Di QIN and Chun LU. The Biology of Kaposi's Sarcoma-Associated Herpesvirus and the Infection of Human Immunodeficiency Virus[J]. Virologica Sinica, 2008, 23(6): 473-485. doi: 10.1007/s12250-008-2996-x
Citation: Di QIN, Chun LU. The Biology of Kaposi's Sarcoma-Associated Herpesvirus and the Infection of Human Immunodeficiency Virus .VIROLOGICA SINICA, 2008, 23(6) : 473-485.  http://dx.doi.org/10.1007/s12250-008-2996-x

The Biology of Kaposi’s Sarcoma-Associated Herpesvirus and the Infection of Human Immunodeficiency Virus*

  • 通讯作者: Chun LU, clu@njmu.edu.cn
  • 收稿日期: 2008-07-30
    录用日期: 2008-09-17
  • Kaposi sarcoma-associated herpesvirus (KSHV), also known as human herpesvirus 8 (HHV-8), is discovered in 1994 from Kaposi’s sarcoma (KS) lesion of an acquired immunodeficiency syndrome (AIDS) patient. In addition to its association with KS, KSHV has also been implicated as the causative agent of two other AIDS-associated malignancies: primary effusion lymphoma (PEL) and multicentric Castleman’s disease (MCD). KSHV is a complex DNA virus that not only has the ability to promote cellular growth and survival for tumor development, but also can provoke deregulated angiogenesis, inflammation, and modulate the patient’s immune system in favor of tumor growth. As KSHV is a necessary but not sufficient etiological factor for KS, human immunodeficiency virus (HIV) is a very important cofactor. Here we review the basic information about the biology of KSHV, development of pathogenesis and interaction between KSHV and HIV.

The Biology of Kaposi's Sarcoma-Associated Herpesvirus and the Infection of Human Immunodeficiency Virus

  • Corresponding author: Chun LU, clu@njmu.edu.cn
  • Received Date: 30 July 2008
    Accepted Date: 17 September 2008

    Fund Project: Fok Ying Tung Education Foundation 101038Program for New Century Excellent Talents in Chinese Universities NCET-05-0506

  • Kaposi sarcoma-associated herpesvirus (KSHV), also known as human herpesvirus 8 (HHV-8), is discovered in 1994 from Kaposi’s sarcoma (KS) lesion of an acquired immunodeficiency syndrome (AIDS) patient. In addition to its association with KS, KSHV has also been implicated as the causative agent of two other AIDS-associated malignancies: primary effusion lymphoma (PEL) and multicentric Castleman’s disease (MCD). KSHV is a complex DNA virus that not only has the ability to promote cellular growth and survival for tumor development, but also can provoke deregulated angiogenesis, inflammation, and modulate the patient’s immune system in favor of tumor growth. As KSHV is a necessary but not sufficient etiological factor for KS, human immunodeficiency virus (HIV) is a very important cofactor. Here we review the basic information about the biology of KSHV, development of pathogenesis and interaction between KSHV and HIV.

  • 加载中
    1. Ablashi D, Chatlynne L, Cooper H, et al. 1999. Seroprevalence of human herpesvirus-8 (HHV-8) in countries of Southeast Asia compared to the USA, the Caribbean and Africa. Br J Cancer, 81: 893-897.
        doi: 10.1038/sj.bjc.6690782

    2. Ablashi D V, Chatlynne L G, Whitman J E Jr, et al. 2002. Spectrum of Kaposi's sarcoma-associated herpes-virus, or human herpesvirus 8, diseases. Clin Microbiol Rev, 15: 439-464.
        doi: 10.1128/CMR.15.3.439-464.2002

    3. Albini A, Paglieri I, Orengo G, et al. 1997. The beta-core fragment of human chorionic gonadotrophin inhibits growth of Kaposi's sarcoma-derived cells and a new immortalized Kaposi's sarcoma cell line. AIDS, 11: 713-721.
        doi: 10.1097/00002030-199706000-00003

    4. Aoki Y, Tosato G. 2003. Targeted inhibition of angio-genic factors in AIDS-related disorders. Curr Drug Targets Infect Disord, 3: 115-128.
        doi: 10.2174/1568005033481222

    5. Aoki Y, Tosato G. 2004. HIV-1 Tat enhances Kaposi sar-coma-associated herpesvirus (KSHV) infectivity. Blood, 104: 810-814.
        doi: 10.1182/blood-2003-07-2533

    6. Atkinson J, Edlin B R, Engels E A, et al. 2003. Seroprevalence of human herpesvirus 8 among injection drug users in San Francisco. J Infect Dis, 187: 974-981.
        doi: 10.1086/jid.2003.187.issue-6

    7. Bechtel J T, Liang Y, Hvidding J, et al. 2003. Host range of Kaposi's sarcoma-associated herpesvirus in cultured cells. J Virol, 77: 6474-6481.
        doi: 10.1128/JVI.77.11.6474-6481.2003

    8. Beral V. 1991. Epidemiology of Kaposi's sarcoma. Cancer Surv, 10: 5-22.

    9. Beral V, Peterman T A, Berkelman R L, et al. 1990. Kaposi's sarcoma among persons with AIDS: a sexually transmitted infection? Lancet, 335: 123-128.
        doi: 10.1016/0140-6736(90)90001-L

    10. Blackbourn D J, Lennette E T, Ambroziak J, et al. 1998. Human herpesvirus 8 detection in nasal secretions and saliva. J Infect Dis, 177: 213-216.
        doi: 10.1086/jid.1998.177.issue-1

    11. Blackbourn D J, Osmond D, Levy J A, et al. 1999. Increased human herpesvirus 8 seroprevalence in young homosexual men who have multiple sex contacts with different partners. J Infect Dis, 179: 237-239.
        doi: 10.1086/jid.1999.179.issue-1

    12. Borkovic S P, Schwartz R A. 1981. Kaposi's sarcoma presenting in the homosexual man --a new and striking phenomenon! Ariz Med, 38: 902-904.

    13. Burysek L, Yeow W S, Lubyova B, 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.

    14. Cannon M J, Dollard S C, Black J B, et al. 2003. Risk factors for Kaposi's sarcoma in men seropositive for both human herpesvirus 8 and human immunodeficiency virus. AIDS, 17: 215-222.
        doi: 10.1097/00002030-200301240-00012

    15. Carbone A, Gaidano G. 1997. HHV-8-positive body-cavity-based lymphoma: a novel lymphoma entity. Br J Haematol, 97: 515-522.
        doi: 10.1046/j.1365-2141.1997.00064.x

    16. Caselli E, Menegazzi P, Bracci A, et al. 2001. Human herpesvirus-8 (Kaposi's sarcoma-associated herpesvirus) ORF50 interacts synergistically with the tat gene product in transactivating the human immunodeficiency virus type 1 LTR. J Gen Virol, 82: 1965-1970.
        doi: 10.1099/0022-1317-82-8-1965

    17. Caselli E, Galvan M, Santoni F, et al. 2003. Human herpesvirus-8 (Kaposi's sarcoma-associated virus) ORF50 increases in vitro cell susceptibility to human immunodeficiency virus type 1 infection. J Gen Virol, 84: 1123-1131.
        doi: 10.1099/vir.0.18799-0

