletion and depletion-complementation experiments revealed that ER-localized

vIL-6 activity via gp130 and gp130-activated STAT signaling, but not ERK activation, were

critical for vIL-6 pro-replication activity. Additionally, vIL-6 and gp130 depletion did not

affect de novo infection or the establishment of latency. These data significantly extend

current understanding of vIL-6 function and associated mechanisms in HHV-8 biology.

82

HUMAN HERPESVIRUS 8 VIRAL INTERLEUKIN-6 SIGNALING THROUGH GP130

PROMOTES VIRUS REPLICATION IN PRIMARY EFFUSION LYMPHOMA AND

ENDOTHELIAL CELLS

This work has been published, and sections are reproduced here (with modifications) with permission from the American Society for Microbiology (Copyright © American Society for Microbiology, [Journal of Virology, 88, 2014, 12167-12172, doi: 10.1128/JVI.01751-14]).

Cousins E, Gao Y, Sandford G, Nicholas J (2014) Human herpesvirus 8 viral interleukin-6 signaling through gp130 promotes virus replication in primary effusion lymphoma and endothelial cells. Journal of Virology 88(20):12167-12172. INTRODUCTION

The previous chapter described the role of vIL-6 in PEL cell growth and viability

and determined that the ER-localized vIL-6/gp130 interaction is critical for these pro-

growth and survival activities of the viral cytokine. Additionally, the elevated levels of

phosphorylated STAT3 commonly observed in PEL could be attributed, in part, to vIL-

6/gp130 signaling. This chapter aims to expand on these findings and to characterize the

role of vIL-6 and its gp130-mediated downstream signaling components in the reactivation

and replication of HHV-8 in both PEL and endothelial cells.

As has been stated previously, HHV-8 is associated with PEL in addition to MCD

and KS [256-258,282]. HHV-8-encoded vIL-6 promotes cell growth and survival,

inflammation, and angiogenesis and is implicated in pathogenesis and promotes disease

progression [115,144,147,169,259]. However, its role in normal virus biology remains

unclear. vIL-6 and hIL-6 share ~25% sequence identity, fold into similar three-dimensional

structures, and initiate signaling through similar complexes [157-159,165,260,261]. While

both hIL-6 and vIL-6 form hexameric signaling complexes comprised of two gp80 receptor

subunits, two gp130 signal transducers, and two IL-6 molecules (gp802-gp1302-IL-62), vIL-

6 is unique in that it can also signal from the ER via tetrameric complexes (gp1302-vIL-62)

[159,165,170]. vIL-6 can be expressed during PEL cell latency in addition to being

83

maximally expressed during the lytic stage of the viral life cycle [115,116]. Furthermore,

latently infected PEL cell viability is maintained in part via vIL-6/gp130 signaling from the

ER and also through its interaction with VKORC1v2 [169,173]. The vIL-6/VKORC1v2

interaction has also been shown to contribute to productive replication [175].

Upon vIL-6-induced gp130 dimerization, phosphorylation of specific gp130

tyrosine residues initiates either STAT1/3 (Y767, Y804, Y905, and Y915) or MAPK (Y759)

signaling [156,161-163]. Activated STAT3 and ERK1/2 have been shown to promote

growth and survival of PEL cells, and vIL-6/gp130 signaling contributes [40,152,169].

However, the role of vIL-6 in HHV-8 lytic replication is unclear. One study utilized a vIL-6-

knockout recombinant virus to assess the role of the viral cytokine in virus replication in

HHV-8 negative BJAB cells, but no effect was observed [283]. A second study found that

vIL-6 depletion in BCBL-1 and JSC-1 PEL cells resulted in less HHV-8 virus production

following chemically-induced reactivation in culture. No other studies have been

conducted to assess whether vIL-6 plays a role in HHV-8 productive replication; thus, we

sought to determine the effects of vIL-6/gp130 signaling on HHV-8 lytic replication in both

PEL and endothelial cells.

METHODS

Cell culture, transfections, and lentivirus production

HEK293T, Vero-r219, TRExBCBL1-RTA, JSC-1, and TIME-TRE/RTA cells were

maintained as described in Chapter 2. For all transfections, HEK293T cells were

transfected using standard calcium-phosphate-DNA co-precipitation methods as detailed

in Chapter 2. Cells were harvested 24 to 48 hrs post-transfection, and lysates were

prepared as outlined previously (Chapter 2). Lysates were analyzed immediately via

western blot analysis or stored at -20°C for later use.

84

To produce lentivirus, HEK293T cells were transfected with pYNC352-encoded

shRNAs (specific for vIL-6, gp130, NS, or STAT3) or pDUET001-encoded plasmids

(empty vector, wild-type [WT] gp130-Flag, gp130.STAT-Ys, gp130.ERK-Y, and

gp130.ΔYs) or pYNC419-encoded plasmids (mvIL-6.KDEL, mvIL-6-W167G-KDEL, or

luciferase) with Pax2 gag/pol and VSV/G expression vectors as described previously and

in Chapter 2 [173]. To harvest lentivirus, medium was collected 2 days post-transfection,

filtered, and centrifuged as detailed in Chapter 2. Viral pellets were resuspended in RPMI

1640 medium and stored at -80°C until needed. Prior to experimental use, lentiviruses

were titered on HEK293T cells to ensure equivalent titers as described previously and in

Chapter 2 [169].

HHV-8 r219 virus was produced as described in Chapter 2 for the infection of naïve

TIME-TRE/RTA cells. Harvested r219 virus was titered on naïve TIME-TRE/RTA cells via

virus-encoded GFP expression prior to infection of TIME-TRE/RTA cells (Chapter 2). After

allowing cultures to recover for two days, r219-infected TIME-TRE/RTA cells were

transduced with lentiviral vectors to assess HHV-8 reactivation (Chapter 2).

Plasmids and Primers

The pYNC352 shRNA-encoded lentiviral vectors (NS, vIL-6, gp130, and STAT3)

have been described [115,169]. The mvIL-6.KDEL, mvIL-6.W167G.KDEL, and luciferase

lentiviral pYNC419 vectors have been reported [115,169]. pSG5-derived vIL-6 and hIL-6

and the pDUET001 vector have also been reported [166,168,270]. To generate the

pDUET001-encoded gp130 WT and signaling mutants that are resistant to depletion by

gp130-specific shRNA, wild-type gp130, gp130.STAT-Ys (Y759F-mutated), gp130.ERK-

Y (Y759 only, all other Ys mutated to F), and gp130.ΔYs (pan Y-to-F-mutated) (described

in [168]) were PCR-amplified from the pEF-BOS vector using gp130-BamHI-FWD (5’

TTGGATCCATGTTGACGTTGCAGACTTGG 3’) and gp130-1205res-REV (5’

85

AGATGGTCTGTCCTCGTACGTAATTCCGCTTGCTTCTTCACTCCAGT 3’) primers to

amplify the 5’ end of the gene. To add a C-terminal FLAG epitope and to amplify the 3’

end of WT gp130 and gp130.STAT-Ys, gp130-1205res-FWD (5’

GAAGAAGCAAGCGGAATTACGTACGAGGACAGACCATCTAAAGCACC 3’) and

gp130-WT-Flag-REV (5’ TTCTCGAGTCACTTGTCATCGTCGTCCTTGTAATCCTGAGG

CATGTAGCCGCCTTG 3’) primers were used for PCR amplification. Similarly, gp130-

1205res-FWD and gp130-deltaY-Flag-REV (5’ TTCTCGAGTCACTTGTCATCGTCGTCC

TTGTAATCCTGAGGCATGAAGCCGCCTTGCCGTACAGT 3’) primers were used to

amplify the 3’ end of gp130.ERK-Y and gp130.ΔYs and to add a FLAG epitope to the C-

terminus. The 5’ and 3’ amplicons for each gp130 gene (WT or signaling mutant) were

then used as templates to amplify across the entire length of gp130-Flag; the resulting

PCR products were resistant to the gp130-specific shRNA and contained C-terminal Flag

epitope tags. Finally, the gp130-shRNA-resistant and Flag epitope-tagged amplicons were

ligated into pDUET001 between the BamHI and XhoI restriction sites, replacing the GFP

ORF. All of the constructions were verified by sequencing.

RT-PCR primers were designed to amplify gp130-Flag sequences in transduced

TRExBCBL1-RTA cells in order to verify the expression of wild-type and gp130 signaling

mutant plasmids. The following primers were used: gp130-Flag-FWD (5’

TAGATGGCGGTGATGGTATTT 3’) and gp130-Flag-REV (5’

CTTGTCATCGTCGTCCTTGTA 3’). Previously validated glyceraldehyde-3-phosphate

dehydrogenase (GAPDH) primers were used to amplify GAPDH as a positive control:

GAPDH-FWD (5’ TTGCCATCAATGACCCCTTCA 3’) and GAPDH-REV (5’

CGCCCCACTTGATTTTGGA 3’); these primers were designed at the RRC Core

Genomics Facility at the University of Illinois at Chicago.

86

Antibodies and Western blotting

The following commercially available antibodies were used: β-actin (Sigma, St.

Louis, MO; catalog number A2228), pSTAT3 (Cell Signaling Technology, Beverly, MA;

catalog number 9131), STAT3 (Santa Cruz Biotechnology, Santa Cruz, CA; catalog

number sc-482), pERK (Santa Cruz Biotechnology; catalog number sc-7383), LANA

(Advanced Biotechnologies, Columbia, MD; catalog number 13-210-100), and gp130

(Santa Cruz Biotechnology; catalog number sc-655). Rabbit anti-serum to vIL-6 was

generated previously in our laboratory [166]. Cy3-goat anti-rat IgG fluorescent antibody

(Life Technologies, Carlsbad, CA; catalog number A10522) was used to detect LANA by

immunofluorescence assay.

Western blotting was conducted by separating heat-denatured and reduced lysate

samples via SDS-PAGE and transferring samples to nitrocellulose membranes using

standard protocols described in Chapter 2. Following antibody incubation, immuno-

targeted proteins were visualized via standard chemiluminescence assay.

Quantification of HHV-8 titers following reactivation

For HHV-8 reactivation experiments in TRExBCBL1-RTA and JSC-1 PEL cells,

cells were transduced with pYNC352-encoded shRNAs as described previously [169] and

in Chapter 2, allowed to recover for 2 days, normalized for cell number, and then induced

with Dox (TREx-BCBL1/RTA cells; 1 μg/μl) or TPA (20 nM) and NaB (0.5 μM) for JSC-1

cells. NaB was removed from JSC-1 cultures after 12 hrs of incubation. Transduced PEL

cells were maintained in Dox or TPA for 4 days, and HHV-8 virus was harvested from

culture media as detailed in Chapter 2. Cells were counted at the time of virus harvest to

ensure no drastic differences in cytotoxicity between cultures. Aliquots of harvested virus

were added to naïve TIME-TRE/RTA cells plated at ~90% confluency in 12-well plates,

87

centrifuged, and incubated for an additional hr as outlined in Chapter 2. One day post-

infection, cells were trypsinized and added to chamber slides for LANA immunostaining.

For reactivation of HHV-8 in TIME-TRE/RTA endothelial cells, cells were seeded

at ~90% confluency in 12-well plates and then infected with HHV-8 r219 virus (described

above and in Chapter 2). After recovering for 48 hrs, TIME-TRE/RTA GFP fluorescence

was assessed to verify a high viral titer (~90% of cells infected). Cells were then infected

with pYNC352-encoded shRNA-expressing lentiviruses as described above and in

Chapter 2. Two days post-shRNA lentiviral infection, cells were induced with Dox (1 μg/μl).

Medium was removed from infected and induced cells every 48 hrs for 6 days, and fresh

medium containing Dox was added to cell cultures. Harvested media containing virus were

stored at -80°C. Reactivated TIME-TRE/RTA cells were also visualized daily for RFP

fluorescence to assess rates of reactivation. Medium collected from each sample was

pooled and centrifuged at 16,000 rpm for 2 hr at 4°C; viral pellets were resuspended in

EMB-2 without serum or supplements. Naïve TIME-TRE/RTA cells were incubated with

an aliquot of collected virus and then transferred to chamber slides for immunostaining as

described above.

For depletion-complementation experiments in TRExBCBL1-RTA cells, cells were

first transduced with pYNC419-expressing mvIL-6.KDEL, mvIL-6.W167G.KDEL, or

luciferase to assess vIL-6 effects on virus replication or were transduced with pDUET001-

expressing empty vector, WT gp130-Flag, gp130.STAT-Ys, gp130.ERK-Y, or gp130.ΔYs

to determine the contribution of STAT and ERK signaling to virus replication. Cells were

allowed to recover for 2 days prior to being transduced with pYNC352-encoded shRNAs

as described in Chapter 2. Dually-transduced cells were induced with Dox; medium was

collected, and virus was harvested as outlined in Chapter 2. Naïve TIME-TRE/RTA cells

were infected with aliquots of harvested virus as described above, and cells were

transferred to chamber slides prior to immunostaining.

88

Chamber slides containing infected TIME-TRE/RTA cells were stained for LANA

expression to quantify HHV-8 infection, as outlined in Chapter 2. Chamber slides were

washed, fixed in ice cold methanol, and incubated with anti-LANA antibody (1:100 dilution

in PBS blocking buffer [Thermo Fisher Scientific, Waltham, MA]) as outlined in Chapter 2.

After thoroughly washing, slides were incubated with Cy3-goat anti-rat secondary antibody

(1:200 dilution in PBS blocking buffer), washed, and then stained with DAPI to detect

nuclei. The percentage of LANA+ cells in the total DAPI+ cell population was determined

from 6 random fields (>50 cells/field) in each duplicate culture.

De Novo Infection

TIME-TRE/RTA cells were transduced with lentivirus-encoding shRNAs specific

for NS, gp130, or vIL-6. Two days post-transduction, cells were infected with a

subsaturating dose of HHV-8 r219 virus. One day post-HHV-8 r219 infection (and 3 days

post-lentiviral transduction), infected and transduced TIME-TRE/RTA cells were washed,

fixed, and stained for LANA and DAPI expression as described in Chapter 2. Multiple (>6)

fields were visualized for each duplicate culture, and the percentage of LANA+ cells in the

total DAPI+ cell population was calculated as outlined above.

Annexin Staining

TIME-TRE/RTA cells were plated on chamber slides at a confluency of ~80%. The

following day, TIME-TRE/RTA cells were transduced with NS-specific or STAT3-specific

lentivirus-encoded shRNAs. One day post-transduction, cells were fixed and stained for

annexin V-Cy3 as described in Chapter 2. Multiple fields were visualized, and mean values

were reported.

89

RESULTS

gp130 and vIL-6 promote HHV-8 replication in PEL and endothelial cells. The

gp130 signal transducer was depleted in TRExBCBL1-RTA and JSC-1 PEL cells

[100,111] and in TIME-TRE/RTA endothelial cells [203] latently infected with HHV-8 r219

[254] using previously established techniques [115,169]. Following depletion, lytic

replication was induced with Dox or TPA and NaB. HHV-8 was collected from PEL and

endothelial culture media, and infectious viral titers were determined as undertaken

previously [175]. Viable (trypan blue-excluding) TRExBCBL1-RTA cell densities at the

time of media harvest were not negatively affected by gp130 transduction (1.31x106 and

1.41x106 cells per ml for NS shRNA-transduced duplicate cultures versus 1.75x106 and

1.83x106 cells per ml for gp130 shRNA-transduced cells). Depletion of gp130 led to

decreased HHV-8 titers in PEL (~3-fold) and endothelial cultures (>8-fold) (Figure 4.1A,

B, C). Similarly, vIL-6 depletion in endothelial cells reduced viral titers by ~3-fold (Figure

4.1D). TIME-TRE/RTA cells were visibly unaffected by gp130 or vIL-6 depletion, and rates

of apoptosis (% annexin V-Cy3+ cells) were equivalent in gp130- and vIL-6-depleted

cultures relative to NS shRNA-transduced controls (data not shown). Rates of reactivation,

measured by ORF59-antigen immunofluorescence (PEL cells, data not shown) and RFP

imaging (Figure 4.1C,D, bottom panels), were equivalent in depleted and control cultures.

