Indian J Med Res 121, April 2005, pp 270-286 270 Upon infection by human immunodeficiency virus type 1 (HIV-1), host cells react with various innate, cellular and humoral immune responses to counteract the viral invasion. Limited and transient restriction of viral infection is normally achieved. However, HIV ultimately overcomes these antiviral responses resulting in successful viral infection and replication. Expression of several HIV-1 regulatory and accessory genes such as vif, vpu and tat is known to regulate some of the host cell innate immune responses to maximize viral infection or replication1-3. For example, a host innate antiviral response mediated by APOBEC3G was recently found to sabotage HIV reverse transcription through cytidine deamination within minus DNA strand4-7.Targeting viral replication by DNA Key words Apoptosis - cell cycle G2/M arrest - disease progression - HIV-1 - host-pathogen interaction - host immune responses - nuclear transport - Vpr - viral pathogenesis HIV-1 viral protein R (Vpr) & host cellular responses Richard Yuqi Zhao*+, Michael Bukrinsky** & Robert T. Elder+ *Department of Pathology, University of Maryland School of Medicine, Baltimore, MD 21201, **Department of Microbiology and Tropical Medicine, The George Washington University, Washington, DC 20037 & +Children’s Memorial Research Center, Northwestern University Feinberg School of Medicine, Chicago, IL 60614, USA Accepted February 21, 2005 During infection of host cells by HIV-1, active host-pathogen interactions take place. The final balance between these interactions determines the efficiency of viral infection and subsequent disease progression. HIV-infected cells respond to viral invasion with various antiviral strategies such as innate, cellular and humoral immune antiviral defense mechanisms. On the other hand, the virus has also developed tactics to suppress these host cellular responses. Among the many viral offensive strategies, viral protein R (Vpr) plays a particularly active role. Vpr involved in nuclear transport of the viral pre-integration complex, activation of viral transcription, induction of cell cycle G2/M arrest and apoptosis of the host cells. However, specific roles of these Vpr activities in viral pathogenesis and their contribution to disease progression are not fully understood. HIV-1 defective for some or all of these Vpr activities have been associated with slow disease progression in some patients. With regard to the host responses to vpr gene expression, studies show that Vpr is specifically targeted by CD8 T-lymphocytes during acute viral infection and that the host innate immune response may also play a crucial role in suppressing the effects of Vpr on various cellular activities. The effect of host cellular responses to vpr gene expression and its roles in nuclear transport, cell cycle G2/M regulation and induction of apoptosis are discussed in this review. Strategies with potential application for future antiviral therapies directed at suppressing Vpr activities are described. 271 deamination seems to be one of the major host innate responses defending against retroviral infections. On the other hand, HIV-1 accessory protein Vif prevents APOBEC3G from entering the HIV-1 virion during viral assembly, thus ensuring viral replication in target cells8,9. HIV-1 viral protein R (Vpr) is another accessory protein. It can be found in the serum of HIV-1 infected patients and in the cerebrospinal fluid (CSF) of AIDS patients with neurological pathologies10,11. Because Vpr is a highly conserved viral protein, it presents a good target for host antiviral responses. However, little is known at present about Vpr-host cell interactions or whether Vpr is involved in suppressing host antiviral responses as Vif does. Numerous reports certainly indicate a dynamic interaction between Vpr and host cellular responses. We will briefly summarize those findings here and will examine the potential contribution of Vpr to viral pathogenesis and disease progression. HIV-1 Vpr is a virion-associated viral gene product with an average length of 96 amino acids, and a calculated molecular weight of 12.7 kDa. However, it typically appears as either a 14 kD or 15 kDa band due to post-translational modifications. Vpr is a highly conserved viral protein among HIV, simian immunodeficiency viruses (SIV) and other lentiviruses12,13. Besides lentiviruses, its protein sequence shares no strong homology with any of the known proteins. A tertiary structure of Vpr proposed on the basis of nuclear magnetic resonance (NMR) analysis consists of an ? -helix-turn-? -helix domain in the amino-terminal half from amino acids 17 to 46 and a long ? -helix from 53 to 78 in the carboxy-terminal half14,15. These three ? -helices are folded around a hydrophobic core in a structure which allows interaction of Vpr with different cellular proteins16. Vpr displays several distinct activities in host cells. These include cytoplasmic-nuclear shuttling 17, induction of cell cycle G2 arrest18 and cell killing19. These three Vpr-specific activities were shown to be functionally independent of each other20, 21 and have been demonstrated in a wide variety of eukaryotic cells ranging from human to yeast, indicating that Vpr most likely affects highly conserved cellular processes. In this review, we describe our current understanding of the host-Vpr interactions and the potential roles of Vpr activities in viral pathogenesis and disease progression. HIV-1 Vpr and host cellular responses All regulatory and accessory HIV-1 viral proteins are being targeted by HIV-1-specific CD8-positive cytotoxic T-lymphocytes (CTLs)22. However, Vpr is being preferentially targeted by the CD8+ T- lymphocytes as compared with other viral proteins, at least during the acute phase of the viral infection23,24, suggesting a salient role of Vpr during the early phase of infection. Consistently, some of the cellular proteins, such as heat shock proteins (HSPs), respond quickly to viral infection or vpr gene expression, which suggests another level of host innate immune response25,26 (our unpublished data). For example, HSP27 and HSP70 mRNA transcription appeared as early as 3-8 h following HIV infection. We now know that some of the small heat shock proteins, such as HSP27 or HSP70, exert effective protective effect against some or all of the Vpr activities27-29. However, responsive elevations of HSPs to HIV-1 infection are transient as the HSP27 and HSP70 mRNA transcripts were significantly downregulated by 24 h after viral infection, concomitant with the first appearance of the full length genomic HIV-1 mRNA30. This observation implies an active interplay between HIV viral proteins and HSP27 or HSP70. Indeed, an active and antagonistic interaction was seen between Vpr and a yeast homologue of HSP2731. Vpr suppresses antigen- specific CD8-mediated CTL and Th1 immune responses32. Consistent with this notion, Rhesus macaques infected with HIV-2 lacking vpr gene had increased antibody titres compared to monkeys infected with the wild-type virus33. Although molecular mechanisms underlying suppression of CTL and antibody production by Vpr are presently unknown, it was surmised that Vpr may prevent antibody production against the virus by inhibiting T-cell clonal expansion through suppressing T-cell proliferation and inducing cell cycle G2/M arrest34. Evidence also ZHAO et al: HIV-1 VPR 272 INDIAN J MED RES, APRIL 2005 suggests that Vpr may suppress host inflammatory responses, which present another level of the host immune responses to viral infections35. Vpr inhibits host inflammatory responses by down regulating pro- inflammatory cytokines (TNF? and IL-12) and chemokines (RANTES, MIP-1? and MIP-1? ) in a manner similar to glucocorticoids36,37; Vpr additionally suppresses host inflammatory response by inhibiting nuclear factor kappa B (NF? B) activity through the induction of I? B37. Therefore, there appears to be at least two levels of host responses to vpr gene expression; one is the cellular immune response mediated by CD8+ CTLs; and another innate immune responses involving some of the cellular chaperone proteins. Conversely, Vpr counteracts those host immune responses. It prevents T-cell proliferation, suppresses host inflammatory responses including production of cytokines and chemokines. Vpr may also have additional mechanisms to counterbalance some of those innate reactions, such as heat shock proteins, that have specific suppressive activities against Vpr. These specific host responses to Vpr and the counteracting effect by Vpr strongly suggest a very dynamic interaction between vpr gene expression and host reactions. Future studies should reveal to what extent these interactions contribute to the success of viral infection and will determine the best way to exploit those specific host responses to design strategies aimed at suppressing Vpr. Potential association of Vpr activities with viral pathogenesis and disease progression The importance of Vpr in viral pathogenesis has been addressed in a number of earlier studies of SIVmac219 infection in Rhesus monkeys. However, these experiments produced conflicting results and interpretations, which were further complicated by the presence of two genes in SIVmac219 homologous to the