Thursday 12 July 2007

Editorial Review: The Role of Human Leukocyte Antigen E and G in HIV Infection



Piyush Tripathi; Suraksha Agrawal
AIDS. 2007;21(11):1395-1404. ©2007 Lippincott Williams & Wilkins
Posted 07/10/2007

Introduction
An important area of HIV research is the immune response and how HIV circumvents it to create a successful and chronic infection. Various studies have provided not only a basic understanding of 'how HIV invades' but also clues for the development of vaccines to fight against AIDS. Although HIV initially evokes an immune response, it later escapes and evades the immune system for a successful infection. Methods of escape from the immune response include rapid mutations altering the organization of cell surface receptors, alterations in the expression profile of human leukocyte antigens (HLA) and destruction of immune effector cells.
HIV infects through exchange of body fluids. The cells mainly infected by HIV are the T helper cells (CD4 T cells), dendritic cells and macrophages. This tropism is generated because HIV utilizes CD4 as a primary receptor plus a coreceptor: CCR5 (expressed on macrophages, dendritic cells and T cells) for the R5 HIV strain and CXCR4 (T cells) for the X4 strain.[1] At the early stages of infection, HIVR5 utilizing CCR5 predominates, whereas at the later stages HIVX4 using CXCR4 is mainly seen.[2,3]


In the early stages of infection, the foremost target is CD4 T cells.[4] These cells along with other putative targets harbour mature virus and be carried in the circulation to lymph nodes and lymphoid organs. Here, the virions continue to infect immune cells, preferentially CD4 cells,[5] in some more vigorous way as the density of target cells are higher at these places. This infection as well as the destruction of CD4 cells later on leads to a profound decrease in CD4 cell count. The sudden depletion in CD4 T cells is unlikely to be caused simply by direct viral-induced lysis as the number of cells infected initially may not be sufficient to account for the massive decrease observed.[6] It has been suggested that bystander apoptosis induced by viral antigens or cytokines,[7,8] and downregulation of CD4 receptor by viral HIV-negative effector (Nef) protein,[9,10] may be involved. Other studies have emphasized apoptosis mediated by CD95 (FAS) and CD95L [FAS ligand (FasL)], which, in turn, are stimulated by increased concentration of viral envelope protein gp120 during infection, as a mechanism to account for the preferential depletion of CD4 T cells.[11,12]


With the continuing decrease in CD4 T cells, there is an explosive increase in virus production, which then evokes and is resisted by cellular immune response. After a peak of viral concentration has been reached, a gradual decrease is observed. Though activated cytotoxic T cells (CTL) can partially check infection,[13] which is evident by the appearance by HIV-specific CTL, this counterattack does not eradicate HIV completely as replicating viruses can escape the CTL response by mutation of their activation markers,[14] and through other mechanisms of immune escape. Some studies have suggested that this decline in virus concentration may be because of 'substrate exhaustion', as it is followed by depletion of CD4 T cells,[15,16] which functions as a reservoir for viral dissemination.


Cellular Immune Response to HIV Infection
HIV infection and intrusion of viral particles are counterattacked by CTL-mediated immune responses. Though the cellular immune response fails to control HIV-1 infection completely in most infected individuals, its occurrence is evident in regulating viral load during infection. During acute infection, reduction in viral load coincides with the appearance of HIV-specific CTL,[17,18] and an inverse relationship is established between viral load and HIV-specific CTL.[19] The initial CTL response may be directed against a few epitopes, which subsequently broadens during prolonged antigen stimulation.[20]


CTL could also be expected to have a role during chronic HIV-1 infection as HIV-1-specific T cells remain at high frequency.[21,22] The high concentration of these T cells may result from continued antigenic stimulation. This observation is supported by the fact that there is a steady decline in CTL as viraemia is reduced by HAART.[23] However, in chronic infection without treatment, a high number of HIV-1-specific CTL is seen. Though the CTL response occurs in early as well as in later stages of infection, the epitopes targeted during acute infection often differ from those recognized during chronic infection.[20,24]


When CTL recognize self-HLA molecules loaded with foreign peptide, they activate Fas and secrete perforins and granzymes, which lyse target cells.[25] The CTL produces cytokines (interferon α and tumour necrosis factor α) that affect viral replication.[26] HIV-1-specific CTL also produce the CC chemokines macrophage inflammatory protein 1α and 1β and RANTES, which suppress HIV-1 replication.[27] Even with these various effector functions, CTL cannot completely check viral intrusion in the immune system.


Immune Escape of Cytotoxic T Cells
As CTL do not carry CD4, the main receptor for viral entry and infection, they are anticipated to be a major player in HIV regulation. CTL can use multiple effector mechanisms to regulate viral replication,[25,28] including lytic mechanisms and CC chemokine-mediated blockade of viral entry.[29,30] The existence of HIV-specific CTL and their successful involvement in protection against disease transmission confirms their importance in disease regulation.[31,32]
As CTL can pose a strong regulatory force against HIV, virions that can escape the CTL response have a selection advantage. HIV has a high mutational rate (1 in 105 bases,[33]) and so can produce many mutants, but only those mutants that do not cost in terms of viral fitness would be selected. These mutational escapes lead to failure of vaccines as well as of immune regulation, as escape variants do not generate specific CTL but keep on eliciting the proliferation of CTL specific for wild type.[34] Escape mutations can work through many mechanisms, including alteration of epitopes presented on HLA for T cell receptors, lack of antigen processing, absence of improper interaction with HLA and finally lack of recognition by T cell receptors. During HIV infection, selective pressure imposed by CTL leads to the generation of various escape mutations and these variants may constitute the majority of the total viral pool. It has been shown that the ratio of non-synonymous substitutions to synonymous substitutions was higher in the CTL epitope. This further confirms the role of CTL selection pressure for occurrence and then for maintenance of these mutations.[35] Later on, evidence of escape mutations in HLA-B8-restricted epitope in Nef, HLA-B44-restricted epitope in Env and HLA-B27-restricted Gag epitope KK10 have supported the CTL-mediated selection of these mutations.[36]


Nef-Mediated Downregulation of Major Histocompatibility Complex Class I
In addition to escape mutations, HIV has strategies that can make the infected cell undetectable by the immune system. The detection of any cell depends on cell surface markers and so effective strategies alter the organization and expression of such markers. To escape from CTL response, HIV inhibits surface expression of the host major histocompatibility complex (MHC) class I, which is most important for CTL recognition; this is achieved through a viral protein called Nef. HIV-1 Nef is a 27-34 kDa multifunctional protein that has no apparent enzymatic activity but functions as an adaptor protein that enters the cell membrane through amino-terminal myristoylation. Though the exact mechanism by which Nef disrupts MHC class I cell surface expression is not clear, the viral protein binds to the cytoplasmic tail of the class I protein and may disrupt class I trafficking.[37] The cytoplasmic domain of class I antigens has a highly conserved region of 33 amino acid residues with nine conserved serine residues; Nef protein interacts with this via amino-terminal α-helix, polyproline and acidic domains.[38]
It was initially thought that Nef reduced MHC class I cell surface expression by accelerating endocytosis and promoting retrograde transport of internalized class I molecules to the trans-Golgi network (TGN).[39] Nef protein can interact with phosphofurin acidic cluster sorting protein 1 and then can activate phosphatidylinositol 3-kinase,[40] guanine exchange factor ARNO and finally ADP ribosylation factor 6.[41] This pathway leads to internalization of MHC class I molecules to 'ADP ribosylation factor compartments' that finally reach the TGN. However, more recent work has shown that Nef disrupts transport of MHC class I in the secretory pathway to the cell surface, rather than causing endocytosis from the cell surface. Further, it has been demonstrated that adaptor protein 1 (AP-1) is necessary for Nef to disrupt class I trafficking.[42] The main function of AP-1 is to sort proteins at the TGN by binding their cytoplasmic tails to clathrin and directing them to endolysosomal pathways.[43] Nef-mediated disruption of class I surface expression may occur by allowing interaction between the cytoplasmic tail of an MHC class I molecule and AP-1, thus redirecting the molecules from the TGN to the lysosomes for degradation,[42] as shown in Fig. 1. Recent work by Kasper et al.,[44] has shown that Nef targets MHC class I in T cells early in the biosynthetic pathway by preferentially binding newly synthesized hypophosphorylated class I molecules. The preferential interaction of Nef prevents phosphorylation of these molecules and so also prevents them reaching the cell surface In summary, the work of Collins and coworkers,[42,44] has demonstrated that Nef preferentially binds hypophosphorylated class I molecules, thus preventing completion of the secretory pathway that would finally provide an antigen-presenting receptor on the cell surface to activate killing of the virus-infected cell. Transport of class I molecules from the cell surface to the TGN occurs normally in infected cells but the class I molecules are then diverted to lysosomes through Nef-assisted binding of AP-1 to their cytoplasmic tail; this further inhibits their phosphorylation as well as their surface expression.


