Interferon-Regulated Expression of Cellular Splicing Factors Modulates Multiple Levels of HIV-1 Gene Expression and Replication
Abstract
:1. The High Complexity of the HIV-1 Genome Requires Exceptional Expression Strategies
2. HIV-1 Depends on Interactions with Host Cell Splicing Factors
2.1. Serine/Arginine-Rich Splicing Factors
2.2. Serine/Arginine-Rich Splicing Factor 1
2.3. Heterogeneous Nuclear Ribonucleoproteins
2.4. Heterogenous Nuclear Ribonucleoprotein A0
3. Cellular Splicing Factors SRSFs and hnRNPs Are Involved in Immunity and Underlie Interferon-Mediated Regulation
4. Impact of Interferon-Regulated Splicing Factors on LTR Transcription
5. Impact of Interferon-Regulated Splicing Factors on HIV-1 Alternative Splicing
6. Impact of Interferon-Regulated Splicing Factors on mRNA Trafficking
7. Impact of Interferon-Regulated Splicing Factors on Programmed Ribosomal Frameshifting
8. RNA-Binding Proteins and Splicing Factors Provide Novel Therapy Strategies
9. Conclusions
Author Contributions
Funding
Conflicts of Interest
References
- Ho, J.S.Y.; Zhu, Z.; Marazzi, I. Unconventional viral gene expression mechanisms as therapeutic targets. Nature 2021, 593, 362–371. [Google Scholar] [CrossRef] [PubMed]
- Ramdas, P.; Sahu, A.K.; Mishra, T.; Bhardwaj, V.; Chande, A. From Entry to Egress: Strategic Exploitation of the Cellular Processes by HIV-1. Front. Microbiol. 2020, 11, 559792. [Google Scholar] [CrossRef] [PubMed]
- Ostermann, P.N.; Ritchie, A.; Ptok, J.; Schaal, H. Let It Go: HIV-1 cis-Acting Repressive Sequences. J. Virol. 2021, 95, e0034221. [Google Scholar] [CrossRef]
- Hulo, C.; de Castro, E.; Masson, P.; Bougueleret, L.; Bairoch, A.; Xenarios, I.; Le Mercier, P. ViralZone: A knowledge resource to understand virus diversity. Nucleic Acids Res. 2011, 39, D576–D582. [Google Scholar] [CrossRef] [PubMed]
- Freed, E.O. HIV-1 replication. Somat. Cell Mol. Genet. 2001, 26, 13–33. [Google Scholar] [CrossRef] [PubMed]
- Rozina, A.; Anisenko, A.; Kikhai, T.; Silkina, M.; Gottikh, M. Complex Relationships between HIV-1 Integrase and Its Cellular Partners. Int. J. Mol. Sci. 2022, 23, 12341. [Google Scholar] [CrossRef] [PubMed]
- Lusic, M.; Siliciano, R.F. Nuclear landscape of HIV-1 infection and integration. Nat. Rev. Microbiol. 2017, 15, 69–82. [Google Scholar] [CrossRef] [PubMed]
- Savoret, J.; Mesnard, J.M.; Gross, A.; Chazal, N. Antisense Transcripts and Antisense Protein: A New Perspective on Human Immunodeficiency Virus Type 1. Front. Microbiol. 2020, 11, 625941. [Google Scholar] [CrossRef] [PubMed]
- Karn, J.; Stoltzfus, C.M. Transcriptional and posttranscriptional regulation of HIV-1 gene expression. Cold Spring Harb. Perspect. Med. 2012, 2, a006916. [Google Scholar] [CrossRef] [PubMed]
- Ott, M.; Geyer, M.; Zhou, Q. The control of HIV transcription: Keeping RNA polymerase II on track. Cell Host Microbe 2011, 10, 426–435. [Google Scholar] [CrossRef] [PubMed]
- Purcell, D.F.; Martin, M.A. Alternative splicing of human immunodeficiency virus type 1 mRNA modulates viral protein expression, replication, and infectivity. J. Virol. 1993, 67, 6365–6378. [Google Scholar] [CrossRef] [PubMed]
- Stoltzfus, C.M. Chapter 1. Regulation of HIV-1 alternative RNA splicing and its role in virus replication. Adv. Virus Res. 2009, 74, 1–40. [Google Scholar] [CrossRef]
- Corbeil, J.; Sheeter, D.; Genini, D.; Rought, S.; Leoni, L.; Du, P.; Ferguson, M.; Masys, D.R.; Welsh, J.B.; Fink, J.L.; et al. Temporal gene regulation during HIV-1 infection of human CD4+ T cells. Genome Res. 2001, 11, 1198–1204. [Google Scholar] [CrossRef] [PubMed]
- Wei, P.; Garber, M.E.; Fang, S.M.; Fischer, W.H.; Jones, K.A. A novel CDK9-associated C-type cyclin interacts directly with HIV-1 Tat and mediates its high-affinity, loop-specific binding to TAR RNA. Cell 1998, 92, 451–462. [Google Scholar] [CrossRef] [PubMed]
- Shilatifard, A.; Conaway, R.C.; Conaway, J.W. The RNA polymerase II elongation complex. Annu. Rev. Biochem. 2003, 72, 693–715. [Google Scholar] [CrossRef] [PubMed]
- Kao, S.Y.; Calman, A.F.; Luciw, P.A.; Peterlin, B.M. Anti-termination of transcription within the long terminal repeat of HIV-1 by tat gene product. Nature 1987, 330, 489–493. [Google Scholar] [CrossRef]
- Tahirov, T.H.; Babayeva, N.D.; Varzavand, K.; Cooper, J.J.; Sedore, S.C.; Price, D.H. Crystal structure of HIV-1 Tat complexed with human P-TEFb. Nature 2010, 465, 747–751. [Google Scholar] [CrossRef] [PubMed]
- Demarchi, F.; d‘Adda di Fagagna, F.; Falaschi, A.; Giacca, M. Activation of transcription factor NF-kappaB by the Tat protein of human immunodeficiency virus type 1. J. Virol. 1996, 70, 4427–4437. [Google Scholar] [CrossRef] [PubMed]
- Westendorp, M.O.; Shatrov, V.A.; Schulze-Osthoff, K.; Frank, R.; Kraft, M.; Los, M.; Krammer, P.H.; Droge, W.; Lehmann, V. HIV-1 Tat potentiates TNF-induced NF-kappa B activation and cytotoxicity by altering the cellular redox state. EMBO J. 1995, 14, 546–554. [Google Scholar] [CrossRef] [PubMed]
- Pahl, H.L. Activators and target genes of Rel/NF-kappaB transcription factors. Oncogene 1999, 18, 6853–6866. [Google Scholar] [CrossRef]
- Cochrane, A.; Kramer, R.; Ruben, S.; Levine, J.; Rosen, C.A. The human immunodeficiency virus rev protein is a nuclear phosphoprotein. Virology 1989, 171, 264–266. [Google Scholar] [CrossRef] [PubMed]
- Malim, M.H.; Bohnlein, S.; Hauber, J.; Cullen, B.R. Functional dissection of the HIV-1 Rev trans-activator--derivation of a trans-dominant repressor of Rev function. Cell 1989, 58, 205–214. [Google Scholar] [CrossRef] [PubMed]
- Fischer, U.; Meyer, S.; Teufel, M.; Heckel, C.; Luhrmann, R.; Rautmann, G. Evidence that HIV-1 Rev directly promotes the nuclear export of unspliced RNA. EMBO J. 1994, 13, 4105–4112. [Google Scholar] [CrossRef] [PubMed]
- Truman, C.T.; Jarvelin, A.; Davis, I.; Castello, A. HIV Revisited. Open Biol. 2020, 10, 200320. [Google Scholar] [CrossRef] [PubMed]
- Cochrane, A.W.; Perkins, A.; Rosen, C.A. Identification of sequences important in the nucleolar localization of human immunodeficiency virus Rev: Relevance of nucleolar localization to function. J. Virol. 1990, 64, 881–885. [Google Scholar] [CrossRef] [PubMed]
- Henderson, B.R.; Percipalle, P. Interactions between HIV Rev and nuclear import and export factors: The Rev nuclear localisation signal mediates specific binding to human importin-beta. J. Mol. Biol. 1997, 274, 693–707. [Google Scholar] [CrossRef] [PubMed]
- Arrigo, S.J.; Chen, I.S. Rev is necessary for translation but not cytoplasmic accumulation of HIV-1 vif, vpr, and env/vpu 2 RNAs. Genes Dev. 1991, 5, 808–819. [Google Scholar] [CrossRef] [PubMed]
- Felber, B.K.; Hadzopoulou-Cladaras, M.; Cladaras, C.; Copeland, T.; Pavlakis, G.N. rev protein of human immunodeficiency virus type 1 affects the stability and transport of the viral mRNA. Proc. Natl. Acad. Sci. USA 1989, 86, 1495–1499. [Google Scholar] [CrossRef] [PubMed]
- Gonzalez-Hernandez, M.J.; Cavalcoli, J.D.; Sartor, M.A.; Contreras-Galindo, R.; Meng, F.; Dai, M.; Dube, D.; Saha, A.K.; Gitlin, S.D.; Omenn, G.S.; et al. Regulation of the human endogenous retrovirus K (HML-2) transcriptome by the HIV-1 Tat protein. J. Virol. 2014, 88, 8924–8935. [Google Scholar] [CrossRef] [PubMed]
- Stove, V.; Van de Walle, I.; Naessens, E.; Coene, E.; Stove, C.; Plum, J.; Verhasselt, B. Human immunodeficiency virus Nef induces rapid internalization of the T-cell coreceptor CD8alphabeta. J. Virol. 2005, 79, 11422–11433. [Google Scholar] [CrossRef] [PubMed]
- Schwartz, O.; Marechal, V.; Le Gall, S.; Lemonnier, F.; Heard, J.M. Endocytosis of major histocompatibility complex class I molecules is induced by the HIV-1 Nef protein. Nat. Med. 1996, 2, 338–342. [Google Scholar] [CrossRef] [PubMed]
- Schindler, M.; Wurfl, S.; Benaroch, P.; Greenough, T.C.; Daniels, R.; Easterbrook, P.; Brenner, M.; Munch, J.; Kirchhoff, F. Down-modulation of mature major histocompatibility complex class II and up-regulation of invariant chain cell surface expression are well-conserved functions of human and simian immunodeficiency virus nef alleles. J. Virol. 2003, 77, 10548–10556. [Google Scholar] [CrossRef] [PubMed]
- Wang, J.K.; Kiyokawa, E.; Verdin, E.; Trono, D. The Nef protein of HIV-1 associates with rafts and primes T cells for activation. Proc. Natl. Acad. Sci. USA 2000, 97, 394–399. [Google Scholar] [CrossRef] [PubMed]