    18. Caselli E, Galvan M, Cassai E, et al. 2003. Transient expression of human herpesvirus-8 (Kaposi's sarcoma-associated herpesvirus) ORF50 enhances HIV-1 replication. Intervirology, 46: 141-149.
        doi: 10.1159/000071454

    19. Caselli E, Galvan M, Cassai E, et al. 2005. Human herpesvirus 8 enhances human immunodeficiency virus replication in acutely infected cells and induces reacti-vation in latently infected cells. Blood, 106: 2790-2797.
        doi: 10.1182/blood-2005-04-1390

    20. Cathomas G, McGandy C E, Terracciano L M, et al. 1996. Detection of herpesvirus-like DNA by nested PCR on archival skin biopsy specimens of various forms of Kaposi sarcoma. J Clin Pathol, 49: 631-633.
        doi: 10.1136/jcp.49.8.631

    21. Cesarman E, Chang Y, Moore P S, et al. 1995. Kaposi's sarcoma-associated herpesvirus-like DNA sequences in AIDS-related body-cavity-based lymphomas. N Engl J Med, 332: 1186-1191.
        doi: 10.1056/NEJM199505043321802

    22. Cesarman E. 2002. The role of Kaposi's sarcoma-associated herpesvirus (KSHV/HHV-8) in lymphoproli-ferative diseases. Recent Results Cancer Res, 159: 27-37.
        doi: 10.1007/978-3-642-56352-2

    23. Chang Y, Cesarman E, Pessin M S, et al. 1994. Identification of herpesvirus-like DNA sequences in AIDS-associated Kaposi's sarcoma. Science, 266: 1865-1869.
        doi: 10.1126/science.7997879

    24. Chatlynne L G, Ablashi D V. 1999. Seroepidemiology of Kaposi's sarcoma-associated herpesvirus (KSHV). Semin Cancer Biol, 9: 175-185.
        doi: 10.1006/scbi.1998.0089

    25. Chatlynne L G, Lapps W, Handy M, et al. 1998. Detection and titration of human herpesvirus-8-specific antibodies in sera from blood donors, acquired immuno-deficiency syndrome patients, and Kaposi's sarcoma patients using a whole virus enzyme-linked immunosor-bent assay. Blood, 92: 53-58.

    26. Cheng E H, Nicholas J, Bellows D S, et al. 1997. A Bcl-2 homolog encoded by Kaposi sarcoma-associated virus, human herpesvirus 8, inhibits apoptosis but does not heterodimerize with Bax or Bak. Proc Natl Acad Sci U S A, 94: 690-694.
        doi: 10.1073/pnas.94.2.690

    27. Child E S, Mann D J. 2001. Novel properties of the cyclin encoded by Human Herpesvirus 8 that facilitate exit from quiescence. Oncogene, 20: 3311-3322.
        doi: 10.1038/sj.onc.1204447

    28. Couty J P, Geras-Raaka E, Weksler B B, et al. 2001. Kaposi's sarcoma-associated herpesvirus G protein-coupled receptor signals through multiple pathways in endothelial cells. J Biol Chem, 276: 33805-33811.
        doi: 10.1074/jbc.M104631200

    29. de-The G, Bestetti G, van Beveren M, et al. 1999. Prevalence of human herpesvirus 8 infection before the acquired immunodeficiency disease syndrome-related epidemic of Kaposi's sarcoma in East Africa. J Natl Cancer Inst, 91: 1888-1889.
        doi: 10.1093/jnci/91.21.1888

    30. Dedicoat M, Newton R, Alkharsah K R, et al. 2004. Mother-to-child transmission of human herpesvirus-8 in South Africa. J Infect Dis, 190: 1068-1075.
        doi: 10.1086/jid.2004.190.issue-6

    31. Duus K M, Lentchitsky V, Wagenaar T, et al. 2004. Wild-type Kaposi's sarcoma-associated herpesvirus isolated from the oropharynx of immune-competent individuals has tropism for cultured oral epithelial cells. J Virol, 78: 4074-4084.
        doi: 10.1128/JVI.78.8.4074-4084.2004

    32. Enbom M, Sheldon J, Lennette E, et al. 2000. Antibodies to human herpesvirus 8 latent and lytic antigens in blood donors and potential high-risk groups in Sweden: variable frequencies found in a multicenter serological study. J Med Virol, 62: 498-504.
        doi: 10.1002/(ISSN)1096-9071

    33. Engels E A, Sinclair M D, Biggar R J, et al. 2000. Latent class analysis of human herpesvirus 8 assay performance and infection prevalence in sub-saharan Africa and Malta. Int J Cancer, 88: 1003-1008.
        doi: 10.1002/(ISSN)1097-0215

    34. Ensoli B, Sgadari C, Barillari G, et al. 2001. Biology of Kaposi's sarcoma. Eur J Cancer, 37: 1251-1269.
        doi: 10.1016/S0959-8049(01)00121-6

    35. Ensoli B, Sturzl M, Monini P. 2001. Reactivation and role of HHV-8 in Kaposi's sarcoma initiation. Adv Cancer Res, 81: 161-200.
        doi: 10.1016/S0065-230X(01)81005-8

    36. Field N, Low W, Daniels M, et al. 2003. KSHV vFLIP binds to IKK-gamma to activate IKK. J Cell Sci, 116: 3721-3728.
        doi: 10.1242/jcs.00691

    37. Friborg J Jr, Kong W, Hottiger M O, et al. 1999. p53 inhibition by the LANA protein of KSHV protects against cell death. Nature, 402: 889-894.

    38. Friedman-Kien A E. 1981. Disseminated Kaposi's sarcoma syndrome in young homosexual men. J Am Acad Dermatol, 5: 468-471.
        doi: 10.1016/S0190-9622(81)80010-2

    39. Gambus G, Bourboulia D, Esteve A, et al. 2001. Prevalence and distribution of HHV-8 in different subpopu-lations, with and without HIV infection, in Spain. AIDS, 15: 1167-1174.
        doi: 10.1097/00002030-200106150-00012

    40. Gao S J, Boshoff C, Jayachandra S, et al. 1997. KSHV ORF K9 (vIRF) is an oncogene which inhibits the interferon signaling pathway. Oncogene, 15: 1979-1985.
        doi: 10.1038/sj.onc.1201571

    41. Gao S J, Kingsley L, Li M, et al. 1996. KSHV antibodies among Americans, Italians and Ugandans with and without Kaposi's sarcoma. Nat Med, 2: 925-928.
        doi: 10.1038/nm0896-925

    42. Gessain A, Mauclere P, van Beveren M, et al. 1999. Human herpesvirus 8 primary infection occurs during childhood in Cameroon, Central Africa. Int J Cancer, 81: 189-192.
        doi: 10.1002/(ISSN)1097-0215

    43. Giraldo G, Beth E, Haguenau F. 1972. Herpes-type virus particles in tissue culture of Kaposi's sarcoma from different geographic regions. J Natl Cancer Inst, 49: 1509-1526.
        doi: 10.1093/jnci/49.6.1509