Figure 1.
Probable mechanism of viral HIV-negative effector (Nef)-mediated major histocompatibility complex class I internalization that may fail with human leukocyte antigen (HLA) G. Nef and adaptor protein (AP) 1 interact with the cytoplasmic domain of the class I in the trans-Golgi network (TGN) and then redirect class I protein-containing vesicles to the endolysosomal pathway. The truncated cytoplasmic domain of HLA-G makes Nef-mediated recycling ineffective as Nef proteins cannot dock on HLA-G and so surface expression of HLA-G does not change during HIV infection.

HLA Genotype and Cytotoxic T Cells
The HLA antigens activate cellular responses by forming the antigen-presenting component on the cell surface that interacts with CTL, directs them against the infected cells and activates natural killer (NK) cells of the innate immune response by interacting with the killer cell immunoglobulin-like receptor (KIR) family of surface molecules. There is substantial evidence that immune responses are effective in challenging the infection and transmission of HIV disease.
Though various genetic factors have been associated with susceptibility to HIV ( Table 1 ), investigations of the role of HLA antigens has concentrated on three areas: zygosity of HLA loci, sharing of alleles, and specific HLA allelic/haplotypic association with the outcome of disease. It has been shown that homozygosity at the class I loci is associated with relatively rapid progression to disease compared with heterozygotes.[54] This heterozygote advantage probably stems from the ability of such individuals to present a wider array of virus-derived epitopes to a more diverse CTL repertoire. This heterozygous repertoire will not only enable recognition and destruction of a greater breadth of infectious agents but will also require many more escape mutations for effective avoidance of the CTL response. Hence, heterozygosity may be associated with delayed progression to AIDS.[50] However, it is also conceivable that virus may become adapted and resistant to highly frequent alleles more easily in that population, and so a rare allele may have selective advantage in HIV disease progression.[55] The rare allele selective advantage may work in conjunction with heterozygote advantage, as the protective rare alleles are more likely to be present as heterozygotes.
Another genetic component that predisposes to the progression of AIDS is HLA sharing. Where the MHC class I is common to the donor and recipient, the basis of successful transplantation, it would lead to increased susceptibility to viral infection. One natural model of viral transmission between HLA-sharing donor and recipient is mother-to-child transmission, which further supports increased transmission of HIV in these circumstances.[56] Further, significant increase in susceptibility to HIV has been shown to be associated with concordance at the HLA-B locus but not at HLA-A or HLA-C.[57]
Knowing that a certain viral escape mechanism is likely to develop under a particular genetic selection pressure, it can be anticipated that an escape variant well adapted to a particular genetic profile and then transmitted to a host of similar genetic set up would be able to escape immunological challenges in the new host also. This may be a mechanism for susceptibility to viral transmission in hosts with HLA alleles in common. By comparison, MHC class I disparity may induce anti-HLA antibodies on passage of the virus and so may prevent HIV infection at early stages. Such a defence would be lacking in HLA concordant individuals, increasing successful transmission of HIV virus.
Previous research in genetic predisposition to viral susceptibility in the context of HLA has concentrated on specific alleles. Various studies have confirmed the contribution of specific class I alleles and more particularly HLA-B alleles in the outcome of disease.[58] This remarkable contribution of HLA-B may be because this group has the highest diversity among the class I antigens: approximately 661 alleles compared with 372 in HLA-A and 190 alleles in HLA-C.[59] Further, substantially greater selection pressure would be imposed on HIV by HLA-B compared with other class I antigens. Consistent association with delayed disease progression has been seen with HLA-B*27 and HLA-B*57.[51] Though the HIV HLA-B*57-specific epitope 'TW 10' may undergo an escape mutation, T242N, under selective pressure, this may cost in terms of viral fitness as the virus reverts after transmission to a new host.[60] Another allele, HLA-B*35, has been implicated as the class I susceptibility allele for AIDS.[52] HLA-B*35 heterozygotes have a rapid progression to AIDS, and homozygotes progress twice as fast as HLA-B*35-negative individuals. The most deleterious effects of HLA-B*35 are seen with its two subtypes, HLA-B*3502 and B*3503, which have proline at anchor position 2 of their loaded peptide and non-tyrosine residue at position 9.[52] By comparison, HLA-B*3501 containing tyrosine at position 9 does not have any substantial effect on disease prognosis. While both HLA-B*35 subtypes can equally induce a CTL response, viral load was cleared less effectively by non-tyrosine-containing HLA-B*3502 and B*3503 compared with HLA-B*3501.[61] It may, therefore, be possible that altered epitope recognition by HLA-B*3502 and B*3503 will induce CTL that may not specifically function against HIV-1-infected cells.
Some HLA-B alleles have been shown to influence the outcome of disease progression by interacting with KIR on NK cells. The Bw4 motif (residues 79-84 of the α3 domain) of various HLA-Bw4 alleles may interact with activating receptors KIR3DS1 of NK cells, thus facilitating clearance of HIV-1-infected lymphocytes and slowing disease progression.[62]
Studies have also been performed to examine particular MHC class II genes, but no consistent effects have been revealed. One recent study implicated the DRβ1*13-DQβ1*06 haplotype in viral suppression during treatment.[63]
Role of HLA-G and HLA-E in Progression of HIV Disease
Among the myriad of mechanisms adopted by HIV to avoid the human immune response is interference with the expression of HLA antigens. One evasion strategy is to downregulate cell surface class I classical antigens (HLA-A and HLA-B) to avoid HIV-specific CTL responses. Normally any change in the self HLA profile of cells is easily detected by immune surveillance and such cells are then subjected to degradation. However, despite reduced expression of class I antigens, HIV-infected cells are resistant to lysis by NK cells. During viraemic HIV-1 infection, there is expansion of an anergic subset of NK cells that do not respond to stimulation with MHC-devoid target cells. These NK cells have increased expression of SHIP (SH2-containing inositol phosphatase), which may be responsible for the reduced functional activity of these cells in chronic HIV-1 infection.[64] Various NK cell receptors that recognize MHC-independent ligands can regulate key cytolytic NK functions. A recent study has demonstrated that these inhibitory receptors recognizing an MHC-independent ligand are overexpressed in SHIP knockout mice and, therefore, may regulate NK cell cytolytic activity. This would suggest that SHIP plays an important role in regulation of this MHC-independent inhibitory NK receptor repertoire, which, in turn, is crucial for NK recognition and cytolysis of various targets.[65] However, this immunoprotection could also be achieved by increased expression of HLA-G and HLA-E during HIV infection. These antigens are less polymorphic than their classical counterparts. Where HIV Nef downregulates surface class I antigens by interacting with their cytoplasmic domain,[66] it may not be able to interact with non-classical HLA-I antigens such as HLA-G, which has a truncated domain,[67] (Fig. 1). Apart from any effects of the shorter cytoplasmic tail in HLA-G, it has been speculated that various mechanisms may upregulate HLA-G and HLA-E. The impact of these non-classical class I antigens on susceptibility to HIV infection is supported by their immunoregulating properties (Fig. 2).
Figure 2.
Upregulation of non-classical major histocompatibility complex class I human leukocyte antigen (HLA) antigens G and E by HIV as a strategy of immunodownregulation. HIV-negative effector (Nef) involved in class I downregulation does not affect HLA-G expression. HIV, through increased interleukin 10 (IL-10) during infection, can upregulate HLA-G expression, and viral peptide p24 amino acids 14-22 can upregulate HLA-E expression. Enhanced HLA-G and HLA-E can regulate natural killer (NK) cells by interacting with their inhibitory receptors. HLA-G can also control T cell response to HIV by directing T cells to apoptosis. KIR, killer cell immunoglobulin-like receptor; NKG2A, an NK-activating receptor.

