• Aucun résultat trouvé

Final thoughts

Dans le document HIV-1 innate immune detection and evasion (Page 145-166)

Chapter 7: General discussion

7.6 Final thoughts

TRIM5α knockdown MDDCs produced 13 times less IL12b after LPS stimulation.

IL12 is the main cytokine directing TH1 development [318]. IL12 production by dendritic cells upon recognition of retroviral capsid would be necessary to induce a robust T cell response able to control HIV-1. The last attempt for a HIV-1 vaccine showed some promising results. The probability of an HIV-1 infection was 26% lower in vaccinated subjects but there was no difference in viral load or CD4+ T cell count in vaccinated individuals after HIV-1 infection [319]. This means that we still do not know what kind of immune response needs to be elicited to either block the initial infection or to control the systemic infection. As mentioned above one of the problems is the ability of HIV-1 to hide from the innate immune system preventing activation of dendritic cells. Our data shows that TRIM5a has the ability to recognize the retroviral capsid, activate dendritic cells and induce the production of IL12 to possibly promote a TH1 response.

Although some secreted IFNβ from MDDCs challenged with N-MLV VLPs can be detected (Figure 3.3 B), we never observed full activation of DCs. This second signal would be necessary for T cell activation. What is missing is the induction of IFNβ after detection of reverse transcription products. As previously mentioned, pDCs are able to detect HIV-1 genomic RNA via TLR7 and 9 and produce IFNβ but conventional DCs lack this ability [143]. One explanation lies in the presence of TREX1. As mentioned before, TREX1 facilitates HIV-1 infection by degrading cytoplasmic DNA, which would otherwise be detected by cGAS and induce an IFNβ response via STING signaling [135, 223]. Under normal circumstances TREX1 probably prevents overstimulation of the innate immune system by endogenous retroviral elements and DNA replication [220, 320]. TREX1 is one of the two

145

nucleases of the SET complex, which fragments DNA after activation by granzyme A induced apoptosis [217].

Interestingly, knockdown of SET complex members can protect HIV-1 from autointegration, a process where HIV-1 IN mistakenly integrates into itself instead of the host genome [222]. It was also shown that uracilated HIV-1 cDNA is less likely to succumb to autointegration because IN preferentially integrate into the host chromosome with no uracils present because uracil causes DNA distortions less favored by IN [321].

Uracilation is especially important in myeloid cells since they have a low dNTP pool. This favors ribonucleotide incorporation by RT, as well as dUTP, a metabolic intermediate during pyrimidine synthesis [254, 255]. This uracil can either protect from autointegration or it can be mutagenic. Usually single base pair misincorporation is repaired by the cellular base excision repair machinery (BER) but macrophages seem to have a lower repair rate because of a reduced RNAse H2 activity [255]. This makes sense since the DNA repair machinery is more important in cycling cells compared to non-dividing cells. The misincorporation of ribonucleotides also slows down the kinetics of reverse transcription. It is worth mentioning that in the absence of SAMHD1 mediated by Vpx, significantly reduces the sensitivity of HIV-1 to nucleoside reverse transcription inhibitors (NRTIs) [322].

Considering these data it makes sense to assume that not only TREX1 is used by HIV-1 as a facilitator to establish an infection without detection but also RNase H2 could play a similar role. As mentioned above RNASEH2B knockout mice are not viable because of an accumulation or incorporate ribonucleotides in the chromosome leading to genomic instability and DNA damage [239]. It is not known if the incorporation of ribonucleotides into the HIV-1 cDNA would cause increased mutations and if the mutations would have an impact on viral replication. But it is reasonable to assume that this could be the case and part of the observed inhibition in the RNASHE2A knockdown experiments in MDDC (chapter 6). This might be another way to hide from detection in dendritic cells by limiting the replication and production of new particles. Another explanation of the observed RNASEH2A phenotype could be that similar to TREX1, RNase H2 complex is able to degrade some aberrant reverse transcription intermediates containing RNA:DNA duplexes.

146

Such a hybrid molecule could serve as a PAMP in the cytoplasm to be sensed by an unknown PRR. Knockdown of RNASHE2A would then interfere with the removal of the potential PAMP and type I IFN could be induced leading to the inhibition of HIV-1 like it is the case for TREX1.

The presence of SAMHD1 in myeloid and dendritic cells inhibits full reverse transcription. This prevents HIV-1 from productively infect these cells on one hand and on the other hand it prevents the production of reverse transcription products with the potential of activating the innate immune system as seen above.