vpr gene of HIV-1. These two genes, vpr and vpx, are thought to have arisen by duplication of the ancestral vpr gene13, although phylogenetic analysis of various lentiviruses later suggested that non- homologous recombination may have played a role in the evolution of vpr and vpx genes12,38. It is also believed that the functions of the single vpr of HIV- 1 have been divided among the vpr and vpx genes of SIV39. In one study, Rhesus monkeys infected with the SIVmac219 defective in vpr had low viral burden and no disease progression while monkeys infected with the wild-type virus or those Vpr-defective viruses spontaneously reverted to the wild-type exhibited high viral burden and rapid disease progression 40. In contrast, in two other studies 41,42, no significant differences in disease progression were found between the vpr-deficient SIVmac219 and the parental wild-type virus, and all of the monkeys developed AIDS41,42. However, in the experiment most relevant to infection by HIV-1 with the single vpr gene, monkeys infected with SIVmac219 defective in both the vpr and vpx genes had severely attenuated infections with much lower viral burden and no evidence of disease progression41,43. At least a 100-fold decrease in pathogenic index was found in the infectivity of these mutant viruses in comparison with the wild-type viruses, confirming the role of Vpr in viral pathogenesis43. Vpr seems to be required for in vivo replication of the virus. An earlier study in Rhesus monkeys showed rapid reversion of a single base pair mutation in the stop codon of the vpr gene40. The requirement for Vpr in vivo by HIV-1 was further illustrated in chimpanzees and an accidentally infected laboratory worker, who was initially infected with a Vpr- defective HIV-1 viral laboratory strain IIIB44. The vpr gene in the viral strain IIIB has a frame shift mutation at codon 73, which results in a truncated Vpr protein with only the first 72 amino acids. However, the mutant vpr gene in the virus initiating the infection reverted to the wild-type Vpr both in this human subject and experimentally infected chimpanzees44. These results clearly demonstrate the necessity of Vpr in vivo. Even though Vpr sequence is one of the most conserved regions in HIV genome, with estimated similarities of 87 per cent between different viral strains45,46, Vpr sequence variations were inevitably found in viruses infecting patients showing fast or slow disease progression. Earlier efforts in studying the role of Vpr in disease progression were mostly based on nucleotide sequence analyses47-51. Even though such analysis is useful in detecting gene deletions or 273 insertions, it cannot be used to predict functional changes derived from amino acid substitutions. As a result, different conclusions were drawn with regard to potential contribution of Vpr to disease progression. A synonymous or non-synonymous amino acid substitution in Vpr could dramatically affect Vpr functions as shown in a number of mutagenesis studies52-55. A potential problem associated with these mutagenesis studies is, however, that the vpr mutants were often artificially created and therefore may not represent the profile of naturally occurring mutations. An alternative method to rapidly determine multiple Vpr activities, such as cell cycle G2 arrest, nuclear localization and induction of cell death, is the use of fission yeast (Schizosaccharomyces pombe) model system, which allows simultaneous determination of these three Vpr activities using an inducible gene expression system20,56,57. These three Vpr activities observed in fission yeast are very similar to that observed in mammalian cells58-60. By using the yeast model system, these Vpr activities were shown, as expected, to be highly conserved in many of the viral laboratory strains and also in viruses isolated from patients progressing to AIDS61. In contrast, a significant proportion (23-50%) of Vpr in viruses isolated from a non-progressing mother-child pair was functionally defective in all three Vpr activities61,62. Variation analyses of the protein sequences indicated that these Vpr proteins carried unique amino acid substitutions that frequently interrupted highly conserved domains, which include a N-terminal ? -turn-? helix and an epitope sequence that is preferentially targeted by the CTL immune response61. Association of slow disease progression with Vpr defect was also found in another HIV- infected non-progressor63. The vpr gene isolated from this patient contained both premature stop codons and an unusual Q3R polymorphism, which significantly impairs the ability of Vpr to confer cytopathicity but has no effect on the efficiency of viral replication. Together, these data support the idea that functional Vpr might be one of the viral factors contributing to disease progression. However, these three reported non-progressors are thus far the only HIV-infected patients known to carry functionally defective Vpr. Whether these are unique cases or a common trend remains to be determined. Based on protein sequence analysis of Vpr deposited in the database, a high frequency of R77Q Vpr mutations were reported to associate with patients with slow disease progression64. Although the authors convincingly demonstrated that this Vpr mutation, when isolated from an HIV-1 viral strain HxBRU, impaired the ability of Vpr to induce apoptosis, it was not present in viruses isolated from some of those long-term non- progressors61,65. Therefore, association of this mutation with slow disease progression remains unproved65,66. It also remains open how strong a correlation is between Vpr functional defects and slow disease progression of HIV-infected patients. Induction of cell cycle G2/M arrest To ensure accurate transmission of the genetic information, eukaryotic cells have developed an elaborate network of checkpoints to monitor the successful completion of every cell cycle step and to respond to cellular abnormalities such as DNA damage and replication inhibition as they arise during cell proliferation. Two of the best characterized G2/M checkpoints, DNA damage and DNA replication67-70 were first characterized in detail by genetic analysis in fission yeast. The G2 to M transition is controlled in fission yeast by the phosphorylation status of Tyr15 on Cdc2, the cyclin- dependent kinase which regulates the cell cycle in all eukaryotic cells71. In fission yeast, Tyr15 is phosphorylated by the Wee1 and Mik1 kinases to hold the cell in G2, and rapid dephosphorylation by the Cdc25 phosphatase triggers the G2 to M transition71-74. The DNA damage checkpoint is activated by ionizing radiation or ultraviolet light, and activation of this checkpoint leads to inhibitory phosphorylation of Cdc2 at Tyr15 by a multi-step pathway75,76. The early genes in the pathway, which include Rad1, Rad3, Rad9, Rad17, Rad26 and Hus1, are thought to sense the DNA damage and lead to phosphorylation of the Chk1 protein77. In response to double strand DNA breaks (DSBs) induced by ionizing radiation, for examples Rad17 acts as a checkpoint-specific loading factor, which responds to the DNA damage by loading a 9-1-1 protein complex onto the sites where DNA is damaged78,79. The 9-1-1 protein complex is also known as the checkpoint clamp complex (CCC), which is composed of Rad1, Rad9 and Hus179. In addition, the Rad3-Rad26 protein complex binds to ZHAO et al: HIV-1 VPR 274 INDIAN J MED RES, APRIL 2005 sites of DNA damage independently of the 9-1-1 protein complex. The independent binding of these two protein complexes to DNA damage to initiate the DNA structure checkpoint is believed to protect the cell against inappropriate checkpoint activation68,79,80. Activation of Chk1 is mediated by Crb2, which may bridge Rad3 and Chk1 81-83. The activated Chk1 kinase then directly phosphorylates the Cdc25 phosphatase84. The phosphorylated Cdc25 binds Rad24/25 protein, and this complex is transported out of the nucleus to render Cdc25 inactive85. The activated Chk1 also regulates the Mik1 kinase to inhibit Cdc286. DNA damage thus initiates a Chk1-mediated protein phosphorylation cascade ending in the inactivation of Cdc25 phosphorylase and activation of Mik1 kinase to increase inhibitory phosphorylation of Tyr15 on Cdc2. The DNA replication checkpoint is activated by treatment with hydroxyurea, which inhibits DNA replication, and this checkpoint also controls the G2 to M transition through inhibitory phosphorylation of Cdc276. Parts of this DNA replication checkpoint are shared with the DNA damage checkpoint as Rad1, Rad3, Rad9, Rad17, Rad26 and Hus1 are required for both checkpoints in fission yeast87. The same 9-1-1 and Rad3-Rad26 checkpoint protein complexes may associate with the DNA replication complex 79. However, the DNA replication checkpoint acts primarily through phosphorylation of Cds1 kinase, which is mediated by another protein Mrc188,89. There is a minor contribution from the Chk1 kinase, and either kinase is sufficient by itself to give cell cycle arrest when DNA synthesis is inhibited90. Activated Cds1 kinase inactivates Cdc25 through the same mechanism as Chk1 and may also activate the Wee1 and Mik1 kinases, which phosphorylate Tyr15 of Cdc267,90. Cell cycle G2/M controls, which were often initially defined in yeast, are highly