Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

A genome-wide CRISPR screen identifies a restricted set of HIV host dependency factors

Abstract

Host proteins are essential for HIV entry and replication and can be important nonviral therapeutic targets. Large-scale RNA interference (RNAi)-based screens have identified nearly a thousand candidate host factors, but there is little agreement among studies and few factors have been validated. Here we demonstrate that a genome-wide CRISPR-based screen identifies host factors in a physiologically relevant cell system. We identify five factors, including the HIV co-receptors CD4 and CCR5, that are required for HIV infection yet are dispensable for cellular proliferation and viability. Tyrosylprotein sulfotransferase 2 (TPST2) and solute carrier family 35 member B2 (SLC35B2) function in a common pathway to sulfate CCR5 on extracellular tyrosine residues, facilitating CCR5 recognition by the HIV envelope. Activated leukocyte cell adhesion molecule (ALCAM) mediates cell aggregation, which is required for cell-to-cell HIV transmission. We validated these pathways in primary human CD4+ T cells through Cas9-mediated knockout and antibody blockade. Our findings indicate that HIV infection and replication rely on a limited set of host-dispensable genes and suggest that these pathways can be studied for therapeutic intervention.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: A pooled, genome-wide CRISPR screen for HIV HDFs.
Figure 2: Loss of TPST2 or SLC35B2 confers strong protection against HIV infection and entry without compromising host cell viability.
Figure 3: SLC35B2 and TPST2 function in a common pathway for the sulfation of CCR5.
Figure 4: Homophilic ALCAM interactions are necessary for GXRCas9 cell aggregation.
Figure 5: Loss of ALCAM protects against cell-to-cell transmission of HIV through disruption of cell aggregation.
Figure 6: CRISPR-based approach for validation of HIV HDFs in primary human CD4+ T cells.

Similar content being viewed by others

References

  1. Friedrich, B.M. et al. Host factors mediating HIV-1 replication. Virus Res. 161, 101–114 (2011).

    CAS  PubMed  Google Scholar 

  2. Fätkenheuer, G. et al. Efficacy of short-term monotherapy with maraviroc, a new CCR5 antagonist, in patients infected with HIV-1. Nat. Med. 11, 1170–1172 (2005).

    PubMed  Google Scholar 

  3. Gulick, R.M. et al. Maraviroc for previously treated patients with R5 HIV-1 infection. N. Engl. J. Med. 359, 1429–1441 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Tebas, P. et al. Gene editing of CCR5 in autologous CD4 T cells of persons infected with HIV. N. Engl. J. Med. 370, 901–910 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Hütter, G. et al. Long-term control of HIV by CCR5 Δ32/Δ32 stem cell transplantation. N. Engl. J. Med. 360, 692–698 (2009).

    PubMed  Google Scholar 

  6. Glass, W.G. et al. CCR5 deficiency increases risk of symptomatic West Nile virus infection. J. Exp. Med. 203, 35–40 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Srivastava, A., Pandey, S.N., Choudhuri, G. & Mittal, B. CCR5-Δ32 polymorphism: associated with gallbladder cancer susceptibility. Scand. J. Immunol. 67, 516–522 (2008).

    CAS  PubMed  Google Scholar 

  8. Singh, H., Sachan, R., Jain, M. & Mittal, B. CCR5-Δ32 polymorphism and susceptibility to cervical cancer: association with early stage of cervical cancer. Oncol. Res. 17, 87–91 (2008).

    PubMed  Google Scholar 

  9. Eri, R. et al. CCR5-Δ32 mutation is strongly associated with primary sclerosing cholangitis. Genes Immun. 5, 444–450 (2004).

    CAS  PubMed  Google Scholar 

  10. König, R. et al. Global analysis of host–pathogen interactions that regulate early-stage HIV-1 replication. Cell 135, 49–60 (2008).

    PubMed  PubMed Central  Google Scholar 

  11. Brass, A.L. et al. Identification of host proteins required for HIV infection through a functional genomic screen. Science 319, 921–926 (2008).

    CAS  PubMed  Google Scholar 

  12. Zhou, H. et al. Genome-scale RNAi screen for host factors required for HIV replication. Cell Host Microbe 4, 495–504 (2008).