HIV and HLA-G With a Truncated Cytoplasmic Domain
HLA G was cloned by Geraghty et al. in 1987.[67] It is less polymorphic, as only 15 alleles are known to date. It has restricted tissue distribution compared with the classical class I antigens. Though initially HLA-G was implicated in the maintenance of tolerance during pregnancy, its role has been explored in the tumour escape mechanism in various cancers and also in organ transplantation.
The exon organization of HLA-G is similar to the classical class I molecules, with three external domains (α1, α2, α3), a transmembrane domain and a cytoplasmic domain, and it is associated with β2-microglobulin to make the complete structure.[68] But HLA-G is more peculiar as it possesses a premature stop codon in exon 6 that results in a truncated cytoplasmic tail (it translates 6 amino acids instead of 30).[67] HLA-G exists in multiple isoforms, created by alternative splicing.[69] Seven different HLA-G transcriptional isoforms have been described; four of these encode membrane-bound forms whereas the remaining three encode soluble isoforms. HLA-G is identified as an immunoregulatory molecule as it can interact with inhibitory KIR of NK cells. So far, three HLA-G specific KIR have been identified: ILT-2 (LIR-1), ILT-4 (LIR-2) and KIR2DL4 ( Table 2 ).[74,75] In addition to acting via the innate mechanisms, HLA-G also provides protection through acquired immunity. HLA-G5 induces apoptosis of activated CD8 cells through activation of the Fas/FasL pathway,[76] whereas HLA-G1 suppresses CD4 lymphocyte proliferation.[77] Interaction of HLA-G1 with KIR of T cells can inhibit the antigen-specific HLA-restricted CTL response,[76] thus confirming the functionality of HLA-G in protecting cells from all possible immune responses.
HIV infection is characterized by loss of HLA-A and HLA-B, but the expression of HLA-G remains unaffected or at least not decreased. Along with inability of viral Nef to downregulate HLA-G, there could be some changes indirectly influencing the expression of HLA-G, particularly increased interleukin 10.[78] It has been shown that this cytokine upregulates expression of HLA-G.[79] Lozano et al.,[80] demonstrated the increased expression of HLA-G in all monocytes and some T lymphocytes after HIV infection. Other evidence had implicated HAART in upregulation of HLA-G,[81] but the study by Lozano et al.,[80] excluded this mechanism by showing elevated levels of HLA-G in untreated HIV-positive individuals. A contradictory report by Derrien et al.,[82] showed downregulation of HLA-G in HIV infection. Though these authors agreed that this was an Nef-independent process, as HLA-G is unable to interact with Nef, they thought it was more likely to be a viral protein U (Vpu)-dependent mechanism as HLA-G possesses a dilysine motif (RKKSSD) at -4 and -5 from the carboxy-terminus,[67] with which Vpu could interact and interfere with further intracellular trafficking of HLA-G. The difference between these two studies may arise for two reasons. First, Derrien et al.,[82] studied expression in cell lines, which would have subtle differences in microenvironment from in vivo. Second, the stage of infection may have a profound effect on the microenvironment, which, in turn, could alter HLA-G expression. Derrien et al.,[82] studied HLA-G expression in acute HIV infection, and their results are similar to other acute viral infections such as human cytomegalovirus and herpes simplex virus. These both decrease cell surface expression of HLA-G1, but the former particularly can increase HLA-G1 expression upon reactivation.[83,84] Possibly the expression of HLA-G could be enhanced in the natural course of HIV infection so that the situation in chronic infection would be as shown by Lozano et al..[80]
Further, HLA-G polymorphism is also associated with the risk of HIV infection. Matte et al.,[85] carried out an extensive study of HLA-G polymorphism in 456 HIV-seropositive and 406 HIV-seronegative African women and found significant association of G*0105N with protection from HIV-1 infection and G*010108 with susceptibility to infection. Allele G*0105N is characterized by deletion of cytosine at position 130 of exon 3, leading to frameshift and introduction of a stop codon in exon 4.[86] Hence allele G*0105N impedes production of a functional HLA-G molecule. The most likely reason for association of G*0105N with protection from HIV infection would be that this impairs the function of HLA-G and so downregulation by HIV would be absent or decreased. Recently Lajoie et al.,[53] presented more extended and explicit data for HLA-G polymorphism in the same cohort. They found that women carrying G*0105N had a 2.2-fold decreased risk of HIV-1 infection compared with women without G*0105N. They also reported an HIV- seronegative woman who was homozygous for G*0105N.
The G*010108 allele, reported to be associated with increased risk of HIV-1 infection,[85] has a synonymous substitution (proline) of G to A at codon 57. Though this mutation does not bring about any change in amino acid sequence, it is in the vicinity of Glu-63, which interacts with the P2 position of loaded peptide.[87] In the mouse homologue Qa-2, P1 arginine of the peptide interacts with Glu-62, Glu-63, Tyr-59 and Trp-167 residues, three out of four of which are in close proximity to Pro-57. Another HLA-G allele, G*010401, shares variation at codon 57 with G*010108, although it also has a non-synonymous substitution at codon 110. The G*010108/G*010401 genotype has been shown to have a greater association with increased risk of HIV infection.[85] However, this may be because G*010401 is a high secretor allele associated with increased secretion of soluble HLA-G molecules, consequently being open to more systemic downregulation. All individuals identified with G*010108/G*010401 were homozygous at codon 57.[85] Though this position is not directly involved in the presentation of peptide, zygosity of HLA at this position could still affect susceptibility to HIV infection. Aikhionbare et al.,[88] have shown that discordance at codon 57 of HLA-G exon 2 was significantly associated with non-transmission of HIV-1 infection in mother-child pairs studied to investigate the risk of perinatal HIV transmission. This is probably in agreement with the observations discussed above that HLA sharing leads to increased susceptibility to HIV-1 transmission. However, more studies are needed to validate these observations, particularly for HLA-G.