Combined, these data identify the importance of gp130 for productive replication in PEL

and endothelial cells and the involvement of vIL-6 in support of productive replication in

endothelial cells in addition to BCBL-1 and JSC-1 cells, reported previously [175]. In

contrast to reactivated productive replication, neither gp130 nor vIL-6 depletion had any

substantial effect on latency titers (measured 3 days post-infection) following de novo

infection of endothelial cells (Figure 4.1E), a process known to involve a burst of

expression of a subset of lytic genes, including vIL-6, prior to rapid latency establishment

[284,285].

90

ER-localized vIL-6/gp130 signaling contributes to productive replication.

Dox-inducible TRExBCBL1-RTA PEL cells were used to test the ability of gp130 signaling-

competent and ER-retained vIL-6 to rescue the vIL-6-depletion phenotype. As before,

lentiviral vectors encoding either control (NS) or vIL-6 mRNA-directed shRNAs were used

for transduction. Depletion of vIL-6 led to reduced virus production (luciferase control),

and this phenotype could be rescued with lentiviral vector-transduced vIL-6.KDEL (ER-

retained [115]) but not vIL-6.W167G.KDEL (ER-retained and gp130-dimerization

incompetent in the ER [166,169]) (Figure 4.2). These data identify the involvement of ER-

localized vIL-6/gp130 signaling in support of HHV-8 productive replication.

gp130-mediated STAT signaling, but not ERK signaling, supports HHV-8

replication. To determine the significance of gp130-activated STAT versus ERK signaling

in HHV-8 replication, we employed signaling-altered gp130 variants in gp130 depletion-

rescue experiments. Four C-tail tyrosines of gp130 initiate STAT1/3 signaling, and one

induces MAPK signaling [156,161-163]. Lentiviral vector-cloned gp130 ORFs for

complementation encoded wild-type gp130 (WT), gp130.STAT-Ys (Y759F-mutated,

ERK1/2 signaling-null), gp130.ERK-Y (Y759 only, STAT1/3 signaling-null), and

gp130.∆Ys (inactive). Functional testing in transfected HEK293T cells verified that in the

presence of vIL-6, introduced WT and gp130.STAT-Ys induced STAT3 phosphorylation

above levels supported by endogenous gp130, and gp130.ERK-Y exclusively induced

pERK levels (Figure 4.3A). Unexpectedly, WT gp130 did not support detectable ERK

signaling in response to vIL-6, but the same result was seen for hIL-6 in the presence of

overexpressed gp80 (data not shown). In TRExBCBL1-RTA-based depletion-

complementation experiments, WT and gp130.STAT-Ys, but not gp130.ERK-Y or

negative controls gp130.ΔY and empty vector, fully rescued replication (Figure 4.3B),

indicating that STAT signaling alone is sufficient for vIL-6/gp130-mediated support of

HHV-8 replication.

91

Contribution of STAT3 to HHV-8 replication. As previous reports have

demonstrated the importance of STAT3 for latent PEL cell proliferation and viability

[152,169], we examined its role in productive replication. STAT3 depletion led to reduced

HHV-8 titers induced from both PEL and endothelial cultures, although effects in

TPA/NaB-induced JSC-1 cultures were modest despite robust STAT3 depletion (Figure

4.4). The latter may reflect different PEL cell-specific thresholds of STAT3 functionality or

different dependencies on STAT3 for support of HHV-8 replication. It is important to note

that viable cell densities at the end of the experiment were equivalent between NS and

STAT3-directed shRNA-transduced cultures (Figure 4.4 panels A and B, bottom),

indicating that pro-proliferative and anti-apoptotic activities of STAT3 operating in latently

infected PEL cells [152,169] do not account directly for diminished replication resulting

from STAT3 depletion in reactivated cultures. Also, STAT3 depletion in TIME-TRE/RTA

cells had no detectable effect on cell viability (Figure 4.4D). Our data show that STAT3

supports HHV-8 replication, and together with the vIL-6 and gp130 depletion-rescue data

(Figures 4.2 and 4.3) indicate that vIL-6/gp130 signaling via STAT3 contributes

significantly to the pro-replication activities of vIL-6 and gp130.

DISCUSSION

The likely role of vIL-6 in pathogenesis has been described in numerous reports

[115,144,147,169,258,259]. However, investigation of the role of vIL-6 in HHV-8

replication is restricted to one study of vIL-6-null virus replication in (HHV-8-negative)

BJAB cells, which noted no phenotype [283], and our own preliminary vIL-6 depletion

experiments in PEL cells [175]. Previously, vIL-6 was shown to promote HHV-8 replication

in BCBL-1 and JSC-1 PEL cells, but its role in endothelial cells and additional PEL cells

was not evaluated. The present study significantly extends these previous reports,

demonstrating that vIL-6 promotes HHV-8 productive replication in endothelial cells in

92

addition to PEL cells, that gp130 contributes significantly to HHV-8 replication in these cell

types, that vIL-6 activity via gp130 can be mediated largely or exclusively through ER-

localized signaling via tetrameric complexes, and that STAT3, but not gp130-activated

MAPK signaling, is likely critical for vIL-6/gp130-enhanced replication.

While it possible that gp130-activated STAT1 signaling could also contribute, lack

of detectable effects of STAT1 depletion on PEL cell growth [169] coupled with the well-

established role of STAT1 in promotion of anti-viral interferon and apoptotic signaling [286]

indicate that this is unlikely. It is noteworthy that previous studies have reported the

importance of STAT3 signaling for replication of viruses, including VZV and hepatitis C

virus [287,288]. For VZV, STAT3-activated survivin was found to be involved in pro-

replication activity, reflecting the pro-survival activities of STAT3 via survivin reported in

PEL cells [152]. However, HCMV blocks STAT3 phosphorylation, though requires it for

optimal replication, and mouse cytomegalovirus induces phosphorylation of STAT3 but

not STAT3-responsive cellular genes, suggesting novel, virus-redirected activities of the

transcription factor [289,290]. Our findings of vIL-6/gp130 and STAT3 involvement in HHV-

8 productive replication could facilitate the development of therapeutic interventions for

the treatment of HHV-8-associated diseases, in which productive replication contributes

significantly.

93

FIGURES

Figure 4.1. Involvement of gp130 and vIL-6 in promotion of HHV-8 replication. (A) HHV-8

titers following gp130 depletion in TRExBCBL1-RTA cells. The gp130 signal transducer

was depleted in TRExBCBL1-RTA cells using lentiviral-encoded shRNAs (NS, control or

gp130-specific) described previously [169]. Cell numbers were normalized, and cultures

were reactivated with Dox. Aliquots of harvested virus were applied to naïve TIME-

TRE/RTA cells to determine infectious virus titers by staining for LANA, as described

previously [175]. Multiple random fields were assessed to quantify the number of LANA-

positive cells in the total DAPI-positive population. (B) Importance of gp130 in HHV-8

replication in JSC-1 cells. JSC-1 cells were depleted of gp130 (as described in 4.1A) and

induced with NaB and TPA. Released virus was pelleted from culture media and quantified

as before. (C) and (D) Influence of gp130 and vIL-6 on HHV-8 replication in endothelial

cells. TIME-TRE/RTA cells were first infected with HHV-8 r219 virus [254], then

transduced with shRNA-encoding lentiviral vectors (NS, control; gp130-specific; or vIL-6-

specific), and finally induced with Dox. Lentiviral vector-encoded GFP, expressed above

levels arising from latent r219 HHV-8 genomes, was used to monitor lentiviral infection,

achieved in >90% of cells. HHV-8 reactivation was monitored by visualizing RTA-induced

RFP expression from r219 viral genomes via fluorescence after Dox treatment (examples

shown). Media containing virus were collected for six days and pooled prior to virus

concentration and titration. (E) NS, gp130, and vIL-6 shRNA-transduced TIME-TRE/RTA

cells were infected with HHV-8 r219 two days post-lentiviral transduction to determine

possible effects of gp130 and vIL-6 depletion on de novo infection and establishment of

latency, as determined by LANA staining three days post-HHV-8 r219 infection. For all

panels, data were derived by counting of LANA+ cells in multiple fields to obtain the

percentage of LANA+/DAPI+ cells. Average values from duplicate cultures are shown; error

bars represent deviations of individual values from the means.

94

Figure 4.1.

95

Figure 4.2. Importance of ER-localized vIL-6/gp130 interactions for HHV-8 replication.

TRExBCBL1-RTA cells were transduced with lentiviral expression vectors encoding KDEL

motif-tagged (ER-retained) and shRNA-resistant forms of vIL-6 (vIL-6.K) or vIL-6.W167G

(W167G.K) [115,169] or with lentivirus encoding luciferase (negative control). Two days

post-transduction, cells were transduced with control (NS) shRNA-encoding lentivirus or

lentivirus expressing vIL-6-specific shRNA. Dually-transduced cells were then normalized

for cell density and induced the following day with Dox. Virus was collected and titered as

before (see Figure 4.1 legend). RNA was harvested from a subset of cells using TRIzol,

and cDNA was amplified for vIL-6.KDEL [169] and GAPDH (normalization and positive

control) mRNA to verify expression of the introduced vIL-6 genes (v, vIL-6 RT-PCR; g,

GAPDH RT-PCR; -RT, no reverse transcriptase). Viable cell densities at the time of Dox

treatment (day 0, normalized) and at virus harvest (day 4) are shown (bottom panel). v,

vIL-6-KDEL; v167.K, vIL-6.W167G-KDEL; luc, luciferase.

96

Figure 4.2

97

Figure 4.3. Contributions of gp130-mediated STAT and ERK signaling to HHV-8

productive replication. (A) Functional assessment of gp130 signaling-tyrosine variants.

HEK293T cells were co-transfected with empty vector (-) or vIL-6 plasmid (+) together with

empty lentiviral vector or lentiviral vectors expressing wild-type gp130, gp130.STAT-Ys

(Y759F-mutated), gp130.ERK-Y (Y759 only, other C-tail tyrosines mutated to F), or

gp130.∆Ys (pan Y-to-F-mutated). Two days post-transduction, cells were lysed for

western blot analysis to detect pSTAT3 and pERK as indicators of gp130-activated STAT

and MAPK pathways. (B) Role of STAT and ERK signaling in vIL-6/gp130-promotion of

HHV-8 replication. TRExBCBL1-RTA PEL cells were transduced with empty lentiviral

vector (pDUET001 [270]) or vectors expressing Flag-tagged wild-type gp130,

gp130.STAT-Ys, gp130.ERK-Y, or gp130.∆Ys. Two days post-transduction, cells were

transduced with either NS (control) or a gp130-specific shRNA-encoding lentiviral vector.

Following Dox-induced HHV-8 reactivation for four days, virus was collected and titered

as described previously. To verify appropriate expression of wild-type gp130 and the

gp130 signaling mutants, a subset of cells was lysed in TRIzol for preparation of RNA

samples. RNA was reverse transcribed and amplified using 3’ Flag- and 5’ gp130-directed

primers to assess expression of transduced gp130-Flag mRNAs. GAPDH-specific primers

were used for normalization and to provide a positive control. Flag, gp130-Flag RT-PCR;

g, GADPH RT-PCR; -RT, no reverse transcriptase.

98

Figure 4.3

99

Figure 4.4. Role of STAT3 in HHV-8 replication in PEL and endothelial cells. (A)

TRExBCBL1-RTA cells were transduced with control (NS) and STAT3-specific shRNA-

encoding lentiviral vectors. After two days, lytic reactivation was induced with Dox for four

days, and virus was then collected and concentrated from culture media for titration.

Experimental procedures and data collection and calculations were as outlined in the

legend to Figure 4.1. Subsets of transduced cells were harvested at the time of induction

and lysed for western blot analysis to verify STAT3 depletion. Densities of viable (trypan

blue-excluding) cells were normalized at the time of lytic induction (day 0) and determined

at the end of the experiment (day 4) (bottom panel). (B) The same analyses were

undertaken for JSC-1 cultures, induced with TPA/NaB and harvested after four days for

virus titration. A subset of cells was lysed for western blotting at the time of induction to

verify STAT3 depletion. Viable cell densities at the start and end of the experiment are

shown (bottom). (C) Dox-induced TIME-TRE/RTA cultures were analyzed in the same

way except that virus was harvested following six days of Dox treatment; parallel cultures

were harvested at the time of induction for western blot analysis. (D) Rates of apoptosis

(% annexin V-Cy3+/DAPI+ cells) in similarly transduced TIME-TRE/RTA cultures were

unaffected by STAT3 depletion.

100

Figure 4.4

101

CHAPTER 5: DISCUSSION AND PERSPECTIVES

102

DISCUSSION OF MAJOR FINDINGS

The goal of this project was to further characterize the function of vIL-6 as it relates

to normal HHV-8 biology and pathogenesis. Many previous studies have sought to

determine the role of vIL-6 in viral pathogenesis and tumorigenesis. While pathogenesis

studies are important, virus-induced tumorigenesis is an unintentional byproduct of HHV-

8 infection and occurs in rare cases. In most immunocompetent individuals, host immune

responses are able to control the virus, and HHV-8-associated malignancies are not

observed. When the immune system is incapable of controlling virus replication and

spread, the resulting increase in infection events provides many more opportunities for the

virus to cause malignant transformation. Thus, determining how vIL-6 facilitates viral

infection under normal physiological conditions is crucial for understanding how the virus

maintains its fitness and expands within the host. The development of therapeutics to

combat viral infection and HHV-8-mediated neoplasia will require a detailed understanding

of virus-host interactions under normal, non-transformative conditions.

Studies presented here indicate that vIL-6 promotes latently infected PEL cell

growth and survival, and they identify an important and therapeutically-targetable

mechanism of PEL cell maintenance. Additionally, these studies demonstrate a role of

vIL-6 signaling in normal virus latency in untransformed B-cells and possibly other cell

types. Thus, disruption of vIL-6 signaling through specific abrogation of its interaction with

gp130 could be a promising means not only to target PEL therapeutically but also to attack

latently infected cells more generally and limit viral load. Furthermore, this project has

identified intracellular signaling by vIL-6 specifically within the ER as leading to high levels

of activated (phosphorylated) STATs 1 and 3 and ERKs 1 and 2. Moreover, STAT3 and

ERK1/2 expression are important for PEL cell growth and survival, and these pathways

could be targeted pharmacologically. These findings are in agreement with a previous

study that determined that STAT3 promotes PEL cell growth and survival via the

103

upregulation of survivin though data described in this dissertation project were generated

using an alternative method to assess the contribution of STAT3 to PEL cell proliferation

and viability. This current work (described in Chapter 3) utilized shRNA-encoding lentiviral

vectors to effectively and specifically deplete STAT3 in multiple PEL cell lines in order to

confirm previous findings and to establish that vIL-6/gp130 signaling from the ER

compartment contributes significantly to the increased levels of phosphorylated STAT3.

Furthermore, a similar approach was undertaken to assess the role of ERK1/2 signaling

as it relates to PEL cell growth and survival. For the first time, ERK1/2 was found to play

a major role in the proliferation and viability of PEL cells in culture. This finding, along with

data demonstrating that depletion of ERK1/2 in HHV-8 negative BJAB cells had no effect

on cell growth, suggests that ERK1/2 activity may be specifically required by HHV-8-

infected PEL cells. The effects of STAT3 and ERK1/2 on PEL cell growth were assessed

by measuring rates of both cell proliferation and apoptosis; depletion of both signaling

molecules led to decreased rates of cell growth and increased rates of apoptosis.

