Human herpesvirus 8 (HHV-8)/Kaposi sarcoma–associated herpesvirus (KSHV) is linked to a number of malignancies thought to be driven by cytokines, including interleukin-6 (IL-6). Rta, a transcriptional activator encoded by HHV-8/KSHV, activates the viral lytic cycle leading to the expression of several viral genes implicated in viral pathogenesis. However, the effect of HHV-8/KSHV Rta on cellular genes has not been reported. We present evidence that the human IL-6(hIL-6) gene is up-regulated by Rta. Rta potently activated (up to 164-fold) the hIL-6 promoter in a dose-dependent manner in a transient transfection reporter system. Rta also induced expression of the endogenous hIL-6 gene, as shown by enzyme-linked immunosorbent assays. Activation of the hIL-6gene by HHV-8/KSHV supports the role of hIL-6 in the development of these malignancies.

Human interleukin-6 (hIL-6) is a multifunctional cytokine, and dysregulation of hIL-6 is implicated in the pathogenesis of several malignancies such as Kaposi sarcoma (KS), primary effusion lymphoma (PEL), and multicentric Castleman disease (MCD). hIL-6 serves as an autocrine growth factor for cultured AIDS-KS cells and may induce endothelial cell proliferation in KS through a paracrine pathway.1,2 Supernatants from PEL-derived cell lines and PEL effusions contain large quantities of hIL-6.3,4Anti–hIL-6 neutralizing antibodies delayed PEL tumor progression in SCID mice.5 Overproduction of IL-6 also reproduced some manifestations of MCD in a mouse model.6 Furthermore, anti–hIL-6 or anti–hIL-6 receptor antibodies exerted a therapeutic effect on MCD patients.7,8 Taken together, these data strongly support the involvement of hIL-6 in the pathogenesis of these malignancies.

Another common feature of KS, PEL, and MCD is their association with human herpesvirus 8 (HHV-8)/Kaposi sarcoma–associated herpesvirus (KSHV).9-11 HHV-8/KSHV encodes a potent transcriptional activator, Rta, which is necessary and sufficient for initiating viral lytic replication.12,13 Among the lytic genes expressed are homologues of cytokines and chemokines, including viral IL-6 (vIL-6) and viral macrophage inflammatory proteins.14,15In particular, vIL-6 has been detected in tumor lesions and sera from KS, PEL, and MCD patients and is thought to play an important role in viral pathogenesis.16-18 In addition to pirating cellular genes, it is likely that HHV-8/KSHV has developed strategies to enhance its replication by modulating the regulation of cellular factors. We are investigating the effect of Rta on cellular genes and report here that hIL-6 expression is up-regulated by Rta.

Plasmid construction

The 1.2-kb hIL-6 promoter region was amplified from total cellular DNA using primers F (5′-GGAAGATCTCTCCTGCAAGAGACACCATCCTGA-3′) and R (5′-CGGGAATTCAGGGCAGAATGAGCCTCAGAGACAT3-3′); the underlined nucleotides represent BglII and EcoRI sites, respectively. The PCR fragment was cloned into pSEAP2-basic (Clontech, Palo Alto, CA) to produce phIL6-1200/SEAP.

Reporter assays

Transfections were performed in 12-well plates using a standard calcium phosphate method for the human embryonic kidney cell line 293T or LipofectAmine PLUS (Invitrogen, Carlsbad, CA) for the immortalized bone marrow stromal cell line R1T.19 At 48 hours after transfection, supernatants and cells were harvested. Supernatants were assayed for secreted alkaline phosphatase (SEAP) activities, using the Great EscAPe SEAP Chemiluminescence Detection Kit (Clontech). Cells were lysed in 1× passive lysis buffer and assayed forRenilla luciferase activities using the Luciferase Reporter Assay System (Promega, Madison, WI).