HIV and the Less Polymorphic HLA-E
HLA-E was initially recognized as HLA-6.2 and was mapped to chromosome 6p21.3 between HLA-C and HLA-A.[89] HLA-E has a wide tissue distribution including T cells, B cells, activated T lymphocytes and various other cells such as placenta cells and trophoblasts.[90,91] HLA-E is less polymorphic, having only three alleles identified so far. These three alleles can be differentiated as HLA-ER (0101) and HLA-EG (01031 and 01032) by a non-synonymous substitution of arginine by glycine at position 107. Alleles 01031 and 01032 differ only by a synonymous mutation at codon 77.
HLA-E also has NK-regulating properties, as HLA-E has been identified as a ligand of a subset of the immunoglobulin superfamily of NK cell receptors, and their interaction with KIR of NK cells may be responsible for inhibition of killer activities in these cells.[92] HLA-E is distinct in that it depends for surface expression on a highly conserved nonamer peptide derived from the signal sequence of other class I molecules including HLA-A, HLA-B, HLA-C and HLA-G, but not HLA-F.[93] The peptide structure is very important, as only appropriate peptide can be loaded onto HLA-E, enabling expression and subsequent protection of target cells by interaction of the HLA-E-peptide complex with the CD94/NKG2 receptor of NK cells.[94]
A potential role for HLA-E in susceptibility to HIV has been neglected until a recent report showed that it was upregulation during p24-positive HIV-1 infection.[95] Though HLA-E has wide tissue distribution, its dependency on peptides derived from MHC class I may affect its expression on HIV-1-infected cells, as they have decreased class I expression. However, HLA-E expression could be supported by peptides derived from HLA-G or of viral origin. There is evidence of HLA-E upregulation by viral peptides: UL40-derived peptide in human cytomegalovirus,[96] and core 35-45 peptide in herpes simplex virus.[97] It has also been shown that HLA-E is upregulated by peptide 14-22 derived from HIV p24. Comparison of the HIV p2414-22 peptide with the sequences of other known HLA-E-specific peptides showed that it was very similar, with only subtle changes, and matched the HLA-E-binding criteria. HIV p2414-22 shares isolucine at position 2, which appears to be essential for HLA-E interactions, and has residues at positions 4, 6 and 7 that are similar to those identified in other HLA-E-specific peptides. HIV p2414-22 peptide has asparagine at position 5, which may be essential for HLA-E-peptide complex interaction with CD94/NKG2A, an HLA-E-specific inhibitory NK cell receptor. It has been reported that upregulation of HLA-E by HIV p2414-22 can inhibit cytolytic function of NK cells by interacting particularly with inhibitory CD94/NKG2A receptor.
Specificity of this HLA-E-peptide complex interaction with CD94/NKG2A, responsible for inhibition of NK cell cytolysis, was further confirmed by studies that restored NK cell cytolytic activity by blocking HLA-E with specific monoclonal antibody 3D12 or blocking CD94/NKG2A with specific anti-NKG2A antibody.[98] The HLA-E-specific HIV p2414-22 peptide is derived from HIV Gag, and as it consists of a putative proteasome cleavage site, it is conceivable that the peptide could be processed by proteasomal cleavage during natural HIV infection.
There are reports relating to HLA-E polymorphism with susceptibility to HIV-1 infection. Lajoie et al.,[53] have demonstrated association of HLA-EG allele with protection against HIV infection. HLA-EG is known to have better immunoregulating properties than HLA-ER. HLA-EG has also been associated with other pathologies, for example nasopharyngeal carcinoma,[98] and affected pregnancy outcome.[99] Strong et al.,[94] have shown that HLA-EG-peptide complex always has higher surface expression than the HLA-ER-peptide complex and HLA-EG is also more thermally stable. When affinity with peptides of various origins was tested, it was found that the relative affinity of HLA-EG for peptide was significantly higher than that of HLA-ER.[94]
As there is substantial evidence for a role for HLA-EG in efficient immunoregulation, its association with better prognosis in HIV infection would be expected; yet the converse is observed, which further suggests that, under cellular stress, HLA-E upregulation instead of immunoprotection supports immunosurveillance. HLA-E can interact with the leader peptide derived from heat shock protein 60 (hsp60),[100] which is generated in response to cellular stress. However, presentation of hsp60-derived peptide on HLA-E would not be sufficient to inhibit NK cell cytolytic activity, as the HLA-E-peptide complex could not interact efficiently with CD94/NKG2A.[100] The same situation might also occur with the HIV-derived peptide. As KIR receptors specifically identify HLA-E complexed with cellular peptide in order to stimulate NK cell inhibition, complexes with non-cellular peptides might interfere with this recognition by the inhibitory receptor. Further HLA-EG has higher stability and affinity with peptide than HLA-ER [94] hence it may have a more rigid three-dimensional conformation - and even subtle changes could be identified by CD94/NKG2A. There is also the possibility that HLA-E could induce virus-specific CTL immune responses, as in the case of cytomegalovirus-derived peptide. However, these assumptions require extensive functional studies to validate the impact of HLA-E and HIV-derived peptide on NK cell receptors.
Table 1. Various Genetic Factors in HIV Susceptibility