    44. Goedert J J, Cote T R, Virgo P, et al. 1998. Spectrum of AIDS-associated malignant disorders. Lancet, 351: 1833-1839.
        doi: 10.1016/S0140-6736(97)09028-4

    45. Gottlieb G J, Ragaz A, Vogel J V, et al. 1981. A preliminary communication on extensively disseminated Kaposi's sarcoma in young homosexual men. Am J Dermatopathol, 3: 111-114.
        doi: 10.1097/00000372-198100320-00002

    46. Grossman Z, Iscovich J, Schwartz F, et al. 2002. Absence of Kaposi sarcoma among Ethiopian immigrants to Israel despite high seroprevalence of human herpesvirus 8. Mayo Clin Proc, 77: 905-909.
        doi: 10.1016/S0025-6196(11)62256-8

    47. Guasparri I, Keller S A, Cesarman E. 2004. KSHV vFLIP is essential for the survival of infected lymphoma cells. J Exp Med, 199: 993-1003.
        doi: 10.1084/jem.20031467

    48. Guo H G, Pati S, Sadowska M, et al. 2004. Tumorigenesis by human herpesvirus 8 vGPCR is accele-rated by human immunodeficiency virus type 1 Tat. J Virol, 78: 9336-9342.
        doi: 10.1128/JVI.78.17.9336-9342.2004

    49. Gwack Y, Hwang S, Byun H, et al. 2001. Kaposi's sarcoma-associated herpesvirus open reading frame 50 represses p53-induced transcriptional activity and apoptosis. J Virol, 75: 6245-6248.
        doi: 10.1128/JVI.75.13.6245-6248.2001

    50. Hengge U R, Ruzicka T, Tyring S K, et al. 2002. Update on Kaposi's sarcoma and other HHV8 associated diseases. Part 2: pathogenesis, Castleman's disease, and pleural effusion lymphoma. Lancet Infect Dis, 2: 344-352.
        doi: 10.1016/S1473-3099(02)00288-8

    51. Hideshima T, Chauhan D, Teoh G, et al. 2000. Characterization of signaling cascades triggered by human interleukin-6 versus Kaposi's sarcoma-associated herpes virus-encoded viral interleukin 6. Clin Cancer Res, 6: 1180-1189.

    52. Huang L M, Chao M F, Chen M Y, et al. 2001. Reciprocal regulatory interaction between human herpes-virus 8 and human immunodeficiency virus type 1. J Biol Chem, 276: 13427-13432.
        doi: 10.1074/jbc.M011314200

    53. Hymes K B, Cheung T, Greene J B, et al. 1981. Kaposi's sarcoma in homosexual men-a report of eight cases. Lancet, 2: 598-600.

    54. Hyun T S, Subramanian C, Cotter M A 2nd, et al. 2001. Latency-associated nuclear antigen encoded by Kaposi's sarcoma-associated herpesvirus interacts with Tat and activates the long terminal repeat of human immunodefi-ciency virus type 1 in human cells. J Virol, 75: 8761-8771.
        doi: 10.1128/JVI.75.18.8761-8771.2001

    55. Iscovich J, Boffetta P, Franceschi S, et al. 2000. Classic kaposi sarcoma: epidemiology and risk factors. Cancer, 88: 500-517.
        doi: 10.1002/(ISSN)1097-0142

    56. Jacobson L P, Jenkins F J, Springer G, et al. 2000. Interaction of human immunodeficiency virus type 1 and human herpesvirus type 8 infections on the incidence of Kaposi's sarcoma. J Infect Dis, 181: 1940-1949.
        doi: 10.1086/jid.2000.181.issue-6

    57. Juhasz A, Remenyik E, Konya J, et al. 2001. Prevalence and age distribution of human herpesvirus-8 specific antibodies in Hungarian blood donors. J Med Virol, 64: 526-530.
        doi: 10.1002/jmv.1081

    58. Juhasz I, Albelda S M, Elder D E, et al. 1993. Growth and invasion of human melanomas in human skin grafted to immunodeficient mice. Am J Pathol, 143: 528-537.

    59. Kaaya S F, Leshabari M T, Mbwambo J K. 1998. Risk behaviors and vulnerability to HIV infection among Tanzanian youth. J Health Popul Dev Ctries, 1: 51-60.

    60. Kaposi M. 1872. Idiopathisches multiples Pigmentsarkom der Haut.Arch Dermatol Syph, 4: 265-273.
        doi: 10.1007/BF01830024

    61. Kirchhoff S, Sebens T, Baumann S, et al. 2002. Viral IFN-regulatory factors inhibit activation-induced cell death via two positive regulatory IFN-regulatory factor 1-dependent domains in the CD95 ligand promoter. J Immunol, 168: 1226-1234.
        doi: 10.4049/jimmunol.168.3.1226

    62. Lang M E, Lottersberger C, Roth B, et al. 1997. Induction of apoptosis in Kaposi's sarcoma spindle cell cultures by the subunits of human chorionic gonadotropin. AIDS, 11: 1333-1340.

    63. Lennette E T, Blackbourn D J, Levy J A. 1996. Antibodies to human herpesvirus type 8 in the general population and in Kaposi's sarcoma patients. Lancet, 348: 858-861.
        doi: 10.1016/S0140-6736(96)03240-0

    64. Li M, Lee H, Yoon D W, et al. 1997. Kaposi's sarcoma-associated herpesvirus encodes a functional cyclin. J Virol, 71: 1984-1991.

    65. Liu L, Eby M T, Rathore N, 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.
        doi: 10.1074/jbc.M110480200

    66. Lubyova B, Pitha P M. 2000. Characterization of a novel human herpesvirus 8-encoded protein, vIRF-3, that shows homology to viral and cellular interferon regulatory factors. J Virol, 74: 8194-8201.
        doi: 10.1128/JVI.74.17.8194-8201.2000

    67. Martin J N, Ganem D E, Osmond D H, et al. 1998. Sexual transmission and the natural history of human herpesvirus 8 infection. N Engl J Med, 338: 948-954.
        doi: 10.1056/NEJM199804023381403

    68. Martin R W 3rd, Hood A F, Farmer E R. 1993. Kaposi sarcoma. Medicine (Baltimore), 72: 245-261.
        doi: 10.1097/00005792-199307000-00004

    69. Matta H, Chaudhary P M. 2004. Activation of alternative NF-kappa B pathway by human herpes virus 8-encoded Fas-associated death domain-like IL-1 beta-converting enzyme inhibitory protein (vFLIP). Proc Natl Acad Sci U S A, 101: 9399-9404.
        doi: 10.1073/pnas.0308016101

    70. Mayama S, Cuevas L E, Sheldon J, et al. 1998. Prevalence and transmission of Kaposi's sarcoma-associated herpesvirus (human herpesvirus 8) in Ugandan children and adolescents. Int J Cancer, 77: 817-820.
        doi: 10.1002/(ISSN)1097-0215

    71. Mbulaiteye S M, Biggar R J, Goedert J J, et al. 2003. Immune deficiency and risk for malignancy among persons with AIDS. J Acquir Immune Defic Syndr, 32: 527-533.
        doi: 10.1097/00126334-200304150-00010