- Raney, A.; Kuo, L.S.; Baugh, L.L.; Foster, J.L.; Garcia, J.V. Reconstitution and molecular analysis of an active human immunodeficiency virus type 1 Nef/p21-activated kinase 2 complex. J. Virol. 2005, 79, 12732–12741. [Google Scholar] [CrossRef] [PubMed]
- Campbell, E.M.; Nunez, R.; Hope, T.J. Disruption of the actin cytoskeleton can complement the ability of Nef to enhance human immunodeficiency virus type 1 infectivity. J. Virol. 2004, 78, 5745–5755. [Google Scholar] [CrossRef] [PubMed]
- Maziere, J.C.; Landureau, J.C.; Giral, P.; Auclair, M.; Fall, L.; Lachgar, A.; Achour, A.; Zagury, D. Lovastatin inhibits HIV-1 expression in H9 human T lymphocytes cultured in cholesterol-poor medium. Biomed. Pharmacother. 1994, 48, 63–67. [Google Scholar] [CrossRef] [PubMed]
- Harris, R.S.; Bishop, K.N.; Sheehy, A.M.; Craig, H.M.; Petersen-Mahrt, S.K.; Watt, I.N.; Neuberger, M.S.; Malim, M.H. DNA deamination mediates innate immunity to retroviral infection. Cell 2003, 113, 803–809. [Google Scholar] [CrossRef] [PubMed]
- Sheehy, A.M.; Gaddis, N.C.; Choi, J.D.; Malim, M.H. Isolation of a human gene that inhibits HIV-1 infection and is suppressed by the viral Vif protein. Nature 2002, 418, 646–650. [Google Scholar] [CrossRef] [PubMed]
- Sheehy, A.M.; Gaddis, N.C.; Malim, M.H. The antiretroviral enzyme APOBEC3G is degraded by the proteasome in response to HIV-1 Vif. Nat. Med. 2003, 9, 1404–1407. [Google Scholar] [CrossRef]
- Lubow, J.; Collins, K.L. Vpr Is a VIP: HIV Vpr and Infected Macrophages Promote Viral Pathogenesis. Viruses 2020, 12, 809. [Google Scholar] [CrossRef] [PubMed]
- Vanegas-Torres, C.A.; Schindler, M. HIV-1 Vpr Functions in Primary CD4(+) T Cells. Viruses 2024, 16, 420. [Google Scholar] [CrossRef] [PubMed]
- West, C.M. Evolutionary and functional implications of the complex glycosylation of Skp1, a cytoplasmic/nuclear glycoprotein associated with polyubiquitination. Cell. Mol. Life Sci. 2003, 60, 229–240. [Google Scholar] [CrossRef] [PubMed]
- Sertznig, H.; Hillebrand, F.; Erkelenz, S.; Schaal, H.; Widera, M. Behind the scenes of HIV-1 replication: Alternative splicing as the dependency factor on the quiet. Virology 2018, 516, 176–188. [Google Scholar] [CrossRef] [PubMed]
- Ocwieja, K.E.; Sherrill-Mix, S.; Mukherjee, R.; Custers-Allen, R.; David, P.; Brown, M.; Wang, S.; Link, D.R.; Olson, J.; Travers, K.; et al. Dynamic regulation of HIV-1 mRNA populations analyzed by single-molecule enrichment and long-read sequencing. Nucleic Acids Res. 2012, 40, 10345–10355. [Google Scholar] [CrossRef] [PubMed]
- Schwartz, S.; Felber, B.K.; Benko, D.M.; Fenyo, E.M.; Pavlakis, G.N. Cloning and functional analysis of multiply spliced mRNA species of human immunodeficiency virus type 1. J. Virol. 1990, 64, 2519–2529. [Google Scholar] [CrossRef] [PubMed]
- Sertznig, H.; Roesmann, F.; Wilhelm, A.; Heininger, D.; Bleekmann, B.; Elsner, C.; Santiago, M.; Schuhenn, J.; Karakoese, Z.; Benatzy, Y.; et al. SRSF1 acts as an IFN-I-regulated cellular dependency factor decisively affecting HIV-1 post-integration steps. Front. Immunol. 2022, 13, 935800. [Google Scholar] [CrossRef] [PubMed]
- Roesmann, F.; Sertznig, H.; Klaassen, K.; Wilhelm, A.; Heininger, D.; Elsner, C.; Marschalek, R.; Santiago, M.; Esser, S.; Sutter, K.; et al. The interferon regulated host factor hnRNP A0 modulates HIV-1 replication by interference with LTR-activity, mRNA trafficking, and programmed ribosomal frameshifting. J. Virol. 2024, article in press. [Google Scholar] [CrossRef]
- Fu, X.D. The superfamily of arginine/serine-rich splicing factors. RNA 1995, 1, 663–680. [Google Scholar] [PubMed]
- Howard, J.M.; Sanford, J.R. The RNAissance family: SR proteins as multifaceted regulators of gene expression. Wiley Interdiscip. Rev. RNA 2015, 6, 93–110. [Google Scholar] [CrossRef] [PubMed]
- Manley, J.L.; Krainer, A.R. A rational nomenclature for serine/arginine-rich protein splicing factors (SR proteins). Genes Dev. 2010, 24, 1073–1074. [Google Scholar] [CrossRef] [PubMed]
- Shepard, P.J.; Hertel, K.J. The SR protein family. Genome Biol. 2009, 10, 242. [Google Scholar] [CrossRef] [PubMed]
- Uhlen, M.; Fagerberg, L.; Hallstrom, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, A.; Kampf, C.; Sjostedt, E.; Asplund, A.; et al. Proteomics. Tissue-based map of the human proteome. Science 2015, 347, 1260419. [Google Scholar] [CrossRef] [PubMed]
- Anko, M.L. Regulation of gene expression programmes by serine-arginine rich splicing factors. Semin. Cell Dev. Biol. 2014, 32, 11–21. [Google Scholar] [CrossRef] [PubMed]
- Kannan, N.; Neuwald, A.F. Evolutionary constraints associated with functional specificity of the CMGC protein kinases MAPK, CDK, GSK, SRPK, DYRK, and CK2alpha. Protein Sci. A Publ. Protein Soc. 2004, 13, 2059–2077. [Google Scholar] [CrossRef] [PubMed]
- Liu, Q.; Dreyfuss, G. In vivo and in vitro arginine methylation of RNA-binding proteins. Mol. Cell. Biol. 1995, 15, 2800–2808. [Google Scholar] [CrossRef] [PubMed]
- Siebel, C.W.; Guthrie, C. The essential yeast RNA binding protein Np13p is methylated. Proc. Natl. Acad. Sci. USA 1996, 93, 13641–13646. [Google Scholar] [CrossRef] [PubMed]
- Yun, C.Y.; Fu, X.D. Conserved SR protein kinase functions in nuclear import and its action is counteracted by arginine methylation in Saccharomyces cerevisiae. J. Cell Biol. 2000, 150, 707–718. [Google Scholar] [CrossRef] [PubMed]
- Choudhary, C.; Kumar, C.; Gnad, F.; Nielsen, M.L.; Rehman, M.; Walther, T.C.; Olsen, J.V.; Mann, M. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science 2009, 325, 834–840. [Google Scholar] [CrossRef] [PubMed]
- Loomis, R.J.; Naoe, Y.; Parker, J.B.; Savic, V.; Bozovsky, M.R.; Macfarlan, T.; Manley, J.L.; Chakravarti, D. Chromatin binding of SRp20 and ASF/SF2 and dissociation from mitotic chromosomes is modulated by histone H3 serine 10 phosphorylation. Mol. Cell 2009, 33, 450–461. [Google Scholar] [CrossRef] [PubMed]
- Ji, X.; Zhou, Y.; Pandit, S.; Huang, J.; Li, H.; Lin, C.Y.; Xiao, R.; Burge, C.B.; Fu, X.D. SR proteins collaborate with 7SK and promoter-associated nascent RNA to release paused polymerase. Cell 2013, 153, 855–868. [Google Scholar] [CrossRef]
- Huang, Y.; Steitz, J.A. Splicing factors SRp20 and 9G8 promote the nucleocytoplasmic export of mRNA. Mol. Cell 2001, 7, 899–905. [Google Scholar] [CrossRef] [PubMed]
- Huang, Y.; Gattoni, R.; Stevenin, J.; Steitz, J.A. SR splicing factors serve as adapter proteins for TAP-dependent mRNA export. Mol. Cell 2003, 11, 837–843. [Google Scholar] [CrossRef] [PubMed]
- Huang, Y.; Steitz, J.A. SRprises along a messenger’s journey. Mol. Cell 2005, 17, 613–615. [Google Scholar] [CrossRef] [PubMed]
- Li, X.; Manley, J.L. Inactivation of the SR protein splicing factor ASF/SF2 results in genomic instability. Cell 2005, 122, 365–378. [Google Scholar] [CrossRef] [PubMed]
- Xiao, S.H.; Manley, J.L. Phosphorylation of the ASF/SF2 RS domain affects both protein-protein and protein-RNA interactions and is necessary for splicing. Genes Dev. 1997, 11, 334–344. [Google Scholar] [CrossRef] [PubMed]
- Lai, M.C.; Lin, R.I.; Tarn, W.Y. Differential effects of hyperphosphorylation on splicing factor SRp55. Biochem. J. 2003, 371, 937–945. [Google Scholar] [CrossRef] [PubMed]
- Colwill, K.; Pawson, T.; Andrews, B.; Prasad, J.; Manley, J.L.; Bell, J.C.; Duncan, P.I. The Clk/Sty protein kinase phosphorylates SR splicing factors and regulates their intranuclear distribution. EMBO J. 1996, 15, 265–275. [Google Scholar] [CrossRef] [PubMed]
- Gui, J.F.; Lane, W.S.; Fu, X.D. A serine kinase regulates intracellular localization of splicing factors in the cell cycle. Nature 1994, 369, 678–682. [Google Scholar] [CrossRef]
- Zhou, Z.; Fu, X.D. Regulation of splicing by SR proteins and SR protein-specific kinases. Chromosoma 2013, 122, 191–207. [Google Scholar] [CrossRef] [PubMed]
- Czubaty, A.; Piekielko-Witkowska, A. Protein kinases that phosphorylate splicing factors: Roles in cancer development, progression and possible therapeutic options. Int. J. Biochem. Cell Biol. 2017, 91, 102–115. [Google Scholar] [CrossRef]
- Duncan, P.I.; Stojdl, D.F.; Marius, R.M.; Bell, J.C. In vivo regulation of alternative pre-mRNA splicing by the Clk1 protein kinase. Mol. Cell. Biol. 1997, 17, 5996–6001. [Google Scholar] [CrossRef] [PubMed]
- Prasad, J.; Colwill, K.; Pawson, T.; Manley, J.L. The protein kinase Clk/Sty directly modulates SR protein activity: Both hyper- and hypophosphorylation inhibit splicing. Mol. Cell. Biol. 1999, 19, 6991–7000. [Google Scholar] [CrossRef] [PubMed]