Moreover, vIL-6-mediated signaling through gp130 was found to be responsible, at least

in part, for high levels of both phosphorylated STAT3 (in all PEL cell lines tested) and

ERK1/2 (in some PEL cell lines). These data also suggest that the constitutively high levels

of phosphorylated STAT3 observed in PEL cells can be attributed largely to vIL-6/gp130

signaling. Finally, additional experiments were performed to evaluate the importance of

ER-localized vIL-6 signaling; data from these depletion-complementation experiments

indicate that vIL-6/gp130 ER-localized signaling is responsible for promoting PEL cell

growth and survival in culture, as a vIL-6 mutant (W167G, incapable of dimerizing gp130 in

the ER) was unable to rescue PEL cell growth following vIL-6 depletion. Thus, the study

described in Chapter 3 adds to the body of work regarding vIL-6 function and intracellular

signaling in the context of PEL cell proliferation and viability and has implications not only

for PEL therapy but also for normal HHV-8 latency.

104

In addition to functioning as a pro-growth and pro-survival viral protein in PEL cells,

we hypothesized that vIL-6 might also play a positive role in productive virus replication

due to its role in prolonging cell survival. Thus, vIL-6 function in HHV-8 lytic replication

was assessed in both PEL and endothelial cells. Using shRNA-encoding lentiviral vectors,

vIL-6 and gp130 were depleted in both cell types. Data from these experiments indicate

that vIL-6 and gp130 are required for efficient HHV-8 replication. Prior to these

experiments, little was known regarding the function of vIL-6 in HHV-8 productive

replication. One study utilized a vIL-6 knock-out recombinant virus to assess replication in

HHV-8 negative BJAB cells, and in this setting, vIL-6 appeared to have no effect on virus

replication [283]. In a second study, the function of vIL-6 in HHV-8 replication was

evaluated in two PEL cell lines (BCBL-1 and JSC-1), and decreased viral titers were

observed following vIL-6 depletion via vIL-6-specific lentivirus-encoded shRNA [175].

These studies also identified vIL-6-targeted VKORC1v2 and associated suppression of

pro-apoptotic CatD were involved in promoting HHV-8 replication. The findings described

in Chapter 4 are novel in that they demonstrate that vIL-6 promotes virus replication in

both endothelial cells and in an additional PEL cell line (TREx-BCBL1/RTA). Additionally,

the contributions of gp130 to HHV-8 productive replication were evaluated using similar

methods, and gp130 depletion led to decreased viral titers in both cell types. Furthermore,

vIL-6 depletion-complementation experiments were conducted using both wild-type and

mutant vIL-6 (W167G) to determine the contribution of vIL-6/gp130 intracellular signaling

on productive replication in PEL cells. This experiment demonstrated that vIL-6/gp130 ER-

localized signaling is required for effective virus replication, as the ER-localized vIL-6

mutant (which was unaffected with respect to VKORC1v2 interaction) was incapable of

rescuing HHV-8 replication following vIL-6 depletion. As expected, ER-directed wild-type

vIL-6 was capable of rescuing replicative titers following lentivirus-mediated shRNA

depletion of vIL-6. STAT and MAPK signaling are activated by vIL-6/gp130 signaling, and

105

these downstream signaling molecules were evaluated to determine their contributions to

HHV-8 productive replication. A set of gp130 signaling mutants were over-expressed

following gp130 depletion in a depletion-complementation assay, and data from this

experiment indicate that vIL-6/gp130-mediated STAT3 signaling is required for optimal

HHV-8 replication and that MAPK signaling is not required for HHV-8 replication. Thus,

vIL-6, gp130, and STAT3 have been identified for the first time as positive regulators of

HHV-8 replication.

The studies described here and in other reports suggest that STAT signaling is

critically important for virus replication and host cell proliferation and survival. STATs are

known to be upregulated following viral infection and the activation of type 1 interferons.

While STATs 1 and 2 facilitate and promote the host IFN response, STAT3 negatively

regulates type 1 interferons [291]. Depletion and/or knockout of STAT3 in cell lines leads

to an enhanced type 1 IFN response, and the addition of STAT3 following depletion

dampens the type 1 IFN response [291]. Additionally, STAT3 knockout mice are

embryonic lethal; conversely, STAT1 knockout mice are viable but are unresponsive to

IFNα and IFNƔ, and STAT2 knockout mice also display greater susceptibility to viral

infections [292-294]. Therefore, the activation of STAT3 following virus infection is

beneficial to the virus because STAT3 attenuates the host antiviral response by

suppressing the pro-type 1 IFN response promoted by STATs 1 and 2.

Many viruses, including HHV-8, modulate STAT3 activity to promote virus

replication and to evade host immune responses. For example, VZV induces STAT3

phosphorylation and upregulates survivin both in cell culture and in virus-infected SCID

mice; inhibition of STAT3 and/or survivin expression limits VZV replication and spread in

cell-based models [287]. VZV also inhibits STAT1, which ultimately attenuates the host

type 1 IFN response [287]. HCMV inhibits STAT3 phosphorylation and sequesters the

transcription factor (in its unphosphorylated form) in the nucleus; consequently, IL-6-

106

mediated gene expression is inhibited [289]. However, STAT3 activity is necessary for

maximal HCMV replication, and inhibition of STAT3 by various chemical inhibitors resulted

in drastically reduced viral titers [289]. Murine CMV (MCMV) inhibits the transcription of

STAT3-responsive genes (including suppressor of cytokine signaling 3 [SOC3], a negative

regulator of STAT3) while simultaneously upregulating STAT3 phosphorylation [290].

HSV1 latency is maintained by STAT3 expression, and the inactivation of STAT3 via

inhibitors or a dominant negative STAT3 induces HSV1 productive replication [295]. In

addition to herpesviruses, the rabies virus P protein interacts with STAT3, inhibiting the

accumulation of the transcription factor in the nucleus [296]. Furthermore, targeting of

STAT3 by the rabies virus abrogates gp130-mediated signaling [296]. The mumps virus,

a member of the paramyxovirus family, also interacts with STAT3 via its V protein; this

interaction leads to the proteasomal degradation of the transcription factor and the

inhibition of IL-6-induced signaling [297]. HIV-1-encoded Nef activates STAT3 by inducing

the expression of soluble factors, including Mip-1α, in monocyte-derived macrophages

[298]. Hepatitis C virus (HCV) also modulates STAT3 activity via interactions with the HCV

core protein [288]. Addition of constitutively active STAT3 leads to increased HCV

replication, and the inhibition of STAT3 via siRNA or chemicals (AG490, STA-21, and S31-

201) results in lower HCV RNA levels [288]. Thus, the targeting and modulation of STAT3

activity is conserved by many viruses across multiple families, suggesting that STAT3

function is crucial to virus biology. Whether STAT3 acts to enhance or inhibit virus

replication and whether a particular virus induces or suppresses STAT3 activation

presumably is dependent on the particular biological requirements of the viruses, their

specific abilities to modify STAT3 activity, and the cell types infected.

STAT3 not only plays a role in regulating virus infection, but it is also involved in

oncogenesis. STAT3 interacts with RelA, a nuclear factor ĸB (NFĸB) family member,

sequestering it within the nucleus; ultimately, this interaction leads to constitutive NFĸB

107

signaling in tumor cells [299]. Furthermore, STAT3 and NFĸB are linked via IL-6. NFĸB

regulates IL-6 expression at the transcriptional level, and IL-6 signaling promotes STAT3

phosphorylation and activation [300]. Thus, NFkB signaling leads to elevated levels of IL-

6, and in turn, high levels of phosphorylated STAT3; this positive feedback loop between

NFkB and STAT3 results in increased inflammation via the upregulation of pro-

inflammatory cytokines [301]. Inflammation is known to contribute to tumorigenesis, and

several cancers have been linked to inflammation caused by virus infection. Virus-induced

inflammatory cancers include HHV-8-associated KS, liver cancer (associated with

hepatitis B and/or C virus), cervical cancer (linked to human papillomavirus [HPV]), and T

cell leukemias and lymphomas (associated with human T lymphotrophic virus [HTLV-1])

[255,302-304]. Viral and cellular cytokines, VEGF, host immune cell extravasation, and

repetitive DNA damage promote inflammation. Therefore, STAT3 activation can lead to

inflammation and a microenvironment that is hospitable for tumor development and

growth.

In immunocompetent individuals, HHV-8 infection does not generally result in

oncogenesis. However, virus infection promotes key pathways, such as angiogenesis and

inflammation, which are often involved in tumor development. HHV-8-encoded vIL-6

increases the expression of VEGF in cell culture and mouse models [147]. VEGF

promotes angiogenesis and causes increased permeability of existing vasculature, which

enhances the infiltration of immune cells into the lesion [305]. HHV-8-encoded vIL-6 also

inhibits neutrophil recruitment during acute inflammation in a murine peritoneal

inflammation model [148]. Data from this mouse model indicate that vIL-6 promotes

STAT3 phosphorylation, CCL-2 production, and inhibition of IL-1β-mediated release of

CXCL8 (IL-8) [148]. CCL-2 recruits monocytes and basophils to the lesion and directs the

immune response away from the anti-viral Th1 response and toward the antibody-

mediated Th2 response [148,306]. HHV-8-encoded viral chemokines (vCCL-1, vCCL-2,

108

and vCCL-3) also promote Th2 polarization, which weakens the host anti-viral immune

response and promotes maintenance of the virus (Table 1.1). IL-8 is a chemotactic factor

for neutrophils and promotes angiogenesis. HHV-8-encoded vFLIP induces IL-8

production that counteracts CCL-2-mediated suppression of IL-8 release, and vFLIP-

regulated induction of IL-8 is dependent on functional NFĸB [226,307]. Thus, STAT3 is a

pleiotropic factor that promotes the growth and viability of virus-infected cells, enhances

inflammation and IL-8 expression via interactions with NFkB, and drives Th2 polarization.

HHV-8 vIL-6 induces STAT3 phosphorylation; therefore, the viral cytokine mediates pro-

inflammatory, pro-angiogenic, and pro-survival signaling in virally infected cells.

IMPLICATIONS FOR CLINICAL RESEARCH

Treatment of viral infections and virus-associated neoplasias has been difficult due

to the fact that some viruses, such as herpesviruses and HIV, have both latent and lytic

phases. Traditional antiviral drugs were designed to target virus-encoded proteins such

as TK, DNA polymerase, reverse transcriptase, and protease. Unfortunately, these drugs

require that the virus be actively replicating. Because the virus is not productively

replicating during latency, many proteins are not expressed and cannot be effectively

targeted during this phase of the viral life cycle. A minority of viral proteins are expressed

during latency. These proteins interact with host proteins to promote latency and genome

maintenance. Therapeutic targeting of intracellular interactions between viral and host

proteins during viral latency may provide a means to target the latent virus population.

Data presented in this body of work suggest means of novel therapeutic

interventions for the treatment of HHV-8-associated disease. Specifically, the

establishment that interactions between vIL-6 and gp130 are critical for viral latency and

viral replication supports the targeting of this protein-protein interaction in future drug

109

design efforts. By preventing this interaction with small molecules or peptides, critical viral

processes might be inhibited. Many experimental approaches were undertaken to

evaluate the vIL-6/gp130 interaction. Specific findings are stated below.

1) Expression of both gp130 and vIL-6 in HHV-8-positive PEL cells is required for cell

growth and survival in culture.

2) Gp130 contributes to elevated levels of active (phosphorylated) STAT1 and STAT3

in PEL cells.

3) Gp130 contributes to levels of active (phosphorylated) ERK1/2 in a subset of PEL

cell lines.

4) STAT3 and ERK1/2 signaling are required for PEL cell proliferation and survival,

but STAT1 is not.

5) ER-localized vIL-6/gp130-mediated signaling promotes PEL cell growth and

survival.

6) Optimal HHV-8 replication requires expression of vIL-6, gp130, and STAT3 in both

PEL and endothelial cells; these proteins do not play a substantial role in de novo

infection (of endothelial cells) or the establishment of latency.

7) The ER-localized vIL-6/gp130 interaction promotes HHV-8 productive replication.

8) Gp130-mediated STAT3 signaling promotes virus replication.

Taken together, these data suggest that the ER-localized vIL-6/gp130 interaction is critical

for multiple viral processes, including growth of latently-infected cells, cell viability, and

productive viral replication. Therapeutic interventions targeting the vIL-6/gp130 interaction

or downstream signaling could theoretically inhibit the growth of latently infected cells

while also limiting the production of virus in lytically infected cells.

110

FUTURE DIRECTIONS

While these studies have shown that vIL-6-mediated signaling through gp130

helps maintain viral latency and is required for optimal lytic replication, many questions

remain. Data indicate that STAT3 and ERK1/2 participate in vIL-6 signaling downstream

of gp130, but the effects of STAT3 and ERK1/2 signaling on cellular gene expression in

virally-infected cells are not known. Future studies should investigate virus-induced

transcriptional changes downstream of STAT3 and ERK1/2 signaling. Results from these

studies may provide additional insight into the ways in which the virus harnesses host

signaling cascades to enhance its own fitness.

Due to the importance of vIL-6/gp130 interactions within the ER, screens should

be conducted to identify small molecule or peptide inhibitors of this interaction within the

ER compartment. Should an in vitro screen identify robust inhibitors of this interaction,

potential inhibitors should be utilized in functional assays to assess their effects on the

growth, survival, and replication potential of virally infected cells. Validation of potential

inhibitors in cell-based assays is the first step in the lengthy process of developing FDA-

approved therapeutics for the treatment of human disease.

111

FIGURES

Figure 5.1. ER-localized vIL-6 and gp130 signaling in PEL cell growth and survival. vIL-6

interacts with both gp130 and VKORC1v2 within the ER. A previous study indicated that

the vIL-6/VKORC1v2 interaction was important for the growth and viability of PEL cells in

culture [173]. The study outlined in Chapter 3 demonstrated that vIL-6 signaling through

gp130 in the ER was critical for the proliferation and survival of PEL cells. Depletion of

gp130 led to reduced cell growth and increased apoptosis in all four PEL cell lines tested

(BCBL-1, JSC-1, BC-1, and BC-3). Furthermore, gp130 was found to contribute to levels

of phosphorylated STATs 1 and 3 (in all PEL cell lines tested) and phosphorylated ERKs

1 and 2 in a subset of PEL cells (BCBL-1 and BC-1). STAT3 had been previously shown

to promote PEL cell growth/viability in culture via the upregulation of survivin [152], but the

role of ERK1/2 in the growth and viability of PEL cells had not been established previously.

Data described in Chapter 3 indicate that ERK1/2 expression is critical for PEL cell

proliferation and survival in all four PEL cell lines tested. The asterisks (*) denote that

STAT3 was required for PEL cell growth and survival in the two cell lines tested (BCBL-1

and JSC-1). To establish that ER-localized vIL-6/gp130 signaling is responsible for vIL-6-

mediated pro-growth and pro-survival functions, two approaches were utilized. First, vIL-

6 depletion-complementation experiments were conducted, and ER-directed and mutated

vIL-6 (W167G mutation, incapable of dimerizing gp130 in the ER though capable of binding

to VKORC1v2) was unable to rescue the growth of vIL-6-depleted PEL cells; ER-directed

and wild-type vIL-6 was able to rescue the vIL-6-depletion phenotype. Secondly,

introduction of an ER-directed soluble gp130 molecule (sgp130.K) into PEL cells inhibited

the vIL-6/gp130 interaction in the ER, leading to decreased PEL cell growth and increased

apoptosis. In summary, ER-localized vIL-6/gp130 signaling promotes cell growth and

survival via the activation of STAT3 and ERK1/2.