Interestingly, when a recombination of SIVRCM and SIVMUS gave rise to SIVCPZ, Vpx from SIVRCM was lost in favor to create a Vif protein, which was able to antagonize the chimp A3G [323]. The SIVCPZ Vif was also able to counteract human A3G, which might have been one of the reasons the virus could cross the species barrier giving rise to HIV-1. Both Vif proteins of SIVRCM and SIVMUS are not able to degrade chimp or human A3G creating the selection pressure for SIVCPZ to use Vpx from SIVRCM to create a new Vif. The Vif of SIVSM - the ancestor of HIV-2 - is already adapted to counteract human A3G and there was no selection pressure to chose between Vpx and Vif and so HIV-2 retained Vpx. SIVSM and HIV-2 are much less pathogenic than HIV-1. One explanation could be that since HIV-2 and SIVSM are able to infect dendritic cells the host can control the virus much better. One report looking at the efficiency of SAMHD1 degradation of HIV-2 Vpx found no correlation with the ability of the infected individuals to control viral replication [324]. When considering that sooty mangabey have high viral load but do not develop AIDS, it seems more likely that it is the host, which tolerates the virus without inducing a strong immune response leading to immune pathologies [18]. One line of evidence is provided by the fact that SAMHD1 restrict HIV-1 reverse transcription in resting T cells [251, 252].

Lymph nodes of HIV-1 infected individuals show a peculiar effect, although only 5%

of CD4+ T cells are actually productively infected, many more T cells go through apoptosis in a by-stander effect. One explanation is that these cells are going through an abortive infection creating reverse transcription intermediates, which are detected by the cells and lead to a caspase 1 depended apoptosis. This process is different from the normal caspase 8-induced apoptosis in that the dying T cell releases inflammatory cytokines causing additional damage on surrounding cells.

This process is called pyrotosis [114]. SAMHD1 could be the reason for the abortive

147

infection in the mostly resting T cells in the lymph node (W.C. Greene, CSHL retroviruses meeting 2013).

Type I IFN is a double-edged sword. On one side it might be one of the causes for HIV-1 induced immune pathologies but on the other hand it is needed to elicit a robust immune response. As mentioned above the last vaccine tested showed first promising results by reducing the risk of infecting by 26% percent. The vaccine was only able to prevent infection but had no effect on viral load or CD4+ T cell count in vaccinated patients and it did not protect from superinfection. For this vaccine a recombinant gp120 was used to induce an immune response [319]. This means even after 30 years of research we still do not know how an effective vaccine has to look like. Clearly the strategy cannot only be to prevent infection at the mucosa but also has to decrease and control viral replication in the patients after infection has taken place. This means we need to find out which epitopes are important to induce a CTL response.

148

The data presented in this thesis helps to better understand the interaction between HIV-1 and the innate immune system. A possible next step would be to combine the findings presented here: We would generate dendritic cells knockdown for TREX1 and possibly RNASH2A and challenge these cells with and HIV-1 virus with a capsid variant restricted by TRIM5α and Vpx in trans or cis (chapter 4). HIV-1 could infect these cells and complete reverse transcription thanks to Vpx. The infected cell would present antigen on MHC I and at the same time TRIM5α capsid binding and recognition and the absence of TREX1 and RNASHE2A would activate the cells, allowing the cross talk to CD8+ T cells. That experimental system might allow us to find epitopes which elicit a robust CD8+ T cell response (Figure 7.1). That knowledge could eventually be transformed into a vaccine approach inducing a robust CTL response enabling the immune system to detect and destroy infected cells.

149

References

1. Centers for Disease, C., Pneumocystis pneumonia--Los Angeles. MMWR Morb Mortal Wkly Rep, 1981. 30(21): p. 250-2.

2. Centers for Disease, C., Kaposi's sarcoma and Pneumocystis pneumonia among homosexual men--New York City and California. MMWR Morb Mortal Wkly Rep, 1981. 30(25): p. 305-8.

3. Gottlieb, M.S., et al., Pneumocystis carinii pneumonia and mucosal candidiasis in previously healthy homosexual men: evidence of a new acquired cellular immunodeficiency. N Engl J Med, 1981. 305(24): p. 1425-31.

4. Masur, H., et al., An outbreak of community-acquired Pneumocystis carinii pneumonia: initial manifestation of cellular immune dysfunction. N Engl J Med, 1981. 305(24): p. 1431-8.