    72. Mbulaiteye S M, Pfeiffer R M, Whitby D, et al. 2003. Human herpesvirus 8 infection within families in rural Tanzania. J Infect Dis, 187: 1780-1785.
        doi: 10.1086/jid.2003.187.issue-11

    73. Mendez J C, Procop G W, Espy M J, et al. 1999. Relationship of HHV8 replication and Kaposi's sarcoma after solid organ transplantation. Transplantation, 67: 1200-1201.
        doi: 10.1097/00007890-199904270-00022

    74. Merat R, Amara A, Lebbe C, et al. 2002. HIV-1 infection of primary effusion lymphoma cell line triggers Kaposi's sarcoma-associated herpesvirus (KSHV) reacti-vation. Int J Cancer, 97: 791-795.
        doi: 10.1002/(ISSN)1097-0215

    75. Mercader M, Taddeo B, Panella J R, et al. 2000. Induction of HHV-8 lytic cycle replication by inflam-matory cytokines produced by HIV-1-infected T cells. Am J Pathol, 156: 1961-1971.
        doi: 10.1016/S0002-9440(10)65069-9

    76. Mercader M, Nickoloff B J, Foreman K E. 2001. Induction of human immunodeficiency virus 1 replication by human herpesvirus 8. Arch Pathol Lab Med, 125: 785-789.

    77. Molden J, Chang Y, You Y, et al. 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.
        doi: 10.1074/jbc.272.31.19625

    78. Montaner S, Sodhi A, Pece S, et al. 2001. The Kaposi's sarcoma-associated herpesvirus G protein-coupled receptor promotes endothelial cell survival through the activation of Akt/protein kinase B. Cancer Res, 61: 2641-2648.

    79. Moore P S, Chang Y. 2001. Molecular virology of Kaposi's sarcoma-associated herpesvirus. Philos Trans R Soc Lond B Biol Sci, 356: 499-516.
        doi: 10.1098/rstb.2000.0777

    80. Moore P S, Kingsley L A, Holmberg S D, et al. 1996. Kaposi's sarcoma-associated herpesvirus infection prior to onset of Kaposi's sarcoma. AIDS, 10: 175-180.
        doi: 10.1097/00002030-199602000-00007

    81. Moore P S, Gao S J, Dominguez G, et al. 1996. Primary characterization of a herpesvirus agent associated with Kaposi's sarcomae. J Virol, 70: 549-558.

    82. Munshi N, Ganju R K, Avraham S, et al. 1999. Kaposi's sarcoma-associated herpesvirus-encoded G protein-coupled receptor activation of c-jun amino-terminal kinase/stress-activated protein kinase and lyn kinase is mediated by related adhesion focal tyrosine kinase/proline-rich tyrosine kinase 2. J Biol Chem, 274: 31863-31867.
        doi: 10.1074/jbc.274.45.31863

    83. Nador R G, Cesarman E, Chadburn A, et al. 1996. Primary effusion lymphoma: a distinct clinicopathologic entity associated with the Kaposi's sarcoma-associated herpes virus. Blood, 88: 645-656.

    84. Nakamura H, Li M, Zarycki J, et al. 2001. Inhibition of p53 tumor suppressor by viral interferon regulatory factor. J Virol, 75: 7572-7582.
        doi: 10.1128/JVI.75.16.7572-7582.2001

    85. Neipel F, Albrecht J C, 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.

    86. O'Brien T R, Kedes D, Ganem D, et al. 1999. Evidence for concurrent epidemics of human herpesvirus 8 and human immunodeficiency virus type 1 in US homosexual men: rates, risk factors, and relationship to Kaposi's sarcoma. J Infect Dis, 180: 1010-1017.
        doi: 10.1086/jid.1999.180.issue-4

    87. Oettle A G. 1962. Geographical and racial differences in the frequency of Kaposi's sarcoma as evidence of environ-mental or genetic causes. Acta Unio Int Contra Cancrum, 18: 330-363.

    88. Olsen S J, Chang Y, Moore P S, et al. 1998. Increasing Kaposi's sarcoma-associated herpesvirus seroprevalence with age in a highly Kaposi's sarcoma endemic region, Zambia in 1985. AIDS, 12: 1921-1925.
        doi: 10.1097/00002030-199814000-00024

    89. Osborne J, Moore P S, Chang Y. 1999. KSHV-encoded viral IL-6 activates multiple human IL-6 signaling path-ways. Hum Immunol, 60: 921-927.
        doi: 10.1016/S0198-8859(99)00083-X

    90. Pati S, Cavrois M, Guo H G, 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.
        doi: 10.1128/JVI.75.18.8660-8673.2001

    91. Pauk J, Huang M L, Brodie S J, et al. 2000. Mucosal shedding of human herpesvirus 8 in men. N Engl J Med, 343: 1369-1377.
        doi: 10.1056/NEJM200011093431904

    92. Plancoulaine S, Abel L, van Beveren M, et al. 2000. Human herpesvirus 8 transmission from mother to child and between siblings in an endemic population. Lancet, 356: 1062-1065.
        doi: 10.1016/S0140-6736(00)02729-X

    93. Plancoulaine S, Abel L, Tregouet D, et al. 2004. Respective roles of serological status and blood specific antihuman herpesvirus 8 antibody levels in human herpesvirus 8 intrafamilial transmission in a highly endemic area. Cancer Res, 64: 8782-8787.
        doi: 10.1158/0008-5472.CAN-04-2000

    94. Preiser W, Szep N I, Lang D, et al. 2001. Kaposi's sarcoma-associated herpesvirus seroprevalence in selected german patients: evaluation by different test systems. Med Microbiol Immunol, 190: 121-127.

    95. Rabkin C S, Janz S, Lash A, et al. 1997. Monoclonal origin of multicentric Kaposi's sarcoma lesions. N Engl J Med, 336: 988-993.
        doi: 10.1056/NEJM199704033361403

    96. Radkov S A, 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.
        doi: 10.1038/80459

    97. Regamey N, Cathomas G, Schwager M, et al. 1998. High human herpesvirus 8 seroprevalence in the homose-xual population in Switzerland. J Clin Microbiol, 36: 1784-1786.

    98. Renne R, Blackbourn D, Whitby D, et al. 1998. Limited transmission of Kaposi's sarcoma-associated herpesvirus in cultured cells. J Virol, 72: 5182-5188.

    99. Renne R, Lagunoff M, Zhong W, et al. 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.

    100. Rivas C, Thlick A E, Parravicini C, et al. 2001. Kaposi's sarcoma-associated herpesvirus LANA2 is a B-cell-specific latent viral protein that inhibits p53. J Virol, 75: 429-438.
        doi: 10.1128/JVI.75.1.429-438.2001

    101. Russo J J, Bohenzky R A, Chien M C, et al. 1996. Nucleotide sequence of the Kaposi sarcoma-associated herpesvirus (HHV8). Proc Natl Acad Sci U S A, 93: 14862-14867.
        doi: 10.1073/pnas.93.25.14862

    102. Samaniego F, Young D, Grimes C, et al. 2002. Vascular endothelial growth factor and Kaposi's sarcoma cells in human skin grafts. Cell Growth Differ, 13: 387-395.