112

Figure 5.1

113

Figure 5.2. Role of vIL-6 and gp130 in HHV-8 replication in PEL and endothelial cells. vIL-

6 and gp130 are required for maximal HHV-8 productive replication in both PEL and

endothelial cells. Depletion of vIL-6 or gp130 results in lower HHV-8 titers following

reactivation. Additionally, STAT3, which is activated by vIL-6/gp130 signaling, promotes

HHV-8 replication; depletion of STAT3 leads to decreased viral titers in both PEL and

endothelial cells. To assess the role of ERK signaling in HHV-8 replication, a series of

gp130 mutants were employed as described in Chapter 4. The gp130.ERK.Y mutant was

capable of initiating ERK signaling but not STAT3 signaling, and gp130.STAT.Ys could

activate STAT3 (and STAT1) signaling but not ERK signaling. In a depletion-

complementation experiment, gp130.STAT.Ys, but not gp130.ERK.Y, was capable of

rescuing HHV-8 titers following depletion of endogenous gp130. These results indicate

that vIL-6/gp130-mediated STAT signaling, but not ERK signaling, is necessary for

maximal HHV-8 replication. Finally, a vIL-6 depletion and complementation experiment

was conducted to determine the contribution of ER-localized vIL-6/gp130 signaling to

productive replication. ER-directed wild-type vIL-6 was capable of rescuing HHV-8 titers

following vIL-6 depletion, but ER-directed and mutated vIL-6 (W167G; incapable of

dimerizing gp130 in the ER) was unable to rescue the vIL-6-depletion phenotype. In

summary, ER-localized vIL-6/gp130-mediated STAT3, but not ERK1/2, signaling is

required for maximal HHV-8 replicative titers.

114

Figure 5.2.

115

REFERENCES

1. Friedman-Kien AE (1981) Disseminated Kaposi's sarcoma syndrome in young

homosexual men. J Am Acad Dermatol 5: 468-471.

2. Haverkos HW, Drotman DP (1985) Prevalence of Kaposi's sarcoma among patients

with AIDS. N Engl J Med 312: 1518.

3. Safai B, Johnson KG, Myskowski PL, Koziner B, Yang SY, et al. (1985) The natural

history of Kaposi's sarcoma in the acquired immunodeficiency syndrome. Ann

Intern Med 103: 744-750.

4. Slavin G, Cameron HM, Singh H (1969) Kaposi's sarcoma in mainland Tanzania: a

report of 117 cases. Br J Cancer 23: 349-357.

5. Taylor JF, Smith PG, Bull D, Pike MC (1972) Kaposi's sarcoma in Uganda: geographic

and ethnic distribution. Br J Cancer 26: 483-497.

6. Geddes M, Franceschi S, Balzi D, Arniani S, Gafa L, et al. (1995) Birthplace and classic

Kaposi's sarcoma in Italy. Associazione Italiana Registri Tumori. J Natl Cancer Inst

87: 1015-1017.

7. Franceschi S, Geddes M (1995) Epidemiology of classic Kaposi's sarcoma, with special

reference to mediterranean population. Tumori 81: 308-314.

8. Penn I (1979) Kaposi's sarcoma in organ transplant recipients: report of 20 cases.

Transplantation 27: 8-11.

9. Kaposi (1872) Idiopathisches multiples Pigmentsarkom der Haut. Archiv für

Dermatologie und Syphilis 4: 265-273.

10. DiGiovanna JJ, Safai B (1981) Kaposi's sarcoma. Retrospective study of 90 cases

with particular emphasis on the familial occurrence, ethnic background and

prevalence of other diseases. Am J Med 71: 779-783.

11. Beral V, Peterman TA, Berkelman RL, Jaffe HW (1990) Kaposi's sarcoma among

persons with AIDS: a sexually transmitted infection? The Lancet 335: 123-128.

116

12. Gingues S, Gill MJ (2006) The impact of highly active antiretroviral therapy on the

incidence and outcomes of AIDS-defining cancers in Southern Alberta. HIV Med

7: 369-377.

13. Beral V, Newton R (1998) Overview of the epidemiology of immunodeficiency-

associated cancers. J Natl Cancer Inst Monogr: 1-6.

14. Stein ME, Spencer D, Ruff P, Lakier R, MacPhail P, et al. (1994) Endemic African

Kaposi's sarcoma: clinical and therapeutic implications. 10-year experience in the

Johannesburg Hospital (1980-1990). Oncology 51: 63-69.

15. Andreoni M, Goletti D, Pezzotti P, Pozzetto A, Monini P, et al. (2001) Prevalence,

incidence and correlates of HHV-8/KSHV infection and Kaposi's sarcoma in renal

and liver transplant recipients. J Infect 43: 195-199.

16. Dupin N, Fisher C, Kellam P, Ariad S, Tulliez M, et al. (1999) Distribution of human

herpesvirus-8 latently infected cells in Kaposi's sarcoma, multicentric Castleman's

disease, and primary effusion lymphoma. Proc Natl Acad Sci U S A 96: 4546-4551.

17. Boshoff C, Schulz TF, Kennedy MM, Graham AK, Fisher C, et al. (1995) Kaposi's

sarcoma-associated herpesvirus infects endothelial and spindle cells. Nat Med 1:

1274-1278.

18. Morris VA, Punjabi AS, Lagunoff M (2008) Activation of Akt through gp130 receptor

signaling is required for Kaposi's sarcoma-associated herpesvirus-induced

lymphatic reprogramming of endothelial cells. J Virol 82: 8771-8779.

19. Hong YK, Foreman K, Shin JW, Hirakawa S, Curry CL, et al. (2004) Lymphatic

reprogramming of blood vascular endothelium by Kaposi sarcoma-associated

herpesvirus. Nat Genet 36: 683-685.

20. Hansen A, Henderson S, Lagos D, Nikitenko L, Coulter E, et al. (2010) KSHV-encoded

miRNAs target MAF to induce endothelial cell reprogramming. Genes Dev 24: 195-

205.

117

21. Duprez R, Lacoste V, Briere J, Couppie P, Frances C, et al. (2007) Evidence for a

multiclonal origin of multicentric advanced lesions of Kaposi sarcoma. J Natl

Cancer Inst 99: 1086-1094.

22. Judde JG, Lacoste V, Briere J, Kassa-Kelembho E, Clyti E, et al. (2000) Monoclonality

or oligoclonality of human herpesvirus 8 terminal repeat sequences in Kaposi's

sarcoma and other diseases. J Natl Cancer Inst 92: 729-736.

23. Corbeil J, Evans LA, Vasak E, Cooper DA, Penny R (1991) Culture and properties of

cells derived from Kaposi sarcoma. J Immunol 146: 2972-2976.

24. Ensoli B, Nakamura S, Salahuddin SZ, Biberfeld P, Larsson L, et al. (1989) AIDS-

Kaposi's sarcoma-derived cells express cytokines with autocrine and paracrine

growth effects. Science 243: 223-226.

25. Miles SA, Rezai AR, Salazar-Gonzalez JF, Vander Meyden M, Stevens RH, et al.

(1990) AIDS Kaposi sarcoma-derived cells produce and respond to interleukin 6.

Proc Natl Acad Sci U S A 87: 4068-4072.

26. Roth WK, Werner S, Schirren CG, Hofschneider PH (1989) Depletion of PDGF from

serum inhibits growth of AIDS-related and sporadic Kaposi's sarcoma cells in

culture. Oncogene 4: 483-487.

27. Sciacca FL, Sturzl M, Bussolino F, Sironi M, Brandstetter H, et al. (1994) Expression

of adhesion molecules, platelet-activating factor, and chemokines by Kaposi's

sarcoma cells. J Immunol 153: 4816-4825.

28. Cesarman E, Chang Y, Moore PS, Said JW, Knowles DM (1995) Kaposi's sarcoma-

associated herpesvirus-like DNA sequences in AIDS-related body-cavity-based

lymphomas. N Engl J Med 332: 1186-1191.

29. Nador RG, Cesarman E, Chadburn A, Dawson DB, Ansari MQ, et al. (1996) Primary

effusion lymphoma: a distinct clinicopathologic entity associated with the Kaposi's

sarcoma-associated herpes virus. Blood 88: 645-656.

118

30. Carbone A, Cilia AM, Gloghini A, Capello D, Perin T, et al. (2000) Primary effusion

lymphoma cell lines harbouring human herpesvirus type-8. Leuk Lymphoma 36:

447-456.

31. Carbone A, Gloghini A (2008) KSHV/HHV8-associated lymphomas. Br J Haematol

140: 13-24.

32. Cesarman E, Knowles DM (1999) The role of Kaposi's sarcoma-associated

herpesvirus (KSHV/HHV-8) in lymphoproliferative diseases. Semin Cancer Biol 9:

165-174.

33. Gaidano G, Gloghini A, Gattei V, Rossi MF, Cilia AM, et al. (1997) Association of

Kaposi's sarcoma-associated herpesvirus-positive primary effusion lymphoma

with expression of the CD138/syndecan-1 antigen. Blood 90: 4894-4900.

34. Renne R, Lagunoff M, Zhong W, Ganem D (1996) The size and conformation of

Kaposi's sarcoma-associated herpesvirus (human herpesvirus 8) DNA in infected

cells and virions. J Virol 70: 8151-8154.

35. Renne R, Zhong W, Herndier B, McGrath M, Abbey N, et al. (1996) Lytic growth of

Kaposi's sarcoma-associated herpesvirus (human herpesvirus 8) in culture. Nat

Med 2: 342-346.

36. Carbone A, Gloghini A, Vaccher E, Cerri M, Gaidano G, et al. (2005) Kaposi's

sarcoma-associated herpesvirus/human herpesvirus type 8-positive solid

lymphomas: a tissue-based variant of primary effusion lymphoma. J Mol Diagn 7:

17-27.

37. Boulanger E, Gerard L, Gabarre J, Molina JM, Rapp C, et al. (2005) Prognostic factors

and outcome of human herpesvirus 8-associated primary effusion lymphoma in

patients with AIDS. J Clin Oncol 23: 4372-4380.

119

38. Soulier J, Grollet L, Oksenhendler E, Cacoub P, Cazals-Hatem D, et al. (1995)

Kaposi's sarcoma-associated herpesvirus-like DNA sequences in multicentric

Castleman's disease. Blood 86: 1276-1280.

39. Radaszkiewicz T, Hansmann ML, Lennert K (1989) Monoclonality and polyclonality of

plasma cells in Castleman's disease of the plasma cell variant. Histopathology 14:

11-24.

40. Yoshizaki K, Matsuda T, Nishimoto N, Kuritani T, Taeho L, et al. (1989) Pathogenic

significance of interleukin-6 (IL-6/BSF-2) in Castleman's disease. Blood 74: 1360-

1367.

41. Nishi J, Arimura K, Utsunomiya A, Yonezawa S, Kawakami K, et al. (1999) Expression

of vascular endothelial growth factor in sera and lymph nodes of the plasma cell

type of Castleman's disease. Br J Haematol 104: 482-485.

42. Nishimoto N, Kanakura Y, Aozasa K, Johkoh T, Nakamura M, et al. (2005) Humanized

anti-interleukin-6 receptor antibody treatment of multicentric Castleman disease.

Blood 106: 2627-2632.

43. Oksenhendler E, Carcelain G, Aoki Y, Boulanger E, Maillard A, et al. (2000) High levels

of human herpesvirus 8 viral load, human interleukin-6, interleukin 10, and C

reactive protein correlate with exacerbation of multicentric Castleman disease in

HIV-infected patients. Blood 96: 2069-2073.

44. Uldrick TS, Wang V, O'Mahony D, Aleman K, Wyvill KM, et al. (2010) An interleukin-

6-related systemic inflammatory syndrome in patients co-infected with Kaposi

sarcoma-associated herpesvirus and HIV but without Multicentric Castleman

disease. Clin Infect Dis 51: 350-358.

45. Polizzotto MN, Uldrick TS, Hu D, Yarchoan R (2012) Clinical Manifestations of Kaposi

Sarcoma Herpesvirus Lytic Activation: Multicentric Castleman Disease (KSHV-

MCD) and the KSHV Inflammatory Cytokine Syndrome. Front Microbiol 3: 73.

120

46. Tamburro KM, Yang D, Poisson J, Fedoriw Y, Roy D, et al. (2012) Vironome of Kaposi

sarcoma associated herpesvirus-inflammatory cytokine syndrome in an AIDS

patient reveals co-infection of human herpesvirus 8 and human herpesvirus 6A.

Virology 433: 220-225.

47. McGeoch DJ, Cook S, Dolan A, Jamieson FE, Telford EA (1995) Molecular phylogeny

and evolutionary timescale for the family of mammalian herpesviruses. J Mol Biol

247: 443-458.

48. Pellett P, Roizman B (2007) The Family Herpesviridae: A Brief Introduction. In: Knipe

D, Howley P, editors. Fields Virology, 5th Edition. 5 ed: Lippincott Williams &

Wilkins.

49. Roizman B, Furlong D (1974) The replication of herpesviruses. 229-403.

50. Whitley R (1996) Herpesviruses. In: Baron S, editor. Medical Microbiology. Galveston,

TX: University of Texas Medical Branch at Galveston.

51. Roizman B, Carmichael LE, Deinhardt F, de-The G, Nahmias AJ, et al. (1981)

Herpesviridae. Definition, provisional nomenclature, and taxonomy. The

Herpesvirus Study Group, the International Committee on Taxonomy of Viruses.

Intervirology 16: 201-217.

52. Crough T, Khanna R (2009) Immunobiology of human cytomegalovirus: from bench

to bedside. Clin Microbiol Rev 22: 76-98.

53. Papazian L, Fraisse A, Garbe L, Zandotti C, Thomas P, et al. (1996) Cytomegalovirus.

An unexpected cause of ventilator-associated pneumonia. Anesthesiology 84:

280-287.

54. Halme L, Lempinen M, Arola J, Sarkio S, Hockerstedt K, et al. (2008) High frequency

of gastroduodenal cytomegalovirus infection in liver transplant patients. APMIS

116: 99-106.

121

55. Rafailidis PI, Kapaskelis A, Falagas ME (2007) Cytomegalovirus meningitis in an

immunocompetent patient. Med Sci Monit 13: CS107-109.

56. Campbell PT, Li JS, Wall TC, O'Connor CM, Van Trigt P, et al. (1995) Cytomegalovirus

pericarditis: a case series and review of the literature. Am J Med Sci 309: 229-234.

57. Braun DK, Dominguez G, Pellett PE (1997) Human herpesvirus 6. Clin Microbiol Rev

10: 521-567.

58. Black JB, Pellett PE (1999) Human herpesvirus 7. Rev Med Virol 9: 245-262.

59. Irving WL, Chang J, Raymond DR, Dunstan R, Grattan-Smith P, et al. (1990) Roseola

infantum and other syndromes associated with acute HHV6 infection. Arch Dis

Child 65: 1297-1300.

60. Yamanishi K, Okuno T, Shiraki K, Takahashi M, Kondo T, et al. (1988) Identification

of human herpesvirus-6 as a causal agent for exanthem subitum. Lancet 1: 1065-

1067.

61. Tanaka K, Kondo T, Torigoe S, Okada S, Mukai T, et al. (1994) Human herpesvirus

7: another causal agent for roseola (exanthem subitum). J Pediatr 125: 1-5.

62. Zahorsky J (1910) Roseola infantilis. Pediatrics 22: 60-64.

63. Chang Y, Cesarman E, Pessin MS, Lee F, Culpepper J, et al. (1994) Identification of

herpesvirus-like DNA sequences in AIDS-associated Kaposi's sarcoma. Science

266: 1865-1869.

64. Damania B (2004) Oncogenic gamma-herpesviruses: comparison of viral proteins

involved in tumorigenesis. Nat Rev Microbiol 2: 656-668.

65. Mocarski Jr. E (2007) Comparative analysis of herpesvirus-common proteins. In: Arvin

A, Campadelli-Fiume G, Mocarski E, Moore PS, Roizman B et al., editors. Human

Herpesviruses: Biology, Therapy, and Immunoprophylaxis. Cambridge.