Enzyme-linked immunosorbent assays

pcDNA3/Rta12 or pcDNA3 was transfected into 293T or R1T cells in 6-well plates using LipofectAmine PLUS. pcDNA3/Rta contained a 3.1-kb genomic sequence encoding Rta, whose expression was driven by the cytomegalovirus immediate-early promoter/enhancer in the vector. Supernatants from transfected cells were collected at 24, 48, and 72 hours after transfection and were assayed for hIL-6 protein levels using an hIL-6 enzyme-linked immunosorbent assay (ELISA) kit (Biosource International).

To investigate the role Rta may play in regulatinghIL-6 gene expression, we first examined whether Rta can activate the hIL-6 promoter in a reporter system. A 1200-bp promoter region upstream of the first hIL-6 exon was cloned into the pSEAP2-basic vector to produce phIL6-1200/SEAP. This reporter plasmid was cotransfected into 293T cells with either pcDNA3/Rta (an Rta expression plasmid) or vector alone. To control for transfection efficiency and other experimental variations, pRL-CMV, which constitutively expresses the Renilla luciferase, was included in each transfection. As shown in Figure1A, phIL6-1200/SEAP was potently activated (164-fold) by Rta. To confirm that activation of the hIL-6 promoter was mediated by the Rta protein, we examined the dose dependence of Rta activation. A fixed amount of the reporter plasmid phIL6-1200/SEAP was cotransfected with increasing amounts of pcDNA3/Rta into 293T cells. As the amount of pcDNA3/Rta in each transfection increased, so did the normalized SEAP activity (Figure 1B), indicating that activation of the hIL-6 promoter by Rta is specific.

Fig. 1.

HHV-8/KSHV Rta activates the hIL-6 promoter in a reporter system.

(A) Activation of the hIL-6 promoter by Rta in 2 different cell lines. Reporter plasmid phIL6-1200/SEAP (20 ng), pRL-CMV (2 ng), filler DNA (720 ng; plasmid DNA that lacks a mammalian promoter/enhancer), and either pcDNA3/Rta or pcDNA3 (50 ng) were transfected into 293T or R1T cells. Supernatants and cells were harvested at 48 hours after transfection and were assayed for SEAP and Renillaluciferase activities, respectively. SEAP activities from the hIL6 promoter were normalized to the corresponding Renillaluciferase activities. Fold activation by Rta was calculated by comparing the normalized SEAP activity stimulated by Rta to that by pcDNA3. (B) Dose-dependent activation of the hIL-6 promoter by Rta in 293T cells. Cells were transfected with 20 ng phIL6-1200/SEAP, 2 ng pRL-CMV, 720 ng filler DNA, and an increasing amount of pcDNA3/Rta (0-50 ng) and a correspondingly decreasing amount of pcDNA3 (50-0 ng) so that the total amount of pcDNA3 vector backbone remained the same. Reporter activities were assayed at 48 hours after transfection; fold activation by different amounts of pcDNA3/Rta was calculated by comparing the normalized SEAP activities to that stimulated by 0 ng pcDNA3/Rta and 50 ng pcDNA3.

Fig. 1.

HHV-8/KSHV Rta activates the hIL-6 promoter in a reporter system.

(A) Activation of the hIL-6 promoter by Rta in 2 different cell lines. Reporter plasmid phIL6-1200/SEAP (20 ng), pRL-CMV (2 ng), filler DNA (720 ng; plasmid DNA that lacks a mammalian promoter/enhancer), and either pcDNA3/Rta or pcDNA3 (50 ng) were transfected into 293T or R1T cells. Supernatants and cells were harvested at 48 hours after transfection and were assayed for SEAP and Renillaluciferase activities, respectively. SEAP activities from the hIL6 promoter were normalized to the corresponding Renillaluciferase activities. Fold activation by Rta was calculated by comparing the normalized SEAP activity stimulated by Rta to that by pcDNA3. (B) Dose-dependent activation of the hIL-6 promoter by Rta in 293T cells. Cells were transfected with 20 ng phIL6-1200/SEAP, 2 ng pRL-CMV, 720 ng filler DNA, and an increasing amount of pcDNA3/Rta (0-50 ng) and a correspondingly decreasing amount of pcDNA3 (50-0 ng) so that the total amount of pcDNA3 vector backbone remained the same. Reporter activities were assayed at 48 hours after transfection; fold activation by different amounts of pcDNA3/Rta was calculated by comparing the normalized SEAP activities to that stimulated by 0 ng pcDNA3/Rta and 50 ng pcDNA3.