    103. Sarek G, Jarviluoma A, Ojala P M. 2006. KSHV viral cyclin inactivates p27KIP1 through Ser10 and Thr187 phosphorylation in proliferating primary effusion lym-phomas. Blood, 107: 725-732.
        doi: 10.1182/blood-2005-06-2534

    104. Sarid R, Flore O, Bohenzky R A, et al. 1998. Transcri-ption mapping of the Kaposi's sarcoma-associated herpes-virus (human herpesvirus 8) genome in a body cavity-based lymphoma cell line (BC-1). J Virol, 72: 1005-1012.

    105. Sarid R, Sato T, Bohenzky R A, et al. 1997. Kaposi's sarcoma-associated herpesvirus encodes a functional bcl-2 homologue. Nat Med, 3: 293-298.
        doi: 10.1038/nm0397-293

    106. Schulz T F. 2000. KSHV (HHV8) infection. J Infect, 41: 125-129.
        doi: 10.1053/jinf.2000.0712

    107. Schwarz M, Murphy P M. 2001. Kaposi's sarcoma-associated herpesvirus G protein-coupled receptor cons-titutively activates NF-kappa B and induces proinflam-matory cytokine and chemokine production via a C-terminal signaling determinant. J Immunol, 167: 505-513.
        doi: 10.4049/jimmunol.167.1.505

    108. Seo T, Lee D, Shim Y S, et al. 2002. Viral interferon regulatory factor 1 of Kaposi's sarcoma-associated herpes-virus interacts with a cell death regulator, GRIM19, and inhibits interferon/retinoic acid-induced cell death. J Virol, 76: 8797-8807.
        doi: 10.1128/JVI.76.17.8797-8807.2002

    109. Seo T, Park J, Lee D, et al. 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.
        doi: 10.1128/JVI.75.13.6193-6198.2001

    110. Shin Y C, Nakamura H, Liang X, 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.
        doi: 10.1128/JVI.80.5.2257-2266.2006

    111. Siegal F P, Lopez C, Hammer G S, et al. 1981. Severe acquired immunodeficiency in male homosexuals, mani-fested by chronic perianal ulcerative herpes simplex lesions. N Engl J Med, 305: 1439-1444.
        doi: 10.1056/NEJM198112103052403

    112. Sitas F, Carrara H, Beral V, et al. 1999. Antibodies against human herpesvirus 8 in black South African patients with cancer. N Engl J Med, 340: 1863-1871.
        doi: 10.1056/NEJM199906173402403

    113. Smit M J, Verzijl D, Casarosa P, et al. 2002. Kaposi's sarcoma-associated herpesvirus-encoded G protein-cou-pled receptor ORF74 constitutively activates p44/p42 MAPK and Akt via G(i) and phospholipase C-dependent signaling pathways. J Virol, 76: 1744-1752.
        doi: 10.1128/JVI.76.4.1744-1752.2002

    114. Soulier J, Grollet L, Oksenhendler E, et al. 1995. Kaposi's sarcoma-associated herpesvirus-like DNA se-quences in multicentric Castleman's disease. Blood, 86: 1276-1280.

    115. Sternbach G, Varon J. 1995. Moritz Kaposi: idiopathic pigmented sarcoma of the skin. J Emerg Med, 13: 671-674.
        doi: 10.1016/0736-4679(95)00077-N

    116. Sturzl M, Hohenadl C, Zietz C, et al. 1999. Expression of K13/v-FLIP gene of human herpesvirus 8 and apo-ptosis in Kaposi's sarcoma spindle cells. J Natl Cancer Inst, 91: 1725-1733.
        doi: 10.1093/jnci/91.20.1725

    117. Sun Q, Matta H, Chaudhary P M. 2005. Kaposi's sarcoma associated herpes virus-encoded viral FLICE inhibitory protein activates transcription from HIV-1 Long Terminal Repeat via the classical NF-kappaB path-way and functionally cooperates with Tat. Retrovirology, 2: 9.
        doi: 10.1186/1742-4690-2-9

    118. 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.

    119. West J T, Wood C. 2003. The role of Kaposi's sarcoma-associated herpesvirus/human herpesvirus-8 regulator of transcription activation (RTA) in control of gene expres-sion.Oncogene, 22: 5150-5163.
        doi: 10.1038/sj.onc.1206555

    120. Whitby D, Howard M R, Tenant-Flowers M, et al. 1995. Detection of Kaposi sarcoma associated herpesvirus in peripheral blood of HIV-infected individuals and progression to Kaposi's sarcoma. Lancet, 346: 799-802.
        doi: 10.1016/S0140-6736(95)91619-9

    121. Whitby D, Luppi M, Barozzi P, et al. 1998. Human herpesvirus 8 seroprevalence in blood donors and lymphoma patients from different regions of Italy. J Natl Cancer Inst, 90: 395-397.
        doi: 10.1093/jnci/90.5.395

    122. Zeng Y, Zhang X, Huang Z, et al. 2007. Intracellular Tat of human immunodeficiency virus type 1 activates lytic cycle replication of Kaposi's sarcoma-associated herpesvirus: role of JAK/STAT signaling. J Virol, 81: 2401-2417.
        doi: 10.1128/JVI.02024-06

    123. Zhu F X, Chong J M, Wu L, et al. 2005. Virion proteins of Kaposi's sarcoma-associated herpesvirus. J Virol, 79: 800-811.
        doi: 10.1128/JVI.79.2.800-811.2005

    124. Ziegler J L, Newton R, Katongole-Mbidde E, et al. 1997. Risk factors for Kaposi's sarcoma in HIV-positive subjects in Uganda. AIDS, 11: 1619-1626.
        doi: 10.1097/00002030-199713000-00011

  • 加载中

Article Metrics

Article views(4875) PDF downloads(14) Cited by(0)

Related
Proportional views
    通讯作者: 陈斌, bchen63@163.com
    • 1. 

      沈阳化工大学材料科学与工程学院 沈阳 110142

    1. 本站搜索
    2. 百度学术搜索
    3. 万方数据库搜索
    4. CNKI搜索

    The Biology of Kaposi's Sarcoma-Associated Herpesvirus and the Infection of Human Immunodeficiency Virus

      Corresponding author: Chun LU, clu@njmu.edu.cn
    • Laboratory of Molecular Virology, Department of Microbiology and Immunology, Nanjing Medical University, Nanjing 210029, China
    Fund Project:  Fok Ying Tung Education Foundation 101038Program for New Century Excellent Talents in Chinese Universities NCET-05-0506

    Abstract: Kaposi sarcoma-associated herpesvirus (KSHV), also known as human herpesvirus 8 (HHV-8), is discovered in 1994 from Kaposi’s sarcoma (KS) lesion of an acquired immunodeficiency syndrome (AIDS) patient. In addition to its association with KS, KSHV has also been implicated as the causative agent of two other AIDS-associated malignancies: primary effusion lymphoma (PEL) and multicentric Castleman’s disease (MCD). KSHV is a complex DNA virus that not only has the ability to promote cellular growth and survival for tumor development, but also can provoke deregulated angiogenesis, inflammation, and modulate the patient’s immune system in favor of tumor growth. As KSHV is a necessary but not sufficient etiological factor for KS, human immunodeficiency virus (HIV) is a very important cofactor. Here we review the basic information about the biology of KSHV, development of pathogenesis and interaction between KSHV and HIV.