122

66. Moore KW, Vieira P, Fiorentino DF, Trounstine ML, Khan TA, et al. (1990) Homology

of cytokine synthesis inhibitory factor (IL-10) to the Epstein-Barr virus gene BCRFI.

Science 248: 1230-1234.

67. Moore PS, Boshoff C, Weiss RA, Chang Y (1996) Molecular mimicry of human

cytokine and cytokine response pathway genes by KSHV. Science 274: 1739-

1744.

68. Neipel F, Albrecht JC, Ensser A, Huang YQ, Li JJ, et al. (1997) Human herpesvirus 8

encodes a homolog of interleukin-6. J Virol 71: 839-842.

69. Nicholas J, Ruvolo VR, Burns WH, Sandford G, Wan X, et al. (1997) Kaposi's

sarcoma-associated human herpesvirus-8 encodes homologues of macrophage

inflammatory protein-1 and interleukin-6. Nat Med 3: 287-292.

70. Chang Y, Moore PS, Talbot SJ, Boshoff CH, Zarkowska T, et al. (1996) Cyclin encoded

by KS herpesvirus. Nature 382: 410.

71. Li M, Lee H, Yoon DW, Albrecht JC, Fleckenstein B, et al. (1997) Kaposi's sarcoma-

associated herpesvirus encodes a functional cyclin. J Virol 71: 1984-1991.

72. McSharry BP, Avdic S, Slobedman B (2012) Human cytomegalovirus encoded

homologs of cytokines, chemokines and their receptors: roles in

immunomodulation. Viruses 4: 2448-2470.

73. Nicholas J (2010) Human herpesvirus 8-encoded cytokines. Future Virology 5: 197-

206.

74. Oudejans JJ, van den Brule AJ, Jiwa NM, de Bruin PC, Ossenkoppele GJ, et al. (1995)

BHRF1, the Epstein-Barr virus (EBV) homologue of the BCL-2 protooncogene, is

transcribed in EBV-associated B-cell lymphomas and in reactive lymphocytes.

Blood 86: 1893-1902.

123

75. Cleary ML, Smith SD, Sklar J (1986) Cloning and structural analysis of cDNAs for bcl-

2 and a hybrid bcl-2/immunoglobulin transcript resulting from the t(14;18)

translocation. Cell 47: 19-28.

76. Henderson S, Huen D, Rowe M, Dawson C, Johnson G, et al. (1993) Epstein-Barr

virus-coded BHRF1 protein, a viral homologue of Bcl-2, protects human B cells

from programmed cell death. Proc Natl Acad Sci U S A 90: 8479-8483.

77. Nicholas J, Ruvolo V, Zong J, Ciufo D, Guo HG, et al. (1997) A single 13-kilobase

divergent locus in the Kaposi sarcoma-associated herpesvirus (human

herpesvirus 8) genome contains nine open reading frames that are homologous to

or related to cellular proteins. J Virol 71: 1963-1974.

78. Thome M, Schneider P, Hofmann K, Fickenscher H, Meinl E, et al. (1997) Viral FLICE-

inhibitory proteins (FLIPs) prevent apoptosis induced by death receptors. Nature

386: 517-521.

79. Kincaid RP, Sullivan CS (2012) Virus-encoded microRNAs: an overview and a look to

the future. PLoS Pathog 8: e1003018.

80. Cesarman E, Moore PS, Rao PH, Inghirami G, Knowles DM, et al. (1995) In vitro

establishment and characterization of two acquired immunodeficiency syndrome-

related lymphoma cell lines (BC-1 and BC-2) containing Kaposi's sarcoma-

associated herpesvirus-like (KSHV) DNA sequences. Blood 86: 2708-2714.

81. Lagunoff M, Ganem D (1997) The structure and coding organization of the genomic

termini of Kaposi's sarcoma-associated herpesvirus. Virology 236: 147-154.

82. Ganem D (2007) Kaposi's Sarcoma-associated Herpesvirus. In: Knipe D, Howley P,

editors. Fields Virology, 5th Edition: Lippincott Williams & Wilkins. pp. 2848-2931.

83. Pfeffer S, Sewer A, Lagos-Quintana M, Sheridan R, Sander C, et al. (2005)

Identification of microRNAs of the herpesvirus family. Nat Methods 2: 269-276.

124

84. Samols MA, Hu J, Skalsky RL, Renne R (2005) Cloning and identification of a

microRNA cluster within the latency-associated region of Kaposi's sarcoma-

associated herpesvirus. J Virol 79: 9301-9305.

85. McAllister SC, Hansen SG, Messaoudi I, Nikolich-Zugich J, Moses AV (2005)

Increased efficiency of phorbol ester-induced lytic reactivation of Kaposi's

sarcoma-associated herpesvirus during S phase. J Virol 79: 2626-2630.

86. Miller G, Heston L, Grogan E, Gradoville L, Rigsby M, et al. (1997) Selective switch

between latency and lytic replication of Kaposi's sarcoma herpesvirus and Epstein-

Barr virus in dually infected body cavity lymphoma cells. J Virol 71: 314-324.

87. Yu Y, Black JB, Goldsmith CS, Browning PJ, Bhalla K, et al. (1999) Induction of human

herpesvirus-8 DNA replication and transcription by butyrate and TPA in BCBL-1

cells. J Gen Virol 80: 83-90.

88. Chang J, Renne R, Dittmer D, Ganem D (2000) Inflammatory cytokines and the

reactivation of Kaposi's sarcoma-associated herpesvirus lytic replication. Virology

266: 17-25.

89. Gregory SM, West JA, Dillon PJ, Hilscher C, Dittmer DP, et al. (2009) Toll-like receptor

signaling controls reactivation of KSHV from latency. Proc Natl Acad Sci U S A

106: 11725-11730.

90. Ye F, Zhou F, Bedolla RG, Jones T, Lei X, et al. (2011) Reactive oxygen species

hydrogen peroxide mediates Kaposi's sarcoma-associated herpesvirus

reactivation from latency. PLoS Pathog 7: e1002054.

91. Yu F, Feng J, Harada JN, Chanda SK, Kenney SC, et al. (2007) B cell terminal

differentiation factor XBP-1 induces reactivation of Kaposi's sarcoma-associated

herpesvirus. FEBS Lett 581: 3485-3488.

92. Sun R, Lin SF, Staskus K, Gradoville L, Grogan E, et al. (1999) Kinetics of Kaposi's

sarcoma-associated herpesvirus gene expression. J Virol 73: 2232-2242.

125

93. Zhu L, Wang R, Sweat A, Goldstein E, Horvat R, et al. (1999) Comparison of human

sera reactivities in immunoblots with recombinant human herpesvirus (HHV)-8

proteins associated with the latent (ORF73) and lytic (ORFs 65, K8.1A, and K8.1B)

replicative cycles and in immunofluorescence assays with HHV-8-infected BCBL-

1 cells. Virology 256: 381-392.

94. Rimessi P, Bonaccorsi A, Sturzl M, Fabris M, Brocca-Cofano E, et al. (2001)

Transcription pattern of human herpesvirus 8 open reading frame K3 in primary

effusion lymphoma and Kaposi's sarcoma. J Virol 75: 7161-7174.

95. Lu M, Suen J, Frias C, Pfeiffer R, Tsai MH, et al. (2004) Dissection of the Kaposi's

sarcoma-associated herpesvirus gene expression program by using the viral DNA

replication inhibitor cidofovir. J Virol 78: 13637-13652.

96. Lukac DM, Renne R, Kirshner JR, Ganem D (1998) Reactivation of Kaposi's sarcoma-

associated herpesvirus infection from latency by expression of the ORF 50

transactivator, a homolog of the EBV R protein. Virology 252: 304-312.

97. Lukac DM, Kirshner JR, Ganem D (1999) Transcriptional activation by the product of

open reading frame 50 of Kaposi's sarcoma-associated herpesvirus is required for

lytic viral reactivation in B cells. J Virol 73: 9348-9361.

98. Sun R, Lin SF, Gradoville L, Yuan Y, Zhu F, et al. (1998) A viral gene that activates

lytic cycle expression of Kaposi's sarcoma-associated herpesvirus. Proc Natl Acad

Sci U S A 95: 10866-10871.

99. Xu Y, AuCoin DP, Huete AR, Cei SA, Hanson LJ, et al. (2005) A Kaposi's sarcoma-

associated herpesvirus/human herpesvirus 8 ORF50 deletion mutant is defective

for reactivation of latent virus and DNA replication. J Virol 79: 3479-3487.

100. Nakamura H, Lu M, Gwack Y, Souvlis J, Zeichner SL, et al. (2003) Global changes

in Kaposi's sarcoma-associated virus gene expression patterns following

expression of a tetracycline-inducible Rta transactivator. J Virol 77: 4205-4220.

126

101. Guito J, Lukac DM (2012) KSHV Rta promoter specification and viral reactivation.

Front Microbiol 3: 30.

102. Bu W, Carroll KD, Palmeri D, Lukac DM (2007) Kaposi's sarcoma-associated

herpesvirus/human herpesvirus 8 ORF50/Rta lytic switch protein functions as a

tetramer. J Virol 81: 5788-5806.

103. Lin CL, Li H, Wang Y, Zhu FX, Kudchodkar S, et al. (2003) Kaposi's sarcoma-

associated herpesvirus lytic origin (ori-Lyt)-dependent DNA replication:

identification of the ori-Lyt and association of K8 bZip protein with the origin. J Virol

77: 5578-5588.

104. Wu FY, Ahn JH, Alcendor DJ, Jang WJ, Xiao J, et al. (2001) Origin-independent

assembly of Kaposi's sarcoma-associated herpesvirus DNA replication

compartments in transient cotransfection assays and association with the ORF-K8

protein and cellular PML. J Virol 75: 1487-1506.

105. Desai PJ, Pryce EN, Henson BW, Luitweiler EM, Cothran J (2012) Reconstitution of

the Kaposi's sarcoma-associated herpesvirus nuclear egress complex and

formation of nuclear membrane vesicles by coexpression of ORF67 and ORF69

gene products. J Virol 86: 594-598.

106. Luitweiler EM, Henson BW, Pryce EN, Patel V, Coombs G, et al. (2013) Interactions

of the Kaposi's Sarcoma-associated herpesvirus nuclear egress complex: ORF69

is a potent factor for remodeling cellular membranes. J Virol 87: 3915-3929.

107. Krishnan HH, Sharma-Walia N, Zeng L, Gao SJ, Chandran B (2005) Envelope

glycoprotein gB of Kaposi's sarcoma-associated herpesvirus is essential for

egress from infected cells. J Virol 79: 10952-10967.

108. Subramanian R, Sehgal I, D'Auvergne O, Kousoulas KG (2010) Kaposi's sarcoma-

associated herpesvirus glycoproteins B and K8.1 regulate virion egress and

127

synthesis of vascular endothelial growth factor and viral interleukin-6 in BCBL-1

cells. J Virol 84: 1704-1714.

109. Mettenleiter TC (2002) Herpesvirus assembly and egress. J Virol 76: 1537-1547.

110. Staskus KA, Zhong W, Gebhard K, Herndier B, Wang H, et al. (1997) Kaposi's

sarcoma-associated herpesvirus gene expression in endothelial (spindle) tumor

cells. J Virol 71: 715-719.

111. Cannon JS, Ciufo D, Hawkins AL, Griffin CA, Borowitz MJ, et al. (2000) A new primary

effusion lymphoma-derived cell line yields a highly infectious Kaposi's sarcoma

herpesvirus-containing supernatant. J Virol 74: 10187-10193.

112. Dittmer D, Lagunoff M, Renne R, Staskus K, Haase A, et al. (1998) A cluster of

latently expressed genes in Kaposi's sarcoma-associated herpesvirus. J Virol 72:

8309-8315.

113. Pearce M, Matsumura S, Wilson AC (2005) Transcripts encoding K12, v-FLIP, v-

cyclin, and the microRNA cluster of Kaposi's sarcoma-associated herpesvirus

originate from a common promoter. J Virol 79: 14457-14464.

114. Talbot SJ, Weiss RA, Kellam P, Boshoff C (1999) Transcriptional analysis of human

herpesvirus-8 open reading frames 71, 72, 73, K14, and 74 in a primary effusion

lymphoma cell line. Virology 257: 84-94.

115. Chen D, Sandford G, Nicholas J (2009) Intracellular signaling mechanisms and

activities of human herpesvirus 8 interleukin-6. J Virol 83: 722-733.

116. Chandriani S, Ganem D (2010) Array-based transcript profiling and limiting-dilution

reverse transcription-PCR analysis identify additional latent genes in Kaposi's

sarcoma-associated herpesvirus. J Virol 84: 5565-5573.

117. Rivas C, Thlick AE, Parravicini C, Moore PS, Chang Y (2001) Kaposi's sarcoma-

associated herpesvirus LANA2 is a B-cell-specific latent viral protein that inhibits

p53. J Virol 75: 429-438.

128

118. Dittmer DP (2003) Transcription profile of Kaposi's sarcoma-associated herpesvirus

in primary Kaposi's sarcoma lesions as determined by real-time PCR arrays.

Cancer Res 63: 2010-2015.

119. Burysek L, Pitha PM (2001) Latently expressed human herpesvirus 8-encoded

interferon regulatory factor 2 inhibits double-stranded RNA-activated protein

kinase. J Virol 75: 2345-2352.

120. Barbera AJ, Ballestas ME, Kaye KM (2004) The Kaposi's sarcoma-associated

herpesvirus latency-associated nuclear antigen 1 N terminus is essential for

chromosome association, DNA replication, and episome persistence. J Virol 78:

294-301.

121. Cotter MA, 2nd, Robertson ES (1999) The latency-associated nuclear antigen tethers

the Kaposi's sarcoma-associated herpesvirus genome to host chromosomes in

body cavity-based lymphoma cells. Virology 264: 254-264.

122. Barbera AJ, Chodaparambil JV, Kelley-Clarke B, Joukov V, Walter JC, et al. (2006)

The nucleosomal surface as a docking station for Kaposi's sarcoma herpesvirus

LANA. Science 311: 856-861.

123. Ballestas ME, Kaye KM (2001) Kaposi's sarcoma-associated herpesvirus latency-

associated nuclear antigen 1 mediates episome persistence through cis-acting

terminal repeat (TR) sequence and specifically binds TR DNA. J Virol 75: 3250-

3258.

124. Hu J, Garber AC, Renne R (2002) The latency-associated nuclear antigen of Kaposi's

sarcoma-associated herpesvirus supports latent DNA replication in dividing cells.

J Virol 76: 11677-11687.

125. Verma SC, Choudhuri T, Kaul R, Robertson ES (2006) Latency-associated nuclear

antigen (LANA) of Kaposi's sarcoma-associated herpesvirus interacts with origin

129

recognition complexes at the LANA binding sequence within the terminal repeats.

J Virol 80: 2243-2256.

126. Grundhoff A, Ganem D (2003) The latency-associated nuclear antigen of Kaposi's

sarcoma-associated herpesvirus permits replication of terminal repeat-containing

plasmids. J Virol 77: 2779-2783.

127. Lim C, Sohn H, Lee D, Gwack Y, Choe J (2002) Functional dissection of latency-

associated nuclear antigen 1 of Kaposi's sarcoma-associated herpesvirus involved

in latent DNA replication and transcription of terminal repeats of the viral genome.

J Virol 76: 10320-10331.

128. Shamay M, Krithivas A, Zhang J, Hayward SD (2006) Recruitment of the de novo

DNA methyltransferase Dnmt3a by Kaposi's sarcoma-associated herpesvirus

LANA. Proc Natl Acad Sci U S A 103: 14554-14559.