Table 2. Interaction of Human Leukocyte Antigen G and E With Different Natural Killer Cell Receptors


References
1. Doms RW, Trono D. The plasma membrane as a combat zone in the HIV battlefield. Genes Dev 2000; 14:2677-2688.
2. Connor RI, Sheridan KE, Ceradini D, Choe S, Landau NR. Change in coreceptor use coreceptor use correlates with disease progression in HIV-1-infected individuals. J Exp Med 1997; 185:621-628.
3. Zhang Y, Lou B, Lal RB, Gettie A, Marx PA, Moore JP. Use of inhibitors to evaluate coreceptor usage by simian and simian/human immunodeficiency viruses and human immunodeficiency virus type 2 in primary cells. J Virol 2000; 74:6893-6910.
4. McDougal JS, Mawle A, Cort SP, Nicholson JK, Cross GD, Scheppler-Campbell JA, et al. Cellular tropism of the human retrovirus HTLV-III/LAV. I. Role of T cell activation and expression of the T4 antigen. J Immunol 1985; 135:3151-3162.
5. Douek DC, Brenchley JM, Betts MR, Ambrozak DR, Hill BJ, Okamoto Y, et al. HIV preferentially infects HIV-specific CD4+ T cells. Nature 2002; 417:95-98.
6. Brinchmann JE, Albert J, Vartdal F. Few infected CD4+ T cells but a high proportion of replication-competent provirus copies in asymptomatic human immunodeficiency virus type 1 infection. J Virol 1991; 65:2019-2023.
7. Finkel TH, Banda NK. Indirect mechanisms of HIV pathogenesis: how does HIV kill T cells? Curr Opin Immunol 1994; 6:605-615.
8. Ameisen JC, Capron A. Cell dysfunction and depletion in AIDS: the programmed cell death hypothesis. Immunol Today 1991; 12:102-105.
9. Greenberg M, DeTulleo L, Rapoport I, Skowronski J, Kirchhausen T. A dileucine motif in HIV-1 Nef is essential for sorting into clathrin-coated pits and for downregulation of CD4. Curr Biol 1998; 8:1239-1242.
10. Greenberg ME, Bronson S, Lock M, Neumann M, Pavlakis GN, Skowronski J. Co-localization of HIV-1 Nef with the AP-2 adaptor protein complex correlates with Nef-induced CD4 down-regulation. EMBO J 1997; 16:6964-6976.
11. Li Q, Duan L, Estes JD, Ma ZM, Rourke T, Wang Y, et al. Peak SIV replication in resting memory CD4+ T cells depletes gut lamina propria CD4+ T cells. Nature 2005; 434:1148-1152.
12. Boirivant M, Viora M, Giordani L, Luzzati AL, Pronio AM, Montesani C, Pugliese O. HIV-1 gp120 accelerates Fas-mediated activation-induced human lamina propria T cell apoptosis. J Clin Immunol 1998; 18:39-47.
13. Koup RA, Safrit JT, Cao Y, Andrews CA, McLeod G, Borkowsky W, et al. Temporal association of cellular immune responses with the initial control of viremia in primary human immunodeficiency virus type 1 syndrome. J Virol 1994; 68:4650-4655.
14. Borrow P, Lewicki H, Wei X, Horwitz MS, Peffer N, Meyers H, et al. Antiviral pressure exerted by HIV-1-specific cytotoxic T lymphocytes (CTLs) during primary infection demonstrated by rapid selection of CTL escape virus. Nat Med 1997; 3:205-211.
15. Phillips AN. Reduction of HIV concentration during acute infection: independence from a specific immune response. Science 1996; 271:497-499.
16. Schwartz EJ, Neumann AU, Teixeira AV, Bruggeman LA, Rappaport J, Perelson AS, et al. Effect of target cell availability on HIV-1 production in vitro. AIDS 2002; 16:341-345.
17. Borrow P, Lewicki H, Hahn BH, Shaw GM, Oldstone MB. Virus-specific CD8+ cytotoxic T-lymphocyte activity associated with control of viremia in primary human immunodeficiency virus type 1 infection. J Virol 1994; 68:6103-6110.
18. Koup RA, Safrit JT, Cao Y, Andrews CA, McLeod G, Borkowsky W, et al. Temporal association of cellular immune responses with the initial control of viremia in primary human immunodeficiency virus type 1 syndrome. J Virol 1994; 68:4650-4655.
19. Ogg GS, Jin X, Bonhoeffer S, Dunbar PR, Nowak MA, Monard S, et al. Quantitation of HIV-1-specific cytotoxic T lymphocytes and plasma load of viral RNA. Science 1998; 279:2103-2106.
20. Yu XG, Addo MM, Rosenberg ES, Rodriguez WR, Lee PK, Fitzpatrick CA, et al. Consistent patterns in the development and immunodominance of human immunodeficiency virus type 1 (HIV-1)-specific CD8+ T-cell responses following acute HIV-1 infection. J Virol 2002; 76:8690-8701.
21. Altman JD, Moss PA, Goulder PJ, Barouch DH, McHeyzer-Williams MG, Bell JI, et al. Phenotypic analysis of antigen-specific T lymphocytes. Science 1996; 274:94-96 [Erratum in Science 1998; 280:1821.].
22. Kuroda MJ, Schmitz JE, Charini WA, Nickerson CE, Lord CI, Forman MA, et al. Comparative analysis of cytotoxic T lymphocytes in lymph nodes and peripheral blood of simian immunodeficiency virus-infected rhesus monkeys. J Virol 1999; 73:1573-1579.
23. Ogg GS, Kostense S, Klein MR, Jurriaans S, Hamann D, McMichael AJ, et al. Longitudinal phenotypic analysis of human immunodeficiency virus type 1-specific cytotoxic T lymphocytes: correlation with disease progression. J Virol 1999; 73:9153-9160.
24. Goulder PJ, Altfeld MA, Rosenberg ES, Nguyen T, Tang Y, Eldridge RL, et al. Substantial differences in specificity of HIV-specific cytotoxic T cells in acute and chronic HIV infection. J Exp Med 2001; 193:181-194.
25. Berke G. The CTL's kiss of death. Cell 1995; 81:9-12.
26. Mosmann TR, Li L, Sad S. Functions of CD8 T-cell subsets secreting different cytokine patterns. Semin Immunol 1997; 9:87-92.
27. Wagner L, Yang OO, Garcia-Zepeda EA, Ge Y, Kalams SA, Walker BD, et al. Beta-chemokines are released from HIV-1-specific cytolytic T-cell granules complexed to proteoglycans. Nature 1998; 391:908-911.
28. Tschopp J, Hofmann K. Cytotoxic cells: more weapons for new targets? Trends Microbiol 1996; 4:91-94.
29. Cocchi F, DeVico AL, Garzino-Demo A, Cara A, Gallo RC, Lusso P. The V3 domain of the HIV-1 gp120 envelope glycoprotein is critical for chemokine-mediated blockade of infection. Nat Med 1996; 2:1244-1247.
30. Cocchi F, DeVico AL, Garzino-Demo A, Arya SK, Gallo RC, Lusso P. Identification of RANTES, MIP-1 alpha, and MIP-1 beta as the major HIV-suppressive factors produced by CD8+ T cells. Science 1995; 270:1811-1815.
31. Rowland-Jones SL, Nixon DF, Aldhous MC, Gotch F, Ariyoshi K, Hallam N, et al. HIV-specific cytotoxic T-cell activity in an HIV-exposed but uninfected infant. Lancet 1993; 341:860-861.
32. Looney DJ. Immune responses to human immunodeficiency virus type 1 in exposed but uninfected individuals: protection or chance? J Clin Invest 1994; 93:920.
33. Coffin JM. HIV population dynamics in vivo: implications for genetic variation, pathogenesis, and therapy. Science 1995; 267:483-489.
34. McAdam S, Klenerman P, Tussey L, Rowland-Jones S, Lalloo D, Phillips R, et al. Immunogenic HIV variant peptides that bind to HLA-B8 can fail to stimulate cytotoxic T lymphocyte responses. J Immunol 1995; 155:2729-2736.
35. Moore CB, John M, James IR, Christiansen FT, Witt CS, Mallal SA. Evidence of HIV-1 adaptation to HLA-restricted immune responses at a population level. Science 2002; 24; 296:1439-1443.