conserved, and most of Table. Human homologues of fission yeast cell cycle control genes Fission yeast Human Putative activity Mitotic regulators: Cdc2 CDK1 Cyclin B-dependent kinase Cdc13 Cyclin B B-type cyclin Wee1 WEE1 Tyrosine kinase Mik1 — Tyrosine kinase Cdc25 CDC25A/B/C Tyrosine phosphatase DNA damage and replication checkpoints: Rad1 hRAD1 Nuclease Rad3 ATM/ATR Protein kinase Rad9 hRAD9 3’-5’ exonuclease Rad17 hRAD17 Unknown Rad24/25 14-3-3 Binds to phosphorylated ser Hus1 hHUS1 A PCNA-related protein Chk1 CHK1 Serine/Threonine Kinase Cds1 CHK2 Serine/Threonine Kinase Crb2 BRCA1 Unknown Mrc1 CLSPN Unknown Cellular proteins involved in Vpr-induced G2 arrest: PP2A PP2A Protein phosphatase 2A Paa1 A A regulatory subunit Pab1 B B regulatory subunit Ppa2 C C catalytic subunit Ppa1 C C catalytic subunit Wos2 P23 Inhibitor of Wee1 —, not found 275 the genes required for the G2/M checkpoints have human homologues (Table). In general, these homologues have similar, although not always identical, roles in the control of the human cell cycle. There is a tendency for multiple, partially redundant checkpoints in human cells compared to simpler checkpoints in yeast probably reflecting the more complex requirements for cell cycle control in multicellular eukaryotes. For example, the single rad3 gene in fission yeast is required for both the DNA damage and replication checkpoints and activation of the chk1 and cds1 checkpoint kinases68,79,80. In human cells, there are two homologues of rad3, ATM and ATR. The primary role of ATM is in the DNA damage checkpoint and activation of CHK2, the human homologue of cds1, whereas the primary role of ATR is in the DNA replication checkpoint and activation of CHK191,92. Similarly, there is only one Cdc25 tyrosine phosphatase that dephosphorylates Cdc2 in fission yeast. In human cells, there are three CDC25 homologues, CDC25A, CDC25B and CDC25C, and each of them can be phosphorylated by CHK193. All three of these phosphatases have been shown to be involved in the control of the G2/M transition, even though their specific roles in this process have not yet been well characterized94-96. The p53 gene is an example of an additional level of cell cycle control in human cells. The p53 transcription factor, which has no homologues in yeast, has multiple roles including regulation of apoptosis and the cell cycle with an essential role in the G1 DNA damage checkpoint 97. It also has important roles in the G2 damage checkpoint. It inhibits Cdc2 through activation of Gadd45, p21, and 14-3-3? . In addition, it is involved in regulatory feedback loops with ATM/ATR and CHK198. The conservation of checkpoints even extends to the regulatory mechanisms as illustrated by the negative regulation of Cdc25 by relocation to the cytoplasm from the nucleus in both fission yeast and human cells. This relocation in both organisms is dependent on 14-3-3 proteins85,99. The HIV-1 Vpr protein induces cell cycle G2 arrest through inhibitory phosphorylation of Cdc2 both in fission yeast and human cells, suggesting that Vpr affects a conserved cellular process. Specifically Vpr induces hyperphosphorylation of fission yeast Cdc2 or human CDK1, the human homologue of Cdc218,56,100. It exerts its inhibitory effect through T14A and Y15F of CDK1 and Y15F of Cdc2, as expression of nonphosphorylatable Cdc2 mutants, T14A Y15F of CDK1 and Y15F of Cdc2, prevents Vpr-induced G2 arrest18,101. Furthermore, Vpr inhibits Cdc25 phosphatase102,103 and activates Wee1 kinase102,104 to promote phosphorylation of Cdc2/Cdk1 during induction of G2 arrest. Consistent with the roles of Wee1 and Cdc25 in Vpr-induced G2 arrest, proteins that are involved in regulation of Cdc25 or Wee1 have also been identified to either augment or alleviate Vpr-induced G2 arrest. Fission yeast Wos2, which is a human p23 homologue and aWee1 inhibitor105, has been shown to be a multicopy Vpr suppressor102. A Cdc25 inhibitor rad2585, which is the human 14-3-3 homologue enhances Vpr-induced G2 arrest when overproduced in fission yeast102. Recent studies further showed that Vpr binds to CDC25C and 14-3-3 in human cells106,107. Given that the DNA checkpoints and Vpr both induce G2 arrest through inhibitory phosphorylation of Cdc2 which is regulated by Wee1 and Cdc25, Vpr might induce G2 arrest through a checkpoint pathway. This possibility has been evaluated in fission yeast by expressing vpr in mutant fission yeast strains defective in early and late steps of the checkpoint pathways. None of the early checkpoint-specific mutants (rad1, rad3, rad9 and rad17) showed a significant effect on the induction of G2 arrest by Vpr 101,102,108. Furthermore, mutations in both chk1 and cds1, which are thought to be the last steps specific for the checkpoint67,84,90, also do not block Vpr-induced G2 arrest102,108. Therefore, Vpr does not appear to use the DNA-damage or DNA-replication checkpoint pathways to induce G2 arrest in fission yeast. Early data in human cells tended to support the conclusion that Vpr does not induce G2 arrest through the DNA damage checkpoint pathways. Vpr still induced G2 arrest in cells from patients with ataxia telangiectasia (AT)103. These AT cells are mutant for the ATM gene, which is a human homologue of fission yeast Rad3, and they do not arrest in G2 in response to DNA damage109-111. However, recent reports show that Vpr activates ATR and a second human homologue of fission yeast Rad3, and other steps in this checkpoint pathway such as Rad17, Hus1, BRCA1 and Y-H2AX112,113. These studies suggest ZHAO et al: HIV-1 VPR 276 INDIAN J MED RES, APRIL 2005 that Vpr induces G2 arrest through either a cellular response to DNA replication stress or to a signal that “mimics” DNA damage. Expression of vpr does not increase gene mutation frequency114 and or changes radiosensitivity of the checkpoint defective mutant102, which argues against the possibility that Vpr actually causes DNA damage. It is thus reasonable to think that other signals other than actual DNA damage triggers DNA damage-like cellular responses. These cellular responses could include the nuclear herniation caused by Vpr115 or cellular stress responses to vpr gene expression27-29. Since ATR and CHK1 are primarily responsive to changes in DNA replication, an alternative possibility is that Vpr may interfere with DNA replication. This possibility is certainly supported by a number of reports showing that Vpr induces genomic instability, formation of micronuclei and aneuploidy116,117. All of these changes in DNA structures could be perceived as replication stresses, which would trigger cell cycle arrest. Considering that G2/M DNA checkpoints are highly conserved between mammalian and fission yeast cells (Table), it is unclear at the moment why human ATR and CHK1 are activated by Vpr but rad3 (the fission yeast homologue of ATR/ATM) or chk1/ cds1 (CHK1/CHK2) double deletion in fission yeast does not block Vpr-induced G2 arrest 101,102. One factor possibly contributing to the observed activation of ATR in mammalian cells is that retroviral integration appears to activate ATR118, and the experiments showing activation of ATR by Vpr were done with lentiviral vectors which might therefore activate ATR to some extent independently of Vpr. In addition, it was noticed that activation of ATR and CHK1 only accounts for only part of the G2 arrest induced by Vpr112. Other as yet unidentified molecular mechanism(s) may explain at least half of the G2 cell population induced by Vpr. Interestingly, Roshal et al119 showed that treatment of Vpr-producing mammalian cells with caffeine completely blocks Vpr-induced G2 arrest. Caffeine is part of the methylxanthine family, and similar to the caffeine effect, another methylxanthine pentoxifylline (PTX) also inhibits Vpr-induced G2 arrest in mammalian cells34. Similarly, both PTX and caffeine suppress Vpr-induced G2 arrest in fission yeast117,120. Since PTX or caffeine inhibits Vpr-induced G2 arrest in fission yeast where the classic DNA checkpoints apparently play no role, these observations suggest molecular mechanism other than the classic DNA checkpoints may be involved in activation of ATR and regulation of Cdc25 and Wee1. The additional molecular mechanism might involve protein phosphatase 2A (PP2A). Although this protein phosphatase has no known role in the DNA checkpoints, it has an important role in Vpr-induced G2 arrest. Okadaic acid is a specific inhibitor of PP2A, and okadaic acid was shown to inhibit Vpr- induced G2 arrest both in human100 and fission yeast cells56. Further evidence for an important role of PP2A comes from PP2A mutant strains. PP2A is composed of three subunits, one catalytic (C) and two regulatory (A and B) subunits. When vpr was expressed in a strain with a deletion for a catalytic subunit (ppa2) or a regulatory subunit ( pab1) of PP2A, Vpr-induced G2 arrest was reduced 108,121. Other evidence supporting involvement of PP2A in Vpr-induced G2 arrest comes from other viral proteins with effects on cell cycle G2/M controls. PP2A appears to be a common viral target since other viruses such as simian virus 40 (SV40), polyoma virus, human T lymphotrophic retrovirus and adenovirus affect the enzymatic activity of at least a subset of PP2A proteins122. Even though these viruses are not