- Dahal, S.; Clayton, K.; Cabral, T.; Cheng, R.; Jahanshahi, S.; Ahmed, C.; Koirala, A.; Villasmil Ocando, A.; Malty, R.; Been, T.; et al. On a path toward a broad-spectrum anti-viral: Inhibition of HIV-1 and coronavirus replication by SR kinase inhibitor harmine. J. Virol. 2023, 97, e0039623. [Google Scholar] [CrossRef] [PubMed]
- Wong, R.; Balachandran, A.; Mao, A.Y.; Dobson, W.; Gray-Owen, S.; Cochrane, A. Differential effect of CLK SR Kinases on HIV-1 gene expression: Potential novel targets for therapy. Retrovirology 2011, 8, 47. [Google Scholar] [CrossRef] [PubMed]
- Ge, H.; Manley, J.L. A protein factor, ASF, controls cell-specific alternative splicing of SV40 early pre-mRNA in vitro. Cell 1990, 62, 25–34. [Google Scholar] [CrossRef] [PubMed]
- Krainer, A.R.; Conway, G.C.; Kozak, D. The essential pre-mRNA splicing factor SF2 influences 5′ splice site selection by activating proximal sites. Cell 1990, 62, 35–42. [Google Scholar] [CrossRef] [PubMed]
- Liu, X.; Kraus, W.L.; Bai, X. Ready, pause, go: Regulation of RNA polymerase II pausing and release by cellular signaling pathways. Trends Biochem. Sci. 2015, 40, 516–525. [Google Scholar] [CrossRef] [PubMed]
- Caputi, M.; Freund, M.; Kammler, S.; Asang, C.; Schaal, H. A bidirectional SF2/ASF- and SRp40-dependent splicing enhancer regulates human immunodeficiency virus type 1 rev, env, vpu, and nef gene expression. J. Virol. 2004, 78, 6517–6526. [Google Scholar] [CrossRef] [PubMed]
- Kammler, S.; Otte, M.; Hauber, I.; Kjems, J.; Hauber, J.; Schaal, H. The strength of the HIV-1 3′ splice sites affects Rev function. Retrovirology 2006, 3, 89. [Google Scholar] [CrossRef] [PubMed]
- Zhou, H.; Bulek, K.; Li, X.; Herjan, T.; Yu, M.; Qian, W.; Wang, H.; Zhou, G.; Chen, X.; Yang, H.; et al. IRAK2 directs stimulus-dependent nuclear export of inflammatory mRNAs. Elife 2017, 6, e29630. [Google Scholar] [CrossRef] [PubMed]
- Michlewski, G.; Sanford, J.R.; Cáceres, J.F. The splicing factor SF2/ASF regulates translation initiation by enhancing phosphorylation of 4E-BP1. Mol. Cell 2008, 30, 179–189. [Google Scholar] [CrossRef] [PubMed]
- Sanford, J.R.; Gray, N.K.; Beckmann, K.; Caceres, J.F. A novel role for shuttling SR proteins in mRNA translation. Genes Dev. 2004, 18, 755–768. [Google Scholar] [CrossRef] [PubMed]
- Aznarez, I.; Nomakuchi, T.T.; Tetenbaum-Novatt, J.; Rahman, M.A.; Fregoso, O.; Rees, H.; Krainer, A.R. Mechanism of Nonsense-Mediated mRNA Decay Stimulation by Splicing Factor SRSF1. Cell Rep. 2018, 23, 2186–2198. [Google Scholar] [CrossRef] [PubMed]
- Sun, S.; Zhang, Z.; Sinha, R.; Karni, R.; Krainer, A.R. SF2/ASF autoregulation involves multiple layers of post-transcriptional and translational control. Nat. Struct. Mol. Biol. 2010, 17, 306–312. [Google Scholar] [CrossRef] [PubMed]
- Zhang, Z.; Krainer, A.R. Involvement of SR proteins in mRNA surveillance. Mol. Cell 2004, 16, 597–607. [Google Scholar] [CrossRef] [PubMed]
- Tacke, R.; Manley, J.L. The human splicing factors ASF/SF2 and SC35 possess distinct, functionally significant RNA binding specificities. EMBO J. 1995, 14, 3540–3551. [Google Scholar] [CrossRef] [PubMed]
- Sanford, J.R.; Wang, X.; Mort, M.; Vanduyn, N.; Cooper, D.N.; Mooney, S.D.; Edenberg, H.J.; Liu, Y. Splicing factor SFRS1 recognizes a functionally diverse landscape of RNA transcripts. Genome Res. 2009, 19, 381–394. [Google Scholar] [CrossRef] [PubMed]
- Wang, J.; Smith, P.J.; Krainer, A.R.; Zhang, M.Q. Distribution of SR protein exonic splicing enhancer motifs in human protein-coding genes. Nucleic Acids Res. 2005, 33, 5053–5062. [Google Scholar] [CrossRef] [PubMed]
- Jeong, S. SR Proteins: Binders, Regulators, and Connectors of RNA. Mol. Cells 2017, 40, 1–9. [Google Scholar] [CrossRef] [PubMed]
- Jiang, L.; Huang, J.; Higgs, B.W.; Hu, Z.; Xiao, Z.; Yao, X.; Conley, S.; Zhong, H.; Liu, Z.; Brohawn, P.; et al. Genomic Landscape Survey Identifies SRSF1 as a Key Oncodriver in Small Cell Lung Cancer. PLoS Genet. 2016, 12, e1005895. [Google Scholar] [CrossRef] [PubMed]
- Karni, R.; de Stanchina, E.; Lowe, S.W.; Sinha, R.; Mu, D.; Krainer, A.R. The gene encoding the splicing factor SF2/ASF is a proto-oncogene. Nat. Struct. Mol. Biol. 2007, 14, 185–193. [Google Scholar] [CrossRef] [PubMed]
- Das, S.; Krainer, A.R. Emerging functions of SRSF1, splicing factor and oncoprotein, in RNA metabolism and cancer. Mol. Cancer Res. 2014, 12, 1195–1204. [Google Scholar] [CrossRef] [PubMed]
- Staffa, A.; Cochrane, A. Identification of positive and negative splicing regulatory elements within the terminal tat-rev exon of human immunodeficiency virus type 1. Mol. Cell. Biol. 1995, 15, 4597–4605. [Google Scholar] [CrossRef] [PubMed]
- Jablonski, J.A.; Caputi, M. Role of cellular RNA processing factors in human immunodeficiency virus type 1 mRNA metabolism, replication, and infectivity. J. Virol. 2009, 83, 981–992. [Google Scholar] [CrossRef] [PubMed]
- Jacquenet, S.; Decimo, D.; Muriaux, D.; Darlix, J.L. Dual effect of the SR proteins ASF/SF2, SC35 and 9G8 on HIV-1 RNA splicing and virion production. Retrovirology 2005, 2, 33. [Google Scholar] [CrossRef] [PubMed]
- Ropers, D.; Ayadi, L.; Gattoni, R.; Jacquenet, S.; Damier, L.; Branlant, C.; Stevenin, J. Differential effects of the SR proteins 9G8, SC35, ASF/SF2, and SRp40 on the utilization of the A1 to A5 splicing sites of HIV-1 RNA. J. Biol. Chem. 2004, 279, 29963–29973. [Google Scholar] [CrossRef] [PubMed]
- Koizumi, J.; Okamoto, Y.; Onogi, H.; Mayeda, A.; Krainer, A.R.; Hagiwara, M. The subcellular localization of SF2/ASF is regulated by direct interaction with SR protein kinases (SRPKs). J. Biol. Chem. 1999, 274, 11125–11131. [Google Scholar] [CrossRef] [PubMed]
- Misteli, T.; Spector, D.L. Serine/threonine phosphatase 1 modulates the subnuclear distribution of pre-mRNA splicing factors. Mol. Biol. Cell 1996, 7, 1559–1572. [Google Scholar] [CrossRef] [PubMed]
- Novoyatleva, T.; Heinrich, B.; Tang, Y.; Benderska, N.; Butchbach, M.E.; Lorson, C.L.; Lorson, M.A.; Ben-Dov, C.; Fehlbaum, P.; Bracco, L.; et al. Protein phosphatase 1 binds to the RNA recognition motif of several splicing factors and regulates alternative pre-mRNA processing. Hum. Mol. Genet. 2008, 17, 52–70. [Google Scholar] [CrossRef] [PubMed]
- Sanford, J.R.; Ellis, J.D.; Cazalla, D.; Caceres, J.F. Reversible phosphorylation differentially affects nuclear and cytoplasmic functions of splicing factor 2/alternative splicing factor. Proc. Natl. Acad. Sci. USA 2005, 102, 15042–15047. [Google Scholar] [CrossRef] [PubMed]
- Sinha, R.; Allemand, E.; Zhang, Z.; Karni, R.; Myers, M.P.; Krainer, A.R. Arginine methylation controls the subcellular localization and functions of the oncoprotein splicing factor SF2/ASF. Mol. Cell. Biol. 2010, 30, 2762–2774. [Google Scholar] [CrossRef] [PubMed]
- Paz, S.; Ritchie, A.; Mauer, C.; Caputi, M. The RNA binding protein SRSF1 is a master switch of gene expression and regulation in the immune system. Cytokine Growth Factor Rev. 2021, 57, 19–26. [Google Scholar] [CrossRef] [PubMed]
- Mo, S.; Ji, X.; Fu, X.D. Unique role of SRSF2 in transcription activation and diverse functions of the SR and hnRNP proteins in gene expression regulation. Transcription 2013, 4, 251–259. [Google Scholar] [CrossRef] [PubMed]
- Yang, J.F.; You, J. Regulation of Polyomavirus Transcription by Viral and Cellular Factors. Viruses 2020, 12, 1072. [Google Scholar] [CrossRef] [PubMed]
- Dreyfuss, G.; Matunis, M.J.; Piñol-Roma, S.; Burd, C.G. hnRNP proteins and the biogenesis of mRNA. Annu. Rev. Biochem. 1993, 62, 289–321. [Google Scholar] [CrossRef] [PubMed]
- Geuens, T.; Bouhy, D.; Timmerman, V. The hnRNP family: Insights into their role in health and disease. Hum. Genet. 2016, 135, 851–867. [Google Scholar] [CrossRef] [PubMed]
- Hamilton, B.J.; Burns, C.M.; Nichols, R.C.; Rigby, W.F. Modulation of AUUUA response element binding by heterogeneous nuclear ribonucleoprotein A1 in human T lymphocytes. The roles of cytoplasmic location, transcription, and phosphorylation. J. Biol. Chem. 1997, 272, 28732–28741. [Google Scholar] [CrossRef] [PubMed]
- Görlach, M.; Wittekind, M.; Beckman, R.A.; Mueller, L.; Dreyfuss, G. Interaction of the RNA-binding domain of the hnRNP C proteins with RNA. EMBO J. 1992, 11, 3289–3295. [Google Scholar] [CrossRef] [PubMed]