    CAS  PubMed  Google Scholar 

  13. Bassik, M.C. et al. Rapid creation and quantitative monitoring of high-coverage shRNA libraries. Nat. Methods 6, 443–445 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Shao, D.D. et al. ATARiS: computational quantification of gene suppression phenotypes from multisample RNAi screens. Genome Res. 23, 665–678 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Zhu, J. et al. Comprehensive identification of host modulators of HIV-1 replication using multiple orthologous RNAi reagents. Cell Rep. 9, 752–766 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Wang, T., Wei, J.J., Sabatini, D.M. & Lander, E.S. Genetic screens in human cells using the CRISPR–Cas9 system. Science 343, 80–84 (2014).

    CAS  PubMed  Google Scholar 

  17. Shalem, O. et al. Genome-scale CRISPR–Cas9 knockout screening in human cells. Science 343, 84–87 (2014).

    CAS  PubMed  Google Scholar 

  18. Wang, T. et al. Identification and characterization of essential genes in the human genome. Science 350, 1096–1101 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Hart, T. et al. High-resolution CRISPR screens reveal fitness genes and genotype-specific cancer liabilities. Cell 163, 1515–1526 (2015).

    CAS  PubMed  Google Scholar 

  20. Keele, B.F. et al. Identification and characterization of transmitted and early-founder virus envelopes in primary HIV-1 infection. Proc. Natl. Acad. Sci. USA 105, 7552–7557 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Ochsenbauer, C. et al. Generation of transmitted and founder HIV-1 infectious molecular clones, and characterization of their replication capacity in CD4 T lymphocytes and monocyte-derived macrophages. J. Virol. 86, 2715–2728 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Lai, M.M. Cellular factors in the transcription and replication of viral RNA genomes: a parallel to DNA-dependent RNA transcription. Virology 244, 1–12 (1998).

    CAS  PubMed  Google Scholar 

  23. Rolando, M. & Buchrieser, C. Legionella pneumophila type IV effectors hijack the transcription and translation machinery of the host cell. Trends Cell Biol. 24, 771–778 (2014).

    CAS  PubMed  Google Scholar 

  24. Egan, E.S. et al. Malaria. A forward genetic screen identifies erythrocyte CD55 as essential for Plasmodium falciparum invasion. Science 348, 711–714 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Cavrois, M., De Noronha, C. & Greene, W.C. A sensitive and specific enzyme-based assay detecting HIV-1 virion fusion in primary T lymphocytes. Nat. Biotechnol. 20, 1151–1154 (2002).

    CAS  PubMed  Google Scholar 

  26. Kamiyama, S. et al. Molecular cloning and identification of 3′-phosphoadenosine 5′-phosphosulfate transporter. J. Biol. Chem. 278, 25958–25963 (2003).

    CAS  PubMed  Google Scholar 

  27. Beisswanger, R. et al. Existence of distinct tyrosylprotein sulfotransferase genes: molecular characterization of tyrosylprotein sulfotransferase 2. Proc. Natl. Acad. Sci. USA 95, 11134–11139 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Baeuerle, P.A. & Huttner, W.B. Chlorate—a potent inhibitor of protein sulfation in intact cells. Biochem. Biophys. Res. Commun. 141, 870–877 (1986).

    CAS  PubMed  Google Scholar 

  29. Rosmarin, D.M. et al. Attachment of Chlamydia trachomatis L2 to host cells requires sulfation. Proc. Natl. Acad. Sci. USA 109, 10059–10064 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Connell, B.J. & Lortat-Jacob, H. Human immunodeficiency virus and heparan sulfate: from attachment to entry inhibition. Front. Immunol. 4, 385 (2013).

    PubMed  PubMed Central  Google Scholar 

  31. Farzan, M. et al. Tyrosine sulfation of the amino terminus of CCR5 facilitates HIV-1 entry. Cell 96, 667–676 (1999).

    CAS  PubMed  Google Scholar 

  32. Seibert, C., Cadene, M., Sanfiz, A., Chait, B.T. & Sakmar, T.P. Tyrosine sulfation of CCR5 N-terminal peptide by tyrosylprotein sulfotransferases 1 and 2 follows a discrete pattern and temporal sequence. Proc. Natl. Acad. Sci. USA 99, 11031–11036 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Wu, L. et al. Interaction of chemokine receptor CCR5 with its ligands: multiple domains for HIV-1 gp120 binding and a single domain for chemokine binding. J. Exp. Med. 186, 1373–1381 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Bowen, M.A. et al. Cloning, mapping, and characterization of activated leukocyte-cell adhesion molecule (ALCAM), a CD6 ligand. J. Exp. Med. 181, 2213–2220 (1995).