36. Goulder PJ, Phillips RE, Colbert RA, McAdam S, Ogg G, Nowak MA, et al. Late escape from an immunodominant cytotoxic T-lymphocyte response associated with progression to AIDS. Nat Med 1997; 3:212-217.
37. Williams M, Roeth JF, Kasper MR, Fleis RI, Przybycin CG, Collins KL. Direct binding of human immunodeficiency virus type 1 Nef to the major histocompatibility complex class I (MHC-I) cytoplasmic tail disrupts MHC-I trafficking. J Virol 2002; 76:12173-12184.
38. Mangasarian A, Piguet V, Wang JK, Chen YL, Trono D. Nef-induced D4 and major histocompatibility complex class I (MHC-I) down-regulation are governed by distinct determinants: N-terminal alpha helix and proline repeat of Nef selectively regulate MHC-I trafficking. J Virol 1999; 73:1964-1973.
39. Le Gall S, Erdtmann L, Benichou S, Berlioz-Torrent C, Liu L, Benarous R, et al. Nef interacts with the mu subunit of clathrin adaptor complexes and reveals a cryptic sorting signal in MHC I molecules. Immunity 1998; 8:483-495.
40. Swann SA, Williams M, Story CM, Bobbitt KR, Fleis R, Collins KL. HIV-1 Nef blocks transport of MHC class I molecules to the cell surface via a PI 3-kinase-dependent pathway. Virology 2001; 282:267-277.
41. Blagoveshchenskaya AD, Thomas L, Feliciangeli SF, Hung CH, Thomas G. HIV-1 Nef downregulates MHC-I by a PACS-1- and PI3K-regulated ARF6 endocytic pathway. Cell 2002; 111:853-866.
42. Roeth JF, Williams M, Kasper MR, Filzen TM, Collins KL. HIV-1 Nef disrupts MHC-I trafficking by recruiting AP-1 to the MHC-I cytoplasmic tail. J Cell Biol 2004; 167:903-913.
43. Robinson MS, Bonifacino JS. Adaptor-related proteins. Curr Opin Cell Biol 2001; 13:444-453.
44. Kasper MR, Roeth JF, Williams M, Filzen TM, Fleis RI, Collins KL. HIV-1 Nef disrupts antigen presentation early in the secretory pathway. J Biol Chem 2005; 280:12840-12848.
45. Samson M, Labbe O, Mollereau C, Vassart G, Parmentier M. Molecular cloning and functional expression of a new human CC-chemokine receptor gene. Biochemistry 1996; 35:3362-3367.
46. Bream JH, Young HA, Rice N, Martin MP, Smith MW, Carrington M, et al. CCR5 promoter alleles and specific DNA binding factors. Science 1999; 284:223.
47. Anzala AO, Ball TB, Rostron T, O'Brien SJ, Plummer FA, Rowland-Jones SL. CCR2-64I allele and genotype association with delayed AIDS progression in African women. University of Nairobi Collaboration for HIV Research. Lancet 1998; 351:1632-1633.
48. An P, Nelson GW, Wang L, Donfield S, Goedert JJ, Phair J, et al. Modulating influence on HIV/AIDS by interacting RANTES gene variants. Proc Natl Acad Sci USA 2002; 99:10002-10007.
49. Winkler C, Modi W, Smith MW, Nelson GW, Wu X, Carrington M, et al. Genetic restriction of AIDS pathogenesis by an SDF-1 chemokine gene variant. ALIVE Study, Hemophilia Growth and Development Study (HGDS), Multicenter AIDS Cohort Study (MACS), Multicenter Hemophilia Cohort Study (MHCS), San Francisco City Cohort (SFCC). Science 1998; 279:389-393.
50. Tang J, Costello C, Keet IP, Rivers C, Leblanc S, Karita E, et al. HLA class I homozygosity accelerates disease progression in human immunodeficiency virus type 1 infection. AIDS Res Hum Retroviruses 1999; 15:317-324.
51. Kaslow RA, Carrington M, Apple R, Park L, Munoz A, Saah AJ, et al. Influence of combinations of human major histocompatibility complex genes on the course of HIV-1 infection. Nat Med 1996; 2:405-411.
52. Gao X, Nelson GW, Karacki P, Martin MP, Phair J, Kaslow R, et al. Effect of a single amino acid change in MHC class I molecules on the rate of progression to AIDS. N Engl J Med 2001; 344:1668-1675.
53. Lajoie J, Hargrove J, Zijenah LS, Humphrey JH, Ward BJ, Roger M. Genetic variants in nonclassical major histocompatibility complex class I human leukocyte antigen (HLA)-E and HLA-G molecules are associated with susceptibility to heterosexual acquisition of HIV-1. J Infect Dis 2006; 193:298-301.
54. Carrington M, Nelson GW, Martin MP, Kissner T, Vlahov D, Goedert JJ, et al. HLA and HIV-1:heterozygote advantage and B*35-Cw*04 disadvantage. Science 1999; 283:1748-1752.
55. Trachtenberg E, Korber B, Sollars C, Kepler TB, Hraber PT, Hayes E, et al. Advantage of rare HLA supertype in HIV disease progression. Nat Med 2003; 9:928-935.
56. Polycarpou A, Ntais C, Korber BT, Elrich HA, Winchester R, Krogstad P, et al. Association between maternal and infant class I and II HLA alleles and of their concordance with the risk of perinatal HIV type 1 transmission. AIDS Res Hum Retroviruses 2002; 18:741-746.
57. Dorak MT, Tang J, Penman-Aguilar A, Westfall AO, Zulu I, Lobashevsky ES, et al. Transmission of HIV-1 and HLA-B allele-sharing within serodiscordant heterosexual Zambian couples. Lancet 2004; 363:2137-2139.
58. Kiepiela P, Leslie AJ, Honeyborne I, Ramduth D, Thobakgale C, Chetty S, et al. Dominant influence of HLA-B in mediating the potential co-evolution of HIV and HLA. Nature 2004; 432:769-775.
59. Marsh SG, Albert ED, Bodmer WF, Bontrop RE, Dupont B, Erlich HA, et al. Nomenclature for factors of the HLA system, 2004. Tissue Antigens 2005; 65:301-369.
60. Leslie AJ, Pfafferott KJ, Chetty P, Draenert R, Addo MM, Feeney M, et al. HIV evolution: CTL escape mutation and reversion after transmission. Nat Med 2004; 10:282-289.
61. Jin X, Gao X, Ramanathan M Jr, Deschenes GR, Nelson GW, O'Brien SJ, et al. Human immunodeficiency virus type 1 (HIV-1)-specific CD8+ T-cell responses for groups of HIV-1-infected individuals with different HLA-B*35 genotypes. J Virol 2002; 76:12603-12610.
62. Martin MP, Gao X, Lee JH, Nelson GW, Detels R, Goedert JJ, et al. Epistatic interaction between KIR3DS1 and HLA-B delays the progression to AIDS. Nat Genet 2002; 31:429-434.
63. Malhotra U, Holte S, Dutta S, Berrey MM, Delpit E, Koelle DM, et al. Role for HLA class II molecules in HIV-1 suppression and cellular immunity following antiretroviral treatment. J Clin Invest 2001; 107:505-517.
64. Alter G, Suscovich TJ, Kleyman M, Teigen N, Streeck H, Zaman MT, et al. Low perforin and elevated SHIP-1 expression is associated with functional anergy of natural killer cells in chronic HIV-1 infection. AIDS 2006; 20:1549-1551.
65. Wahle JA, Paraiso KH, Costello AL, Goll EL, Sentman CL, Kerr WG. Cutting edge: dominance by an MHC-independent inhibitory receptor compromises NK killing of complex targets. J Immunol 2006; 176:7165-7169.
66. Schwartz EJ, Neumann AU, Teixeira AV, Bruggeman LA, Rappaport J, Perelson AS, et al. Effect of target cell availability on HIV-1 production in vitro. AIDS 2002; 16:341-345.
67. Geraghty DE, Koller BH, Orr HT. A human major histocompatibility complex class I gene that encodes a protein with a shortened cytoplasmic segment. Proc Natl Acad Sci USA 1987; 84:9145-9149.