otherwise related, they all seem to have adapted a similar strategy to affect cellular processes by direct interaction with PP2A. Similar to the Vpr effects, both adenoviral E4orf4123-125 and HTLV Tax protein induce cell cycle G2 arrest 118. These two viral proteins both bind to PP2A and affect its enzymatic activity123,126. Interestingly, similar to Vpr, Tax- induced G2 arrest is reversible by caffeine 118. Further examinations indicated that Tax binds to CHK2 in Jurkat T-cells 118 but it complexes with CHK1 in other T-cells127. Taken together, it is possible that a concerted cellular mechanism interlinks PP2A and ATR/CHK1 in the cellular response to vpr gene expression during the induction of G2 arrest. Nuclear transport of viral preintegration complex Ability to replicate in non-dividing cells (terminally differentiated macrophages and incompletely 277 activated CD4+ T lymphocytes) is the characteristic feature of HIV-1 which determines to a large extent its high replicative capacity and pathogenesis. To replicate in non-dividing cells, HIV-1 needs to transport its genomic DNA [in the context of the viral pre-integration complex (PIC)] from the cytoplasm into the nucleus of a target cell. Vpr is believed to be among the main regulators of HIV-1 nuclear import17,128. Proteins engaging in nuclear transport typically contain a classical nuclear localization sequence (cNLS)129,130, which is a short amino acid region rich in basic amino acids (lysines and arginines) and binds to the adaptor protein importin ? . The complex of cNLS-importin ? then binds to the receptor importin ? through the importin ? -binding region on importin ? . Importin ? interacts with components of the nuclear pore complex (NPC) as an essential part of the nuclear translocation process. The NPC is a large structure composed of 50 to 100 proteins called nucleoporins, which contains a central 10 nm aqueous channel through which proteins are actively transported. Directionality of this translocation process is ensured by Ran. A high concentration of RanGTP inside the nucleus stimulates binding of RanGTP to cNLS-importin ? -importin ? complex and disassembles it to release the protein carrying the cNLS into the nucleoplasm. Importin ? and importin ? -RanGTP are then exported out of the nucleus to be reused in another round of nuclear transport. This model for cNLS translocation is partially based on work done in budding yeast, and the high degree of conservation is demonstrated by functional complementation of many budding yeast mutants in nuclear transport proteins by human homologues131. Most studies have reported that Vpr expressed without other viral proteins localizes predominantly to the nuclear envelope in human, fission yeast and budding yeast20,132,133. Two hypotheses (not necessarily mutually exclusive) for the mode of action of Vpr in HIV-1 nuclear import have been proposed: (i) that Vpr targets the HIV-1 PIC to the nucleus via a distinct, importin-independent pathway134,135; or (ii) that Vpr modifies cellular importin-dependent import machinery136,137. The first model was based on the observation that in the in vitro nuclear import assay Vpr can enter nuclei in the absence of soluble import factors135. Consistent with this concept, Vpr was shown to induce dynamic disruptions in the nuclear envelope115 which may serve as entry points for isolated Vpr and for the PICs. However, based on our results and reports from several other laboratories, we favour a hypothesis that Vpr uses a modification of the importin ? pathway to enter the nucleus138,139. Vpr was shown to bind to importin ? both from human and budding yeast, but the binding site is different from the binding site for cNLS133,136,137. Vpr also binds to nucleoporins, thus performing activity normally attributed to importin ? 133,140,141, suggesting that Vpr might mediate binding of the PIC to the nuclear pore. However, another study showed that importin ? was necessary for nuclear localization of Vpr and that importin ? , importin ? , cNLS and Vpr form a ternary complex136. Thus, Vpr appears to bind to a previously unknown site on importin ? 142, and this complex in turn binds to importin ? which mediates transport through the nuclear pore. The effect of Ran-GTP binding to importin ? on the ternary complex has not been reported and it is not understood why Vpr is frequently observed to localize at the nuclear envelope, although this may be related to described above binding of Vpr to nucleoporins. One study found Vpr to be at the inside of the nuclear envelope suggesting that Vpr is transported through the pore but then stops at the inside of the nuclear pore rather than being released into the nucleoplasm133. In the nuclear transport of the viral PIC, Vpr appears to cooperate with cNLSs present on other HIV-1 proteins, in particular matrix protein (MA). There has been disagreement over the role and identity of cNLS in the nuclear transport of PIC, and at least one reason for this disagreement is the presence of two cNLS in MA143. Deletion of either MA cNLS does not prevent nuclear translocation of PIC in macrophage cells144,145, but when both are deleted, nuclear transport of the PIC no longer occurs even when Vpr is present143. Thus, Vpr does not function as an independent nuclear transport factor. However, when one or both MA cNLS are present in the PIC, nuclear transport occurs efficiently only in the presence of Vpr133,136,137,146, suggesting Vpr is ZHAO et al: HIV-1 VPR 278 INDIAN J MED RES, APRIL 2005 an important player in the nuclear transport of PIC. Vpr increases the binding of the MA cNLSs to importin ? 137, and this increased affinity may at least in part account for Vpr’s activity in HIV-1 nuclear transport. In general, the major role of Vpr may be to stimulate the nuclear import of unusually large complexes carrying relatively weak NLSs, as the effect of Vpr on nuclear import of the model substrate, serum albumin coupled to cNLS, decreased with an increase in the number of coupled cNLS copies 136. One interesting implication of a conserved Vpr- binding site present both on human and budding yeast importin ? is that this binding site plays some important cellular function in nuclear transport and that a cellular protein may bind to this site. Agostini et al142 have identified this cellular protein as Hsp70, a highly conserved heat shock protein, which competes with Vpr for binding to importin ? . Hsp70 can in fact replace Vpr in the nuclear transport of PIC and similar to Vpr also strengthens the binding of MA cNLS to importin ? 142. One cellular role of this Vpr/Hsp70 binding site thus appears to be in strengthening the interaction of a weak cNLS with importin ? . Therefore, one possible role for Hsp70 may be to stimulate efficient translocation of large complexes through the nuclear pore, similar to the role of Vpr in the nuclear import of the HIV-1 PIC. Induction of apoptosis A major pathway for the induction of apoptosis by Vpr is through the mitochondria. This intrinsic pathway for apoptosis is initiated by mitochondrial outer membrane permeabilization (MOMP)147. The release of proteins from the space between the inner and outer mitochondrial membranes ultimately leads to apoptosis. Cytochrome c is particularly important in this process since it combines in the cytoplasm with Apaf-1 to activate procaspase 9, the initiating caspase for the intrinsic pathway. Activated caspase 9 in turn activates the downstream caspases such as caspase 3 which carry out many of the apoptotic events147. Vpr is thought to lead to MOMP by virtue of binding to ANT (adenine nucleotide transporter) protein of the inner mitochondrial membrane148-150. After crossing the outer mitochondrial membrane, possibly through VDA (the Voltage Dependent Anion Channel), binding of Vpr to ANT leads to depolarization of the inner mitochrondrial membrane, swelling of the inner mitochondria and ultimately MOMP with release of the apoptosis factors. Among the considerable evidence supporting this model are depolarization of the inner mitochondrial membrane by Vpr in intact cells, depolarization of isolated mitoZHAO et al: HIV-1 VPRchondria by purified Vpr, strong binding between Vpr and ANT shown by several methods, reduced cell killing when ANT levels are decreased148-150 and activation of caspase 9 by Vpr151,152. While there is considerable evidence supporting the idea that Vpr induces apoptosis through MOMP, there are other reports that do not readily fit into this model and which raise the possibility that Vpr kills cells through other redundant pathways. The localization of Vpr raises one question about the MOMP model since Vpr has been consistently reported to be in the nucleus or at the nuclear membrane20,53,132,133,153,154 rather than in the mitochondria150. It may be that only a small fraction of Vpr molecules localizes to the mitochondria, which is sufficient to induce apoptosis and methods used to visualize Vpr may have overlooked this small amount. However, the predominant nuclear localization of Vpr and the association of nuclear localization with cell killing in Vpr mutants20,154 suggest that Vpr located in the nucleus may sometimes play a role in initiating cell killing. Other observations seemingly inconsistent with the MOMP model concern