Close modal

These results from the reporter system indicate that Rta activates the hIL-6 promoter in the absence of chromatin structure. We next examined whether Rta also activates the endogenous hIL-6 gene. pcDNA3/Rta or pcDNA3 was transfected into 293T cells, and supernatants were harvested at different time points after transfection. The hIL-6 protein levels in these samples were then assayed by ELISA. Consistent with the lack of endogenous hIL-6 expression in 293T cells, the hIL-6 protein levels were low (less than 7.8 pg/mL, the detection limit of the kit) in pcDNA3-transfected cells (Figure2A). However, the expression of Rta in 293T cells stimulated hIL-6 expression and resulted in progressively higher amounts of hIL-6 protein accumulating in the supernatant at 48 and 72 hours after transfection (54.0 and 84.5 pg/mL, respectively).

Fig. 2.

HHV-8/KSHV Rta activates the endogenous

hIL-6 gene. pcDNA3/Rta or pcDNA3 (1 μg) was transfected into 293T (A) or R1T (B) cells. Supernatants were harvested 24, 48, and 72 hours later, diluted where appropriate, and assayed for hIL-6 protein levels by ELISA. Average hIL-6 protein concentrations in the supernatants are indicated by horizontal bars, with the numbers (pg/mL) shown. The detection range of the ELISA kit is 7.8 to 500 pg/mL. Dotted line in panel A indicates the lowest detection limit of the kit. When the hIL-6 level in one or more experiments was lower than 7.8 pg/mL, the average for that time point was not calculated. A logarithmic scale is used in panel B.

Fig. 2.

HHV-8/KSHV Rta activates the endogenous

hIL-6 gene. pcDNA3/Rta or pcDNA3 (1 μg) was transfected into 293T (A) or R1T (B) cells. Supernatants were harvested 24, 48, and 72 hours later, diluted where appropriate, and assayed for hIL-6 protein levels by ELISA. Average hIL-6 protein concentrations in the supernatants are indicated by horizontal bars, with the numbers (pg/mL) shown. The detection range of the ELISA kit is 7.8 to 500 pg/mL. Dotted line in panel A indicates the lowest detection limit of the kit. When the hIL-6 level in one or more experiments was lower than 7.8 pg/mL, the average for that time point was not calculated. A logarithmic scale is used in panel B.

Close modal

To further establish the ability of Rta to activate the hIL-6 promoter and to induce hIL-6 protein expression, we performed similar experiments in R1T cells. R1T cells manifest a significant level of basal hIL-6 expression19 and thus complement the use of 293T cells. The reporter plasmid phIL6-1200/SEAP was activated 27-fold by Rta in transient transfection reporter assays in R1T cells (Figure1A). The fold activation in R1T cells was lower than that in 293T cells because of the higher basal level of the reporter plasmid. Moreover, transfection of pcDNA3/Rta stimulated the expression of endogenous hIL-6 in R1T cells, when compared to transfection of pcDNA3, and resulted in hIL-6 levels of 1209, 5762, and 21 447 pg/mL at 24, 48, and 72 hours after transfection, respectively (Figure 2B).

Up-regulation of the hIL-6 gene has emerged as a common theme among herpesvirus infections, and multiple mechanisms may be involved.20-22 In the case of HHV-8/KSHV, latently infected B-cell lines (eg, BC-1 and KS-1) express hIL-6 at high levels.3,4 This is attributed in part to the responsiveness of the hIL-6 promoter to an HHV-8/KSHV-latent gene product, the latency-associated nuclear antigen.19 Because HHV-8/KSHV exists predominantly in a latent state in KS and PEL lesions, the induction of hIL-6 expression by the latency-associated nuclear antigen may play a critical role in the development of these malignancies. Here we have demonstrated that HHV-8/KSHV also stimulates hIL-6 expression through its lytic transcriptional activator, Rta. We hypothesize that activation of hIL-6 by Rta plays an important role in lytic infections. This is especially relevant in patients with HHV-8/KSHV-associated MCD. Our results are consistent with the high plasma hIL-6 levels observed in MCD patients and with the fact that most HHV-8/KSHV–infected cells in MCD lesions express the viral lytic gene expression program driven by Rta.16,17 