68. Lala PK, Zdravkovic M, Aboagye-Mathiesen G, Guimond MJ. Functional roles of HLA-G in the human placenta: fact and fancies. In: Gupta SK, editor. Reproductive Immunology. New Delhi: Narosa; 1999. pp. 233-241.
69. Ishitani A, Geraghty DE. Alternative splicing of HLA-G transcripts yields proteins with primary structures resembling both class I and class II antigens. Proc Natl Acad Sci USA 1992; 89:3947-3951.
70. Navarro F, Llano M, Bellon T, Colonna M, Geraghty DE, Lopez-Botet M. The ILT2(LIR1) and CD94/NKG2A NK cell receptors respectively recognize HLA-G1 and HLA-E molecules co-expressed on target cells. Eur J Immunol 1999; 29:277-283.
71. Colonna M, Samaridis J, Cella M, Angman L, Allen RL, O'Callaghan CA, et al. Human myelomonocytic cells express an inhibitory receptor for classical and nonclassical MHC class I molecules. J Immunol 1998; 160:3096-3100.
72. Borrego F, Ulbrecht M, Weiss EH, Coligan JE, Brooks AG. Recognition of human histocompatibility leukocyte antigen (HLA)-E complexed with HLA class I signal sequence-derived peptides by CD94/NKG2 confers protection from natural killer cell-mediated lysis. J Exp Med 1998; 187:813-818.
73. Braud VM, Allan DS, O'Callaghan CA, Soderstrom K, D'Andrea A, Ogg GS, et al. HLA-E binds to natural killer cell receptors CD94/NKG2A, B and C. Nature 1998; 391:795-799.
74. Rajagopalan S, Long EO. A human histocompatibility leukocyte antigen (HLA)-G-specific receptor expressed on all natural killer cells. J Exp Med 1999; 189:1093-1100.
75. Ponte M, Cantoni C, Biassoni R, Tradori-Cappai A, Bentivoglio G, Vitale C, et al. Inhibitory receptors sensing HLA-G1 molecules in pregnancy: decidua-associated natural killer cells express LIR-1 and CD94/NKG2A and acquire p49, an HLA-G1-specific receptor. Proc Natl Acad Sci USA 1999; 96:5674-5679.
76. Fournel S, Aguerre-Girr M, Huc X, Lenfant F, Alam A, Toubert A, et al. Cutting edge: soluble HLA-G1 triggers CD95/CD95 ligand-mediated apoptosis in activated CD8+ cells by interacting with CD8. J Immunol 2000; 164:6100-6104.
77. Bainbridge DR, Ellis SA, Sargent IL. HLA-G suppresses proliferation of CD4(+) T-lymphocytes. J Reprod Immunol 2000; 48:17-26.
78. Navikas V, Link J, Persson C, Olsson T, Hojeberg B, Ljungdahl A, et al. Increased mRNA expression of IL-6, IL-10, TNF-alpha, and perforin in blood mononuclear cells in human HIV infection. J Acquir Immune Defic Syndr Hum Retroviruses 1995; 9:484-489.
79. Moreau P, Adrian-Cabestre F, Menier C, Guiard V, Gourand L, Dausset J, et al. IL-10 selectively induces HLA-G expression in human trophoblasts and monocytes. Int Immunol 1999; 11:803-811.
80. Lozano JM, Gonzalez R, Kindelan JM, Rouas-Freiss N, Caballos R, Dausset J, et al. Monocytes and T lymphocytes in HIV-1-positive patients express HLA-G molecule. AIDS 2002; 16:347-351.
81. Cabello A, Rivero A, Garcia MJ, Lozano JM, Torre-Cisneros J, Gonzalez R, et al. HAART induces the expression of HLA-G on peripheral monocytes in HIV-1 infected individuals. Hum Immunol 2003; 64:1045-1049.
82. Derrien M, Pizzato N, Dolcini G, Menu E, Chaouat G, Lenfant F, et al. Human immunodeficiency virus 1 downregulates cell surface expression of the non-classical major histocompatibility class I molecule HLA-G1. J Gen Virol 2004; 85:1945-1954.
83. Fisher S, Genbacev O, Maidji E, Pereira L. Human cytomegalovirus infection of placental cytotrophoblasts in vitro and in utero: implications for transmission and pathogenesis. J Virol 2000; 74:6808-6820.
84. Onno M, Pangault C, Le Friec G, Guilloux V, Andre P, Fauchet R. Modulation of HLA-G antigens expression by human cytomegalovirus: specific induction in activated macrophages harboring human cytomegalovirus infection. J Immunol 2000; 164:6426-6434.
85. Matte C, Lajoie J, Lacaille J, Zijenah LS, Ward BJ, Roger M. Functionally active HLA-G polymorphisms are associated with the risk of heterosexual HIV-1 infection in African women. AIDS 2004; 18:427-431.
86. Ober C, Aldrich C, Rosinsky B, Robertson A, Walker MA, Willadsen S, et al. HLA-G1 protein expression is not essential for fetal survival. Placenta 1998; 19:127-132.
87. Clements CS, Kjer-Nielsen L, Kostenko L, Hoare HL, Dunstone MA, Moses E, et al. Crystal structure of HLA-G: a nonclassical MHC class I molecule expressed at the fetal-maternal interface. Proc Natl Acad Sci USA 2005; 102:3360-3365.
88. Aikhionbare FO, Hodge T, Kuhn L, Bulterys M, Abrams EJ, Bond VC. Mother-to-child discordance in HLA-G exon 2 is associated with a reduced risk of perinatal HIV-1 transmission. AIDS 2001; 15:2196-2198.
89. Koller BH, Geraghty DE, Shimizu Y, DeMars R, Orr HT. H.L.A.-E. A novel HLA class I gene expressed in resting T lymphocytes. J Immunol 1988; 141:897-904.
90. Boucraut J, Guillaudeux T, Alizadeh M, Boretto J, Chimini G, Malecaze F, et al. HLA-E is the only class I gene that escapes CpG methylation and is transcriptionally active in the trophoblast-derived human cell line JAR. Immunogenetics 1993; 38:117-130.
91. Houlihan JM, Biro PA, Harper HM, Jenkinson HJ, Holmes CH. The human amnion is a site of MHC class Ib expression: evidence for the expression of HLA-E and HLA-G. J Immunol 1995; 154:5665-5674.
92. Vales-Gomez M, Reyburn H, Strominger J. Molecular analyses of the interactions between human NK receptors and their HLA ligands. Hum Immunol 2000; 61:28-38.
93. Lee N, Goodlett DR, Ishitani A, Marquardt H, Geraghty DE. HLA-E surface expression depends on binding of TAP-dependent peptides derived from certain HLA class I signal sequences. J Immunol 1998; 160:4951-4960.
94. Strong RK, Holmes MA, Li P, Braun L, Lee N, Geraghty DE. HLA-E allelic variants. Correlating differential expression, peptide affinities, crystal structures, and thermal stabilities. J Biol Chem 2003; 278:5082-5090.
95. Nattermann J, Nischalke HD, Hofmeister V, Kupfer B, Ahlenstiel G, Feldmann G, et al. HIV-1 infection leads to increased HLA-E expression resulting in impaired function of natural killer cells. Antivir Ther 2005; 10:95-107.
96. Wang EC, McSharry B, Retiere C, Tomasec P, Williams S, Borysiewicz LK, et al. UL40-mediated NK evasion during productive infection with human cytomegalovirus. Proc Natl Acad Sci USA 2002; 99:7570-7575.
97. Nattermann J, Nischalke HD, Hofmeister V, Ahlenstiel G, Zimmermann H, Leifeld L, et al. The HLA-A2 restricted T cell epitope HCV core 35-44 stabilizes HLA-E expression and inhibits cytolysis mediated by natural killer cells. Am J Pathol 2005; 166:443-453.
98. Hirankarn N, Kimkong I, Mutirangura A. HLA-E polymorphism in patients with nasopharyngeal carcinoma. Tissue Antigens 2004; 64:588-592.
99. Tripathi P, Naik S, Agrawal S. HLA-E and immunobiology of pregnancy. Tissue Antigens 2006; 67:207-213.
100. Michaelsson J, Teixeira de Matos C, Achour A, Lanier LL, Karre K, Soderstrom K. A signal peptide derived from hsp60 binds HLA-E and interferes with D94/NKG2A recognition. J Exp Med 2002; 196:1403-1414.