    CAS  PubMed  Google Scholar 

  35. Swart, G.W. Activated leukocyte cell adhesion molecule (CD166 or ALCAM): developmental and mechanistic aspects of cell clustering and cell migration. Eur. J. Cell Biol. 81, 313–321 (2002).

    CAS  PubMed  Google Scholar 

  36. Iolyeva, M. et al. Novel role for ALCAM in lymphatic network formation and function. FASEB J. 27, 978–990 (2013).

    CAS  PubMed  Google Scholar 

  37. Williams, D.W., Anastos, K., Morgello, S. & Berman, J.W. JAM-A and ALCAM are therapeutic targets to inhibit diapedesis across the BBB of CD14+CD16+ monocytes in HIV-infected individuals. J. Leukoc. Biol. 97, 401–412 (2015).

    PubMed  Google Scholar 

  38. Te Riet, J. et al. Dynamic coupling of ALCAM to the actin cortex strengthens cell adhesion to CD6. J. Cell Sci. 127, 1595–1606 (2014).

    CAS  PubMed  Google Scholar 

  39. van Kempen, L.C. et al. Molecular basis for the homophilic activated leukocyte cell adhesion molecule (ALCAM)–ALCAM interaction. J. Biol. Chem. 276, 25783–25790 (2001).

    CAS  PubMed  Google Scholar 

  40. Gartner, Z.J. & Bertozzi, C.R. Programmed assembly of 3-dimensional microtissues with defined cellular connectivity. Proc. Natl. Acad. Sci. USA 106, 4606–4610 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Schumann, K. et al. Generation of knock-in primary human T cells using Cas9 ribonucleoproteins. Proc. Natl. Acad. Sci. USA 112, 10437–10442 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Dustin, M.L. & Springer, T.A. T cell receptor cross-linking transiently stimulates adhesiveness through LFA-1. Nature 341, 619–624 (1989).

    CAS  PubMed  Google Scholar 

  43. Sabatos, C.A. et al. A synaptic basis for paracrine interleukin-2 signaling during homotypic T cell interaction. Immunity 29, 238–248 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Zumwalde, N.A., Domae, E., Mescher, M.F. & Shimizu, Y. ICAM-1-dependent homotypic aggregates regulate CD8 T cell effector function and differentiation during T cell activation. J. Immunol. 191, 3681–3693 (2013).

    CAS  PubMed  Google Scholar 

  45. Kaelin, W.G. Jr. Use and abuse of RNAi to study mammalian gene function. Science 337, 421–422 (2012).

    PubMed  PubMed Central  Google Scholar 

  46. Sasaki, N. et al. A mutation in Tpst2, encoding tyrosylprotein sulfotransferase, causes dwarfism associated with hypothyroidism. Mol. Endocrinol. 21, 1713–1721 (2007).

    CAS  PubMed  Google Scholar 

  47. Ding, Z.M. et al. Relative contribution of LFA-1 and Mac-1 to neutrophil adhesion and migration. J. Immunol. 163, 5029–5038 (1999).

    CAS  PubMed  Google Scholar 

  48. Ghosh, S., Chackerian, A.A., Parker, C.M., Ballantyne, C.M. & Behar, S.M. The LFA-1 adhesion molecule is required for protective immunity during pulmonary Mycobacterium tuberculosis infection. J. Immunol. 176, 4914–4922 (2006).

    CAS  PubMed  Google Scholar 

  49. Clément, A. et al. Regulation of zebrafish skeletogenesis by ext2 (dackel) and papst1 (pinscher). PLoS Genet. 4, e1000136 (2008).

    PubMed  PubMed Central  Google Scholar 

  50. Jäger, S. et al. Global landscape of HIV–human protein complexes. Nature 481, 365–370 (2011).

    PubMed  PubMed Central  Google Scholar 

  51. Singh, P.K. et al. LEDGF (p75) interacts with mRNA splicing factors and targets HIV-1 integration to highly spliced genes. Genes Dev. 29, 2287–2297 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Sowd, G.A. et al. A critical role for alternative polyadenylation factor CPSF6 in targeting HIV-1 integration to transcriptionally active chromatin. Proc. Natl. Acad. Sci. USA 113, E1054–E1063 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Desfosses, Y. et al. Regulation of human immunodeficiency virus type 1 gene expression by clade-specific Tat proteins. J. Virol. 79, 9180–9191 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Gardner, M.R. et al. AAV-expressed eCD4–Ig provides durable protection from multiple SHIV challenges. Nature 519, 87–91 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Seibert, C. et al. Sequential tyrosine sulfation of CXCR4 by tyrosylprotein sulfotransferases. Biochemistry 47, 11251–11262 (2008).