the activation of caspases by Vpr. While activation of caspase-9 with no activation of caspase-8 supports the role of MOMP in the induction of apoptosis by Vpr151, there have been other conflicting reports that Vpr does activate capase-864,155. Caspase-8 activation is thought to be a hallmark of the extrinsic pathway for apoptosis induction by death receptors such as FAS and TNF 156. It has also been reported that a fragment of Vpr is able to induce cell death without caspase activation157, and even that Vpr induces a necrotic type of cell death in neurons158. The observation that Vpr is able to kill fission yeast cells56, where caspases play at most a minor role in cell death159, also suggests that there may be a caspase- and mitochondria- independent pathway for cell killing by Vpr. 279 A prospective role of HIV-1 in future antiviral therapy Evidence described in this review suggests that Vpr plays a pivotal role in viral pathogenesis and disease progression. Specifically Vpr activities are linked to promotion of viral infection in non-dividing macrophages and monocytes, activation of viral transcription and replication, and depletion of CD4 T-lymphocytes. Therefore, strategies that can be used to inhibit these adverse Vpr effects could potentially alleviate the impact of the virus and benefit infected patients. It is desirable then to identify Vpr-specific inhibitors for the design of future anti-Vpr regimens. A number of hexameric peptides with a di-tryptophan motif were found by genetic selection in budding yeast to suppress G2 arrest and apoptosis in T-cells160. So far a fission yeast small heat shock protein 16 (Hsp16) is the only protein known to specifically block all Vpr activities27. However, its utility in suppressing Vpr- mediated viral infection in macrophages/monocytes or its potential beneficial effect for HIV-infected patients has yet to be determined. Vpr is a virion-associated protein that is packaged in the viral particle in quantities similar to those of the major structural proteins132. This unique property of Vpr could potentially be used to target antiviral proteins to the virus. A proposed strategy is to create a chimeric protein that fuses Vpr with a functional protein capable of interfering with the progeny virion production161,162. Various Vpr fusion proteins have thus far been created. In all case, Vpr fusion proteins are efficiently incorporated into the viral particles, confirming feasibility of Vpr-mediated virion incorporation. Since HIV-1 protease is required for the processing of HIV-1 viral precursor proteins in order to generate infectious viral particles, (Gag and Gag-Pol), one approach is to inhibit viral maturation by sequestering protease162, i.e., to use a Vpr fusion protein containing the HIV-1 protease cleavage sites. A chimeric Vpr containing the cleavage sequences from the junction of p24 and p2 completely abolished viral infectivity162. A similar approach was used in targeting dimerization of the protease163. A four amino acid peptide from the C-terminal end of the protease, which is the dimmer interface region, was fused to Vpr. Incorporation of this chimeric protein results in reduced viral infectivity163. The HIV-1 integrase was targeted by fusing Vpr with a single-chain antibody (scAb) or a single-chain variable antibody fragments (SFvs) against the integrase164,165. Suppression of progeny virion production was observed with both Vpr fusions164,165. These results indicate that the inactivation of progeny virions by the use of chimeric Vpr fusion proteins represents a promising strategy in developing future antiviral therapies. The unique properties of Vpr could also be used for other therapeutic purposes. The ability of Vpr to arrest the cell cycle in G2 provides an opportunity to explore its potential as a cell cycle ”G2 blocker” to inhibit growth of the cancerous cells. Furthermore, since mammalian cells residing in the G2 phase of the cell cycle are most sensitive to radiation, Vpr- induced G2 arrest could conceivably sensitize tumour cells to radiation. The combination of gene and radiotherapies based on Vpr and the radiosensitivity of G2 cells may lead to the development of new methods to treat cancer. Several earlier studies have shown that expression of vpr can effectively suppress tumour growth in vitro and in vivo166-169. Vpr also induces apoptosis in various tumour cells regardless of p53 status166,168. Interestingly, one report shows that Vpr not only reduces tumor growth in a mouse but also triggers sufficient immune responses that the mouse is protected against a second oncogenic challenge170. Because Vpr is capable of suppressing various host inflammatory responses, it also could potentially be used as a therapeutic tool for treating diseases with hyperinflammatory conditions such as rheumatoid arthritis, sepsis and Crohn’s disease35. This strategy may have advantages over other conventional anti- inflammatory drugs since Vpr specifically inhibits pro- inflammatory cytokines such as TNF? and IL-12. Moreover, it limits chemokine production and suppresses NF? B-mediated transcription36,37. In addition, Vpr was shown to be more effective in reducing inflammatory cytokines than some of the commonly prescribed anti-inflammatory drugs35 indicating that the use of Vpr as an anti-inflammatory agent merits further investigation. ZHAO et al: HIV-1 VPR 280 INDIAN J MED RES, APRIL 2005 Acknowledgment Authors acknowledge the financial support received from the National Institutes of Health, USA. References 1. Rodman TC, Sullivan JJ, Bai X, Winston R. The human uniqueness of HIV: innate immunity and the viral Tat protein. Hum Immunol 1999; 60 : 631-9. 2. Sheehy AM, Gaddis NC, Choi JD, Malim MH. Isolation of a human gene that inhibits HIV-1 infection and is suppressed by the viral Vif protein. Nature 2002; 418 : 646-50. 3. Leulier F, Marchal C, Miletich I, Limbourg-Bouchon B, Benarous R, Lemaitre B. Directed expression of the HIV- 1 accessory protein Vpu in Drosophila fat-body cells inhibits Toll-dependent immune responses. EMBO Rep 2003; 4 : 976-81. 4. Harris RS, Bishop KN, Sheehy AM, Craig HM, Petersen- Mahrt SK, Watt IN, et al. DNA deamination mediates innate immunity to retroviral infection. Cell 2003; 113 : 803-9. 5. Zhang H, Yang B, Pomerantz RJ, Zhang C, Arunachalam SC, Gao L. The cytidine deaminase CEM15 induces hypermutation in newly synthesized HIV-1 DNA. Nature 2003; 424 : 94-8. 6. Mangeat B, Turelli P, Caron G, Friedli M, Perrin L, Trono D. Broad antiretroviral defence by human APOBEC3G through lethal editing of nascent reverse transcripts. Nature 2003; 424 : 99-103. 7. Mariani R, Chen D, Schrofelbauer B, Navarro F, Konig R, Bollman B, et al. Species-specific exclusion of APOBEC3G from HIV-1 virions by Vif. Cell 2003; 114 : 21-31. 8. Vartanian JP, Sommer P, Wain-Hobson S. Death and the retrovirus. Trends Mol Med 2003; 9 : 409-13. 9. Schrofelbauer B, Yu Q, Landau NR. New insights into the role of Vif in HIV-1 replication. AIDS Rev 2004; 6 : 34-9. 10. Levy DN, Refaeli Y, MacGregor RR, Weiner DB. Serum Vpr regulates productive infection and latency of human immunodeficiency virus type 1. Proc Natl Acad Sci USA 1994; 91 : 10873-7. 11. Tungaturthi PK, Sawaya BE, Singh SP, Tomkowicz B, Ayyavoo V, Khalili K, et al. Role of HIV-1 Vpr in AIDS pathogenesis: relevance and implications of intravirion, intracellular and free Vpr. Biomed Pharmacother 2003; 57 : 20-4. 12. Tristem M, Purvis A, Quicke DL. Complex evolutionary history of primate lentiviral vpr genes. Virology 1998; 240 : 232-7. 13. Tristem M, Marshall C, Karpas A, Hill F. Evolution of the primate lentiviruses: evidence from vpx and vpr. EMBO J 1992; 11 : 3405-12. 14. Schuler W, Wecker K, de Rocquigny H, Baudat Y, Sire J, Roques BP. NMR structure of the (52-96) C-terminal domain of the HIV-1 regulatory protein Vpr: molecular insights into its biological functions. J Mol Biol 1999; ZHAO et al: HIV-1 VPR285 : 2105-17. 15. Wecker K, Roques BP. NMR structure of the (1-51) N-terminal domain of the HIV-1 regulatory protein Vpr. Eur J Biochem 1999; 266 : 359-69. 16. Morellet N, Bouaziz S, Petitjean P, Roques BP. NMR structure of the HIV-1 regulatory protein Vpr. J Mol Biol 2003; 327 : 215-27. 17. Heinzinger N, Bukinsky M, Haggerty S, Ragland A, Kewalramani V, Lee M, et al. The Vpr protein of human immunodeficiency virus type 1 influences nuclear localization of viral nucleic acids in nondividing host cells. Proc Natl Acad Sci USA 1994; 91 :7311-5. 18. He J, Choe S, Walker R, Di Marzio PD, Morgan DO, Landau NR. Human immunodeficiency virus type 1 viral protein R (Vpr) arrests cells in the G2 phase of the cell cycle by inhibiting p34cdc2 activity. J Virol 1995; 69 : 6705-11. 19. Stewart SA, Poon B, Jowett JB, Chen IS. Human immunodeficiency virus type 1 Vpr induces apoptosis following cell cycle arrest. J Virol 1997; 71 : 5579-92. 20. Chen M, Elder RT, Yu M, O’Gorman MG, Selig L, Benarous R, et al. Mutational analysis of Vpr-induced G2 arrest, nuclear localization, and cell death in fission yeast. J Virol 1999; 73 : 3236-45. 21. Elder RT, Yu M, Chen M, Edelson S, Zhao Y. Cell cycle G2 arrest induced by HIV-1 Vpr in fission yeast (Schizosaccharomyces pombe) is independent of cell death and early genes in the DNA damage checkpoint. Viral Res 2000; 68 : 161-73. 