    • In 1872, the famed Hungarian dermatologist Moritz Kaposi described the purple-coloured lesions found in the skin of 5 male patients as "idiopathic multiple pigmented sarcoma of the skin" (60), an entity that was later on known as classic Kaposi's sarcoma (KS), which predominantly affected older men of Mediterranean and eastern European descent. In the 1920s, it was observed that KS occurred more frequently in East and Central Africa. African endemic KS was more rapidly progressive over several months to years. The uneven geographical distribution led to the hypothesis that KS might be caused by an in-fectious agent (87). In 1972 Giraldo et al. identified herpesvirus-like particles in the tissue culture of KS (43). An abrupt increase in the number of cases of KS among previously healthy young homosexual men was first reported in 1981, ushering in a new era of aggressive, rapidly fatal KS (12, 38, 45, 53, 111). Since approximately 30% of acquired immunodefi-ciency syndrome (AIDS) patients presented with KS as their initial symptom of human immunodeficiency virus (HIV) infection, KS evolved into a defining characteristic of one of the most devastating infectious diseases in history (8, 44). In 1990, a landmark epide-miological study from Beral et al. reported that KS was 20 000 times more likely to occur in people with HIV than in the general population (9). KS was more common in those who had acquired HIV sexually than in those who had acquired it via other routes. The incidence of KS was not related to age or race, but showed a definite geographical distribution, with the highest prevalence in the areas that were the initial foci of the AIDS epidemic. The accumulation of the epidemiological evidence suggested the involvement of a sexually transmissible agent in the development of KS, which in western countries had spread mainly among homosexual men. Several groups attempted to identify the unknown agent, and in 1994, Chang et al. identified a new human herpsvirus called Kaposi's sarcoma-associated herpesvirus (KSHV) or human herpesvirus 8 (HHV-8) from an AIDS-KS lesion using representational difference analysis, which showed the presence of unique herpesviral sequences in the KS tissue (23).

    • Based on biological characteristics and genomic organization, herpesviruses are classified into three subfamilies: alpha (α), beta (β), and gamma (γ). The gammaherpesvirinae are grouped into two classes: lymphocryptoviruses (γ-1) and rhadinoviruses (γ-2). Epstein-Barr virus (EBV) is a lymphocryptovirus while KSHV is a rhadinovirus (81). KSHV contains a large double-stranded DNA which is a closed circular episome in the nucleus during latency but is linear during lytic replication. Over 90 genes are encoded by a ~140 kilobase (kb) long unique region (LUR) with 53.5% G+C content, which is flanked by 20-35 kb terminal repeat regions composed of 801 base pair (bp) terminal repeat units (TR) with 84.5% G+C content (99, 101). The viral genes encoded by KSHV can be divided into three classes: 1) genes common to all herpesviruses, 2) genes unique to KSHV (these are generally given a "K" designation followed by number, and 3) KSHV-encoded genes that are homologous to cellular genes (these may be unique to KSHV or shared with other herpesviruses), and are likely to have usurped from the host genome during the course of evolution (85). Because of the KSHV genome complexity, the exact number of genes in the genome remain unknown.

    • Similar to other herpesviruses, KSHV displays two patterns of infection: latent and lytic phase. KSHV genome contains genes coding for different proteins of the lytic and latent phases, which have functional roles in virion structure, viral replication, host cell gene regulation, and immune response (79).

      During the latent phase of KSHV infection, 6 proteins are expressed from the viral episome: viral cyclin D/ vCyclinD (ORF72), latency-associated nuclear antigen 1/ LANA1 (ORF73), Kaposin A and B, the viral Fas-ligand interleukin-1β-converting enzyme (vFLICE) inhibitory protein/ vFLIP (K13), and the viral interferon regulatory factor 3/ vIRF3/ LANA2 (79, 123). The switch from latency to replicative phase is under tight control by a regulator of transcription [replication and transcription activator (RTA/ORF50)] (119). KSHV expresses 3 categories of lytic genes: immediate early (IE), early, and late genes (79). Early genes are expressed after IE genes and correspond to different ORF including K3, K5, K8, viral interleukin (vIL)-6, viral macrophage inhibitory protein (vMIP)-Ⅰ, vMIP-Ⅱ, vMIP-Ⅲ, vBCL-2, viral G protein-coupled receptor (vGPCR)/vIL-8R, and thymidilate cyclase (104, 119). Late lytic genes code for capsid, teguments, and membrane proteins (123). Lytic reactivation pro-motes the spread of the virions from the lymphoid reservoir to other target cell types such as endothelial cells (119).

    • Many studies have been done to measure the preva-lence of antibodies to KSHV (24). A variety of techni-ques have been used in these studies, but to date there is no gold standard against which to measure the efficacy of these techniques, so some may overesti-mate the number of positives, and some may underesti-mate the number. Nonetheless, certain trends are clear (106). Unlike most human herpesviruses, KSHV is not spread universally among all human populations. The seroprevalence rates of infection with KSHV in the general population vary depending on the geographic origin of the subjects: it is relatively low in Asia (India, 3.7%; Thailand, Malaysia, and Sri Lanka less than 4.4%) (1), in the United States (5%-20%, but 20%-50% in populations infected with HIV-1), and in Europe [Italy, 2%-6.5%; The exception is Sweden, where 20% of blood donors were reported to test positive for KSHV antibodies (a wide variety of results depending on the assay used)] (25, 32, 39, 57, 94), and is high in the south of Italy (24%) (121), and in sub-Saharan Africa ( > 36%) (1, 33, 88). In East African countries the prevalence of serum antibodies to KSHV is invariably high (Tanzania, 33%-84%; Uganda, 53%-77%; and Zambia, 37.5%) (1, 29, 46, 63, 88). In black South African blood donors and patients with cancers other than KS, the seroprevalence is 32%; in KS patients, it rises to 83% (112). African countries have the highest prevalence of KS in the world (124).

      KSHV seroprevalence is higher among HIV-serone-gative homosexual subjects compared with the serone-gative heterosexual population, ranging from 20% to 38% (6, 67, 86), and is even higher in HIV-seroposi-tive homosexual populations, ranging from 30% to 48% (1, 41, 67, 97). HIV-seropositive subjects with KS have at least a 95% KSHV seropositivity com-pared with 30% in HIV-seropositive subjects without KS (55). In homosexual men who are seropositive for both KSHV and HIV, the presence of KSHV DNA in peripheral blood mononuclear cells (PBMC) precedes and predicts the development of KS (80, 120), and between 30% and 50% of subjects in this group will develop KS within 10 years of contracting the dual infection (14, 56, 86).