- Dominguez, C.; Fisette, J.F.; Chabot, B.; Allain, F.H. Structural basis of G-tract recognition and encaging by hnRNP F quasi-RRMs. Nat. Struct. Mol. Biol. 2010, 17, 853–861. [Google Scholar] [CrossRef] [PubMed]
- Samatanga, B.; Dominguez, C.; Jelesarov, I.; Allain, F.H. The high kinetic stability of a G-quadruplex limits hnRNP F qRRM3 binding to G-tract RNA. Nucleic Acids Res. 2013, 41, 2505–2516. [Google Scholar] [CrossRef] [PubMed]
- Burd, C.G.; Dreyfuss, G. Conserved structures and diversity of functions of RNA-binding proteins. Science 1994, 265, 615–621. [Google Scholar] [CrossRef] [PubMed]
- Kirking, D.M.; Thomas, J.W.; Ascione, F.J.; Boyd, E.L. Detecting and preventing adverse drug interactions: The potential contribution of computers in pharmacies. Soc. Sci. Med. 1986, 22, 1–8. [Google Scholar] [CrossRef] [PubMed]
- Marenda, M.; Lazarova, E.; Gilbert, N. The role of SAF-A/hnRNP U in regulating chromatin structure. Curr. Opin. Genet. Dev. 2022, 72, 38–44. [Google Scholar] [CrossRef] [PubMed]
- Shishkin, S.S.; Kovalev, L.I.; Pashintseva, N.V.; Kovaleva, M.A.; Lisitskaya, K. Heterogeneous Nuclear Ribonucleoproteins Involved in the Functioning of Telomeres in Malignant Cells. Int. J. Mol. Sci. 2019, 20, 745. [Google Scholar] [CrossRef] [PubMed]
- Ford, L.P.; Wright, W.E.; Shay, J.W. A model for heterogeneous nuclear ribonucleoproteins in telomere and telomerase regulation. Oncogene 2002, 21, 580–583. [Google Scholar] [CrossRef] [PubMed]
- Liu, Y.; Abula, A.; Xiao, H.; Guo, H.; Li, T.; Zheng, L.; Chen, B.; Nguyen, H.C.; Ji, X. Structural Insight Into hnRNP A2/B1 Homodimerization and DNA Recognition. J. Mol. Biol. 2023, 435, 167920. [Google Scholar] [CrossRef] [PubMed]
- Cartegni, L.; Maconi, M.; Morandi, E.; Cobianchi, F.; Riva, S.; Biamonti, G. hnRNP A1 selectively interacts through its Gly-rich domain with different RNA-binding proteins. J. Mol. Biol. 1996, 259, 337–348. [Google Scholar] [CrossRef] [PubMed]
- Biamonti, G.; Riva, S. New insights into the auxiliary domains of eukaryotic RNA binding proteins. FEBS Lett. 1994, 340, 1–8. [Google Scholar] [CrossRef]
- Dreyfuss, G.; Kim, V.N.; Kataoka, N. Messenger-RNA-binding proteins and the messages they carry. Nat. Rev. Mol. Cell Biol. 2002, 3, 195–205. [Google Scholar] [CrossRef] [PubMed]
- Busch, A.; Hertel, K.J. Evolution of SR protein and hnRNP splicing regulatory factors. Wiley Interdiscip. Rev. RNA 2012, 3, 1–12. [Google Scholar] [CrossRef] [PubMed]
- Black, D.L. Mechanisms of alternative pre-messenger RNA splicing. Annu. Rev. Biochem. 2003, 72, 291–336. [Google Scholar] [CrossRef] [PubMed]
- Martinez-Contreras, R.; Cloutier, P.; Shkreta, L.; Fisette, J.F.; Revil, T.; Chabot, B. hnRNP proteins and splicing control. Adv. Exp. Med. Biol. 2007, 623, 123–147. [Google Scholar] [PubMed]
- Kutluay, S.B.; Emery, A.; Penumutchu, S.R.; Townsend, D.; Tenneti, K.; Madison, M.K.; Stukenbroeker, A.M.; Powell, C.; Jannain, D.; Tolbert, B.S.; et al. Genome-Wide Analysis of Heterogeneous Nuclear Ribonucleoprotein (hnRNP) Binding to HIV-1 RNA Reveals a Key Role for hnRNP H1 in Alternative Viral mRNA Splicing. J. Virol. 2019, 93, e01048-19. [Google Scholar] [CrossRef] [PubMed]
- Kim, J.H.; Hahm, B.; Kim, Y.K.; Choi, M.; Jang, S.K. Protein-protein interaction among hnRNPs shuttling between nucleus and cytoplasm. J. Mol. Biol. 2000, 298, 395–405. [Google Scholar] [CrossRef] [PubMed]
- Zhang, X.; Flavell, R.A.; Li, H.B. hnRNPA2B1: A nuclear DNA sensor in antiviral immunity. Cell Res. 2019, 29, 879–880. [Google Scholar] [CrossRef] [PubMed]
- Hallay, H.; Locker, N.; Ayadi, L.; Ropers, D.; Guittet, E.; Branlant, C. Biochemical and NMR study on the competition between proteins SC35, SRp40, and heterogeneous nuclear ribonucleoprotein A1 at the HIV-1 Tat exon 2 splicing site. J. Biol. Chem. 2006, 281, 37159–37174. [Google Scholar] [CrossRef] [PubMed]
- Krecic, A.M.; Swanson, M.S. hnRNP complexes: Composition, structure, and function. Curr. Opin. Cell Biol. 1999, 11, 363–371. [Google Scholar] [CrossRef] [PubMed]
- Fuentes, Y.; Olguín, V.; López-Ulloa, B.; Mendonça, D.; Ramos, H.; Abdalla, A.L.; Guajardo-Contreras, G.; Niu, M.; Rojas-Araya, B.; Mouland, A.J.; et al. Heterogeneous nuclear ribonucleoprotein K promotes cap-independent translation initiation of retroviral mRNAs. Nucleic Acids Res. 2024, 52, 2625–2647. [Google Scholar] [CrossRef] [PubMed]
- Piñol-Roma, S.; Dreyfuss, G. Shuttling of pre-mRNA binding proteins between nucleus and cytoplasm. Nature 1992, 355, 730–732. [Google Scholar] [CrossRef] [PubMed]
- England, W.E.; Wang, J.; Chen, S.; Baldi, P.; Flynn, R.A.; Spitale, R.C. An atlas of posttranslational modifications on RNA binding proteins. Nucleic Acids Res. 2022, 50, 4329–4339. [Google Scholar] [CrossRef] [PubMed]
- Thibault, P.A.; Ganesan, A.; Kalyaanamoorthy, S.; Clarke, J.W.E.; Salapa, H.E.; Levin, M.C. hnRNP A/B Proteins: An Encyclopedic Assessment of Their Roles in Homeostasis and Disease. Biology 2021, 10, 712. [Google Scholar] [CrossRef] [PubMed]
- Jean-Philippe, J.; Paz, S.; Caputi, M. hnRNP A1: The Swiss army knife of gene expression. Int. J. Mol. Sci. 2013, 14, 18999–19024. [Google Scholar] [CrossRef] [PubMed]
- Rousseau, S.; Morrice, N.; Peggie, M.; Campbell, D.G.; Gaestel, M.; Cohen, P. Inhibition of SAPK2a/p38 prevents hnRNP A0 phosphorylation by MAPKAP-K2 and its interaction with cytokine mRNAs. EMBO J. 2002, 21, 6505–6514. [Google Scholar] [CrossRef] [PubMed]
- Cannell, I.G.; Merrick, K.A.; Morandell, S.; Zhu, C.Q.; Braun, C.J.; Grant, R.A.; Cameron, E.R.; Tsao, M.S.; Hemann, M.T.; Yaffe, M.B. A Pleiotropic RNA-Binding Protein Controls Distinct Cell Cycle Checkpoints to Drive Resistance of p53-Defective Tumors to Chemotherapy. Cancer Cell 2015, 28, 831. [Google Scholar] [CrossRef] [PubMed]
- Konishi, H.; Fujiya, M.; Kashima, S.; Sakatani, A.; Dokoshi, T.; Ando, K.; Ueno, N.; Iwama, T.; Moriichi, K.; Tanaka, H.; et al. A tumor-specific modulation of heterogeneous ribonucleoprotein A0 promotes excessive mitosis and growth in colorectal cancer cells. Cell Death Dis. 2020, 11, 245. [Google Scholar] [CrossRef] [PubMed]
- Chen, C.Y.; Shyu, A.B. AU-rich elements: Characterization and importance in mRNA degradation. Trends Biochem. Sci. 1995, 20, 465–470. [Google Scholar] [CrossRef] [PubMed]
- Myer, V.E.; Steitz, J.A. Isolation and characterization of a novel, low abundance hnRNP protein: A0. RNA 1995, 1, 171–182. [Google Scholar] [PubMed]
- Young, D.J.; Stoddart, A.; Nakitandwe, J.; Chen, S.C.; Qian, Z.; Downing, J.R.; Le Beau, M.M. Knockdown of Hnrnpa0, a del(5q) gene, alters myeloid cell fate in murine cells through regulation of AU-rich transcripts. Haematologica 2014, 99, 1032–1040. [Google Scholar] [CrossRef] [PubMed]
- Schneider, R.; Campbell, M.; Nasioulas, G.; Felber, B.K.; Pavlakis, G.N. Inactivation of the human immunodeficiency virus type 1 inhibitory elements allows Rev-independent expression of Gag and Gag/protease and particle formation. J. Virol. 1997, 71, 4892–4903. [Google Scholar] [CrossRef] [PubMed]
- de Veer, M.J.; Holko, M.; Frevel, M.; Walker, E.; Der, S.; Paranjape, J.M.; Silverman, R.H.; Williams, B.R. Functional classification of interferon-stimulated genes identified using microarrays. J. Leukoc. Biol. 2001, 69, 912–920. [Google Scholar] [CrossRef] [PubMed]
- Billiau, A.; Matthys, P. Interferon-gamma: A historical perspective. Cytokine Growth Factor Rev. 2009, 20, 97–113. [Google Scholar] [CrossRef] [PubMed]
- McKenzie, A.N.J.; Spits, H.; Eberl, G. Innate lymphoid cells in inflammation and immunity. Immunity 2014, 41, 366–374. [Google Scholar] [CrossRef] [PubMed]
- Kang, S.; Brown, H.M.; Hwang, S. Direct Antiviral Mechanisms of Interferon-Gamma. Immune Netw. 2018, 18, e33. [Google Scholar] [CrossRef] [PubMed]
- Weissmann, C.; Weber, H. The interferon genes. Prog. Nucleic Acid Res. Mol. Biol. 1986, 33, 251–300. [Google Scholar] [CrossRef] [PubMed]
- Hardy, M.P.; Owczarek, C.M.; Jermiin, L.S.; Ejdeback, M.; Hertzog, P.J. Characterization of the type I interferon locus and identification of novel genes. Genomics 2004, 84, 331–345. [Google Scholar] [CrossRef] [PubMed]
- Guo, K.; Shen, G.; Kibbie, J.; Gonzalez, T.; Dillon, S.M.; Smith, H.A.; Cooper, E.H.; Lavender, K.; Hasenkrug, K.J.; Sutter, K.; et al. Qualitative Differences Between the IFNalpha subtypes and IFNbeta Influence Chronic Mucosal HIV-1 Pathogenesis. PLoS Pathog. 2020, 16, e1008986. [Google Scholar] [CrossRef] [PubMed]
- Karakoese, Z.; Ingola, M.; Sitek, B.; Dittmer, U.; Sutter, K. IFNalpha Subtypes in HIV Infection and Immunity. Viruses 2024, 16, 364. [Google Scholar] [CrossRef] [PubMed]