    CAS  PubMed  Google Scholar 

  56. Kajumo, F., Thompson, D.A., Guo, Y. & Dragic, T. Entry of R5X4 and X4 human immunodeficiency virus type 1 strains is mediated by negatively charged and tyrosine residues in the amino-terminal domain and the second extracellular loop of CXCR4. Virology 271, 240–247 (2000).

    CAS  PubMed  Google Scholar 

  57. Lin, G., Baribaud, F., Romano, J., Doms, R.W. & Hoxie, J.A. Identification of gp120-binding sites on CXCR4 by using CD4-independent human immunodeficiency virus type 2 Env proteins. J. Virol. 77, 931–942 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Ingulli, E., Mondino, A., Khoruts, A. & Jenkins, M.K. In vivo detection of dendritic cell antigen presentation to CD4+ T cells. J. Exp. Med. 185, 2133–2141 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Hommel, M. & Kyewski, B. Dynamic changes during the immune response in T cell–antigen-presenting cell clusters isolated from lymph nodes. J. Exp. Med. 197, 269–280 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Inaba, K., Witmer, M.D. & Steinman, R.M. Clustering of dendritic cells, helper T lymphocytes, and histocompatible B cells during primary antibody responses in vitro. J. Exp. Med. 160, 858–876 (1984).

    CAS  PubMed  Google Scholar 

  61. Sourisseau, M., Sol-Foulon, N., Porrot, F., Blanchet, F. & Schwartz, O. Inefficient human immunodeficiency virus replication in mobile lymphocytes. J. Virol. 81, 1000–1012 (2007).

    CAS  PubMed  Google Scholar 

  62. Agosto, L.M., Uchil, P.D. & Mothes, W. HIV cell-to-cell transmission: effects on pathogenesis and antiretroviral therapy. Trends Microbiol. 23, 289–295 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Chen, P., Hübner, W., Spinelli, M.A. & Chen, B.K. Predominant mode of human immunodeficiency virus transfer between T cells is mediated by sustained Env-dependent neutralization-resistant virological synapses. J. Virol. 81, 12582–12595 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Sigal, A. et al. Cell-to-cell spread of HIV permits ongoing replication despite antiretroviral therapy. Nature 477, 95–98 (2011).

    CAS  PubMed  Google Scholar 

  65. Doitsh, G. et al. Abortive HIV infection mediates CD4 T cell depletion and inflammation in human lymphoid tissue. Cell 143, 789–801 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Santangelo, P.J. et al. Whole-body immunoPET reveals active SIV dynamics in viremic and antiretroviral-therapy-treated macaques. Nat. Methods 12, 427–432 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Lorenzo-Redondo, R. et al. Persistent HIV-1 replication maintains the tissue reservoir during therapy. Nature 530, 51–56 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Yusuf-Makagiansar, H., Anderson, M.E., Yakovleva, T.V., Murray, J.S. & Siahaan, T.J. Inhibition of LFA-1–ICAM-1 and VLA-4–VCAM-1 as a therapeutic approach to inflammation and autoimmune diseases. Med. Res. Rev. 22, 146–167 (2002).

    CAS  PubMed  Google Scholar 

  69. Hourmant, M. et al. A randomized multicenter trial comparing leukocyte function-associated antigen 1 monoclonal antibody with rabbit antithymocyte globulin as induction treatment in first kidney transplantations. Transplantation 62, 1565–1570 (1996).

    CAS  PubMed  Google Scholar 

  70. Brockman, M.A., Tanzi, G.O., Walker, B.D. & Allen, T.M. Use of a novel GFP reporter cell line to examine replication capacity of CXCR4- and CCR5-tropic HIV-1 by flow cytometry. J. Virol. Methods 131, 134–142 (2006).

    CAS  PubMed  Google Scholar 

  71. Gibson, D.G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–345 (2009).

    CAS  PubMed  Google Scholar 

  72. Wang, T., Lander, E.S. & Sabatini, D.M. Large-scale single guide RNA library construction and use for CRISPR–Cas9-based genetic screens. Cold Spring Harb. Protoc. 2016, pdb.top086892 (2016).