      Transmission of KSHV most likely includes both sexual and nonsexual routes. Sexual route is the most common route in lowprevalence developed countries, where homosexual men have the highest risk of contracting the virus (11, 67). In the endemic areas of Africa infection rates are high and both sexual and nonsexual transmission occurs. Most individuals get infected during childhood and high prevalence rates have been reported in infants and children reaching ≥50% in some studies (42, 70, 72). Epidemiological studies support the role of horizontal transmission of KSHV from mother to child and between siblings during close non-sexual contacts (30, 70, 72, 92, 93). KSHV DNA has been detected frequently in saliva, where virus titers are higher than in samples from other anatomic sites such as genital secretions (10, 91). Interestingly, also unaffected carriers can harbor infectious KSHV in their oral cavities and the virus from these carriers is able to infect cultured primary epithelial cells (31).

    • The spindle cells isolated from KS lesions of HIV-1-infected individuals are generally hyperplastic non-transformed cells that proliferate in culture with inflammatory cytokines. These cells initially contain KSHV that is rapidly lost as the cells are passaged in culture. Cell lines composed of transformed cells have also been isolated from advanced KS lesions (3, 62). These cell lines, unlike AIDS-KS spindle cells, are transformed cells that grow in the absence of inflam-matory cytokines, contain cytogenetic abnormalities, and induce durable tumor lesions when inoculated into nude mice (58, 102). Similar to the hyperplastic KS spindle cells, the KS transformed cells have also lost the KSHV genomes. One possible explanation is that when cultured ex-vivo, KSHV-infected cells undergo epigenetic mutations or changes enabling them to grow without the virus. In addition, other host-specific factors not present in explants might be required for the maintenance of the virus. In cell culture, KSHV is capable of infecting a diverse range of human and animal cell lines including primary CD19+ B cells, human papilloma virus (HPV)-transformed human brain endothelial cells BB18 and 181GB1-4, primary neonatal capillary endothelial cells, human embryonic kidney 293 cells, Ln-Cap cells, human lung carcinoma A549 cells, CHELI (Chediak-Higashi syndrome) cells, squamous cell carcinoma SCC15 cells, human fibro-blast cells, human bladder carcinoma T24 cells, human prostate carcinoma DU145 cells, human cervical carcinoma HeLa cells, baby hamster kidney BHK-21 cells, owl monkey kidney OMK637 cells, green monkey fibroblasts (Vero) cells, green monkey kidney CV-1 cells, SLK cells (KS-spindle cells), and murine fibroblast 3T3 cells (7, 98).

    • Four distinct clinical forms of KS have been described. Classic KS is a disease of elderly Mediterranean and Eastern European men (115). Endemic KS is found in parts of equatorial Africa such as Uganda and Zambia (67), where it is one of the most frequently occurring tumors. Endemic KS tends to be more aggressive than classic KS. Iatrogenic KS occurs after solid-organ transplantation in patients receiving immunosuppres-sive therapy and KS comprises an estimated 3% of all tumors associated with transplantation (73). Epidemic AIDS-KS is the most common neoplastic mani-festation of AIDS in the United States and Europe and is one of the diagnostic criteria for AIDS (68). Highly active antiretroviral therapy (HAART) has led to a substantial decline of AIDS-KS in the United States. However, even in the current post-HAART era, stan-dardized incidence rates for KS are higher than that of any other AIDS-defining or AIDS-associated cancers (71). This suggests that KS will remain a permanent health problem for years to come. In contrast to the indolent course of classic KS, AIDS-KS takes a much more aggressive course. AIDS-KS tends to dissemi-nate widely to mucous membranes and the visceral organs. Despite the different clinical manifestations of KS, the histology of lesions from skin, lymph nodes, respiratory tract, and intestines is very similar. The KS lesion is highly angiogenic and comprises spindle-shaped cells, slitlike endothelium-lined vasculature, and infiltrating blood cells. The spindle cells form the majority of the cell population, and are thought to arise from lymphatic endothelial cells. KS lesions are divided into patch, plaque, and nodular stages. KS lesions range from patches or plaques to nodules and there is evidence for both polyclonality and mono-clonality of the lesions (59, 95). It is thought that KS probably initiates as a polyclonal hyperplasia and develops into a clonal neoplasia.

      In addition to KS, KSHV is also found in B lym-phoproliferative diseases: PEL and multicentric Castleman's disease (MCD) (21, 23, 114). PEL, also called body cavity-based lymphoma, is a rare non-Hodgkin's B-cell lymphoma commonly found in HIV-infected patients (22). This type of lymphoma is characterized as a malignant effusion in the pleural, pericardial, or peritoneal space, usually without a contiguous tumor mass (15, 22). Most PELs are KSHV-positive, and are often coinfected with EBV as well. These tumors are typically large-cell immuno-blastic or anaplastic large-cell lymphomas that express CD45, clonal immunoglobulin gene rearrangements, and lack c-myc, bcl-2, ras, and p53 gene alterations (2, 83). MCD is a B-cell lymphoproliferative disorder that is sometimes referred to as multicentric angiofolli-cular hyperplasia. There are two forms of MCD: 1) a plasmablastic variant form that is associated with lymphadenopathy and immune dysregulation and 2) a hyaline vascular form, which presents as a solid mass. KSHV-associated MCD belongs to the plasma cell variant subgroup of MCD. MCD is a polyclonal tumor and is highly dependent on cytokines such as IL-6. Contrary to KS and PELs, KSHV DNA sequences have been detected only in a subset of MCD-patients, most often those co-infected with HIV (114).

    • KSHV is considered to be a human oncogenic virus. Its genome encodes not only proteins necessary for the reproduction and persistence of the virus but also pathogenic factors to promote tumor cell proliferation and survival, regulate angiogenesis, and modulate host immune response in favor of tumor growth.

      KSHV uses multiple mechanisms to promote cell survival and facilitate tumor development. Like other oncogenic DNA viruses, KSHV genes LANA-1, LANA-2, RTA, and vIRF-1 target both p53 and retinoblastoma (Rb) tumor suppressor pathways (37, 49, 84, 96, 100, 109, 110). Furthermore, KSHV can promote cell cycle progression from G1 to S-phase and accelerate cellular proliferation through vCyclin (27, 64, 103). The KSHV genome encodes several virus homologues of human antiapoptosis proteins. For instance, vBcl-2 is a KSHV homolog of human antiapoptosis protein Bcl-2 (26, 105). KSHV also encodes several cellular IRF homologues vIRFs that inhibit apoptosis (13, 40, 61, 66, 108, 109). Another unique mechanism that KSHV utilizes to promote tumor growth and cell survival is through activation of the nuclear factor κB (NF-κB) pathway. By activation of NF-κB, apoptosis is suppressed and regulation of the cell cycle is disturbed. Two KSHV genes, vFLIP and vGPCR have been shown to enhance cell growth and survival by activating the NF-κB pathway (36, 65, 69, 107). vFLIP is associated with anti-apoptotic activity, and activation of NF-kB by vFLIP is associated with prolonged survival of KSHV infected PEL cell lines (47, 116). vGPCR is a member of the CXC chemokine GPCR family that has significant homology to IL-8 receptor, but unlike the cellular GPCR, vGPCR exhibits ligand independent activity. vGPCR can activate PKC (protein kinase C), Akt [protein kinase B (PKB)], NF-κB, and MAPK (mitogen-activated protein kinase) to regulate the expression of growth factors [VEGF (vascular endothelial growth factor), bFGF (basic fibroblast growth factor)], cytokines (IL-1β, TNF-α, IL-6), and chemokines [IL-8, monocyte chemoattractant protein 1 (MCP-1)] (28, 78, 82, 90, 107, 113). The secretion of various factors could promote cell proliferation and stimulate angioproliferation through autocrine/paracrine signaling. KSHV-encoded vIL-6 also promotes cell survival. By coupling to the glycoprotein-130-receptor, the signal transducer and activator of transcription (STAT) cascade is activated and thus autocrine me-chanisms of tumorigenes is involved (51, 77, 89, 118).