- Schneider, W.M.; Chevillotte, M.D.; Rice, C.M. Interferon-stimulated genes: A complex web of host defenses. Annu. Rev. Immunol. 2014, 32, 513–545. [Google Scholar] [CrossRef] [PubMed]
- Gad, H.H.; Hamming, O.J.; Hartmann, R. The structure of human interferon lambda and what it has taught us. J. Interferon Cytokine Res. Off. J. Int. Soc. Interferon Cytokine Res. 2010, 30, 565–571. [Google Scholar] [CrossRef] [PubMed]
- Kotenko, S.V.; Gallagher, G.; Baurin, V.V.; Lewis-Antes, A.; Shen, M.; Shah, N.K.; Langer, J.A.; Sheikh, F.; Dickensheets, H.; Donnelly, R.P. IFN-lambdas mediate antiviral protection through a distinct class II cytokine receptor complex. Nat. Immunol. 2003, 4, 69–77. [Google Scholar] [CrossRef] [PubMed]
- Sheppard, P.; Kindsvogel, W.; Xu, W.; Henderson, K.; Schlutsmeyer, S.; Whitmore, T.E.; Kuestner, R.; Garrigues, U.; Birks, C.; Roraback, J.; et al. IL-28, IL-29 and their class II cytokine receptor IL-28R. Nat. Immunol. 2003, 4, 63–68. [Google Scholar] [CrossRef] [PubMed]
- Lazear, H.M.; Nice, T.J.; Diamond, M.S. Interferon-lambda: Immune Functions at Barrier Surfaces and Beyond. Immunity 2015, 43, 15–28. [Google Scholar] [CrossRef] [PubMed]
- Josephson, K.; Logsdon, N.J.; Walter, M.R. Crystal structure of the IL-10/IL-10R1 complex reveals a shared receptor binding site. Immunity 2001, 15, 35–46. [Google Scholar] [CrossRef]
- Logsdon, N.J.; Jones, B.C.; Allman, J.C.; Izotova, L.; Schwartz, B.; Pestka, S.; Walter, M.R. The IL-10R2 binding hot spot on IL-22 is located on the N-terminal helix and is dependent on N-linked glycosylation. J. Mol. Biol. 2004, 342, 503–514. [Google Scholar] [CrossRef] [PubMed]
- Gad, H.H.; Dellgren, C.; Hamming, O.J.; Vends, S.; Paludan, S.R.; Hartmann, R. Interferon-lambda is functionally an interferon but structurally related to the interleukin-10 family. J. Biol. Chem. 2009, 284, 20869–20875. [Google Scholar] [CrossRef] [PubMed]
- Mesev, E.V.; LeDesma, R.A.; Ploss, A. Decoding type I and III interferon signalling during viral infection. Nat. Microbiol. 2019, 4, 914–924. [Google Scholar] [CrossRef] [PubMed]
- Lavender, K.J.; Gibbert, K.; Peterson, K.E.; Van Dis, E.; Francois, S.; Woods, T.; Messer, R.J.; Gawanbacht, A.; Muller, J.A.; Munch, J.; et al. Interferon Alpha Subtype-Specific Suppression of HIV-1 Infection In Vivo. J. Virol. 2016, 90, 6001–6013. [Google Scholar] [CrossRef] [PubMed]
- Harper, M.S.; Guo, K.; Gibbert, K.; Lee, E.J.; Dillon, S.M.; Barrett, B.S.; McCarter, M.D.; Hasenkrug, K.J.; Dittmer, U.; Wilson, C.C.; et al. Interferon-alpha Subtypes in an Ex Vivo Model of Acute HIV-1 Infection: Expression, Potency and Effector Mechanisms. PLoS Pathog. 2015, 11, e1005254. [Google Scholar] [CrossRef] [PubMed]
- Li, Y.; Sun, B.; Esser, S.; Jessen, H.; Streeck, H.; Widera, M.; Yang, R.; Dittmer, U.; Sutter, K. Expression Pattern of Individual IFNA Subtypes in Chronic HIV Infection. J. Interferon Cytokine Res. Off. J. Int. Soc. Interferon Cytokine Res. 2017, 37, 541–549. [Google Scholar] [CrossRef] [PubMed]
- Moll, H.P.; Maier, T.; Zommer, A.; Lavoie, T.; Brostjan, C. The differential activity of interferon-alpha subtypes is consistent among distinct target genes and cell types. Cytokine 2011, 53, 52–59. [Google Scholar] [CrossRef] [PubMed]
- Dolken, L.; Ruzsics, Z.; Radle, B.; Friedel, C.C.; Zimmer, R.; Mages, J.; Hoffmann, R.; Dickinson, P.; Forster, T.; Ghazal, P.; et al. High-resolution gene expression profiling for simultaneous kinetic parameter analysis of RNA synthesis and decay. RNA 2008, 14, 1959–1972. [Google Scholar] [CrossRef] [PubMed]
- Trilling, M.; Bellora, N.; Rutkowski, A.J.; de Graaf, M.; Dickinson, P.; Robertson, K.; Prazeres da Costa, O.; Ghazal, P.; Friedel, C.C.; Alba, M.M.; et al. Deciphering the modulation of gene expression by type I and II interferons combining 4sU-tagging, translational arrest and in silico promoter analysis. Nucleic Acids Res. 2013, 41, 8107–8125. [Google Scholar] [CrossRef] [PubMed]
- Brass, A.L.; Dykxhoorn, D.M.; Benita, Y.; Yan, N.; Engelman, A.; Xavier, R.J.; Lieberman, J.; Elledge, S.J. Identification of host proteins required for HIV infection through a functional genomic screen. Science 2008, 319, 921–926. [Google Scholar] [CrossRef] [PubMed]
- Konig, R.; Zhou, Y.; Elleder, D.; Diamond, T.L.; Bonamy, G.M.; Irelan, J.T.; Chiang, C.Y.; Tu, B.P.; De Jesus, P.D.; Lilley, C.E.; et al. Global analysis of host-pathogen interactions that regulate early-stage HIV-1 replication. Cell 2008, 135, 49–60. [Google Scholar] [CrossRef] [PubMed]
- Zhou, H.; Xu, M.; Huang, Q.; Gates, A.T.; Zhang, X.D.; Castle, J.C.; Stec, E.; Ferrer, M.; Strulovici, B.; Hazuda, D.J.; et al. Genome-scale RNAi screen for host factors required for HIV replication. Cell Host Microbe 2008, 4, 495–504. [Google Scholar] [CrossRef] [PubMed]
- Megger, D.A.; Philipp, J.; Le-Trilling, V.T.K.; Sitek, B.; Trilling, M. Deciphering of the Human Interferon-Regulated Proteome by Mass Spectrometry-Based Quantitative Analysis Reveals Extent and Dynamics of Protein Induction and Repression. Front. Immunol. 2017, 8, 1139. [Google Scholar] [CrossRef] [PubMed]
- Lehmann, C.; Taubert, D.; Jung, N.; Fatkenheuer, G.; van Lunzen, J.; Hartmann, P.; Romerio, F. Preferential upregulation of interferon-alpha subtype 2 expression in HIV-1 patients. AIDS Res. Hum. Retroviruses 2009, 25, 577–581. [Google Scholar] [CrossRef] [PubMed]
- Butic, A.B.; Spencer, S.A.; Shaheen, S.K.; Lukacher, A.E. Polyomavirus Wakes Up and Chooses Neurovirulence. Viruses 2023, 15, 2112. [Google Scholar] [CrossRef] [PubMed]
- Rocchi, A.; Sariyer, I.K.; Berger, J.R. Revisiting JC virus and progressive multifocal leukoencephalopathy. J. Neurovirology 2023, 29, 524–537. [Google Scholar] [CrossRef] [PubMed]
- Guo, L.; Liu, J.J.; Long, S.Y.; Wang, P.Y.; Li, S.; Wang, J.L.; Wei, X.F.; Li, J.; Lei, L.; Huang, A.L.; et al. TIM22 and TIM29 inhibit HBV replication by up-regulating SRSF1 expression. J. Med. Virol. 2024, 96, e29439. [Google Scholar] [CrossRef] [PubMed]
- Xiong, Z.; Shaibani, A.; Li, Y.P.; Yan, Y.; Zhang, S.; Yang, Y.; Yang, F.; Wang, H.; Yang, X.F. Alternative splicing factor ASF/SF2 is down regulated in inflamed muscle. J. Clin. Pathol. 2006, 59, 855–861. [Google Scholar] [CrossRef] [PubMed]
- Kontoyiannis, D.; Pasparakis, M.; Pizarro, T.T.; Cominelli, F.; Kollias, G. Impaired on/off regulation of TNF biosynthesis in mice lacking TNF AU-rich elements: Implications for joint and gut-associated immunopathologies. Immunity 1999, 10, 387–398. [Google Scholar] [CrossRef] [PubMed]
- The ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef] [PubMed]
- Luo, Y.; Hitz, B.C.; Gabdank, I.; Hilton, J.A.; Kagda, M.S.; Lam, B.; Myers, Z.; Sud, P.; Jou, J.; Lin, K.; et al. New developments on the Encyclopedia of DNA Elements (ENCODE) data portal. Nucleic Acids Res. 2020, 48, D882–D889. [Google Scholar] [CrossRef] [PubMed]
- Lopez-Huertas, M.R.; Palladino, C.; Garrido-Arquero, M.; Esteban-Cartelle, B.; Sanchez-Carrillo, M.; Martinez-Roman, P.; Martin-Carbonero, L.; Ryan, P.; Dominguez-Dominguez, L.; Santos, I.L.; et al. HCV-coinfection is related to an increased HIV-1 reservoir size in cART-treated HIV patients: A cross-sectional study. Sci. Rep. 2019, 9, 5606. [Google Scholar] [CrossRef] [PubMed]
- Martinez-Roman, P.; Lopez-Huertas, M.R.; Crespo-Bermejo, C.; Arca-Lafuente, S.; Cortegano, I.; Valle-Millares, D.; Gaspar, M.L.; Martin-Carbonero, L.; Dominguez-Dominguez, L.; Ryan, P.; et al. Hepatitis C Virus Influences HIV-1 Viral Splicing in Coinfected Patients. J. Clin. Med. 2020, 9, 2091. [Google Scholar] [CrossRef] [PubMed]
- Jarboui, M.A.; Bidoia, C.; Woods, E.; Roe, B.; Wynne, K.; Elia, G.; Hall, W.W.; Gautier, V.W. Nucleolar protein trafficking in response to HIV-1 Tat: Rewiring the nucleolus. PLoS ONE 2012, 7, e48702. [Google Scholar] [CrossRef] [PubMed]
- Dowling, D.; Nasr-Esfahani, S.; Tan, C.H.; O‘Brien, K.; Howard, J.L.; Jans, D.A.; Purcell, D.F.; Stoltzfus, C.M.; Sonza, S. HIV-1 infection induces changes in expression of cellular splicing factors that regulate alternative viral splicing and virus production in macrophages. Retrovirology 2008, 5, 18. [Google Scholar] [CrossRef] [PubMed]
- Maldarelli, F.; Xiang, C.; Chamoun, G.; Zeichner, S.L. The expression of the essential nuclear splicing factor SC35 is altered by human immunodeficiency virus infection. Virus Res. 1998, 53, 39–51. [Google Scholar] [CrossRef] [PubMed]