129. An FQ, Compitello N, Horwitz E, Sramkoski M, Knudsen ES, et al. (2005) The

latency-associated nuclear antigen of Kaposi's sarcoma-associated herpesvirus

modulates cellular gene expression and protects lymphoid cells from p16 INK4A-

induced cell cycle arrest. J Biol Chem 280: 3862-3874.

130. Friborg J, Jr., Kong W, Hottiger MO, Nabel GJ (1999) p53 inhibition by the LANA

protein of KSHV protects against cell death. Nature 402: 889-894.

131. Radkov SA, Kellam P, Boshoff C (2000) The latent nuclear antigen of Kaposi

sarcoma-associated herpesvirus targets the retinoblastoma-E2F pathway and with

the oncogene Hras transforms primary rat cells. Nat Med 6: 1121-1127.

132. Jeong JH, Orvis J, Kim JW, McMurtrey CP, Renne R, et al. (2004) Regulation and

autoregulation of the promoter for the latency-associated nuclear antigen of

Kaposi's sarcoma-associated herpesvirus. J Biol Chem 279: 16822-16831.

133. Lan K, Kuppers DA, Verma SC, Robertson ES (2004) Kaposi's sarcoma-associated

herpesvirus-encoded latency-associated nuclear antigen inhibits lytic replication

130

by targeting Rta: a potential mechanism for virus-mediated control of latency. J

Virol 78: 6585-6594.

134. Lan K, Kuppers DA, Robertson ES (2005) Kaposi's sarcoma-associated herpesvirus

reactivation is regulated by interaction of latency-associated nuclear antigen with

recombination signal sequence-binding protein Jkappa, the major downstream

effector of the Notch signaling pathway. J Virol 79: 3468-3478.

135. Fujimuro M, Hayward SD (2003) The latency-associated nuclear antigen of Kaposi's

sarcoma-associated herpesvirus manipulates the activity of glycogen synthase

kinase-3beta. J Virol 77: 8019-8030.

136. Marcu KB, Bossone SA, Patel AJ (1992) myc function and regulation. Annu Rev

Biochem 61: 809-860.

137. Liu J, Martin HJ, Liao G, Hayward SD (2007) The Kaposi's sarcoma-associated

herpesvirus LANA protein stabilizes and activates c-Myc. J Virol 81: 10451-10459.

138. Bubman D, Guasparri I, Cesarman E (2007) Deregulation of c-Myc in primary

effusion lymphoma by Kaposi's sarcoma herpesvirus latency-associated nuclear

antigen. Oncogene 26: 4979-4986.

139. Verma SC, Borah S, Robertson ES (2004) Latency-associated nuclear antigen of

Kaposi's sarcoma-associated herpesvirus up-regulates transcription of human

telomerase reverse transcriptase promoter through interaction with transcription

factor Sp1. J Virol 78: 10348-10359.

140. Fakhari FD, Jeong JH, Kanan Y, Dittmer DP (2006) The latency-associated nuclear

antigen of Kaposi sarcoma-associated herpesvirus induces B cell hyperplasia and

lymphoma. J Clin Invest 116: 735-742.

141. Neipel F, Albrecht JC, Fleckenstein B (1997) Cell-homologous genes in the Kaposi's

sarcoma-associated rhadinovirus human herpesvirus 8: determinants of its

pathogenicity? J Virol 71: 4187-4192.

131

142. Heinrich PC, Castell JV, Andus T (1990) Interleukin-6 and the acute phase response.

Biochem J 265: 621-636.

143. Aoki Y, Tosato G, Fonville TW, Pittaluga S (2001) Serum viral interleukin-6 in AIDS-

related multicentric Castleman disease. Blood 97: 2526-2527.

144. Jones KD, Aoki Y, Chang Y, Moore PS, Yarchoan R, et al. (1999) Involvement of

interleukin-10 (IL-10) and viral IL-6 in the spontaneous growth of Kaposi's sarcoma

herpesvirus-associated infected primary effusion lymphoma cells. Blood 94: 2871-

2879.

145. Zhang YJ, Patel D, Nan Y, Fan S (2011) Inhibition of primary effusion lymphoma

engraftment in SCID mice by morpholino oligomers against early lytic genes of

Kaposi's sarcoma-associated herpesvirus. Antivir Ther 16: 657-666.

146. Aoki Y, Tosato G (1999) Role of vascular endothelial growth factor/vascular

permeability factor in the pathogenesis of Kaposi's sarcoma-associated

herpesvirus-infected primary effusion lymphomas. Blood 94: 4247-4254.

147. Aoki Y, Jaffe ES, Chang Y, Jones K, Teruya-Feldstein J, et al. (1999) Angiogenesis

and hematopoiesis induced by Kaposi's sarcoma-associated herpesvirus-encoded

interleukin-6. Blood 93: 4034-4043.

148. Fielding CA, McLoughlin RM, Colmont CS, Kovaleva M, Harris DA, et al. (2005) Viral

IL-6 blocks neutrophil infiltration during acute inflammation. J Immunol 175: 4024-

4029.

149. Mori Y, Nishimoto N, Ohno M, Inagi R, Dhepakson P, et al. (2000) Human

herpesvirus 8-encoded interleukin-6 homologue (viral IL-6) induces endogenous

human IL-6 secretion. J Med Virol 61: 332-335.

150. Kovaleva M, Bussmeyer I, Rabe B, Grotzinger J, Sudarman E, et al. (2006)

Abrogation of viral interleukin-6 (vIL-6)-induced signaling by intracellular retention

132

and neutralization of vIL-6 with an anti-vIL-6 single-chain antibody selected by

phage display. J Virol 80: 8510-8520.

151. Zhang YJ, Bonaparte RS, Patel D, Stein DA, Iversen PL (2008) Blockade of viral

interleukin-6 expression of Kaposi's sarcoma-associated herpesvirus. Mol Cancer

Ther 7: 712-720.

152. Aoki Y, Feldman GM, Tosato G (2003) Inhibition of STAT3 signaling induces

apoptosis and decreases survivin expression in primary effusion lymphoma. Blood

101: 1535-1542.

153. Nakajima K, Yamanaka Y, Nakae K, Kojima H, Ichiba M, et al. (1996) A central role

for Stat3 in IL-6-induced regulation of growth and differentiation in M1 leukemia

cells. EMBO J 15: 3651-3658.

154. Ambrosini G, Adida C, Altieri DC (1997) A novel anti-apoptosis gene, survivin,

expressed in cancer and lymphoma. Nat Med 3: 917-921.

155. Bartoli M, Gu X, Tsai NT, Venema RC, Brooks SE, et al. (2000) Vascular endothelial

growth factor activates STAT proteins in aortic endothelial cells. J Biol Chem 275:

33189-33192.

156. Heinrich PC, Behrmann I, Muller-Newen G, Schaper F, Graeve L (1998) Interleukin-

6-type cytokine signalling through the gp130/Jak/STAT pathway. Biochem J 334 (

Pt 2): 297-314.

157. Heinrich PC, Behrmann I, Haan S, Hermanns HM, Muller-Newen G, et al. (2003)

Principles of interleukin (IL)-6-type cytokine signalling and its regulation. Biochem

J 374: 1-20.

158. Chow D, He X, Snow AL, Rose-John S, Garcia KC (2001) Structure of an extracellular

gp130 cytokine receptor signaling complex. Science 291: 2150-2155.

133

159. Boulanger MJ, Chow DC, Brevnova E, Martick M, Sandford G, et al. (2004) Molecular

mechanisms for viral mimicry of a human cytokine: activation of gp130 by HHV-8

interleukin-6. J Mol Biol 335: 641-654.

160. Kishimoto T, Akira S, Narazaki M, Taga T (1995) Interleukin-6 family of cytokines

and gp130. Blood 86: 1243-1254.

161. Yamanaka Y, Nakajima K, Fukada T, Hibi M, Hirano T (1996) Differentiation and

growth arrest signals are generated through the cytoplasmic region of gp130 that

is essential for Stat3 activation. EMBO J 15: 1557-1565.

162. Schiemann WP, Bartoe JL, Nathanson NM (1997) Box 3-independent signaling

mechanisms are involved in leukemia inhibitory factor receptor alpha- and gp130-

mediated stimulation of mitogen-activated protein kinase. Evidence for

participation of multiple signaling pathways which converge at Ras. J Biol Chem

272: 16631-16636.

163. Schaper F, Gendo C, Eck M, Schmitz J, Grimm C, et al. (1998) Activation of the

protein tyrosine phosphatase SHP2 via the interleukin-6 signal transducing

receptor protein gp130 requires tyrosine kinase Jak1 and limits acute-phase

protein expression. Biochem J 335: 557-565.

164. Aoki Y, Narazaki M, Kishimoto T, Tosato G (2001) Receptor engagement by viral

interleukin-6 encoded by Kaposi sarcoma-associated herpesvirus. Blood 98: 3042-

3049.

165. Chen D, Nicholas J (2006) Structural requirements for gp80 independence of human

herpesvirus 8 interleukin-6 (vIL-6) and evidence for gp80 stabilization of gp130

signaling complexes induced by vIL-6. J Virol 80: 9811-9821.

166. Wan X, Wang H, Nicholas J (1999) Human herpesvirus 8 interleukin-6 (vIL-6) signals

through gp130 but has structural and receptor-binding properties distinct from

those of human IL-6. J Virol 73: 8268-8278.

134

167. Hu F, Nicholas J (2006) Signal transduction by human herpesvirus 8 viral interleukin-

6 (vIL-6) is modulated by the nonsignaling gp80 subunit of the IL-6 receptor

complex and is distinct from signaling induced by human IL-6. J Virol 80: 10874-

10878.

168. Li H, Wang H, Nicholas J (2001) Detection of direct binding of human herpesvirus 8-

encoded interleukin-6 (vIL-6) to both gp130 and IL-6 receptor (IL-6R) and

identification of amino acid residues of vIL-6 important for IL-6R-dependent and -

independent signaling. J Virol 75: 3325-3334.

169. Cousins E, Nicholas J (2013) Role of human herpesvirus 8 interleukin-6-activated

gp130 signal transducer in primary effusion lymphoma cell growth and viability. J

Virol 87: 10816-10827.

170. Meads MB, Medveczky PG (2004) Kaposi's sarcoma-associated herpesvirus-

encoded viral interleukin-6 is secreted and modified differently than human

interleukin-6: evidence for a unique autocrine signaling mechanism. J Biol Chem

279: 51793-51803.

171. Chen D, Choi YB, Sandford G, Nicholas J (2009) Determinants of secretion and

intracellular localization of human herpesvirus 8 interleukin-6. J Virol 83: 6874-

6882.

172. Dela Cruz CS, Lee Y, Viswanathan SR, El-Guindy AS, Gerlach J, et al. (2004) N-

linked glycosylation is required for optimal function of Kaposi's sarcoma

herpesvirus-encoded, but not cellular, interleukin 6. J Exp Med 199: 503-514.

173. Chen D, Cousins E, Sandford G, Nicholas J (2012) Human herpesvirus 8 viral

interleukin-6 (vIL-6) interacts with splice variant-2 of vitamin K epoxide reductase

complex subunit 1. J Virol 86: 1577-1588.

135

174. Oldenburg J, Bevans CG, Muller CR, Watzka M (2006) Vitamin K epoxide reductase

complex subunit 1 (VKORC1): the key protein of the vitamin K cycle. Antioxid

Redox Signal 8: 347-353.

175. Chen D, Gao Y, Nicholas J (2014) Human herpesvirus 8 interleukin-6 contributes to

primary effusion lymphoma cell viability via suppression of proapoptotic cathepsin

D, a cointeraction partner of vitamin K epoxide reductase complex subunit 1 variant

2. J Virol 88: 1025-1038.

176. Benes P, Vetvicka V, Fusek M (2008) Cathepsin D--many functions of one aspartic

protease. Crit Rev Oncol Hematol 68: 12-28.

177. Prakash O, Tang ZY, Peng X, Coleman R, Gill J, et al. (2002) Tumorigenesis and

aberrant signaling in transgenic mice expressing the human herpesvirus-8 K1

gene. J Natl Cancer Inst 94: 926-935.

178. Lee BS, Lee SH, Feng P, Chang H, Cho NH, et al. (2005) Characterization of the

Kaposi's sarcoma-associated herpesvirus K1 signalosome. J Virol 79: 12173-

12184.

179. Tomlinson CC, Damania B (2004) The K1 protein of Kaposi's sarcoma-associated

herpesvirus activates the Akt signaling pathway. J Virol 78: 1918-1927.

180. Wang L, Dittmer DP, Tomlinson CC, Fakhari FD, Damania B (2006) Immortalization

of primary endothelial cells by the K1 protein of Kaposi's sarcoma-associated

herpesvirus. Cancer Res 66: 3658-3666.

181. Lee BS, Alvarez X, Ishido S, Lackner AA, Jung JU (2000) Inhibition of intracellular

transport of B cell antigen receptor complexes by Kaposi's sarcoma-associated

herpesvirus K1. J Exp Med 192: 11-21.

182. Prakash O, Swamy OR, Peng X, Tang ZY, Li L, et al. (2005) Activation of Src kinase

Lyn by the Kaposi sarcoma-associated herpesvirus K1 protein: implications for

lymphomagenesis. Blood 105: 3987-3994.

136

183. Cousins E, Nicholas J (2014) Molecular biology of human herpesvirus 8: novel

functions and virus-host interactions implicated in viral pathogenesis and

replication. Recent Results Cancer Res 193: 227-268.

184. Coscoy L, Ganem D (2000) Kaposi's sarcoma-associated herpesvirus encodes two

proteins that block cell surface display of MHC class I chains by enhancing their

endocytosis. Proc Natl Acad Sci U S A 97: 8051-8056.

185. Ishido S, Wang C, Lee BS, Cohen GB, Jung JU (2000) Downregulation of major

histocompatibility complex class I molecules by Kaposi's sarcoma-associated

herpesvirus K3 and K5 proteins. J Virol 74: 5300-5309.

186. Ishido S, Choi JK, Lee BS, Wang C, DeMaria M, et al. (2000) Inhibition of natural

killer cell-mediated cytotoxicity by Kaposi's sarcoma-associated herpesvirus K5

protein. Immunity 13: 365-374.

187. Choi YB, Nicholas J (2008) Autocrine and paracrine promotion of cell survival and

virus replication by human herpesvirus 8 chemokines. J Virol 82: 6501-6513.

188. Boshoff C, Endo Y, Collins PD, Takeuchi Y, Reeves JD, et al. (1997) Angiogenic and

HIV-inhibitory functions of KSHV-encoded chemokines. Science 278: 290-294.

189. Nakano K, Isegawa Y, Zou P, Tadagaki K, Inagi R, et al. (2003) Kaposi's sarcoma-

associated herpesvirus (KSHV)-encoded vMIP-I and vMIP-II induce signal

transduction and chemotaxis in monocytic cells. Arch Virol 148: 871-890.

190. Chen S, Bacon KB, Li L, Garcia GE, Xia Y, et al. (1998) In vivo inhibition of CC and

CX3C chemokine-induced leukocyte infiltration and attenuation of

glomerulonephritis in Wistar-Kyoto (WKY) rats by vMIP-II. J Exp Med 188: 193-

198.

191. Kledal TN, Rosenkilde MM, Coulin F, Simmons G, Johnsen AH, et al. (1997) A broad-

spectrum chemokine antagonist encoded by Kaposi's sarcoma-associated

herpesvirus. Science 277: 1656-1659.

137

192. Luttichau HR, Lewis IC, Gerstoft J, Schwartz TW (2001) The herpesvirus 8-encoded

chemokine vMIP-II, but not the poxvirus-encoded chemokine MC148, inhibits the

CCR10 receptor. Eur J Immunol 31: 1217-1220.

193. Luttichau HR, Johnsen AH, Jurlander J, Rosenkilde MM, Schwartz TW (2007) Kaposi

sarcoma-associated herpes virus targets the lymphotactin receptor with both a

broad spectrum antagonist vCCL2 and a highly selective and potent agonist

vCCL3. J Biol Chem 282: 17794-17805.