22. Addo MM, Yu XG, Rosenberg ES, Walker BD, Altfeld M. Cytotoxic T-lymphocyte (CTL) responses directed against regulatory and accessory proteins in HIV-1 infection. DNA Cell Biol 2002; 21 : 671-8. 23. Altfeld M, Addo MM, Eldridge RL, Yu XG, Thomas S, Khatri A, et al. Vpr is preferentially targeted by CTL during HIV-1 infection. J Immunol 2001; 167 : 2743-52. 24. Mothe BR, Horton H, Carter DK, Allen TM, Liebl ME, Skinner P, et al. Dominance of CD8 responses specific for epitopes bound by a single major histocompatibility complex class I molecule during the acute phase of viral infection. J Virol 2002; 76 : 875-84. 281 25. Brenner BG, Tao Y, Pearson E, Remer I, Wainberg MA. Altered constitutive and stress-regulated heat shock protein 27 expression in HIV type 1-infected cell lines. AIDS Res Hum Retroviruses 1995; 11 : 713-7. 26. Wainberg Z, Oliveira M, Lerner S, Tao Y, Brenner BG. Modulation of stress protein (hsp27 and hsp70) expression in CD4+ lymphocytic cells following acute infection with human immunodeficiency virus type-1. Virology 1997; 233 : 364-73. 27. Benko Z, Liang D, Agbottah E, Hou J, Chiu K, Yu M, et al. Anti-Vpr activity of a yeast chaperone protein. J Virol 2004; 78 : 11016-29. 28. Iordanskiy S, Zhao Y, DiMarzio P, Agostini I, Dubrovsky L, Bukrinsky M. Heat-shock protein 70 exerts opposing effects on Vpr-dependent and Vpr-independent HIV-1 replication in macrophages. Blood 2004; 104 : 1867-72. 29. Iordanskiy S, Zhao Y, Dubrovsky L, Iordanskaya T, Chen M, Liang D, et al. Heat shock protein 70 protects cells from cell cycle arrest and apoptosis induced by human immunodeficiency virus type 1 viral protein R. J Virol 2004; 78 : 9697-704. 30. Wainberg Z, Oliveira M, Lerner S, Tao Y, Brenner BG. Modulation of stress protein (hsp27 and hsp70) expression in CD4+ lymphocytic cells following acute infection with human immunodeficiency virus type-1. Virology 1997; 233 : 364-73. 31. Zhao Y, Benko Z, Liang D, Hou J, Chiu K, Agbottah E, et al. Small heat shock proteins as innate antiviral factors counteracted by HIV-1 viral protein R. Paper presented at: The Eleventh Conference on Retroviruses and Opportunistic Infections.; October, 2004; San Francisco. D132. 32. Ayyavoo V, Muthumani K, Kudchodkar S, Zhang D, Ramanathan P, Dayes NS, et al. HIV-1 viral protein R compromises cellular immune function in vivo. Int Immunol 2002; 14 : 13-22. 33. Abimiku AG, Franchini G, Aldrich K, Myagkikh M, Markham P, Gard E, et al. Humoral and cellular immune responses in rhesus macaques infected with human immunodeficiency virus type 2. AIDS Res Hum Retroviruses 1995; 11 : 383-93. 34. Poon B, Grovit-Ferbas K, Stewart SA, Chen ISY. Cell cycle arrest by Vpr in HIV-1 virions and insensitivity to antiretroviral agents. Science 1998; 281 : 266-9. 35. Muthumani K, Desai BM, Hwang DS, Choo AY, Laddy DJ, Thieu KP, et al. HIV-1 Vpr and anti- inflammatory activity. DNA Cell Biol 2004; 23 : 239-47. 36. Refaeli Y, Levy DN, Weiner DB. The glucocorticoid receptor type II complex is a target of the HIV-1 vpr gene product. Proc Natl Acad Sci USA 1995; 92 : 3621-5. 37. Ayyavoo V, Mahboubi A, Mahalingam S, Ramalingam R, Kudchodkar S, Williams MV, et al. HIV-1 Vpr suppresses immune activation and apoptosis through regulation of nuclear factor kappa B. Nat Med 1997; 3 : 1117-23. 38. Sharp PM, Bailes E, Stevenson M, Emerman M, Hahn BH. Gene acquisition in HIV and SIV [letter]. Nature 1996; 383 : 586-7. 39. Fletcher TM, 3rd, Brichacek B, Sharova N, Newman MA, Stivahtis G, Sharp PM, et al. Nuclear import and cell cycle arrest functions of the HIV-1 Vpr protein are encoded by two separate genes in HIV-2/SIV(SM). EMBO J 1996; 15 : 6155-65. 40. Lang SM, Weeger M, Stahl-Hennig C, Coulibaly C, Hunsmann G, Muller J, et al. Importance of vpr for infection of rhesus monkeys with simian immunodeficiency virus. J Virol 1993; 67 : 902-12. 41. Gibbs JS, Lackner AA, Lang SM, Simon MA, Sehgal PK, Daniel MD, et al. Progression to AIDS in the absence of a gene for vpr or vpx. J Virol 1995; 69 : 2378-83. 42. Hoch J, Lang SM, Weeger M, Stahl-Henning C, Coulibaly C, Dittmer U, et al. vpr deletion mutant of simian immunodeficiency virus induces AIDS in rhesus monkeys. J Virol 1995; 69 : 4807-13. 43. Desrosiers RC. Prospects for live attenuated HIV [letter]. Nat Med 1998; 4 : 982. 44. Goh WC, Rogel ME, Kinsey CM, Michael SF, Fultz PN, Nowak MA, et al. HIV-1 Vpr increases viral expression by manipulation of the cell cycle: a mechanism for selection of Vpr in vivo. Nat Med 1998; 4 : 65-71. 45. Kuiken C, Foley B, Hahn BH, Marx P, McCutchan F, Mellors J, et al. HIV sequence compendium, 2nd ed. Los Alamos National Laboratory; 2000. 46. Levy JA. HIV and the Pathogenesis of AIDS. 2nd ed. Washington DC: ASM Press; 1998. 47. Huang Y, Zhang L, Ho DD. Characterization of gag and pol sequences from long-term survivors of human immunodeficiency virus type 1 infection. Virology 1998; 240 : 36-49. 48. Michael NL, Chang G, d’Arcy LA, Ehrenberg PK, Mariani R, Busch MP, et al. Defective accessory genes in a human immunodeficiency virus type 1-infected long- term survivor lacking recoverable virus. J Virol 1995; 69 : 4228-36. 49. Wang B, Ge Y, Palasanthiran P, Xiang S, Ziegler J, Dwyer D, et al. Gene defects clustered at the C-terminus of the vpr gene of HIV-1 in long-term nonprogressing mother and child pair: in vivo evolution of vpr quasispecies in blood and plasma. Virology 1996; 223 : 224-32. ZHAO et al: HIV-1 VPR 282 INDIAN J MED RES, APRIL 2005 50. Yedavalli VR, Ahmad N. Low conservation of functional domains of HIV type 1 vif and vpr genes in infected mothers correlates with lack of vertical transmission. AIDS Res Hum Retroviruses 2001; 17 : 911-23. 51. Zhang L, Huang Y, Yuan H, Tuttleton S, Ho DD. Genetic characterization of vif, vpr, and vpu sequences from long- term survivors of human immunodeficiency virus type 1 infection. Virology 1997; 228 : 340-9. 52. Chen M, Elder RT, Yu M, O’Gorman M, Selig L, Benerous R, et al. Mutational analysis of Vpr-induced G2 arrest, nuclear localization and cell death in fission yeast. J Virol 1999; 73 : 3236-45. 53. Di Marzio P, Choe S, Ebright M, Knoblauch R, Landau NR. Mutational analysis of cell cycle arrest, nuclear localization and virion packaging of human immunodeficiency virus type 1 Vpr. J Virol 1995; 69 : 7909-16. 54. Yao XJ, Subbramanian RA, Rougeau N, Boisvert F, Bergeron D, Cohen EA. Mutagenic analysis of human immunodeficiency virus type 1 Vpr: role of a predicted N-terminal alpha-helical structure in Vpr nuclear localization and virion incorporation. J Virol 1995; 69 : 7032-44. 55. Zhou Y, Ratner L. A novel inducible expression system to study transdominant mutants of HIV-1 vpr. Virology 2001; 287 : 133-42. 56. Zhao Y, Cao J, O’Gorman MRG, Yu M, Yogev R. Effect of human immunodeficiency virus Type 1 protein R (vpr) gene expression on basic cellular functions of fission yeast Schizosaccharomyces pombe. J Virol 1996; 70 : 5821-6. 57. Zhao Y, Yu M, Chen M, Elder RT, Yamamoto A, Cao J. Pleiotropic effects of HIV-1 protein R (Vpr) on morphogenesis and cell survival in fission yeast and antagonism by pentoxifylline. Virology 1998; 246 : 266-76. 58. Elder RT, Benko Z, Zhao Y. HIV-1 VPR modulates cell cycle G2/M transition through an alternative cellular mechanism other than the classic mitotic checkpoints. Front Biosci 2002; 7 : 349-57. 59. Zhao Y, Elder RT. Yeast perspectives on HIV-1 VPR. Front Biosci 2000; 5 : 905-16. 60. Bukrinsky M, Zhao Y. Heat-shock proteins reverse the G2 arrest caused by HIV-1 viral protein R. DNA Cell Biol 2004; 23 : 223-5. 61. Zhao Y, Chen M, Wang B, Yang J, Elder RT, Song XQ, et al. Functional conservation of HIV-1 Vpr and variability in a mother-child pair of long-term non-progressors. Virus Res 2002; 89 : 103-21. 62. Chen M, Yang J, Wang B, Elder RT, Song X-Q, Saksena N, et al. Functional conservation of HIV-1 viral protein R (Vpr) and its association with disease progression of HIV-infected patients. Paper presented at: The Eighth Conference on Retroviruses and Opportunistic Infections, 2001; Chicago, 17 : 138. 63. Somasundaran M, Sharkey M, Brichacek B, Luzuriaga K, Emerman M, Sullivan JL, et al. Evidence for a cytopathogenicity determinant in HIV-1 Vpr. Proc Natl Acad Sci USA 2002; 99 : 9503-8. 64. Lum JJ, Cohen OJ, Nie Z, Weaver JG, Gomez TS, Yao XJ, et al. Vpr R77Q is associated with long-term nonprogressive HIV infection and impaired induction of apoptosis. J Clin Invest 2003; 111 : 1547-54. 65. Zhao Y. Association overstated (letter to the editor). Re: Vpr R77Q is associated with long-term nonprogressive HIV infection and impaired induction of apoptosis. J Clin Invest 2003; 111 : 1547-54. 66. Fischer A, Lejczak C, Lambert C, Roman F, Servais J, Karita E, et al. Is the Vpr R77Q mutation associated with long-term non-progression of HIV infection? AIDS 2004; 18 : 1346-7. 67. Boddy M, Furnari B, Mondesert O, Russell P. Replication checkpoint enforced by kinases Cds1 and Chk1. Science 1998; 280 : 909-12. 68. Caspari T, Carr AM. DNA structure checkpoint pathways in Schizosaccharomyces pombe. Biochimie 1999; 81 : 173-81. 69. Elledge SJ. Cell cycle checkpoints: preventing an identity crisis. Science 1996; 274 : 1664-72. 70. Rhind N, Russell P. Mitotic DNA damage and replication checkpoints in yeast. Curr Opin Cell Biol 1998; 10 : 749-58. 