- Paz, S.; Krainer, A.R.; Caputi, M. HIV-1 transcription is regulated by splicing factor SRSF1. Nucleic Acids Res. 2014, 42, 13812–13823. [Google Scholar] [CrossRef] [PubMed]
- D‘Orso, I.; Frankel, A.D. RNA-mediated displacement of an inhibitory snRNP complex activates transcription elongation. Nat. Struct. Mol. Biol. 2010, 17, 815–821. [Google Scholar] [CrossRef]
- Barboric, M.; Yik, J.H.; Czudnochowski, N.; Yang, Z.; Chen, R.; Contreras, X.; Geyer, M.; Matija Peterlin, B.; Zhou, Q. Tat competes with HEXIM1 to increase the active pool of P-TEFb for HIV-1 transcription. Nucleic Acids Res. 2007, 35, 2003–2012. [Google Scholar] [CrossRef] [PubMed]
- Sedore, S.C.; Byers, S.A.; Biglione, S.; Price, J.P.; Maury, W.J.; Price, D.H. Manipulation of P-TEFb control machinery by HIV: Recruitment of P-TEFb from the large form by Tat and binding of HEXIM1 to TAR. Nucleic Acids Res. 2007, 35, 4347–4358. [Google Scholar] [CrossRef] [PubMed]
- Muniz, L.; Egloff, S.; Ughy, B.; Jady, B.E.; Kiss, T. Controlling cellular P-TEFb activity by the HIV-1 transcriptional transactivator Tat. PLoS Pathog. 2010, 6, e1001152. [Google Scholar] [CrossRef] [PubMed]
- Sobhian, B.; Laguette, N.; Yatim, A.; Nakamura, M.; Levy, Y.; Kiernan, R.; Benkirane, M. HIV-1 Tat assembles a multifunctional transcription elongation complex and stably associates with the 7SK snRNP. Mol. Cell 2010, 38, 439–451. [Google Scholar] [CrossRef] [PubMed]
- AJ, C.Q.; Bugai, A.; Barboric, M. Cracking the control of RNA polymerase II elongation by 7SK snRNP and P-TEFb. Nucleic Acids Res. 2016, 44, 7527–7539. [Google Scholar] [CrossRef] [PubMed]
- Paz, S.; Caputi, M. SRSF1 inhibition of HIV-1 gene expression. Oncotarget 2015, 6, 19362–19363. [Google Scholar] [CrossRef] [PubMed]
- Krueger, B.J.; Jeronimo, C.; Roy, B.B.; Bouchard, A.; Barrandon, C.; Byers, S.A.; Searcey, C.E.; Cooper, J.J.; Bensaude, O.; Cohen, E.A.; et al. LARP7 is a stable component of the 7SK snRNP while P-TEFb, HEXIM1 and hnRNP A1 are reversibly associated. Nucleic Acids Res. 2008, 36, 2219–2229. [Google Scholar] [CrossRef] [PubMed]
- Van Herreweghe, E.; Egloff, S.; Goiffon, I.; Jady, B.E.; Froment, C.; Monsarrat, B.; Kiss, T. Dynamic remodelling of human 7SK snRNP controls the nuclear level of active P-TEFb. EMBO J. 2007, 26, 3570–3580. [Google Scholar] [CrossRef] [PubMed]
- Burd, C.G.; Dreyfuss, G. RNA binding specificity of hnRNP A1: Significance of hnRNP A1 high-affinity binding sites in pre-mRNA splicing. EMBO J. 1994, 13, 1197–1204. [Google Scholar] [CrossRef] [PubMed]
- Jeang, K.T.; Xiao, H.; Rich, E.A. Multifaceted activities of the HIV-1 transactivator of transcription, Tat. J. Biol. Chem. 1999, 274, 28837–28840. [Google Scholar] [CrossRef] [PubMed]
- Li, L.; Dahiya, S.; Kortagere, S.; Aiamkitsumrit, B.; Cunningham, D.; Pirrone, V.; Nonnemacher, M.R.; Wigdahl, B. Impact of Tat Genetic Variation on HIV-1 Disease. Adv. Virol. 2012, 2012, 123605. [Google Scholar] [CrossRef] [PubMed]
- Erkelenz, S.; Hillebrand, F.; Widera, M.; Theiss, S.; Fayyaz, A.; Degrandi, D.; Pfeffer, K.; Schaal, H. Balanced splicing at the Tat-specific HIV-1 3′ss A3 is critical for HIV-1 replication. Retrovirology 2015, 12, 29. [Google Scholar] [CrossRef] [PubMed]
- Erkelenz, S.; Theiss, S.; Otte, M.; Widera, M.; Peter, J.O.; Schaal, H. Genomic HEXploring allows landscaping of novel potential splicing regulatory elements. Nucleic Acids Res. 2014, 42, 10681–10697. [Google Scholar] [CrossRef] [PubMed]
- Zahler, A.M.; Damgaard, C.K.; Kjems, J.; Caputi, M. SC35 and heterogeneous nuclear ribonucleoprotein A/B proteins bind to a juxtaposed exonic splicing enhancer/exonic splicing silencer element to regulate HIV-1 tat exon 2 splicing. J. Biol. Chem. 2004, 279, 10077–10084. [Google Scholar] [CrossRef] [PubMed]
- Amendt, B.A.; Hesslein, D.; Chang, L.J.; Stoltzfus, C.M. Presence of negative and positive cis-acting RNA splicing elements within and flanking the first tat coding exon of human immunodeficiency virus type 1. Mol. Cell. Biol. 1994, 14, 3960–3970. [Google Scholar] [PubMed]
- Caputi, M.; Mayeda, A.; Krainer, A.R.; Zahler, A.M. hnRNP A/B proteins are required for inhibition of HIV-1 pre-mRNA splicing. EMBO J. 1999, 18, 4060–4067. [Google Scholar] [CrossRef] [PubMed]
- Si, Z.; Amendt, B.A.; Stoltzfus, C.M. Splicing efficiency of human immunodeficiency virus type 1 tat RNA is determined by both a suboptimal 3′ splice site and a 10 nucleotide exon splicing silencer element located within tat exon 2. Nucleic Acids Res. 1997, 25, 861–867. [Google Scholar] [CrossRef] [PubMed]
- Jacquenet, S.; Mereau, A.; Bilodeau, P.S.; Damier, L.; Stoltzfus, C.M.; Branlant, C. A second exon splicing silencer within human immunodeficiency virus type 1 tat exon 2 represses splicing of Tat mRNA and binds protein hnRNP H. J. Biol. Chem. 2001, 276, 40464–40475. [Google Scholar] [CrossRef] [PubMed]
- Tange, T.O.; Damgaard, C.K.; Guth, S.; Valcarcel, J.; Kjems, J. The hnRNP A1 protein regulates HIV-1 tat splicing via a novel intron silencer element. EMBO J. 2001, 20, 5748–5758. [Google Scholar] [CrossRef] [PubMed]
- Huelga, S.C.; Vu, A.Q.; Arnold, J.D.; Liang, T.Y.; Liu, P.P.; Yan, B.Y.; Donohue, J.P.; Shiue, L.; Hoon, S.; Brenner, S.; et al. Integrative genome-wide analysis reveals cooperative regulation of alternative splicing by hnRNP proteins. Cell Rep. 2012, 1, 167–178. [Google Scholar] [CrossRef] [PubMed]
- Buratti, E.; Brindisi, A.; Giombi, M.; Tisminetzky, S.; Ayala, Y.M.; Baralle, F.E. TDP-43 binds heterogeneous nuclear ribonucleoprotein A/B through its C-terminal tail: An important region for the inhibition of cystic fibrosis transmembrane conductance regulator exon 9 splicing. J. Biol. Chem. 2005, 280, 37572–37584. [Google Scholar] [CrossRef] [PubMed]
- Hutchison, S.; LeBel, C.; Blanchette, M.; Chabot, B. Distinct sets of adjacent heterogeneous nuclear ribonucleoprotein (hnRNP) A1/A2 binding sites control 5′ splice site selection in the hnRNP A1 mRNA precursor. J. Biol. Chem. 2002, 277, 29745–29752. [Google Scholar] [CrossRef] [PubMed]
- Blanchette, M.; Green, R.E.; MacArthur, S.; Brooks, A.N.; Brenner, S.E.; Eisen, M.B.; Rio, D.C. Genome-wide analysis of alternative pre-mRNA splicing and RNA-binding specificities of the Drosophila hnRNP A/B family members. Mol. Cell 2009, 33, 438–449. [Google Scholar] [CrossRef] [PubMed]
- Pasquereau, S.; Kumar, A.; Herbein, G. Targeting TNF and TNF Receptor Pathway in HIV-1 Infection: From Immune Activation to Viral Reservoirs. Viruses 2017, 9, 64. [Google Scholar] [CrossRef] [PubMed]
- Dutilleul, A.; Rodari, A.; Van Lint, C. Depicting HIV-1 Transcriptional Mechanisms: A Summary of What We Know. Viruses 2020, 12, 1385. [Google Scholar] [CrossRef] [PubMed]
- Sonza, S.; Mutimer, H.P.; O‘Brien, K.; Ellery, P.; Howard, J.L.; Axelrod, J.H.; Deacon, N.J.; Crowe, S.M.; Purcell, D.F. Selectively reduced tat mRNA heralds the decline in productive human immunodeficiency virus type 1 infection in monocyte-derived macrophages. J. Virol. 2002, 76, 12611–12621. [Google Scholar] [CrossRef] [PubMed]
- Moron-Lopez, S.; Telwatte, S.; Sarabia, I.; Battivelli, E.; Montano, M.; Macedo, A.B.; Aran, D.; Butte, A.J.; Jones, R.B.; Bosque, A.; et al. Human splice factors contribute to latent HIV infection in primary cell models and blood CD4+ T cells from ART-treated individuals. PLoS Pathog. 2020, 16, e1009060. [Google Scholar] [CrossRef] [PubMed]
- Scalabrin, M.; Frasson, I.; Ruggiero, E.; Perrone, R.; Tosoni, E.; Lago, S.; Tassinari, M.; Palu, G.; Richter, S.N. The cellular protein hnRNP A2/B1 enhances HIV-1 transcription by unfolding LTR promoter G-quadruplexes. Sci. Rep. 2017, 7, 45244. [Google Scholar] [CrossRef] [PubMed]
- Dahal, S.; Clayton, K.; Been, T.; Fernet-Brochu, R.; Ocando, A.V.; Balachandran, A.; Poirier, M.; Maldonado, R.K.; Shkreta, L.; Boligan, K.F.; et al. Opposing roles of CLK SR kinases in controlling HIV-1 gene expression and latency. Retrovirology 2022, 19, 18. [Google Scholar] [CrossRef] [PubMed]
- Pasternak, A.O.; Berkhout, B. The Splice of Life: Does RNA Processing Have a Role in HIV-1 Persistence? Viruses 2021, 13, 1751. [Google Scholar] [CrossRef] [PubMed]
- Erkelenz, S.; Mueller, W.F.; Evans, M.S.; Busch, A.; Schoneweis, K.; Hertel, K.J.; Schaal, H. Position-dependent splicing activation and repression by SR and hnRNP proteins rely on common mechanisms. RNA 2013, 19, 96–102. [Google Scholar] [CrossRef] [PubMed]