194. Stine JT, Wood C, Hill M, Epp A, Raport CJ, et al. (2000) KSHV-encoded CC

chemokine vMIP-III is a CCR4 agonist, stimulates angiogenesis, and selectively

chemoattracts TH2 cells. Blood 95: 1151-1157.

195. Haque M, Chen J, Ueda K, Mori Y, Nakano K, et al. (2000) Identification and analysis

of the K5 gene of Kaposi's sarcoma-associated herpesvirus. J Virol 74: 2867-2875.

196. Dairaghi DJ, Fan RA, McMaster BE, Hanley MR, Schall TJ (1999) HHV8-encoded

vMIP-I selectively engages chemokine receptor CCR8. Agonist and antagonist

profiles of viral chemokines. J Biol Chem 274: 21569-21574.

197. Wang HW, Sharp TV, Koumi A, Koentges G, Boshoff C (2002) Characterization of

an anti-apoptotic glycoprotein encoded by Kaposi's sarcoma-associated

herpesvirus which resembles a spliced variant of human survivin. EMBO J 21:

2602-2615.

198. Feng P, Park J, Lee BS, Lee SH, Bram RJ, et al. (2002) 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 76: 11491-11504.

199. Feng P, Scott CW, Cho NH, Nakamura H, Chung YH, et al. (2004) Kaposi's sarcoma-

associated herpesvirus K7 protein targets a ubiquitin-like/ubiquitin-associated

138

domain-containing protein to promote protein degradation. Mol Cell Biol 24: 3938-

3948.

200. Chandran B, Bloomer C, Chan SR, Zhu L, Goldstein E, et al. (1998) Human

herpesvirus-8 ORF K8.1 gene encodes immunogenic glycoproteins generated by

spliced transcripts. Virology 249: 140-149.

201. Wang FZ, Dahl H, Ljungman P, Linde A (1999) Lymphoproliferative responses to

human herpesvirus-6 variant A and variant B in healthy adults. J Med Virol 57: 134-

139.

202. Cunningham C, Barnard S, Blackbourn DJ, Davison AJ (2003) Transcription mapping

of human herpesvirus 8 genes encoding viral interferon regulatory factors. J Gen

Virol 84: 1471-1483.

203. Choi YB, Sandford G, Nicholas J (2012) Human herpesvirus 8 interferon regulatory

factor-mediated BH3-only protein inhibition via Bid BH3-B mimicry. PLoS Pathog

8: e1002748.

204. Burysek L, Yeow WS, Lubyova B, Kellum M, Schafer SL, et al. (1999) Functional

analysis of human herpesvirus 8-encoded viral interferon regulatory factor 1 and

its association with cellular interferon regulatory factors and p300. J Virol 73: 7334-

7342.

205. Li M, Damania B, Alvarez X, Ogryzko V, Ozato K, et al. (2000) Inhibition of p300

histone acetyltransferase by viral interferon regulatory factor. Mol Cell Biol 20:

8254-8263.

206. Lin R, Genin P, Mamane Y, Sgarbanti M, Battistini A, et al. (2001) HHV-8 encoded

vIRF-1 represses the interferon antiviral response by blocking IRF-3 recruitment

of the CBP/p300 coactivators. Oncogene 20: 800-811.

207. Nakamura H, Li M, Zarycki J, Jung JU (2001) Inhibition of p53 tumor suppressor by

viral interferon regulatory factor. J Virol 75: 7572-7582.

139

208. Seo T, Park J, Lee D, Hwang SG, Choe J (2001) Viral interferon regulatory factor 1

of Kaposi's sarcoma-associated herpesvirus binds to p53 and represses p53-

dependent transcription and apoptosis. J Virol 75: 6193-6198.

209. Seo T, Lee D, Shim YS, Angell JE, Chidambaram NV, et al. (2002) Viral interferon

regulatory factor 1 of Kaposi's sarcoma-associated herpesvirus interacts with a cell

death regulator, GRIM19, and inhibits interferon/retinoic acid-induced cell death. J

Virol 76: 8797-8807.

210. Seo T, Park J, Choe J (2005) Kaposi's sarcoma-associated herpesvirus viral IFN

regulatory factor 1 inhibits transforming growth factor-beta signaling. Cancer Res

65: 1738-1747.

211. Shin YC, Nakamura H, Liang X, Feng P, Chang H, et al. (2006) Inhibition of the

ATM/p53 signal transduction pathway by Kaposi's sarcoma-associated

herpesvirus interferon regulatory factor 1. J Virol 80: 2257-2266.

212. Lee HR, Toth Z, Shin YC, Lee JS, Chang H, et al. (2009) Kaposi's sarcoma-

associated herpesvirus viral interferon regulatory factor 4 targets MDM2 to

deregulate the p53 tumor suppressor pathway. J Virol 83: 6739-6747.

213. Lee HR, Choi WC, Lee S, Hwang J, Hwang E, et al. (2011) Bilateral inhibition of

HAUSP deubiquitinase by a viral interferon regulatory factor protein. Nat Struct

Mol Biol 18: 1336-1344.

214. Joo CH, Shin YC, Gack M, Wu L, Levy D, et al. (2007) Inhibition of interferon

regulatory factor 7 (IRF7)-mediated interferon signal transduction by the Kaposi's

sarcoma-associated herpesvirus viral IRF homolog vIRF3. J Virol 81: 8282-8292.

215. Munoz-Fontela C, Marcos-Villar L, Gallego P, Arroyo J, Da Costa M, et al. (2007)

Latent protein LANA2 from Kaposi's sarcoma-associated herpesvirus interacts

with 14-3-3 proteins and inhibits FOXO3a transcription factor. J Virol 81: 1511-

1516.

140

216. Shin YC, Joo CH, Gack MU, Lee HR, Jung JU (2008) Kaposi's sarcoma-associated

herpesvirus viral IFN regulatory factor 3 stabilizes hypoxia-inducible factor-1 alpha

to induce vascular endothelial growth factor expression. Cancer Res 68: 1751-

1759.

217. Wies E, Mori Y, Hahn A, Kremmer E, Sturzl M, et al. (2008) The viral interferon-

regulatory factor-3 is required for the survival of KSHV-infected primary effusion

lymphoma cells. Blood 111: 320-327.

218. Burysek L, Yeow WS, Pitha PM (1999) Unique properties of a second human

herpesvirus 8-encoded interferon regulatory factor (vIRF-2). J Hum Virol 2: 19-32.

219. Sadler R, Wu L, Forghani B, Renne R, Zhong W, et al. (1999) A complex translational

program generates multiple novel proteins from the latently expressed kaposin

(K12) locus of Kaposi's sarcoma-associated herpesvirus. J Virol 73: 5722-5730.

220. Kliche S, Nagel W, Kremmer E, Atzler C, Ege A, et al. (2001) Signaling by human

herpesvirus 8 kaposin A through direct membrane recruitment of cytohesin-1. Mol

Cell 7: 833-843.

221. McCormick C, Ganem D (2005) The kaposin B protein of KSHV activates the

p38/MK2 pathway and stabilizes cytokine mRNAs. Science 307: 739-741.

222. McCormick C, Ganem D (2006) Phosphorylation and function of the kaposin B direct

repeats of Kaposi's sarcoma-associated herpesvirus. J Virol 80: 6165-6170.

223. Yoo J, Kang J, Lee HN, Aguilar B, Kafka D, et al. (2010) Kaposin-B enhances the

PROX1 mRNA stability during lymphatic reprogramming of vascular endothelial

cells by Kaposi's sarcoma herpes virus. PLoS Pathog 6: e1001046.

224. Belanger C, Gravel A, Tomoiu A, Janelle ME, Gosselin J, et al. (2001) Human

herpesvirus 8 viral FLICE-inhibitory protein inhibits Fas-mediated apoptosis

through binding and prevention of procaspase-8 maturation. J Hum Virol 4: 62-73.

141

225. Liu L, Eby MT, Rathore N, Sinha SK, Kumar A, et al. (2002) The human herpes virus

8-encoded viral FLICE inhibitory protein physically associates with and persistently

activates the Ikappa B kinase complex. J Biol Chem 277: 13745-13751.

226. Sun Q, Matta H, Lu G, Chaudhary PM (2006) Induction of IL-8 expression by human

herpesvirus 8 encoded vFLIP K13 via NF-kappaB activation. Oncogene 25: 2717-

2726.

227. Foster-Cuevas M, Wright GJ, Puklavec MJ, Brown MH, Barclay AN (2004) Human

herpesvirus 8 K14 protein mimics CD200 in down-regulating macrophage

activation through CD200 receptor. J Virol 78: 7667-7676.

228. Brinkmann MM, Glenn M, Rainbow L, Kieser A, Henke-Gendo C, et al. (2003)

Activation of mitogen-activated protein kinase and NF-kappaB pathways by a

Kaposi's sarcoma-associated herpesvirus K15 membrane protein. J Virol 77:

9346-9358.

229. Lim CS, Seet BT, Ingham RJ, Gish G, Matskova L, et al. (2007) The K15 protein of

Kaposi's sarcoma-associated herpesvirus recruits the endocytic regulator

intersectin 2 through a selective SH3 domain interaction. Biochemistry 46: 9874-

9885.

230. Pietrek M, Brinkmann MM, Glowacka I, Enlund A, Havemeier A, et al. (2010) Role of

the Kaposi's sarcoma-associated herpesvirus K15 SH3 binding site in

inflammatory signaling and B-cell activation. J Virol 84: 8231-8240.

231. Sharp TV, Wang HW, Koumi A, Hollyman D, Endo Y, et al. (2002) K15 protein of

Kaposi's sarcoma-associated herpesvirus is latently expressed and binds to HAX-

1, a protein with antiapoptotic function. J Virol 76: 802-816.

232. Gottwein E, Cullen BR (2010) A human herpesvirus microRNA inhibits p21

expression and attenuates p21-mediated cell cycle arrest. J Virol 84: 5229-5237.

142

233. Lei X, Bai Z, Ye F, Xie J, Kim CG, et al. (2010) Regulation of NF-kappaB inhibitor

IkappaBalpha and viral replication by a KSHV microRNA. Nat Cell Biol 12: 193-

199.

234. Suffert G, Malterer G, Hausser J, Viiliainen J, Fender A, et al. (2011) Kaposi's

sarcoma herpesvirus microRNAs target caspase 3 and regulate apoptosis. PLoS

Pathog 7: e1002405.

235. Samols MA, Skalsky RL, Maldonado AM, Riva A, Lopez MC, et al. (2007)

Identification of cellular genes targeted by KSHV-encoded microRNAs. PLoS

Pathog 3: e65.

236. Catrina Ene AM, Borze I, Guled M, Costache M, Leen G, et al. (2014) MicroRNA

expression profiles in Kaposi's sarcoma. Pathol Oncol Res 20: 153-159.

237. Lu CC, Li Z, Chu CY, Feng J, Feng J, et al. (2010) MicroRNAs encoded by Kaposi's

sarcoma-associated herpesvirus regulate viral life cycle. EMBO Rep 11: 784-790.

238. Lu F, Stedman W, Yousef M, Renne R, Lieberman PM (2010) Epigenetic regulation

of Kaposi's sarcoma-associated herpesvirus latency by virus-encoded microRNAs

that target Rta and the cellular Rbl2-DNMT pathway. J Virol 84: 2697-2706.

239. Ziegelbauer JM, Sullivan CS, Ganem D (2009) Tandem array-based expression

screens identify host mRNA targets of virus-encoded microRNAs. Nat Genet 41:

130-134.

240. Abend JR, Ramalingam D, Kieffer-Kwon P, Uldrick TS, Yarchoan R, et al. (2012)

Kaposi's sarcoma-associated herpesvirus microRNAs target IRAK1 and MYD88,

two components of the toll-like receptor/interleukin-1R signaling cascade, to

reduce inflammatory-cytokine expression. J Virol 86: 11663-11674.

241. Nachmani D, Stern-Ginossar N, Sarid R, Mandelboim O (2009) Diverse herpesvirus

microRNAs target the stress-induced immune ligand MICB to escape recognition

by natural killer cells. Cell Host Microbe 5: 376-385.

143

242. Lin X, Liang D, He Z, Deng Q, Robertson ES, et al. (2011) miR-K12-7-5p encoded

by Kaposi's sarcoma-associated herpesvirus stabilizes the latent state by targeting

viral ORF50/RTA. PLoS One 6: e16224.

243. Bellare P, Ganem D (2009) Regulation of KSHV lytic switch protein expression by a

virus-encoded microRNA: an evolutionary adaptation that fine-tunes lytic

reactivation. Cell Host Microbe 6: 570-575.

244. Abend JR, Uldrick T, Ziegelbauer JM (2010) Regulation of tumor necrosis factor-like

weak inducer of apoptosis receptor protein (TWEAKR) expression by Kaposi's

sarcoma-associated herpesvirus microRNA prevents TWEAK-induced apoptosis

and inflammatory cytokine expression. J Virol 84: 12139-12151.

245. Liu Y, Sun R, Lin X, Liang D, Deng Q, et al. (2012) Kaposi's sarcoma-associated

herpesvirus-encoded microRNA miR-K12-11 attenuates transforming growth

factor beta signaling through suppression of SMAD5. J Virol 86: 1372-1381.

246. Gottwein E, Mukherjee N, Sachse C, Frenzel C, Majoros WH, et al. (2007) A viral

microRNA functions as an orthologue of cellular miR-155. Nature 450: 1096-1099.

247. Skalsky RL, Samols MA, Plaisance KB, Boss IW, Riva A, et al. (2007) Kaposi's

sarcoma-associated herpesvirus encodes an ortholog of miR-155. J Virol 81:

12836-12845.

248. Qin Z, Freitas E, Sullivan R, Mohan S, Bacelieri R, et al. (2010) Upregulation of xCT

by KSHV-encoded microRNAs facilitates KSHV dissemination and persistence in

an environment of oxidative stress. PLoS Pathog 6: e1000742.

249. Cai X, Lu S, Zhang Z, Gonzalez CM, Damania B, et al. (2005) Kaposi's sarcoma-

associated herpesvirus expresses an array of viral microRNAs in latently infected

cells. Proc Natl Acad Sci U S A 102: 5570-5575.

144

250. Lin YT, Kincaid RP, Arasappan D, Dowd SE, Hunicke-Smith SP, et al. (2010) Small

RNA profiling reveals antisense transcription throughout the KSHV genome and

novel small RNAs. RNA 16: 1540-1558.

251. Alexander L, Denekamp L, Knapp A, Auerbach MR, Damania B, et al. (2000) The

primary sequence of rhesus monkey rhadinovirus isolate 26-95: sequence

similarities to Kaposi's sarcoma-associated herpesvirus and rhesus monkey

rhadinovirus isolate 17577. J Virol 74: 3388-3398.

252. Searles RP, Bergquam EP, Axthelm MK, Wong SW (1999) Sequence and genomic

analysis of a Rhesus macaque rhadinovirus with similarity to Kaposi's sarcoma-

associated herpesvirus/human herpesvirus 8. J Virol 73: 3040-3053.

253. Estep RD, Hansen SG, Rogers KS, Axthelm MK, Wong SW (2013) Genomic

characterization of Japanese macaque rhadinovirus, a novel herpesvirus isolated

from a nonhuman primate with a spontaneous inflammatory demyelinating

disease. J Virol 87: 512-523.

254. Vieira J, O'Hearn PM (2004) Use of the red fluorescent protein as a marker of

Kaposi's sarcoma-associated herpesvirus lytic gene expression. Virology 325:

225-240.

255. Ensoli B, Sturzl M, Monini P (2001) Reactivation and role of HHV-8 in Kaposi's

sarcoma initiation. Adv Cancer Res 81: 161-200.