      KSHV utilizes two major strategies to evade the host immune response and achieve successful in-fections. The first is the so-called passive strategy. After a successful entry into host cell, KSHV establi-shes latency, during which only a minimum number of virus genes are expressed, thus reducing the number of antigens that are exposed to the immune systems. Secondly, during lytic replication or de novo infection, when most viral proteins are expressed and are susceptible to immune surveillance, the virus utilizes an active strategy that involves a number of its own unique genes to modulate the host immune response.

    • KSHV is a large (~170 kb) double-stranded DNA herpesvirus (99) and HIV-1 is a much smaller (~10 kb) retrovirus. Despite rather dramatic molecular dissimi-larities between the two viral genomes, many findings have provided compelling evidence that these two viruses interact to promote KS pathogenesis in dually infected individuals. Indeed, it is increasingly clear that, although the KSHV genome is invariably present in all described cases of KS (20), only a small proportion of infected people ever develop KS or KSHV-induced lymphomas. On the other hand, among KSHV-infected individuals, the risk of KS is much higher in those with HIV-1 infection than among those with other types of immunosuppression. Thus, a reasonable deduction is that KSHV is neces-sary but insufficient for producing KS and that HIV-1 is an important cofactor, which promotes KSHV induced KS.

      HIV-1 probably exacerbates KSHV pathogenesis at multiple levels, including through immunosuppression, by priming of target cells and the tissue microenviron-ment for KSHV infection and replication, and by exerting direct effects on KSHV gene expression and viral replication. Direct and reciprocal interactions between HIV-1 and KSHV at the molecular level have received a lot of recent attention in the literature, they suggest mutual positive-feedback loops during replication of both viruses.

    • Studies have demonstrated that KSHV can modu-late HIV-1 replication and KSHV might be a cofactor for HIV progression. HIV-1 replication significantly increases in the presence of KSHV both in vitro and in vivo model systems (76). KSHV coinfection markedly increases HIV replication in primary or transformed monocytic and endothelial cells (19). KSHV infection also induces HIV reactivation in chronically infected cell lines and in PBMCs from patients with asympto-matic HIV (19). ORF50, an IE gene of KSHV, encodes RTA necessary for virus reactivation and lytic replication. KSHV ORF50 can enhance HIV repli-cation in T and B cells (Jurkat and BC-3 cells) (17, 18). Transfection of ORF50 into nonsusceptible B and glial cells [BCBL-1 (body cavity-based lymphoma 1, latently infected with KSHV) and A172, respectively] increases cell susceptibility to infection and results in transient permissiveness to HIV replication (17, 18). KSHV ORF50 also interacts synergistically with HIV-1 transactivative transcription protein (Tat) in the transactivation of HIV-1 long terminal repeat (LTR) both in BCBL-1 cells and in HL3T1 cells (an epithelial cell line non-permissive to KSHV infection) (16). In addition, KSHV-encoded vFLIP (K13) is able to activate HIV-1 LTR via the activation of the classical NF-κB pathway and coexpression of HIV-1 Tat with K13 leads to synergistic activation of HIV-1 LTR (117). KSHV LANA can also transactivate the HIV-1 LTR in the human B-cell line BJAB, human monocytic cell line U937, and the human embryonic kidney fibroblast cell line 293T. Moreover, LANA cooperates with HIV Tat in activation of the LTR in a dose-response fashion with increasing amounts of LANA (54). KSHV ORF57, encoding a post-trans-criptional regulator, also enhances Tat-induced trans-activation of HIV-1 LTR in BCBL-1 cells (16).

    • HIV-1 has an active role in the development of KSHV-associated diseases. HIV-1 not only activates lytic replication of KSHV but also induces secretion of a number of cytokines and growth factors, further promoting tumor cell proliferation and survival, as well as tumor angiogenesis. In BC-3 cells, HIV-1 infection leads to reactivation of latent KSHV genomes (74). Although recombinant HIV-1 gp120 fails to enhance herpesvirus expression, transient transfection of the HIV-1 Tat suffices to reactivate latent KSHV (74). HIV-1 Tat activates lytic cycle replication of KSHV, JAK/STAT signaling partially contributes to this process (122). HIV-1-encoded viral protein r (Vpr) also increases the expression of KSHV genes (52). Full-length HIV-1 Tat and a 13-amino-acid peptide corresponding to the basic region of Tat can specifically enhance the entry of KSHV into endothelial and other cells, probably by concentrating virions on cell surface (5). The specific cytokines produced by or in response to HIV-1-infected T cells in the coculture system (HIV-1-infected T cells and the KSHV-infected BCBL-1 cell line), particularly Oncostatin M, hepatocyte growth factor/scatter factor, and interferon-gamma, can play an important role in the initiation and progression of KS through reactivation of KSHV (75). HIV and its Tat protein induced inflammatory cytokines including TNF-α, IL-1β, IL-8, and IL-6 can trigger lesion formation by inducing the activation of endothelial cells that leads to adhesion and tissue extravasation of lymphomono-cytes, spindle cell formation, angiogenesis and KSHV reactivation that, in turn, leads to virus spread to all circulating cell types and virus dissemination into tissues (34, 35, 75). The other viral proteins of HIV, e. g. gp120 and nef also lead to paracrine secretion of cytokines and thus to stimulation of tumor growth and angiogenesis (4). HIV-1 Tat stimulates angiogenesis via specifically binding and activating the Flk-1/ kinase insert domain receptor (Flk-1/KDR), a VEGF-A tyrosine kinase receptor (50). The combination of inflammatory cytokines (IL-1β, TNF-α, and IFN-γ) increased in KS lesions synergizes with Tat to promote the development of angioproliferative KS-like lesions in nude mice. Inflammatory cytokines induce the tissue expression of both VEGF and bFGF, two angiogenic molecules highly produced in primary KS lesions. Tumorigenesis by KSHV vGPCR is accelerated by HIV-1 Tat via activating NF-κB and NF-AT (48).

    • As discussed above, KSHV is associated with KS, PEL, and MCD, malignancies occurring more fre-quently in AIDS patients. The aggressive nature of KSHV in the context of HIV infection suggests that interactions between the two viruses enhance patho-genesis. KSHV encodes several genes that have the capacity to stimulate angiogenesis by inducing an angiogenic phenotype of KSHV-infected cells, or by the production of inflammatory cytokines, growth factors and chemokines, that can activate and promote angiogenesis. However, our current recognition and understanding of KSHV-related malignancies and its molecular mechanisms are still poorly developed. Further studies are needed to reveal the unique mechanisms used by this viral pathogen, which is crucial for the development of targeted-therapies for KSHV-associated malignancies.

    Reference (124) Relative (20)

    目录

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return