- Fu, X.D.; Ares, M., Jr. Context-dependent control of alternative splicing by RNA-binding proteins. Nat. Rev. Genet. 2014, 15, 689–701. [Google Scholar] [CrossRef] [PubMed]
- Tomezsko, P.J.; Corbin, V.D.A.; Gupta, P.; Swaminathan, H.; Glasgow, M.; Persad, S.; Edwards, M.D.; McIntosh, L.; Papenfuss, A.T.; Emery, A.; et al. Determination of RNA structural diversity and its role in HIV-1 RNA splicing. Nature 2020, 582, 438–442. [Google Scholar] [CrossRef] [PubMed]
- Barash, Y.; Calarco, J.A.; Gao, W.; Pan, Q.; Wang, X.; Shai, O.; Blencowe, B.J.; Frey, B.J. Deciphering the splicing code. Nature 2010, 465, 53–59. [Google Scholar] [CrossRef] [PubMed]
- Wang, Z.; Burge, C.B. Splicing regulation: From a parts list of regulatory elements to an integrated splicing code. RNA 2008, 14, 802–813. [Google Scholar] [CrossRef] [PubMed]
- Widera, M.; Erkelenz, S.; Hillebrand, F.; Krikoni, A.; Widera, D.; Kaisers, W.; Deenen, R.; Gombert, M.; Dellen, R.; Pfeiffer, T.; et al. An Intronic G Run within HIV-1 Intron 2 Is Critical for Splicing Regulation of vif mRNA. J. Virol. 2013, 87, 2707–2720. [Google Scholar] [CrossRef] [PubMed]
- Widera, M.; Hillebrand, F.; Erkelenz, S.; Vasudevan, A.; Munk, C.; Schaal, H. A functional conserved intronic G run in HIV-1 intron 3 is critical to counteract APOBEC3G-mediated host restriction. Retrovirology 2014, 11, 72. [Google Scholar] [CrossRef]
- Exline, C.M.; Feng, Z.; Stoltzfus, C.M. Negative and positive mRNA splicing elements act competitively to regulate human immunodeficiency virus type 1 vif gene expression. J. Virol. 2008, 82, 3921–3931. [Google Scholar] [CrossRef]
- Wentz, M.P.; Moore, B.E.; Cloyd, M.W.; Berget, S.M.; Donehower, L.A. A naturally arising mutation of a potential silencer of exon splicing in human immunodeficiency virus type 1 induces dominant aberrant splicing and arrests virus production. J. Virol. 1997, 71, 8542–8551. [Google Scholar] [CrossRef] [PubMed]
- Madsen, J.M.; Stoltzfus, C.M. An exonic splicing silencer downstream of the 3′ splice site A2 is required for efficient human immunodeficiency virus type 1 replication. J. Virol. 2005, 79, 10478–10486. [Google Scholar] [CrossRef]
- Asang, C.; Erkelenz, S.; Schaal, H. The HIV-1 major splice donor D1 is activated by splicing enhancer elements within the leader region and the p17-inhibitory sequence. Virology 2012, 432, 133–145. [Google Scholar] [CrossRef] [PubMed]
- Mikaelian, I.; Krieg, M.; Gait, M.J.; Karn, J. Interactions of INS (CRS) elements and the splicing machinery regulate the production of Rev-responsive mRNAs. J. Mol. Biol. 1996, 257, 246–264. [Google Scholar] [CrossRef] [PubMed]
- Schaub, M.C.; Lopez, S.R.; Caputi, M. Members of the heterogeneous nuclear ribonucleoprotein H family activate splicing of an HIV-1 splicing substrate by promoting formation of ATP-dependent spliceosomal complexes. J. Biol. Chem. 2007, 282, 13617–13626. [Google Scholar] [CrossRef] [PubMed]
- Brillen, A.L.; Walotka, L.; Hillebrand, F.; Muller, L.; Widera, M.; Theiss, S.; Schaal, H. Analysis of competing HIV-1 splice donor sites uncovers a tight cluster of splicing regulatory elements within exon 2/2b. J. Virol. 2017, 91, e00389-17. [Google Scholar] [CrossRef]
- Bilodeau, P.S.; Domsic, J.K.; Mayeda, A.; Krainer, A.R.; Stoltzfus, C.M. RNA splicing at human immunodeficiency virus type 1 3′ splice site A2 is regulated by binding of hnRNP A/B proteins to an exonic splicing silencer element. J. Virol. 2001, 75, 8487–8497. [Google Scholar] [CrossRef]
- Domsic, J.K.; Wang, Y.; Mayeda, A.; Krainer, A.R.; Stoltzfus, C.M. Human immunodeficiency virus type 1 hnRNP A/B-dependent exonic splicing silencer ESSV antagonizes binding of U2AF65 to viral polypyrimidine tracts. Mol. Cell. Biol. 2003, 23, 8762–8772. [Google Scholar] [CrossRef] [PubMed]
- Hillebrand, F.; Peter, J.O.; Brillen, A.L.; Otte, M.; Schaal, H.; Erkelenz, S. Differential hnRNP D isoform incorporation may confer plasticity to the ESSV-mediated repressive state across HIV-1 exon 3. Biochim. Biophys. Acta 2017, 1860, 205–217. [Google Scholar] [CrossRef] [PubMed]
- Erkelenz, S. Regulation of Human and HIV-1 Splice Sites. Ph.D. Thesis, Heinrich-Heine-Universität Düsseldorf, 500 Naturwissenschaften und Mathematik » 570 Biowissenschaften, Biologie, Düsseldorf, Germany, 2013. [Google Scholar]
- Tsuruno, C.; Ohe, K.; Kuramitsu, M.; Kohma, T.; Takahama, Y.; Hamaguchi, Y.; Hamaguchi, I.; Okuma, K. HMGA1a is involved in specific splice site regulation of human immunodeficiency virus type 1. Biochem. Biophys. Res. Commun. 2011, 406, 512–517. [Google Scholar] [CrossRef] [PubMed]
- Tange, T.; Kjems, J. SF2/ASF binds to a splicing enhancer in the third HIV-1 tat exon and stimulates U2AF binding independently of the RS domain. J. Mol. Biol. 2001, 312, 649–662. [Google Scholar] [CrossRef] [PubMed]
- Amendt, B.A.; Si, Z.H.; Stoltzfus, C.M. Presence of exon splicing silencers within human immunodeficiency virus type 1 tat exon 2 and tat-rev exon 3: Evidence for inhibition mediated by cellular factors. Mol. Cell. Biol. 1995, 15, 4606–4615. [Google Scholar] [CrossRef] [PubMed]
- Si, Z.H.; Rauch, D.; Stoltzfus, C.M. The exon splicing silencer in human immunodeficiency virus type 1 Tat exon 3 is bipartite and acts early in spliceosome assembly. Mol. Cell. Biol. 1998, 18, 5404–5413. [Google Scholar] [CrossRef] [PubMed]
- Balachandran, A.; Ming, L.; Cochrane, A. Teetering on the Edge. In Retrovirus-Cell Interactions; Parent, L.J., Ed.; Academic Press: Cambridge, MA, USA, 2018; pp. 229–251. [Google Scholar]
- Jordan-Paiz, A.; Nevot, M.; Lamkiewicz, K.; Lataretu, M.; Franco, S.; Marz, M.; Martinez, M.A. HIV-1 Lethality and Loss of Env Protein Expression Induced by Single Synonymous Substitutions in the Virus Genome Intronic-Splicing Silencer. J. Virol. 2020, 94, e01108-20. [Google Scholar] [CrossRef] [PubMed]
- Muller, L.; Moskorz, W.; Brillen, A.L.; Hillebrand, F.; Ostermann, P.N.; Kiel, N.; Walotka, L.; Ptok, J.; Timm, J.; Lubke, N.; et al. Altered HIV-1 mRNA Splicing Due to Drug-Resistance-Associated Mutations in Exon 2/2b. Int. J. Mol. Sci. 2021, 23, 156. [Google Scholar] [CrossRef] [PubMed]
- Doi, N.; Koma, T.; Adachi, A.; Nomaguchi, M. Expression Level of HIV-1 Vif Can Be Fluctuated by Natural Nucleotide Variations in the vif-Coding and Regulatory SA1D2prox Sequences of the Proviral Genome. Front. Microbiol. 2019, 10, 2758. [Google Scholar] [CrossRef] [PubMed]
- Nguyen Quang, N.; Goudey, S.; Segeral, E.; Mohammad, A.; Lemoine, S.; Blugeon, C.; Versapuech, M.; Paillart, J.C.; Berlioz-Torrent, C.; Emiliani, S.; et al. Dynamic nanopore long-read sequencing analysis of HIV-1 splicing events during the early steps of infection. Retrovirology 2020, 17, 25. [Google Scholar] [CrossRef] [PubMed]
- Koma, T.; Doi, N.; Le, B.Q.; Kondo, T.; Ishizue, M.; Tokaji, C.; Tsukada, C.; Adachi, A.; Nomaguchi, M. Involvement of a Rarely Used Splicing SD2b Site in the Regulation of HIV-1 vif mRNA Production as Revealed by a Growth-Adaptive Mutation. Viruses 2023, 15, 2424. [Google Scholar] [CrossRef]
- Engeland, C.E.; Brown, N.P.; Borner, K.; Schumann, M.; Krause, E.; Kaderali, L.; Muller, G.A.; Krausslich, H.G. Proteome analysis of the HIV-1 Gag interactome. Virology 2014, 460–461, 194–206. [Google Scholar] [CrossRef] [PubMed]
- Grewe, B.; Vogt, C.; Horstkotter, T.; Tippler, B.; Xiao, H.; Muller, B.; Uberla, K.; Wagner, R.; Asbach, B.; Bohne, J. The HIV 5′ Gag Region Displays a Specific Nucleotide Bias Regulating Viral Splicing and Infectivity. Viruses 2021, 13, 997. [Google Scholar] [CrossRef] [PubMed]
- Kim, G.N.; Yu, K.L.; Kim, H.I.; You, J.C. Investigation of the effect of SRSF9 overexpression on HIV-1 production. BMB Rep. 2022, 55, 639–644. [Google Scholar] [CrossRef] [PubMed]
- Asang, C.; Hauber, I.; Schaal, H. Insights into the selective activation of alternatively used splice acceptors by the human immunodeficiency virus type-1 bidirectional splicing enhancer. Nucleic Acids Res. 2008, 36, 1450–1463. [Google Scholar] [CrossRef]
- Robberson, B.L.; Cote, G.J.; Berget, S.M. Exon definition may facilitate splice site selection in RNAs with multiple exons. Mol. Cell. Biol. 1990, 10, 84–94. [Google Scholar]
- Taniguchi, I.; Mabuchi, N.; Ohno, M. HIV-1 Rev protein specifies the viral RNA export pathway by suppressing TAP/NXF1 recruitment. Nucleic Acids Res. 2014, 42, 6645–6658. [Google Scholar] [CrossRef] [PubMed]