5. Barre-Sinoussi, F., et al., Isolation of a T-lymphotropic retrovirus from a patient at risk for acquired immune deficiency syndrome (AIDS). Science, 1983. 220(4599): p. 868-71.

6. Gallo, R.C., et al., Isolation of human T-cell leukemia virus in acquired immune deficiency syndrome (AIDS). Science, 1983. 220(4599): p. 865-7.

7. Gallo, R.C., Historical essay. The early years of HIV/AIDS. Science, 2002. 298(5599): p. 1728-30.

8. Gallo, R.C. and L. Montagnier, Historical essay. Prospects for the future. Science, 2002.

298(5599): p. 1730-1.

9. Montagnier, L., Historical essay. A history of HIV discovery. Science, 2002. 298(5599): p. 1727-10. 8. Levy, J.A., et al., Isolation of lymphocytopathic retroviruses from San Francisco patients with

AIDS. Science, 1984. 225(4664): p. 840-2.

11. Sarngadharan, M.G., et al., Antibodies reactive with human T-lymphotropic retroviruses (HTLV-III) in the serum of patients with AIDS. Science, 1984. 224(4648): p. 506-8.

12. Fauci, A.S., HIV and AIDS: 20 years of science. Nat Med, 2003. 9(7): p. 839-43.

13. Gao, F., et al., Origin of HIV-1 in the chimpanzee Pan troglodytes troglodytes. Nature, 1999.

397(6718): p. 436-41.

14. Plantier, J.C., et al., A new human immunodeficiency virus derived from gorillas. Nat Med, 2009. 15(8): p. 871-2.

15. Gao, F., et al., Human infection by genetically diverse SIVSM-related HIV-2 in west Africa.

Nature, 1992. 358(6386): p. 495-9.

16. Ayouba, A., et al., Evidence for continuing cross-species transmission of SIVsmm to humans:

characterization of a new HIV-2 lineage in rural Cote d'Ivoire. AIDS, 2013.

17. Santiago, M.L., et al., Simian immunodeficiency virus infection in free-ranging sooty mangabeys (Cercocebus atys atys) from the Tai Forest, Cote d'Ivoire: implications for the origin of epidemic human immunodeficiency virus type 2. J Virol, 2005. 79(19): p. 12515-27.

18. Silvestri, G., et al., Nonpathogenic SIV infection of sooty mangabeys is characterized by limited bystander immunopathology despite chronic high-level viremia. Immunity, 2003.

18(3): p. 441-52.

19. Letvin, N.L., et al., Induction of AIDS-like disease in macaque monkeys with T-cell tropic retrovirus STLV-III. Science, 1985. 230(4721): p. 71-3.

20. Keele, B.F., et al., Increased mortality and AIDS-like immunopathology in wild chimpanzees infected with SIVcpz. Nature, 2009. 460(7254): p. 515-9.

21. Temin, H.M. and S. Mizutani, RNA-dependent DNA polymerase in virions of Rous sarcoma virus. Nature, 1970. 226(5252): p. 1211-3.

22. Baltimore, D., RNA-dependent DNA polymerase in virions of RNA tumour viruses. Nature, 1970. 226(5252): p. 1209-11.

23. Coffin, J.M., S.H. Hughes, and H.E. Varmus, The Interactions of Retroviruses and their Hosts, in Retroviruses, J.M. Coffin, S.H. Hughes, and H.E. Varmus, Editors. 1997: Cold Spring Harbor (NY).

150

24. Rai, K.R., et al., Treatment of acute myelocytic leukemia: a study by cancer and leukemia group B. Blood, 1981. 58(6): p. 1203-12.

25. Wei, X., et al., Antibody neutralization and escape by HIV-1. Nature, 2003. 422(6929): p. 307-26. 12. Gottlieb, M. and G.B. Brandt, Temperature sensing in optical fibers using cladding and jacket

loss effects. Appl Opt, 1981. 20(22): p. 3867-73.

27. Finzi, A., et al., Topological layers in the HIV-1 gp120 inner domain regulate gp41 interaction and CD4-triggered conformational transitions. Mol Cell, 2010. 37(5): p. 656-67.

28. Dalgleish, A.G., et al., The CD4 (T4) antigen is an essential component of the receptor for the AIDS retrovirus. Nature, 1984. 312(5996): p. 763-7.

29. Scarlatti, G., et al., In vivo evolution of HIV-1 co-receptor usage and sensitivity to chemokine-mediated suppression. Nat Med, 1997. 3(11): p. 1259-65.