Interestingly, in a separate study, we demonstrated that Rta also strongly activates the HHV-8/KSHV vIL-6 gene.23Like hIL-6, vIL-6 promotes the growth of IL-6–dependent B cells and activates signal transduction pathways. However, vIL-6 may stimulate a broader spectrum of target cells because it requires only the ubiquitously expressed gp130 receptor, whereas hIL-6 requires both gp130 and IL-6Rα for signal transduction.14,15,24 On the other hand, the amount of vIL-6 required to stimulate the growth of IL-6–dependent B cells was greater than that of hIL-6, and the binding affinity of vIL-6 for soluble gp130 was determined to be 1000-fold lower than that of hIL-6/soluble IL-6R complex for gp130.25 Therefore, hIL-6 and vIL-6 may both be important in HHV-8/KSHV replication and pathogenesis, but they may play overlapping yet different roles.

We thank Mike Johnson and Jiabin An for excellent technical assistance, Dr Tonia Symensma for critical reading of the manuscript, and members of the Sun and Martinez-Maza laboratories for discussion.

Prepublished online as Blood First Edition Paper, May 13, 2002; DOI 10.1182/blood-2002-01-0015.

Supported by National Institutes of Health grants CA91791, CA83525, DE14153, and CA57152, the Jonsson Cancer Center Foundation, the Stop Cancer Foundation, and the Concern Foundation. H.D. is a Lymphoma Research Foundation Fellow.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 U.S.C. section 1734.