Reprint Address
Professor Suraksha Agrawal, Department of Medical Genetics, Sanjay Gandhi Post Graduate Institute of Medical Sciences, Raebareli Road, Lucknow (UP) 226014, India. E-mail: suraksha@sgpgi.ac.in
Piyush Tripathi, Suraksha AgrawalFrom the Department of Medical Genetics, Sanjay Gandhi Post Graduate Institute of Medical Sciences, Raebareli Road, Lucknow (UP) 226014, India.


Saturday 7 July 2007

Increased prevalence of HIV: Not a casualty of war

The Lancet 2007; 369:2187-2195, DOI:10.1016/S0140-6736(07)61015-0

Prevalence of HIV infection in conflict-affected and displaced people in seven sub-Saharan African countries: a systematic review

Dr Paul B Spiegel MD a , Anne Rygaard Bennedsen BSc b, Johanna Claass MD a, Laurie Bruns MA a, Njogu Patterson MD a, Dieudonne Yiweza MD a and Marian Schilperoord MA a

Summary

Background
Violence and rape are believed to fuel the HIV epidemic in countries affected by conflict. We compared HIV prevalence in populations directly affected by conflict with that in those not directly affected and in refugees versus the nearest surrounding host communities in sub-Saharan African countries.

Methods
Seven countries affected by conflict (Democratic Republic of Congo, southern Sudan, Rwanda, Uganda, Sierra Leone, Somalia, and Burundi) were chosen since HIV prevalence surveys within the past 5 years had been done and data, including original antenatal-care sentinel surveillance data, were available. We did a systematic and comprehensive literature search using Medline and Embase. Only articles and reports that contained original data for prevalence of HIV infection were included. All survey reports were independently evaluated by two epidemiologists to assess internationally accepted guidelines for HIV sentinel surveillance and population-based
surveys. Whenever possible, data from the nearest antenatal care and host country sentinel site of the neighbouring countries were presented. 95% CIs were provided when available.

Findings
Of the 295 articles that met our search criteria, 88 had original prevalence data and 65 had data from the seven selected countries. Data from these countries did not show an increase in prevalence of HIV infection during periods of conflict, irrespective of prevalence when conflict began. Prevalence in urban areas affected by conflict decreased in Burundi, Rwanda, and Uganda at similar rates to urban areas unaffected by conflict in their respective countries. Prevalence in conflict-affected rural areas remained low and fairly stable in these countries. Of the 12 sets of refugee camps, nine had a lower prevalence of HIV infection, two a similar prevalence, and one a higher prevalence than their respective host communities. Despite wide-scale rape in many countries, there are no data to show that rape increased prevalence of HIV infection at the population level.

Interpretation
We have shown that there is a need for mechanisms to provide time-sensitive information on the effect of conflict on incidence of HIV infection, since we found insufficient data to support the assertions that conflict, forced displacement, and wide-scale rape increase prevalence or that refugees spread HIV infection in host communities.

Affiliations

a. UN High Commissioner for Refugees, Geneva, Switzerland
b. University of Copenhagen, Copenhagen, Denmark

Correspondence to: Dr Paul B Spiegel, UN High Commissioner for Refugees, 94 rue de Montbrillant, 1202 Geneva, Switzerland. e-mail:
SPIEGEL@...

http://www.thelancet.com/journals/lancet/article/PIIS0140673607610150/fulltext

Tuesday 3 July 2007

Half Of HIV Spread By Newly Infected



Source:McGill UniversityDate:March 7, 2007

http://www.sciencedaily.com/releases/2007/03/070305084232.htm

Science Daily A new study led by McGill University researchers shows that half of all HIV transmissions happen when newly infected people don’t know they are carrying the virus and may not even test positive for it.

The study, published in the April edition of the Journal of Infectious Diseases and already available online, followed 2,500 patients in eight Montreal HIV clinics over eight years. It showed that newly infected patients are eight times more likely to transmit the virus than those in the chronic stage of AIDS given the same behaviour.

Dr. Mark Wainberg, Director of the McGill AIDS Centre and internationally respected AIDS researcher, presented the findings at an academic AIDS conference in Los Angeles March 1 with lead author Dr. Bluma Brenner of the McGill Faculty of Medicine and the Jewish General Hospital.

“The most alarming thing is the confluence of a highly infectious state and the lack of awareness of that state,” said Dr. Wainberg. “It means we have to reconsider a lot of what we’re doing, both on the public education front and on the early intervention front.”

McGill Professor of Medicine and McGill University Health Centre (MUHC) AIDS researcher Dr. Jean-Pierre Routy, who was also instrumental in the study, said the Montreal urban population provided the ideal sample for the groundbreaking survey. “We had the infrastructure and the data here to get a comprehensive picture.” The study also involved researchers at Université de Montréal and at private and public AIDS clinics in the city.

The findings could change not only how soon people get tested after engaging in high-risk behaviour, but how they view that behaviour. “It has been shown that an HIV-positive diagnosis modifies high-risk behaviour,” said Dr. Wainberg. “So the more actively we can seek out and find newly infected people for testing and counselling, the better.”

Note: This story has been adapted from a news release issued by McGill University.

Are We Spending Too Much On HIV?


Source:BMJ-British Medical JournalDate:February 19, 2007


http://www.sciencedaily.com/releases/2007/02/070216113829.htm

Science Daily Billions of pounds are being spent on the fight against AIDS in developing countries. In this week's British Medical Journal, two experts go head to head over whether we are spending too much.

HIV is receiving relatively too much money, with much of it used inefficiently and sometimes counterproductively, argues Roger England, Chairman of Health Systems Workshop.

Data show that 21% of health aid was allocated to HIV in 2004, up from 8% in 2000. It could now exceed a quarter. Yet HIV constitutes only 5% of the burden of disease in low and middle income countries as measured by disability adjusted life years lost (DALYs). It causes 2.8 million deaths a year worldwide -- fewer than the number of stillbirths, and much less than half the number of infant deaths. More deaths are attributable to diabetes than to HIV.

Furthermore, HIV interventions are not cost effective enough to justify this disproportionate spending, he writes. Much HIV money could be spent with more certain benefits on, for example, bed nets, immunisation, or family planning. Money is also wasted in areas that reflect the interests of those on the AIDS industry payroll more than evidence.

He believes that the money could be more effective if used to strengthen public health systems rather than focusing on disease-specific programmes.

AIDS is widely acknowledged as a public health crisis and current spending is woefully inadequate, argue Paul de Lay and colleagues at the Joint United Nations Programme on HIV and AIDS (UNAIDS).

Resources currently pledged are only half of what is needed for a comprehensive response. For instance, in 2006, $9bn was available for the AIDS response but the real need was estimated at $15bn. Poor coordination between different stakeholders in affected countries also impedes effective spending. This is compounded by weak institutions and regulatory policies, poor governance, and in some cases corruption.

They argue that the response to AIDS needs to be seen in the context of international commitments to the millennium development goals, which also call for progress across many other developmental priorities. HIV threatens many of these goals, especially those related to poverty and health.

The cost of inaction against AIDS is huge, far greater than for any other public health crisis, they say. Current costs are so high because of the inadequacy of previous investments, but they will be higher tomorrow if we continue to underinvest.

Note: This story has been adapted from a news release issued by BMJ-British Medical Journal.

DNA Vaccine Protects Against AIDS, Not HIV

Source: American Society For Microbiology
Date: April 20, 2005

http://www.sciencedaily.com/releases/2005/04/050419105400.htm

Science Daily While a new DNA vaccine may not be able to prevent HIV infection, it could protect against progression to full-blown AIDS. Researchers from Kansas report their findings in the March 2005 issue of the Journal of Virology.

Developing a vaccine to protect against HIV in attempt to gain control of the AIDS pandemic is a top priority for researchers throughout the world. Extensive testing has been conducted with live vaccines to determine if immunization would be effective at prevention, but they are not suitable for human use due to the potential that the vaccine viruses could mutate and reacquire the ability to cause disease.

DNA vaccines offer a new possibility for treatment. The have the advantages of safety, low cost of production, and ease of use in field conditions due to their minimal need for refrigeration.

In the study the DNA of a simian/human immunodeficiency virus (SHIV) was made noninfectious by removing the gene that makes reverse transciptase (a protein the virus requires to replicate). Four macaques were injected with the noninfectious vaccine, while two control animals remained unvaccinated. Both groups were challenged with SHIV. All four of the immunized macaques became infected with the challenge virus, but three survived. The two control subjects died.

"The results showed strong evidence that this type of vaccine could prevent AIDS and established that a DNA vaccine, such as this one, could be used alone, without the need for booster doses with viral proteins, for large-scale immunization programs," say the researchers.

(D.K. Singh, Z. Liu, D. Sheffer, G.A. Mackay, M. Smith, S. Dhillon, R. Hegde, F. Jia, I. Adany, O. Narayan. 2005. A noninfectious simian/human immunodeficiency virus DNA vaccine that protects macaques against AIDS. Journal of Virology, 79. 6: 3419-3428.)

Note: This story has been adapted from a news release issued by American Society For Microbiology.