30. Deng, H., et al., Identification of a major co-receptor for primary isolates of HIV-1. Nature, 1996. 381(6584): p. 661-6.

31. Liu, R., et al., Homozygous defect in HIV-1 coreceptor accounts for resistance of some multiply-exposed individuals to HIV-1 infection. Cell, 1996. 86(3): p. 367-77.

32. Alfadhli, A., R.L. Barklis, and E. Barklis, HIV-1 matrix organizes as a hexamer of trimers on membranes containing phosphatidylinositol-(4,5)-bisphosphate. Virology, 2009. 387(2): p.

466-72.

33. Ono, A., J.M. Orenstein, and E.O. Freed, Role of the Gag matrix domain in targeting human immunodeficiency virus type 1 assembly. J Virol, 2000. 74(6): p. 2855-66.

34. Freed, E.O. and M.A. Martin, Domains of the human immunodeficiency virus type 1 matrix and gp41 cytoplasmic tail required for envelope incorporation into virions. J Virol, 1996.

70(1): p. 341-51.

35. Briggs, J.A., et al., Structural organization of authentic, mature HIV-1 virions and cores. EMBO J, 2003. 22(7): p. 1707-15.

36. Pornillos, O., B.K. Ganser-Pornillos, and M. Yeager, Atomic-level modelling of the HIV capsid.

Nature, 2011. 469(7330): p. 424-7.

37. Zhao, G., et al., Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature, 2013. 497(7451): p. 643-6.

38. Levin, J.G., et al., Role of HIV-1 nucleocapsid protein in HIV-1 reverse transcription. RNA Biol, 2010. 7(6): p. 754-74.

39. Tsuchihashi, Z. and P.O. Brown, DNA strand exchange and selective DNA annealing promoted by the human immunodeficiency virus type 1 nucleocapsid protein. J Virol, 1994. 68(9): p.

5863-70.

40. Wu, T., et al., Fundamental differences between the nucleic acid chaperone activities of HIV-1 nucleocapsid protein and Gag or Gag-derived proteins: biological implications. Virology, 2010. 405(2): p. 556-67.

41. Gottlinger, H.G., et al., Effect of mutations affecting the p6 gag protein on human immunodeficiency virus particle release. Proc Natl Acad Sci U S A, 1991. 88(8): p. 3195-9.

42. Paxton, W., R.I. Connor, and N.R. Landau, Incorporation of Vpr into human immunodeficiency virus type 1 virions: requirement for the p6 region of gag and mutational analysis. J Virol, 1993. 67(12): p. 7229-37.

43. Moulard, M. and E. Decroly, Maturation of HIV envelope glycoprotein precursors by cellular endoproteases. Biochim Biophys Acta, 2000. 1469(3): p. 121-32.

44. Vollenweider, F., et al., Comparative cellular processing of the human immunodeficiency virus (HIV-1) envelope glycoprotein gp160 by the mammalian subtilisin/kexin-like convertases.

Biochem J, 1996. 314 ( Pt 2): p. 521-32.

45. Tristem, M., et al., Evolution of the primate lentiviruses: evidence from vpx and vpr. EMBO J, 1992. 11(9): p. 3405-12.

46. Eckert, D.M. and P.S. Kim, Mechanisms of viral membrane fusion and its inhibition. Annu Rev Biochem, 2001. 70: p. 777-810.

151

47. Tilton, J.C. and R.W. Doms, Entry inhibitors in the treatment of HIV-1 infection. Antiviral Res, 2010. 85(1): p. 91-100.

48. Champoux, J.J. and S.J. Schultz, Ribonuclease H: properties, substrate specificity and roles in retroviral reverse transcription. FEBS J, 2009. 276(6): p. 1506-16.

49. Herschhorn, A. and A. Hizi, Retroviral reverse transcriptases. Cell Mol Life Sci, 2010. 67(16): p.

2717-47.

50. Hu, W.S. and S.H. Hughes, HIV-1 reverse transcription. Cold Spring Harb Perspect Med, 2012.

2(10).

51. Forshey, B.M., et al., Formation of a human immunodeficiency virus type 1 core of optimal stability is crucial for viral replication. J Virol, 2002. 76(11): p. 5667-77.

52. Hulme, A.E., O. Perez, and T.J. Hope, Complementary assays reveal a relationship between HIV-1 uncoating and reverse transcription. Proc Natl Acad Sci U S A, 2011. 108(24): p. 9975-80.