    PubMed  PubMed Central  Google Scholar 

  73. McKinley, K.L. et al. The CENP–L–N complex forms a critical node in an integrated meshwork of interactions at the centromere–kinetochore interface. Mol. Cell 60, 886–898 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Weber, K., Bartsch, U., Stocking, C. & Fehse, B. A multicolor panel of novel lentiviral 'gene ontology' (LeGO) vectors for functional gene analysis. Mol. Ther. 16, 698–706 (2008).

    CAS  PubMed  Google Scholar 

  75. Salzberger, W. et al. Influence of glycosylation inhibition on the binding of KIR3DL1 to HLA-B*57:01. PLoS One 10, e0145324 (2015).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank the Ragon Institute Virology, Imaging, and Flow Cytometry cores, as well as the Center for Computational and Integrative Biology (CCIB) DNA Core Facility at Massachusetts General Hospital. We would like to thank A. McKeon, P. Jani, N.W. Hughes, and B.X. Liu for superb technical assistance, and A. Brass, G. Gaiha, and J.S. Park for helpful discussions. pMM310 was obtained through the AIDS Reagent Program, Division of AIDS, NIAID, NIH from M. Miller (Merck Research Laboratories). All plasmid reagents generated in this study have been deposited in Addgene. This work was supported by the Howard Hughes Medical Institute (D.M.S. and B.D.W.), the National Institutes of Health (grants CA103866 (D.M.S.), F31 CA189437 (T.W.), P50 GM082250 (A.M. and N.J.K.), U19 AI106754 (J.F.H. and N.J.K.), and P01 AI090935 (N.J.K.)), the National Human Genome Research Institute (grant 2U54HG003067-10; E.S.L.), the National Science Foundation (T.W.), the MIT Whitaker Health Sciences Fund (T.W.), the UCSF Sandler Fellowship (A.M.), a gift from J. Aronov (A.M.), the UCSF MPHD T32 Training Grant (J.F.H.), and the Deutsche Forschungsgemeinschaft (grant SCHU3020/2-1; K.S.). Support was also provided by NIH-funded Centers for AIDS Research (grant P30 AI027763, UCSF Center for AIDS Research (N.J.K.) and grant P30 AI060354, Harvard University Center for AIDS Research (B.D.W.)), which are supported by the following NIH co-funding and participating Institutes and Centers: NIAID, NCI, NICHD, NHLBI, NIDA, NIMH, NIA, FIC, and OAR. D.M.S. and B.D.W. are investigators of the Howard Hughes Medical Institute. R.J.P. is a Howard Hughes Medical Institute Research Fellow.

Author information

Authors and Affiliations

Authors

Contributions

R.J.P., T.W., A.M., N.J.K., D.M.S., E.S.L., N.H., and B.D.W. designed research. R.J.P., T.W., D.K., J.F.H., P.L.-M., B.M., K.S., H.Y., and K.M.K. conducted experiments. W.G.-B. and A.P.-T. contributed methods and reagents. R.J.P., T.W., D.K., and J.F.H. analyzed data. R.J.P., T.W., N.H., and B.D.W. wrote the manuscript.

Corresponding authors

Correspondence to David M Sabatini, Eric S Lander, Nir Hacohen or Bruce D Walker.

Ethics declarations

Competing interests

T.W., D.M.S., and E.S.L. are inventors on a patent application for functional genomics using the CRISPR–Cas system (US 15/141,348), T.W. and D.M.S. are founders of KSQ Therapeutics, a CRISPR functional genomics company, and D.M.S. is a scientific advisor for KSQ Therapeutics. A patent has been filed on the use of Cas9–RNPs to edit the genome of human primary T cells (A.M. and K.S.). A.M. serves as an advisor to Juno Therapeutics, and the laboratory of A.M. has had sponsored research agreements with Juno Therapeutics and Epinomics.

Integrated supplementary information

Supplementary Figure 1 Alternative gene ranking method reveals an identical set of HIV HDFs.

Gene scores were calculated by using mean log2 fold change in the abundance of all sgRNAs for each gene.

Supplementary Figure 2 mRNA expression analysis of SLC35B2 and TPST2 in GXRCas9 cells subcloned following CRISPR-based knockout.