256. Schulz TF (2006) The pleiotropic effects of Kaposi's sarcoma herpesvirus. J Pathol

208: 187-198.

257. Uldrick TS, Polizzotto MN, Yarchoan R (2012) Recent advances in Kaposi sarcoma

herpesvirus-associated multicentric Castleman disease. Curr Opin Oncol 24: 495-

505.

258. Nicholas J (2007) Human herpesvirus 8-encoded proteins with potential roles in

virus-associated neoplasia. Frontiers in Bioscience 12: 265-281.

145

259. Sakakibara S, Tosato G (2011) Viral interleukin-6: role in Kaposi's sarcoma-

associated herpesvirus: associated malignancies. J Interferon Cytokine Res 31:

791-801.

260. Molden J, Chang Y, You Y, Moore PS, Goldsmith MA (1997) A Kaposi's sarcoma-

associated herpesvirus-encoded cytokine homolog (vIL-6) activates signaling

through the shared gp130 receptor subunit. J Biol Chem 272: 19625-19631.

261. Adam N, Rabe B, Suthaus J, Grotzinger J, Rose-John S, et al. (2009) Unraveling

viral interleukin-6 binding to gp130 and activation of STAT-signaling pathways

independently of the interleukin-6 receptor. J Virol 83: 5117-5126.

262. Verzijl D, Pardo L, van Dijk M, Gruijthuijsen YK, Jongejan A, et al. (2006) Helix 8 of

the viral chemokine receptor ORF74 directs chemokine binding. J Biol Chem 281:

35327-35335.

263. Benekli M, Baumann H, Wetzler M (2009) Targeting signal transducer and activator

of transcription signaling pathway in leukemias. J Clin Oncol 27: 4422-4432.

264. Bowman T, Garcia R, Turkson J, Jove R (2000) STATs in oncogenesis. Oncogene

19: 2474-2488.

265. Hodge DR, Hurt EM, Farrar WL (2005) The role of IL-6 and STAT3 in inflammation

and cancer. Eur J Cancer 41: 2502-2512.

266. Ho HH, Ivashkiv LB (2006) Role of STAT3 in type I interferon responses. Negative

regulation of STAT1-dependent inflammatory gene activation. J Biol Chem 281:

14111-14118.

267. Huang F, Tong X, Fu L, Zhang R (2008) Knockdown of STAT3 by shRNA inhibits the

growth of CAOV3 ovarian cancer cell line in vitro and in vivo. Acta Biochim Biophys

Sin (Shanghai) 40: 519-525.

268. Chatterjee M, Stuhmer T, Herrmann P, Bommert K, Dorken B, et al. (2004) Combined

disruption of both the MEK/ERK and the IL-6R/STAT3 pathways is required to

146

induce apoptosis of multiple myeloma cells in the presence of bone marrow

stromal cells. Blood 104: 3712-3721.

269. Chang LJ, Zaiss AK (2003) Self-inactivating lentiviral vectors and a sensitive Cre-

loxP reporter system. Methods Mol Med 76: 367-382.

270. Zhou BY, Ye Z, Chen G, Gao ZP, Zhang YA, et al. (2007) Inducible and reversible

transgene expression in human stem cells after efficient and stable gene transfer.

Stem Cells 25: 779-789.

271. Uddin S, Hussain AR, Manogaran PS, Al-Hussein K, Platanias LC, et al. (2005)

Curcumin suppresses growth and induces apoptosis in primary effusion

lymphoma. Oncogene 24: 7022-7030.

272. Gururajan M, Chui R, Karuppannan AK, Ke J, Jennings CD, et al. (2005) c-Jun N-

terminal kinase (JNK) is required for survival and proliferation of B-lymphoma cells.

Blood 106: 1382-1391.

273. Cohen A, Brodie C, Sarid R (2006) An essential role of ERK signalling in TPA-induced

reactivation of Kaposi's sarcoma-associated herpesvirus. J Gen Virol 87: 795-802.

274. Ford PW, Bryan BA, Dyson OF, Weidner DA, Chintalgattu V, et al. (2006)

Raf/MEK/ERK signalling triggers reactivation of Kaposi's sarcoma-associated

herpesvirus latency. J Gen Virol 87: 1139-1144.

275. Qin Z, DeFee M, Isaacs JS, Parsons C (2010) Extracellular Hsp90 serves as a co-

factor for MAPK activation and latent viral gene expression during de novo

infection by KSHV. Virology 403: 92-102.

276. Xie J, Ajibade AO, Ye F, Kuhne K, Gao SJ (2008) Reactivation of Kaposi's sarcoma-

associated herpesvirus from latency requires MEK/ERK, JNK and p38 multiple

mitogen-activated protein kinase pathways. Virology 371: 139-154.

147

277. Pan H, Xie J, Ye F, Gao SJ (2006) Modulation of Kaposi's sarcoma-associated

herpesvirus infection and replication by MEK/ERK, JNK, and p38 multiple mitogen-

activated protein kinase pathways during primary infection. J Virol 80: 5371-5382.

278. Sharma-Walia N, Krishnan HH, Naranatt PP, Zeng L, Smith MS, et al. (2005) ERK1/2

and MEK1/2 induced by Kaposi's sarcoma-associated herpesvirus (human

herpesvirus 8) early during infection of target cells are essential for expression of

viral genes and for establishment of infection. J Virol 79: 10308-10329.

279. Haura EB, Turkson J, Jove R (2005) Mechanisms of disease: Insights into the

emerging role of signal transducers and activators of transcription in cancer. Nat

Clin Pract Oncol 2: 315-324.

280. Hirano T, Ishihara K, Hibi M (2000) Roles of STAT3 in mediating the cell growth,

differentiation and survival signals relayed through the IL-6 family of cytokine

receptors. Oncogene 19: 2548-2556.

281. Punjabi AS, Carroll PA, Chen L, Lagunoff M (2007) Persistent activation of STAT3

by latent Kaposi's sarcoma-associated herpesvirus infection of endothelial cells. J

Virol 81: 2449-2458.

282. Ensoli B, Sturzl M (1998) Kaposi's sarcoma: a result of the interplay among

inflammatory cytokines, angiogenic factors and viral agents. Cytokine Growth

Factor Rev 9: 63-83.

283. Chen L, Lagunoff M (2007) The KSHV viral interleukin-6 is not essential for latency

or lytic replication in BJAB cells. Virology 359: 425-435.

284. Krishnan HH, Naranatt PP, Smith MS, Zeng L, Bloomer C, et al. (2004) Concurrent

expression of latent and a limited number of lytic genes with immune modulation

and antiapoptotic function by Kaposi's sarcoma-associated herpesvirus early

during infection of primary endothelial and fibroblast cells and subsequent decline

of lytic gene expression. J Virol 78: 3601-3620.

148

285. Toth Z, Brulois K, Lee HR, Izumiya Y, Tepper C, et al. (2013) Biphasic euchromatin-

to-heterochromatin transition on the KSHV genome following de novo infection.

PLoS Pathog 9: e1003813.

286. Najjar I, Fagard R (2010) STAT1 and pathogens, not a friendly relationship. Biochimie

92: 425-444.

287. Sen N, Che X, Rajamani J, Zerboni L, Sung P, et al. (2012) Signal transducer and

activator of transcription 3 (STAT3) and survivin induction by varicella-zoster virus

promote replication and skin pathogenesis. Proc Natl Acad Sci U S A 109: 600-

605.

288. McCartney EM, Helbig KJ, Narayana SK, Eyre NS, Aloia AL, et al. (2013) Signal

transducer and activator of transcription 3 is a proviral host factor for hepatitis C

virus. Hepatology 58: 1558-1568.

289. Reitsma JM, Sato H, Nevels M, Terhune SS, Paulus C (2013) Human

cytomegalovirus IE1 protein disrupts interleukin-6 signaling by sequestering

STAT3 in the nucleus. J Virol 87: 10763-10776.

290. Trilling M, Le VT, Rashidi-Alavijeh J, Katschinski B, Scheller J, et al. (2014)

"Activated" STAT proteins: a paradoxical consequence of inhibited JAK-STAT

signaling in cytomegalovirus-infected cells. J Immunol 192: 447-458.

291. Wang WB, Levy DE, Lee CK (2011) STAT3 negatively regulates type I IFN-mediated

antiviral response. J Immunol 187: 2578-2585.

292. Meraz MA, White JM, Sheehan KC, Bach EA, Rodig SJ, et al. (1996) Targeted

disruption of the Stat1 gene in mice reveals unexpected physiologic specificity in

the JAK-STAT signaling pathway. Cell 84: 431-442.

293. Park C, Li S, Cha E, Schindler C (2000) Immune response in Stat2 knockout mice.

Immunity 13: 795-804.

149

294. Takeda K, Noguchi K, Shi W, Tanaka T, Matsumoto M, et al. (1997) Targeted

disruption of the mouse Stat3 gene leads to early embryonic lethality. Proc Natl

Acad Sci U S A 94: 3801-3804.

295. Du T, Zhou G, Roizman B (2013) Modulation of reactivation of latent herpes simplex

virus 1 in ganglionic organ cultures by p300/CBP and STAT3. Proc Natl Acad Sci

U S A 110: E2621-2628.

296. Lieu KG, Brice A, Wiltzer L, Hirst B, Jans DA, et al. (2013) The rabies virus interferon

antagonist P protein interacts with activated STAT3 and inhibits Gp130 receptor

signaling. J Virol 87: 8261-8265.

297. Ulane CM, Rodriguez JJ, Parisien JP, Horvath CM (2003) STAT3 ubiquitylation and

degradation by mumps virus suppress cytokine and oncogene signaling. J Virol

77: 6385-6393.

298. Percario Z, Olivetta E, Fiorucci G, Mangino G, Peretti S, et al. (2003) Human

immunodeficiency virus type 1 (HIV-1) Nef activates STAT3 in primary human

monocyte/macrophages through the release of soluble factors: involvement of Nef

domains interacting with the cell endocytotic machinery. J Leukoc Biol 74: 821-

832.

299. Lee H, Herrmann A, Deng JH, Kujawski M, Niu G, et al. (2009) Persistently activated

Stat3 maintains constitutive NF-kappaB activity in tumors. Cancer Cell 15: 283-

293.

300. Yu H, Pardoll D, Jove R (2009) STATs in cancer inflammation and immunity: a

leading role for STAT3. Nat Rev Cancer 9: 798-809.

301. Mantovani A, Allavena P, Sica A, Balkwill F (2008) Cancer-related inflammation.

Nature 454: 436-444.

150

302. Yoshizawa H (2002) Hepatocellular carcinoma associated with hepatitis C virus

infection in Japan: projection to other countries in the foreseeable future. Oncology

62 Suppl 1: 8-17.

303. Boccardo E, Lepique AP, Villa LL (2010) The role of inflammation in HPV

carcinogenesis. Carcinogenesis 31: 1905-1912.

304. Morrison WB (2012) Inflammation and cancer: a comparative view. J Vet Intern Med

26: 18-31.

305. Croll SD, Ransohoff RM, Cai N, Zhang Q, Martin FJ, et al. (2004) VEGF-mediated

inflammation precedes angiogenesis in adult brain. Exp Neurol 187: 388-402.

306. Luther SA, Cyster JG (2001) Chemokines as regulators of T cell differentiation. Nat

Immunol 2: 102-107.

307. Pati S, Cavrois M, Guo HG, Foulke JS, Jr., Kim J, et al. (2001) Activation of NF-

kappaB by the human herpesvirus 8 chemokine receptor ORF74: evidence for a

paracrine model of Kaposi's sarcoma pathogenesis. J Virol 75: 8660-8673.

151

CURRICULUM VITAE

Emily M. Cousins EDUCATION 2014 PhD, Cellular and Molecular Medicine

Johns Hopkins University School of Medicine, Baltimore, MD 2008 B.S. Biochemistry

Calhoun Honors College, Clemson University, Clemson, SC RESEARCH EXPERIENCE 2009–2014 Graduate Student, Cellular and Molecular Medicine

Johns Hopkins University School of Medicine, Baltimore, MD Advisor: John Nicholas, PhD

Research: Viral Oncology. Described the role of endoplasmic reticulum-localized HHV-8 viral interleukin-6 (vIL-6) on cell growth, viability, cell signaling, and virus production. Developed an assay to determine infectious viral titers from cells following depletion or overexpression of viral and/or host proteins. Research findings will facilitate the search for small molecule inhibitors of viral-host interactions within the endoplasmic reticulum, advancing therapeutic options for those suffering from HHV-8 infection.

2008 Summer Intern, Pediatric Oncology Education

St. Jude Children’s Research Hospital, Memphis, TN Advisor: Phil Potter, MD Research: Molecular Pharmacology research. Mutagenized human cholinesterase I at key catalytic residues to evaluate substrate specificity. Research findings will help to elucidate the differences in human cholinesterase specificities for chemotherapeutic pro-drugs, impacting future pro-drug design.

2007–2008 Undergraduate Research

Clemson University, Clemson, SC Advisor: Michael Sehorn, PhD Research: DNA repair research. Designed and cloned bacterial DNA repair protein homologs to determine key protein-protein interaction residues. Research findings from the bacterial system have driven experiments utilizing repair proteins in an effort to characterize homologous recombination repair pathways.

152

TEACHING EXPERIENCE 2011 Pollard Scholar – Genetics JHU School of Medicine 2007-2008 Student Athlete Tutor Clemson University SELECTED HONORS AND AWARDS 2008 General and Departmental Honors Clemson University 2005-2008 Out of State Tuition Scholarship Clemson University 2005-2008 Presidential Scholar Clemson University 2005-2008 Philip Prince Alumni Presid. Scholar Clemson University 2006 President’s List Clemson University 2005 Dean’s List Clemson University 2007, 2008 Dean’s List Clemson University 2006-2008 Dixon Fellow Clemson University 2005-2008 Alpha Lambda Delta Natl Hon Soc Clemson University PUBLICATIONS Peer-reviewed scientific articles: 1. Chen D, Cousins E, Sandford G, and Nicholas J. 2012. Human herpesvirus 8 viral

interleukin-6 interacts with splice variant 2 of vitamin K epoxide reductase complex subunit 1. J. Virol. 2012; 86(3):1577-1588

2. Cousins E and Nicholas J. 2013. Role of human herpesvirus 8 interleukin-6-activated gp130 signal transducer in primary effusion lymphoma cell growth and viability. J. Virol. 2013; 87(19):10816-10827

3. Cousins E, Gao Y, Sandford G, and Nicholas J. 2014. Human herpesvirus 8 viral

interleukin-6 signaling through gp130 promotes virus replication in primary effusion lymphoma and endothelial cells. J. Virol. 2014; 88(20):12167-12172

Book chapter 1. Cousins E and Nicholas J. 2014. Molecular biology of human herpesvirus 8: Novel

functions and virus-host interactions implicated in viral pathogenesis and replication. Recent Results Cancer Res. 2014; 193:227-68. doi:10.1007/978-3-642-38965-8_13

Abstracts 1. Cousins E and Nicholas J. 2013. Role of human herpesvirus 8 interleukin-6-

activated gp130 signal transducer in primary effusion lymphoma cell growth and viability. Abstract #31. 16th International Workshop on Kaposi’s Sarcoma Herpesvirus (KSHV) and Related Agents. Oral presentation. Puerto Vallarta, Mexico. July 2013

153

SERVICE AND LEADERSHIP 2011–2014 Contract editor American Journal Exp., Durham, NC 2010–2014 Mentor Incentive Mentoring Program, Baltimore, MD PROFESSIONAL SOCIETY MEMBERSHIPS 2011– American Association for the Advancement of Science (AAAS) FUNDING 2012–2014 Ruth Kirschstein Fellowship F31-CA171933, NIH NCI 2008 Calhoun Honors College Research Grant, Clemson University


Recommended