71. Morgan DO. Principles of CDK regulation. Nature 1995; 374 : 131-4. 72. Gould KL, Nurse P. Tyrosine phosphorylation of the fission yeast cdc2+ protein kinase regulates entry into mitosis. Nature 1989; 342 : 39-45. 73. Krek W, Nigg EA. Differential phosphorylation of vertebrate p34cdc2 kinase at the G1/S and G2/M transitions of the cell cycle: identification of major phosphorylation sites. EMBO J 1991; 10 : 305-16. 74. Norbury C, Blow J, Nurse P. Regulatory phosphorylation of the p34cdc2 protein kinase in vertebrates. EMBO J 1991; 10 : 3321-9. 75. Nurse P. Checkpoint pathways come of age. Cell 1997; 91 : 865-7. 76. Rhind N, Russell P. Tyrosine phosphorylation of cdc2 is required for the replication checkpoint in Schizosaccharomyces pombe. Mol Cell Biol 1998; 18 : 3782-7. 283 77. Walworth NC, Bernards R. rad-dependent response of the chk1-encoded protein kinase at the DNA damage checkpoint. Science 1996; 271 : 353-6. 78. Burtelow MA, Roos-Mattjus PM, Rauen M, Babendure JR, Karnitz LM. Reconstitution and molecular analysis of the hRad9-hHus1-hRad1 (9-1-1) DNA damage responsive checkpoint complex. J Biol Chem 2001; 276 : 25903-9. 79. Carr AM. DNA structure dependent checkpoints as regulators of DNA repair. DNA Repair 2002; 1 : 983-94. 80. Caspari T, Carr AM. Checkpoints: how to flag up double- strand breaks. Curr Biol 2002; 12 : R105-7. 81. Saka Y, Esashi F, Matsusaka T, Mochida S, Yanagida M. Damage and replication checkpoint control in fission yeast is ensured by interactions of Crb2, a protein with BRCT motif, with Cut5 and Chk1. Genes Dev Dec 1997; 11 : 3387-400. 82. Francesconi S, Smeets M, Grenon M, Tillit J, Blaisonneau J, Baldacci G. Fission yeast chk1 mutants show distinct responses to different types of DNA damaging treatments. Genes Cells 2002; 7 : 663-73. 83. Esashi F, Mochida S, Matsusaka T, Obara T, Ogawa A, Tamai K, et al. Establishment of and recovery from damage checkpoint requires sequential interactions of Crb2 with protein kinases Rad3, Chk1, and Cdc2. Cold Spring Harb Symp Quant Biol 2000; 65 : 443-9. 84. Furnari B, Rhind N, Russell P. Cdc25 mitotic inducer targeted by Chk1 DNA damage checkpoint kinase. Science 1997; 277 : 1495-7. 85. Lopez-Girona A, Furnari B, Mondesert O, Russell P. Nuclear localization of Cdc25 is regulated by DNA damage and a 14-3-3 protein. Nature 1999; 397 : 172-5. 86. Baber-Furnari BA, Rhind N, Boddy MN, Shanahan P, Lopez-Girona A, Russell P. Regulation of mitotic inhibitor Mik1 helps to enforce the DNA damage checkpoint. Mol Biol Cell 2000; 11 : 1-11. 87. al-Khodairy F, Carr AM. DNA repair mutants defining G2 checkpoint pathways in Schizosaccharomyces pombe. EMBO J 1992; 11 : 1343-50. 88. Alcasabas AA, Osborn AJ, Bachant J, Hu F, Werler PJ, Bousset K, et al. Mrc1 transduces signals of DNA replication stress to activate Rad53. Nat Cell Biol 2001; 3 : 958-65. 89. Tanaka K, Russell P. Mrc1 channels the DNA replication arrest signal to checkpoint kinase Cds1. Nat Cell Biol 2001; 3 : 966-972. 90. Zeng Y, Forbes KC, Wu Z, Moreno S, Piwnica-Worms H, Enoch T. Replication checkpoint requires phosphorylation of the phosphatase Cdc25 by Cds1 or Chk1. Nature 1998; 395 : 507-10. 91. Shiloh Y. ATM and ATR: networking cellular responses to DNA damage. Curr Opin Genet Dev 2001; 11 : 71-7. 92. Abraham RT. Cell cycle checkpoint signaling through the ATM and ATR kinases. Genes Dev 2001; 15 : 2177-96. 93. Sanchez Y, Wong C, Thoma RS, Richman R, Wu Z, Piwnica- Worms H, et al. Conservation of the Chk1 checkpoint pathway in mammals: linkage of DNA damage to Cdk regulation through Cdc25. Science 1997; 277 : 1497-501. 94. Cans C, Ducommun B, Baldin V. Proteasome-dependent degradation of human CDC25B phosphatase. Mol Biol Rep 1999; 26 : 53-7. 95. Lammer C, Wagerer S, Saffrich R, Mertens D, Ansorge W, Hoffmann I. The cdc25B phosphatase is essential for the G2/M phase transition in human cells. J Cell Sci 1998; 111 : 2445-53. 96. Mils V, Baldin V, Goubin F, Pinta I, Papin C, Waye M, et al. Specific interaction between 14-3-3 isoforms and the human CDC25B phosphatase. Oncogene 2000; 19 : 1257-65. 97. Bartek J, Lukas J. Pathways governing G1/S transition and their response to DNA damage. FEBS Lett 2001; 490 : 117-22. 98. Taylor WR, Stark GR. Regulation of the G2/M transition by p53. Oncogene 2001; 20 : 1803-15. 99. Graves PR, Lovly CM, Uy GL, Piwnica-Worms H. Localization of human Cdc25C is regulated both by nuclear export and 14-3-3 protein binding. Oncogene 2001; 20 : 1839-51. 100. Re F, Braaten D, Franke EK, Luban J. Human immunodeficiency virus type 1 Vpr arrests the cell cycle in G2 by inhibiting the activation of p34cdc2-cyclin B. J Virol 1995; 69 : 6859-64. 101. Elder RT, Yu M, Chen M, Edelson S, Zhao Y. Cell cycle G2 arrest induced by HIV-1 vpr in fission yeast (Schizosaccharomyces pombe) is independent of cell death and early genes in the DNA damage checkpoint. Virus Res 2000; 68 : 161-73. 102. Elder RT, Yu M, Chen M, Zhu X, Yanagida M, Zhao Y. HIV-1 Vpr induces cell cycle G2 arrest in fission yeast (Schizosaccharomyces pombe) through a pathway involving regulatory and catalytic subunits of PP2A and acting on both Wee1 and Cdc25. Virology 2001; 287 : 359-70. 103. Bartz SR, Rogel ME, Emerman M. Human immunodeficiency virus type 1 cell cycle control: Vpr is cytostatic and mediates G2 accumulation by a mechanism which differs from DNA damage checkpoint control. J Virol 1996; 70 : 2324-31. ZHAO et al: HIV-1 VPR 284 INDIAN J MED RES, APRIL 2005 104. Yuan H, Kamata M, Xie YM, Chen IS. Increased levels of Wee-1 kinase in G(2) are necessary for Vpr- and gamma irradiation-induced G(2) arrest. J Virol 2004; 78 : 8183-90. 105. Munoz MJ, Bejarano ER, Daga RR, Jimenez J. The identification of Wos2, a p23 homologue that interacts with Wee1 and Cdc2 in the mitotic control of fission yeasts. Genetics 1999; 153 : 1561-72. 106. Goh WC, Manel N, Emerman M. The human immunodeficiency virus Vpr protein binds Cdc25C: implications for G2 arrest. Virology 2004; 318 : 337-49. 107. Kino T, Pavlakis GN. Partner molecules of accessory protein Vpr of the human immunodeficiency virus type 1. DNA Cell Biol 2004; 23 : 193-205. 108. Masuda M, Nagai Y, Oshima N, Tanaka K, Murakami H, Igarashi H, et al. Genetic studies with the fission yeast Schizosaccharomyces pombe suggest involvement of wee1, ppa2, and rad24 in induction of cell cycle arrest by human immunodeficiency virus type 1 Vpr. J Virol 2000; 74 : 2636-46. 109. Matsuoka S, Huang M, Elledge SJ. Linkage of ATM to cell cycle regulation by the Chk2 protein kinase. Science 1998; 282 : 1893-7. 110. Bentley NJ, Holtzman DA, Flaggs G, Keegan KS, DeMaggio A, Ford JC, et al. The Schizosaccharomyces pombe rad3 checkpoint gene. EMBO J 1996; 15 : 6641-51. 111. Savitsky K, Bar-Shira A, Gilad S, Rotman G, Ziv Y, Vanagaite L, et al. A single ataxia telangiectasia gene with a product similar to PI-3 kinase. Science 1995; 268 : 1749-53. 112. Zhu Y, Roshal M, Li F, Blackett J, Planelles V. Upregulation of survivin by HIV-1 Vpr. Apoptosis 2003; 8 : 71-9. 113. Zimmerman ES, Chen J, Andersen JL, Ardon O, Dehart JL, Blackett J, et al. Human immunodeficiency virus type 1 Vpr-mediated G2 arrest requires Rad17 and Hus1 and induces nuclear BRCA1 and gamma-H2AX focus formation. Mol Cell Biol 2004; 24 : 9286-94. 114. Mansky LM. The mutation rate of human immunodeficiency virus type 1 is influenced by the vpr gene. Virology 1996; 222 : 391-400. 115. de Noronha CM, Sherman MP, Lin HW, Cavrois MV, Moir RD, Goldman RD, et al. Dynamic disruptions in nuclear envelope architecture and integrity induced by HIV-1 Vpr. Science 2001; 294 : 1105-8. 116. Shimura M, Tanaka Y, Nakamura S, Minermoto Y, Yamashita K, Hatake K, et al. Micronuclei formation and aneuploidy induced by vpr, an accessory gene of human immunodeficiency virus type 1. FASEB J 1999; 13 : 621-37. 117. Zhao Y, Yu M, Chen M, Elder RT, Yamamoto A, Cao J. Pleiotropic effects of HIV-1 protein R (Vpr) on morphogenesis and cell survival in fission yeast and antagonism by pentoxifylline. Virology 1998; 246 : 266-76. 118. Haoudi A, Daniels RC, Wong E, Kupfer G, Semmes OJ. Human T-cell leukemia virus-I tax oncoprotein functionally targets a subnuclear complex involved in cellular DNA damage-response. J Biol Chem 2003; 278 : 37736-44. 119. Roshal M, Kim B, Zhu Y, Nghiem P, Planelles V. Activation of the ATR-mediated DNA damage response by the HIV-1 viral protein R. J Biol Chem 2003; 278 : 25879-86. 120. Elder RT, Yu M, Chen M, Zhu X, Yanagida M, Zhao Y. HIV-1 Vpr induces cell cycle G2 arrest in fission yeast (Schizosaccharomyces pombe) through a pathway involving regulatory and catalytic subunits of PP2A and acting on both Wee1 and Cdc25. Virology 2001; 287 : 359-70. 121. Elder RT, Yu M, Chen M, Zhu X, Yanagida M, Zhao Y. HIV-1 Vpr induces cell cycle G2 arrest in fission yeast (Schizosaccharomyces pombe) through a pathway involving regulatory and catalytic subunits of PP2A and acting on both Wee1 and Cdc25. Virology 2001; 287 : 359-70. 122. Janssens V, Goris J. Protein phosphatase 2A: a highly regulated family of serine/threonine phosphatases implicated in cell growth and signalling. Biochem J 2001; 353 : 417-39. 123. Kornitzer D, Sharf R, Kleinberger T. Adenovirus E4orf4 protein induces PP2A-dependent growth arrest in Saccharomyces cerevisiae and interacts with the anaphase-promoting complex/cyclosome. J Cell Biol 2001; 154 : 331-44. 