Wild-type SLC35B2 and TPST2 mRNA expression levels in WT GXRCas9 cells and TPST2- and SLC35B2-knockout clones as assessed by qRT-PCR, using primers that overlap the sgRNA target site to selectively amplify wild-type cDNA. Error bars, s.d. from triplicate reactions.

Supplementary Figure 3 ALCAM-null GXRCas9 cells are protected from a multi-round, spreading JR-CSF infection.

Low MOI (MOI = 0.1) virus challenge. Six days following JR-CSF infection, viable, GFP cells were counted and cell number was normalized to that under a mock-infected condition. Error bars, s.d. from triplicate wells; *P < 0.0001, Welch’s t test.

Supplementary Figure 4 Validation of SLC35B2 as an HIV HDF in primary human CD4+ T cells.

Sulfation of surface CCR5 in primary CD4+ T cells following JR-CSF or Rejo.C challenge. Intracellular HIV Gag (p24) and total and sulfated surface CCR5 expression are shown as assessed by flow cytometry. Error bars, s.d. from triplicate wells; *P < 0.01, Welch’s t test. All P < 0.0001, except as follows: Donor 1 uninfected vs. Rejo.C p24+, P = 0.0005; Donor 2 uninfected vs. JRCSF p24, P = 0.0003; uninfected vs. Rejo.C p24, P = 0.0033; Rejo.C p24 vs. p24+, P = 0.0001.

Supplementary Figure 5 mRNA expression analysis of selected genes involved in T cell adhesion.

mRNA expression of ALCAM, LFA-1, and the ICAM family in primary CD4+ T cells and GXRCas9 cells as assessed by RNA sequencing.

Supplementary Figure 6 Antibody blockade of cell adhesion factors attenuates HIV spread in primary human CD4+ T cells.

(a) CRISPR-mediated knockout of the LFA-1 subunit (encoded by ITGAL) only blocks cell-to-cell transmission if donor and acceptor cells are both CD11a-null. (b,c) Cell-to-cell HIV transmission assay in primary CD4+ T cells following blockade with antibody to ICAM-1/LFA-1 (b) or CD45, as a control (c). The assay is set up as in Figure 5 except that donor cells are infected 36 h prior to co-culture and co-culture is for 2 d. Readout is by flow cytometry following intracellular staining for HIV Gag. Antibodies against ICAM-1, CD11a, and CD18 are added 15 min prior to co-culture. Error bars, s.d. from triplicate wells; *P < 0.01, Welch’s t test. P values were as follows: (a) n.s., P = 0.836; *P = 0.0036; (b) *P = 0.002; (c) n.s., P = 0.72.

Supplementary Figure 7 Loss of RELA, a candidate HDF identified by three previous RNAi screens, does not protect GXR cells from JR-CSF viral challenge.

Virus challenge assay (JR-CSF, MOI = 1) of GXRCas9 cells transduced with sgRELA (left) and immunoblot demonstrating depletion of RelA (right). RagC is a loading control. Error bars, s.d. from triplicate wells; n.s., P = 0.787, Welch’s t test.

Supplementary Figure 8 Essentiality analysis of screen hits and selected putative HIV HDFs.

Screen for essential genes in the Raji B cell line (Wang et al., 2015). For every gene in the human genome, the mean of the individual log2-transformed fold change values in the abundance of each of the sgRNAs targeting that gene is shown. Screen hits and selected putative HIV HDFs that are among the 10% most depleted (i.e., cell-essential) genes are highlighted.

Supplementary Figure 9 mRNA expression analysis of paralogous sulfation pathway genes.

(a,b) mRNA expression of TPST2, SLC35B2, PAPSS1, and select paralogs in GXRCas9 cells (a) and activated, primary CD4+ T cells and GXRCas9 cells (b) as assessed by RNA sequencing.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–9 and Supplementary Note. (PDF 1104 kb)

Supplementary Table 1

Genome-wide human sgRNA library annotation. (XLSX 10841 kb)

Supplementary Table 2

CRISPR gene scores. (XLSX 518 kb)

Supplementary Table 3

RNA sequencing analysis of GXRCas9 and activated, primary CD4+ T cells. (XLSX 1261 kb)

Supplementary Table 4

Nucleotide sequences. (XLSX 8 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Park, R., Wang, T., Koundakjian, D. et al. A genome-wide CRISPR screen identifies a restricted set of HIV host dependency factors. Nat Genet 49, 193–203 (2017). https://doi.org/10.1038/ng.3741

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3741

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research