53. De Iaco, A., et al., TNPO3 protects HIV-1 replication from CPSF6-mediated capsid stabilization in the host cell cytoplasm. Retrovirology, 2013. 10: p. 20.

54. Li, Y., A.K. Kar, and J. Sodroski, Target cell type-dependent modulation of human immunodeficiency virus type 1 capsid disassembly by cyclophilin A. J Virol, 2009. 83(21): p.

10951-62.

55. Stremlau, M., et al., Specific recognition and accelerated uncoating of retroviral capsids by the TRIM5alpha restriction factor. Proc Natl Acad Sci U S A, 2006. 103(14): p. 5514-9.

56. Lewis, P.F. and M. Emerman, Passage through mitosis is required for oncoretroviruses but not for the human immunodeficiency virus. J Virol, 1994. 68(1): p. 510-6.

57. Lewis, P., M. Hensel, and M. Emerman, Human immunodeficiency virus infection of cells arrested in the cell cycle. EMBO J, 1992. 11(8): p. 3053-8.

58. Weinberg, J.B., et al., Productive human immunodeficiency virus type 1 (HIV-1) infection of nonproliferating human monocytes. J Exp Med, 1991. 174(6): p. 1477-82.

59. Fujiwara, T. and K. Mizuuchi, Retroviral DNA integration: structure of an integration intermediate. Cell, 1988. 54(4): p. 497-504.

60. Brown, P.O., et al., Retroviral integration: structure of the initial covalent product and its precursor, and a role for the viral IN protein. Proc Natl Acad Sci U S A, 1989. 86(8): p. 2525-9.

61. Craigie, R. and F.D. Bushman, HIV DNA Integration. Cold Spring Harb Perspect Med, 2012.

2(7): p. a006890.

62. Farnet, C.M. and W.A. Haseltine, Circularization of human immunodeficiency virus type 1 DNA in vitro. J Virol, 1991. 65(12): p. 6942-52.

63. Bukrinsky, M.I., et al., Active nuclear import of human immunodeficiency virus type 1 preintegration complexes. Proc Natl Acad Sci U S A, 1992. 89(14): p. 6580-4.

64. Wang, G.P., et al., HIV integration site selection: analysis by massively parallel pyrosequencing reveals association with epigenetic modifications. Genome Res, 2007. 17(8):

p. 1186-94.

65. Santoni, F.A., O. Hartley, and J. Luban, Deciphering the code for retroviral integration target site selection. PLoS Comput Biol, 2010. 6(11): p. e1001008.

66. Maertens, G., et al., LEDGF/p75 is essential for nuclear and chromosomal targeting of HIV-1 integrase in human cells. J Biol Chem, 2003. 278(35): p. 33528-39.

67. Berkhout, B., R.H. Silverman, and K.T. Jeang, Tat trans-activates the human immunodeficiency virus through a nascent RNA target. Cell, 1989. 59(2): p. 273-82.

68. Purcell, D.F. and M.A. Martin, Alternative splicing of human immunodeficiency virus type 1 mRNA modulates viral protein expression, replication, and infectivity. J Virol, 1993. 67(11): p.

6365-78.

69. Ocwieja, K.E., et al., Dynamic regulation of HIV-1 mRNA populations analyzed by single-molecule enrichment and long-read sequencing. Nucleic Acids Res, 2012. 40(20): p. 10345-55.

152

70. Emerman, M., R. Vazeux, and K. Peden, The rev gene product of the human immunodeficiency virus affects envelope-specific RNA localization. Cell, 1989. 57(7): p. 1155-71. 65. Malim, M.H., et al., Functional comparison of the Rev trans-activators encoded by different

primate immunodeficiency virus species. Proc Natl Acad Sci U S A, 1989. 86(21): p. 8222-6.

72. Chang, D.D. and P.A. Sharp, Regulation by HIV Rev depends upon recognition of splice sites.

Cell, 1989. 59(5): p. 789-95.

73. Strebel, K., Virus-host interactions: role of HIV proteins Vif, Tat, and Rev. AIDS, 2003. 17 Suppl 4: p. S25-34.

74. Berkowitz, R.D., et al., Retroviral nucleocapsid domains mediate the specific recognition of genomic viral RNAs by chimeric Gag polyproteins during RNA packaging in vivo. J Virol, 1995.

69(10): p. 6445-56.