124. Roopchand DE, Lee JM, Shahinian S, Paquette D, Bussey H, Branton PE. Toxicity of human adenovirus E4orf4 protein in Saccharomyces cerevisiae results from interactions with the Cdc55 regulatory B subunit of PP2A. Oncogene 2001; 20 : 5279-90. 125. Shtrichman R, Sharf R, Barr H, Dobner T, Kleinberger T. Induction of apoptosis by adenovirus E4orf4 protein is specific to transformed cells and requires an interaction with protein phosphatase 2A. Proc Natl Acad Sci USA 1999; 96 : 10080-5. 126. Fu DX, Kuo YL, Liu BY, Jeang KT, Giam CZ. Human T- lymphotropic virus type I tax activates I-kappa B kinase by inhibiting I-kappa B kinase-associated serine/threonine protein phosphatase 2A. J Biol Chem 2003; 278 : 1487-93. 127. Park HU, Jeong JH, Chung JH, Brady JN. Human T-cell leukemia virus type 1 Tax interacts with Chk1 and attenuates DNA-damage induced G2 arrest mediated by Chk1. Oncogene 2004; 23 : 4966-74. 285 128. Connor RI, Chen BK, Choe S, Landau NR. Vpr is required for efficient replication of human immunodeficiency virus type-1 in mononuclear phagocytes. Virology 1995; 206 : 935-44. 129. Nakielny S, Dreyfuss G. Transport of proteins and RNAs in and out of the nucleus. Cell 1999; 99 : 677-90. 130. Wente SR. Gatekeepers of the nucleus. Science 2000; 288 : 1374-7. 131. Corbett AH, Silver PA. Nucleocytoplasmic transport of macromolecules. Microbiol Mol Biol Rev 1997; 61 : 193-211. 132. Lu YL, Spearman P, Ratner L. Human immunodeficiency virus type 1 viral protein R localization in infected cells and virions. J Virol 1993; 67 : 6542-50. 133. Vodicka MA, Koepp DM, Silver PA, Emerman M. HIV-1 Vpr interacts with the nuclear transport pathway to promote macrophage infection. Genes Dev 1998; 12 : 175-85. 134. Gallay P, Stitt V, Mundy C, Oettinger M, Trono D. Role of the karyopherin pathway in human immunodeficiency virus type 1 nuclear import. J Virol 1996; 70 : 1027-32. 135. Jenkins Y, McEntee M, Weis K, Greene WC. Characterization of HIV-1 vpr nuclear import: analysis of signals and pathways. J Cell Biol 1998; 143 : 875-85. 136. Popov S, Rexach M, Ratner L, Blobel G, Bukrinsky M. Viral protein R regulates docking of the HIV-1 preintegration complex to the nuclear pore complex. J Biol Chem 1998; 273 : 13347-52. 137. Popov S, Rexach M, Zybarth G, Reiling N, Lee MA, Ratner L, et al. Viral protein R regulates nuclear import of the HIV-1 pre-integration complex. EMBO J 1998; 17 : 909-17. 138. Bukrinsky MI, Haffar OK. HIV-1 nuclear import: matrix protein is back on center stage, this time together with Vpr. Mol Med 1998; 4 : 138-43. 139. Bukrinsky MI, Haffar OK. HIV-1 nuclear import: in search of a leader. Front Biosci 1999; 4 : 772-81. 140. Fouchier RA, Meyer BE, Simon JH, Fischer U, Albright AV, Gonzalez-Scarano F, et al. Interaction of the human immunodeficiency virus type 1 Vpr protein with the nuclear pore complex. J Virol 1998; 72 : 6004-13. 141. Le Rouzic E, Mousnier A, Rustum C, Stutz F, Hallberg E, Dargemont C, et al. Docking of HIV-1 Vpr to the nuclear envelope is mediated by the interaction with the nucleoporin hCG1. J Biol Chem 2002; 277 : 45091-8. 142. Agostini I, Popov S, Li J, Dubrovsky L, Hao T, Bukrinsky M. Heat-shock protein 70 can replace viral protein R of HIV-1 during nuclear import of the viral preintegration complex. Exp Cell Res 2000; 259 : 398-403. 143. Haffar OK, Popov S, Dubrovsky L, Agostini I, Tang H, Pushkarsky T, et al. Two nuclear localization signals in the HIV-1 matrix protein regulate nuclear import of the HIV-1 pre-integration complex. J Mol Biol 2000; 299 : 359-68. 144. Fouchier RA, Meyer BE, Simon JH, Fischer U, Malim MH. HIV-1 infection of non-dividing cells: evidence that the amino-terminal basic region of the viral matrix protein is important for Gag processing but not for post- entry nuclear import. EMBO J 1997; 16 : 4531-9. 145. Freed EO, Englund G, Martin MA. Role of the basic domain of human immunodeficiency virus type 1 matrix in macrophage infection. J Virol 1995; 69 : 3949-54. 146. Subbramanian RA, Kessous-Elbaz A, Lodge R, Forget J, Yao XJ, Bergeron D, et al. Human immunodeficiency virus type 1 Vpr is a positive regulator of viral transcription and infectivity in primary human macrophages. J Exp Med 1998; 187 : 1103-11. 147. Green DR, Kroemer G. The pathophysiology of mitochondrial cell death. Science 2004; 305 : 626-9. 148. Brenner C, Kroemer G. The mitochondriotoxic domain of Vpr determines HIV-1 virulence. J Clin Invest 2003; 111 : 1455-7. 149. Jacotot E, Ferri KF, El Hamel C, Brenner C, Druillennec S, Hoebeke J, et al. Control of mitochondrial membrane permeabilization by adenine nucleotide translocator interacting with HIV-1 viral protein rR and Bcl-2. J Exp Med 2001; 193 : 509-19. 150. Jacotot E, Ravagnan L, Loeffler M, Ferri KF, Vieira HL, Zamzami N, et al. The HIV-1 viral protein R induces apoptosis via a direct effect on the mitochondrial permeability transition pore. J Exp Med 2000; 191 : 33-46. 151. Muthumani K, Hwang DS, Desai BM, Zhang D, Dayes N, Green DR, et al. HIV-1 Vpr induces apoptosis through caspase 9 in T cells and peripheral blood mononuclear cells. J Biol Chem 2002; 277 : 37820-31. 152. Muthumani K, Zhang D, Hwang DS, Kudchodkar S, Dayes NS, Desai BM, et al. Adenovirus encoding HIV-1 Vpr activates caspase 9 and induces apoptotic cell death in both p53 positive and negative human tumor cell lines. Oncogene 2002; 21 : 4613-25. 153. Mahalingam S, Collman RG, Patel M, Monken CE, Srinivasan A. Functional analysis of HIV-1 Vpr: identification of determinants essential for subcellular localization. Virology 1995; 212 : 331-9. 154. Waldhuber MG, Bateson M, Tan J, Greenway AL, McPhee DA. Studies with GFP-Vpr fusion proteins: induction of apoptosis but ablation of cell-cycle arrest despite nuclear membrane or nuclear localization. Virology 2003; 313 : 91-104. ZHAO et al: HIV-1 VPR 286 INDIAN J MED RES, APRIL 2005 155. Patel CA, Mukhtar M, Pomerantz RJ. Human immunodeficiency virus type 1 Vpr induces apoptosis in human neuronal cells. J Virol 2000; 74 : 9717-26. 156. Barnhart BC, Peter ME. The TNF receptor 1: a split personality complex. Cell 2003; 114 : 148-50. 157. Roumier T, Vieira HL, Castedo M, Ferri KF, Boya P, Andreau K, et al. The C-terminal moiety of HIV-1 Vpr induces cell death via a caspase-independent mitochondrial pathway. Cell Death Differ 2002; 9 : 1212-9. 158. Huang MB, Weeks O, Zhao LJ, Saltarelli M, Bond VC. Effects of extracellular human immunodeficiency virus type 1 vpr protein in primary rat cortical cell cultures. J Neurovirol 2000; 6 : 202-20. 159. Madeo F, Herker E, Maldener C, Wissing S, Lachelt S, Herlan M, et al. A caspase-related protease regulates apoptosis in yeast. Mol Cell 2002; 9 : 911-7. 160. Yao XJ, Lemay J, Rougeau N, Clement M, Kurtz S, Belhumeur P, et al. Genetic selection of peptide inhibitors of human immunodeficiency virus type 1 Vpr. J Biol Chem 2002; 277 : 48816-26. 161. Wu X, Liu H, Xiao H, Kim J, Seshaiah P, Natsoulis G, et al. Targeting foreign proteins to human immunodeficiency virus particles via fusion with Vpr and Vpx. J Virol 1995; 69 : 3389-98. 162. Serio D, Rizvi TA, Cartas M, Kalyanaraman VS, Weber IT, Koprowski H, et al. Development of a novel anti- HIV-1 agent from within: effect of chimeric Vpr-containing protease cleavage site residues on virus replication. Proc Natl Acad Sci USA 1997; 94 : 3346-51. 163. Cartas M, Singh SP, Serio D, Rizvi TA, Kalyanaraman VS, Goldsmith CS, et al. Intravirion display of a peptide corresponding to the dimer structure of protease attenuates HIV-1 replication. DNA Cell Biol 2001; 20 : 797-805. 164. Okui N, Sakuma R, Kobayashi N, Yoshikura H, Kitamura T, Chiba J, et al. Packageable antiviral therapeutics against human immunodeficiency virus type 1: virion-targeted virus inactivation by incorporation of a single-chain antibody against viral integrase into progeny virions. Hum Gene Ther 2000; 11 : 537-46. 165. BouHamdan M, Kulkosky J, Duan LX, Pomerantz RJ. Inhibition of HIV-1 replication and infectivity by expression of a fusion protein, VPR-anti-integrase single- chain variable fragment (SFv): intravirion molecular therapies. J Hum Virol 2000; 3 : 6-15. 166. Mahalingam S, MacDonald B, Ugen KE, Ayyavoo V, Agadjanyan MG, Williams WV, et al. In vitro and in vivo tumor growth suppression by HIV-1 Vpr. DNA Cell Biol 1997; 16 : 137-43. 167. Toy EP, Rodriguez-Rodriguez L, McCance D, Ludlow J, Planelles V. Induction of cell-cycle arrest in cervical cancer cells by the human immunodeficiency virus type 1 viral protein R. Obstet Gynecol 2000; 95 : 141-6. 168. Stewart SA, Poon B, Jowett JB, Xie Y, Chen IS. Lentiviral delivery of HIV-1 Vpr protein induces apoptosis in transformed cells. Proc Natl Acad Sci USA 1999; 96 : 12039-43. 169. Yang J, Yi Y, Wang J, Hung T, Zhao Y. Cell cycle G2 arrest, cell death and nuclear localization induced by human immunodeficiency virus type 1 protein R (VPR) in cervical cancer cells. Chinese J Clin Exp Virol 2000; 14 : 223-6, 301. 170. Pang S, Kang MK, Kung S, Yu D, Lee A, Poon B, et al. Anticancer effect of a lentiviral vector capable of expressing HIV-1 Vpr. Clin Cancer Res 2001; 7 : 3567-73. Reprint requests: Dr Richard Y. Zhao, Department of Pathology, University of Maryland School of Medicine 10 South Pine Street, MSTF 600, Baltimore, MD 21201-1192, USA e-mail: RZhao@som.umaryland.edu