1
Miles
SA
Rezai
AR
Salazar-Gonzalez
JF
et al
AIDS Kaposi sarcoma-derived cells produce and respond to interleukin 6.
Proc Natl Acad Sci U S A.
87
1990
4068
4072
2
Corbeil
J
Evans
LA
Vasak
E
Cooper
DA
Penny
R
Culture and properties of cells derived from Kaposi sarcoma.
J Immunol.
146
1991
2972
2976
3
Asou
H
Said
JW
Yang
R
et al
Mechanisms of growth control of Kaposi's sarcoma-associated herpes virus-associated primary effusion lymphoma cells.
Blood.
91
1998
2475
2481
4
Aoki
Y
Yarchoan
R
Braun
J
Iwamoto
A
Tosato
G
Viral and cellular cytokines in AIDS-related malignant lymphomatous effusions.
Blood.
96
2000
1599
1601
5
Foussat
A
Wijdenes
J
Bouchet
L
et al
Human interleukin-6 is in vivo an autocrine growth factor for human herpesvirus-8-infected malignant B lymphocytes.
Eur Cytokine Netw.
10
1999
501
508
6
Brandt
SJ
Bodine
DM
Dunbar
CE
Nienhuis
AW
Dysregulated interleukin 6 expression produces a syndrome resembling Castleman's disease in mice.
J Clin Invest.
86
1990
592
599
7
Beck
JT
Hsu
SM
Wijdenes
J
et al
Brief report: alleviation of systemic manifestations of Castleman's disease by monoclonal anti-interleukin-6 antibody.
N Engl J Med.
330
1994
602
605
8
Nishimoto
N
Sasai
M
Shima
Y
et al
Improvement in Castleman's disease by humanized anti-interleukin-6 receptor antibody therapy.
Blood.
95
2000
56
61
9
Cesarman
E
Chang
Y
Moore
PS
Said
JW
Knowles
DM
Kaposi's sarcoma-associated herpesvirus-like DNA sequences in AIDS-related body-cavity-based lymphomas.
N Engl J Med.
332
1995
1186
1191
10
Chang
Y
Cesarman
E
Pessin
MS
et al
Identification of herpesvirus-like DNA sequences in AIDS-associated Kaposi's sarcoma.
Science.
266
1994
1865
1869
11
Soulier
J
Grollet
L
Oksenhendler
E
et al
Kaposi's sarcoma-associated herpesvirus-like DNA sequences in multicentric Castleman's disease.
Blood.
86
1995
1276
1280
12
Sun
R
Lin
SF
Gradoville
L
Yuan
Y
Zhu
F
Miller
G
A viral gene that activates lytic cycle expression of Kaposi's sarcoma-associated herpesvirus.
Proc Natl Acad Sci U S A.
95
1998
10866
10871
13
Lukac
DM
Kirshner
JR
Ganem
D
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
1999
9348
9361
14
Moore
PS
Boshoff
C
Weiss
RA
Chang
Y
Molecular mimicry of human cytokine and cytokine response pathway genes by KSHV.
Science.
274
1996
1739
1744
15
Nicholas
J
Ruvolo
VR
Burns
WH
et al
Kaposi's sarcoma-associated human herpesvirus-8 encodes homologues of macrophage inflammatory protein-1 and interleukin-6.
Nat Med.
3
1997
287
292
16
Staskus
KA
Sun
R
Miller
G
et al
Cellular tropism and viral interleukin-6 expression distinguish human herpesvirus 8 involvement in Kaposi's sarcoma, primary effusion lymphoma, and multicentric Castleman's disease.
J Virol.
73
1999
4181
4187
17
Parravicini
C
Chandran
B
Corbellino
M
et al
Differential viral protein expression in Kaposi's sarcoma-associated herpesvirus-infected diseases: Kaposi's sarcoma, primary effusion lymphoma, and multicentric Castleman's disease.
Am J Pathol.
156
2000
743
749
18
Aoki
Y
Yarchoan
R
Wyvill
K
Okamoto
S
Little
RF
Tosato
G
Detection of viral interleukin-6 in Kaposi sarcoma-associated herpesvirus-linked disorders.
Blood.
97
2001
2173
2176
19
An
J
Lichtenstein
AK
Brent
G
Rettig
MB
Kaposi's sarcoma-associated herpesvirus (KSHV) induces cellular interleukin-6 expression: role of the KSHV latency-associated nuclear antigen and AP-1 response element.
Blood.
99
2002
649
654
20
D'Addario
M
Libermann
TA
Xu
J
Ahmad
A
Menezes
J
Epstein-Barr virus and its glycoprotein-350 up-regulate IL-6 in human B-lymphocytes via CD21, involving activation of NF-κB and different signaling pathways.
J Mol Biol.
308
2001
501
514
21
Eliopoulos
AG
Stack
M
Dawson
CW
et al
Epstein-Barr virus-encoded LMP1 and CD40 mediate IL-6 production in epithelial cells via an NF-κB pathway involving TNF receptor-associated factors.
Oncogene.
14
1997
2899
2916
22
Carlquist
JF
Edelman
L
Bennion
DW
Anderson
JL
Cytomegalovirus induction of interleukin-6 in lung fibroblasts occurs independently of active infection and involves a G protein and the transcription factor, NF-κB.
J Infect Dis.
179
1999
1094
1100
23
Deng
H
Song
MJ
Chu
JT
Sun
R
Transcriptional regulation of the interleukin-6 gene of human herpesvirus 8/Kaposi's sarcoma-associated herpesvirus.
J Virol.
76
2002
8252
8264
24
Molden
J
Chang
Y
You
Y
Moore
PS
Goldsmith
MA
A Kaposi's sarcoma-associated herpesvirus-encoded cytokine homolog (vIL-6) activates signaling through the shared gp130 receptor subunit.
J Biol Chem.
272
1997
19625
19631
25
Aoki
Y
Narazaki
M
Kishimoto
T
Tosato
G
Receptor engagement by viral interleukin-6 encoded by Kaposi sarcoma-associated herpesvirus.
Blood.
98
2001
3042
3049

Author notes

Ren Sun, Department of Molecular and Medical Pharmacology, University of California at Los Angeles, Los Angeles, CA 90095-1735; e-mail: rsun@mednet.ucla.edu.

Sign in via your Institution