75. De Guzman, R.N., et al., Structure of the HIV-1 nucleocapsid protein bound to the SL3 psi-RNA recognition element. Science, 1998. 279(5349): p. 384-8.

76. Moore, M.D., et al., Probing the HIV-1 genomic RNA trafficking pathway and dimerization by genetic recombination and single virion analyses. PLoS Pathog, 2009. 5(10): p. e1000627.

77. Morita, E., et al., ESCRT-III protein requirements for HIV-1 budding. Cell Host Microbe, 2011.

9(3): p. 235-42.

78. Sundquist, W.I. and H.G. Krausslich, HIV-1 assembly, budding, and maturation. Cold Spring Harb Perspect Med, 2012. 2(7): p. a006924.

79. Hladik, F. and M.J. McElrath, Setting the stage: host invasion by HIV. Nat Rev Immunol, 2008.

8(6): p. 447-57.

80. Keele, B.F., et al., Identification and characterization of transmitted and early founder virus envelopes in primary HIV-1 infection. Proc Natl Acad Sci U S A, 2008. 105(21): p. 7552-7.

81. Hu, J., M.B. Gardner, and C.J. Miller, Simian immunodeficiency virus rapidly penetrates the cervicovaginal mucosa after intravaginal inoculation and infects intraepithelial dendritic cells.

J Virol, 2000. 74(13): p. 6087-95.

82. Kaizu, M., et al., Repeated intravaginal inoculation with cell-associated simian immunodeficiency virus results in persistent infection of nonhuman primates. J Infect Dis, 2006. 194(7): p. 912-6.

83. Khanna, K.V., et al., Vaginal transmission of cell-associated HIV-1 in the mouse is blocked by a topical, membrane-modifying agent. J Clin Invest, 2002. 109(2): p. 205-11.

84. Zhu, T., et al., Genetic characterization of human immunodeficiency virus type 1 in blood and genital secretions: evidence for viral compartmentalization and selection during sexual transmission. J Virol, 1996. 70(5): p. 3098-107.

85. Fox, J. and S. Fidler, Sexual transmission of HIV-1. Antiviral Res, 2010. 85(1): p. 276-85.

86. Miller, C.J., M. McChesney, and P.F. Moore, Langerhans cells, macrophages and lymphocyte subsets in the cervix and vagina of rhesus macaques. Lab Invest, 1992. 67(5): p. 628-34.

87. Hu, Q., et al., Blockade of attachment and fusion receptors inhibits HIV-1 infection of human cervical tissue. J Exp Med, 2004. 199(8): p. 1065-75.

88. Geijtenbeek, T.B., et al., Identification of DC-SIGN, a novel dendritic cell-specific ICAM-3 receptor that supports primary immune responses. Cell, 2000. 100(5): p. 575-85.

89. McGreal, E.P., J.L. Miller, and S. Gordon, Ligand recognition by antigen-presenting cell C-type lectin receptors. Curr Opin Immunol, 2005. 17(1): p. 18-24.

90. Geijtenbeek, T.B., et al., DC-SIGN, a dendritic cell-specific HIV-1-binding protein that enhances trans-infection of T cells. Cell, 2000. 100(5): p. 587-97.

91. Gupta, P., et al., Memory CD4(+) T cells are the earliest detectable human immunodeficiency virus type 1 (HIV-1)-infected cells in the female genital mucosal tissue during HIV-1 transmission in an organ culture system. J Virol, 2002. 76(19): p. 9868-76.

92. Greenhead, P., et al., Parameters of human immunodeficiency virus infection of human cervical tissue and inhibition by vaginal virucides. J Virol, 2000. 74(12): p. 5577-86.

153

93. Sharova, N., et al., Macrophages archive HIV-1 virions for dissemination in trans. EMBO J, 2005. 24(13): p. 2481-9.

94. Groot, F., S. Welsch, and Q.J. Sattentau, Efficient HIV-1 transmission from macrophages to T cells across transient virological synapses. Blood, 2008. 111(9): p. 4660-3.

95. Cameron, P.U., et al., Dendritic cells exposed to human immunodeficiency virus type-1 transmit a vigorous cytopathic infection to CD4+ T cells. Science, 1992. 257(5068): p. 383-7.

95. Cameron, P.U., et al., Dendritic cells exposed to human immunodeficiency virus type-1 transmit a vigorous cytopathic infection to CD4+ T cells. Science, 1992. 257(5068): p. 383-7.

Dans le document HIV-1 innate immune detection and evasion (Page 145-166)