- Hautbergue, G.M.; Hung, M.L.; Golovanov, A.P.; Lian, L.Y.; Wilson, S.A. Mutually exclusive interactions drive handover of mRNA from export adaptors to TAP. Proc. Natl. Acad. Sci. USA 2008, 105, 5154–5159. [Google Scholar] [CrossRef] [PubMed]
- Chen, S.; Wang, R.; Zheng, D.; Zhang, H.; Chang, X.; Wang, K.; Li, W.; Fan, J.; Tian, B.; Cheng, H. The mRNA Export Receptor NXF1 Coordinates Transcriptional Dynamics, Alternative Polyadenylation, and mRNA Export. Mol. Cell 2019, 74, 118–131.e7. [Google Scholar] [CrossRef] [PubMed]
- Pacheco-Fiallos, B.; Vorlander, M.K.; Riabov-Bassat, D.; Fin, L.; O‘Reilly, F.J.; Ayala, F.I.; Schellhaas, U.; Rappsilber, J.; Plaschka, C. mRNA recognition and packaging by the human transcription-export complex. Nature 2023, 616, 828–835. [Google Scholar] [CrossRef] [PubMed]
- Puhringer, T.; Hohmann, U.; Fin, L.; Pacheco-Fiallos, B.; Schellhaas, U.; Brennecke, J.; Plaschka, C. Structure of the human core transcription-export complex reveals a hub for multivalent interactions. Elife 2020, 9, e61503. [Google Scholar] [CrossRef]
- Viphakone, N.; Hautbergue, G.M.; Walsh, M.; Chang, C.T.; Holland, A.; Folco, E.G.; Reed, R.; Wilson, S.A. TREX exposes the RNA-binding domain of Nxf1 to enable mRNA export. Nat. Commun. 2012, 3, 1006. [Google Scholar] [CrossRef] [PubMed]
- Strasser, K.; Masuda, S.; Mason, P.; Pfannstiel, J.; Oppizzi, M.; Rodriguez-Navarro, S.; Rondon, A.G.; Aguilera, A.; Struhl, K.; Reed, R.; et al. TREX is a conserved complex coupling transcription with messenger RNA export. Nature 2002, 417, 304–308. [Google Scholar] [CrossRef]
- Viphakone, N.; Sudbery, I.; Griffith, L.; Heath, C.G.; Sims, D.; Wilson, S.A. Co-transcriptional Loading of RNA Export Factors Shapes the Human Transcriptome. Mol. Cell 2019, 75, 310–323.e8. [Google Scholar] [CrossRef] [PubMed]
- Singh, G.; Kucukural, A.; Cenik, C.; Leszyk, J.D.; Shaffer, S.A.; Weng, Z.; Moore, M.J. The cellular EJC interactome reveals higher-order mRNP structure and an EJC-SR protein nexus. Cell 2012, 151, 750–764. [Google Scholar] [CrossRef]
- Muller-McNicoll, M.; Botti, V.; de Jesus Domingues, A.M.; Brandl, H.; Schwich, O.D.; Steiner, M.C.; Curk, T.; Poser, I.; Zarnack, K.; Neugebauer, K.M. SR proteins are NXF1 adaptors that link alternative RNA processing to mRNA export. Genes Dev. 2016, 30, 553–566. [Google Scholar] [CrossRef] [PubMed]
- Zhong, X.Y.; Ding, J.H.; Adams, J.A.; Ghosh, G.; Fu, X.D. Regulation of SR protein phosphorylation and alternative splicing by modulating kinetic interactions of SRPK1 with molecular chaperones. Genes Dev. 2009, 23, 482–495. [Google Scholar] [CrossRef]
- Botti, V.; McNicoll, F.; Steiner, M.C.; Richter, F.M.; Solovyeva, A.; Wegener, M.; Schwich, O.D.; Poser, I.; Zarnack, K.; Wittig, I.; et al. Cellular differentiation state modulates the mRNA export activity of SR proteins. J. Cell Biol. 2017, 216, 1993–2009. [Google Scholar] [CrossRef] [PubMed]
- Huang, Y.; Yario, T.A.; Steitz, J.A. A molecular link between SR protein dephosphorylation and mRNA export. Proc. Natl. Acad. Sci. USA 2004, 101, 9666–9670. [Google Scholar] [CrossRef] [PubMed]
- Twyffels, L.; Gueydan, C.; Kruys, V. Shuttling SR proteins: More than splicing factors. FEBS J. 2011, 278, 3246–3255. [Google Scholar] [CrossRef] [PubMed]
- Tintaru, A.M.; Hautbergue, G.M.; Hounslow, A.M.; Hung, M.L.; Lian, L.Y.; Craven, C.J.; Wilson, S.A. Structural and functional analysis of RNA and TAP binding to SF2/ASF. EMBO Rep. 2007, 8, 756–762. [Google Scholar] [CrossRef] [PubMed]
- Mahiet, C.; Swanson, C.M. Control of HIV-1 gene expression by SR proteins. Biochem. Soc. Trans. 2016, 44, 1417–1425. [Google Scholar] [CrossRef] [PubMed]
- Bejarano, D.A.; Peng, K.; Laketa, V.; Borner, K.; Jost, K.L.; Lucic, B.; Glass, B.; Lusic, M.; Muller, B.; Krausslich, H.G. HIV-1 nuclear import in macrophages is regulated by CPSF6-capsid interactions at the nuclear pore complex. Elife 2019, 8, e41800. [Google Scholar] [CrossRef] [PubMed]
- Jang, S.; Cook, N.J.; Pye, V.E.; Bedwell, G.J.; Dudek, A.M.; Singh, P.K.; Cherepanov, P.; Engelman, A.N. Differential role for phosphorylation in alternative polyadenylation function versus nuclear import of SR-like protein CPSF6. Nucleic Acids Res. 2019, 47, 4663–4683. [Google Scholar] [CrossRef]
- Xiao, H.; Wyler, E.; Milek, M.; Grewe, B.; Kirchner, P.; Ekici, A.; Silva, A.; Jungnickl, D.; Full, F.; Thomas, M.; et al. CRNKL1 Is a Highly Selective Regulator of Intron-Retaining HIV-1 and Cellular mRNAs. mBio 2021, 12, e02525-20. [Google Scholar] [CrossRef] [PubMed]
- Fischer, U.; Huber, J.; Boelens, W.C.; Mattaj, I.W.; Luhrmann, R. The HIV-1 Rev activation domain is a nuclear export signal that accesses an export pathway used by specific cellular RNAs. Cell 1995, 82, 475–483. [Google Scholar] [CrossRef] [PubMed]
- Fernandes, J.; Jayaraman, B.; Frankel, A. The HIV-1 Rev response element: An RNA scaffold that directs the cooperative assembly of a homo-oligomeric ribonucleoprotein complex. RNA Biol. 2012, 9, 6–11. [Google Scholar] [CrossRef]
- Askjaer, P.; Jensen, T.H.; Nilsson, J.; Englmeier, L.; Kjems, J. The specificity of the CRM1-Rev nuclear export signal interaction is mediated by RanGTP. J. Biol. Chem. 1998, 273, 33414–33422. [Google Scholar] [CrossRef]
- Booth, D.S.; Cheng, Y.; Frankel, A.D. The export receptor Crm1 forms a dimer to promote nuclear export of HIV RNA. Elife 2014, 3, e04121. [Google Scholar] [CrossRef] [PubMed]
- Port, S.A.; Monecke, T.; Dickmanns, A.; Spillner, C.; Hofele, R.; Urlaub, H.; Ficner, R.; Kehlenbach, R.H. Structural and Functional Characterization of CRM1-Nup214 Interactions Reveals Multiple FG-Binding Sites Involved in Nuclear Export. Cell Rep. 2015, 13, 690–702. [Google Scholar] [CrossRef] [PubMed]
- Banerjee, A.; Benjamin, R.; Banerjee, S. Impact of viral factors on subcellular distribution and RNA export activity of HIV-1 rev in astrocytes 1321N1. PLoS ONE 2013, 8, e72905. [Google Scholar] [CrossRef] [PubMed]
- Brighty, D.W.; Rosenberg, M. A cis-acting repressive sequence that overlaps the Rev-responsive element of human immunodeficiency virus type 1 regulates nuclear retention of env mRNAs independently of known splice signals. Proc. Natl. Acad. Sci. USA 1994, 91, 8314–8318. [Google Scholar] [CrossRef] [PubMed]
- Sodroski, J.; Goh, W.C.; Rosen, C.; Dayton, A.; Terwilliger, E.; Haseltine, W. A second post-transcriptional trans-activator gene required for HTLV-III replication. Nature 1986, 321, 412–417. [Google Scholar] [CrossRef] [PubMed]
- Terwilliger, E.; Burghoff, R.; Sia, R.; Sodroski, J.; Haseltine, W.; Rosen, C. The art gene product of human immunodeficiency virus is required for replication. J. Virol. 1988, 62, 655–658. [Google Scholar] [CrossRef] [PubMed]
- Suh, D.; Seguin, B.; Atkinson, S.; Ozdamar, B.; Staffa, A.; Emili, A.; Mouland, A.; Cochrane, A. Mapping of determinants required for the function of the HIV-1 env nuclear retention sequence. Virology 2003, 310, 85–99. [Google Scholar] [CrossRef] [PubMed]
- Pinol-Roma, S. HnRNP proteins and the nuclear export of mRNA. Semin. Cell Dev. Biol. 1997, 8, 57–63. [Google Scholar] [CrossRef] [PubMed]
- Nakielny, S.; Dreyfuss, G. The hnRNP C proteins contain a nuclear retention sequence that can override nuclear export signals. J. Cell Biol. 1996, 134, 1365–1373. [Google Scholar] [CrossRef] [PubMed]
- Gordon, H.; Ajamian, L.; Valiente-Echeverria, F.; Levesque, K.; Rigby, W.F.; Mouland, A.J. Depletion of hnRNP A2/B1 overrides the nuclear retention of the HIV-1 genomic RNA. RNA Biol. 2013, 10, 1714–1725. [Google Scholar] [CrossRef] [PubMed]
- Ryan, V.H.; Watters, S.; Amaya, J.; Khatiwada, B.; Venditti, V.; Naik, M.T.; Fawzi, N.L. Weak binding to the A2RE RNA rigidifies hnRNPA2 RRMs and reduces liquid-liquid phase separation and aggregation. Nucleic Acids Res. 2020, 48, 10542–10554. [Google Scholar] [CrossRef] [PubMed]
- Najera, I.; Krieg, M.; Karn, J. Synergistic stimulation of HIV-1 rev-dependent export of unspliced mRNA to the cytoplasm by hnRNP A1. J. Mol. Biol. 1999, 285, 1951–1964. [Google Scholar] [CrossRef] [PubMed]
- Izaurralde, E.; Jarmolowski, A.; Beisel, C.; Mattaj, I.W.; Dreyfuss, G.; Fischer, U. A role for the M9 transport signal of hnRNP A1 in mRNA nuclear export. J. Cell Biol. 1997, 137, 27–35. [Google Scholar] [CrossRef] [PubMed]
- Siomi, M.C.; Eder, P.S.; Kataoka, N.; Wan, L.; Liu, Q.; Dreyfuss, G. Transportin-mediated nuclear import of heterogeneous nuclear RNP proteins. J. Cell Biol. 1997, 138, 1181–1192. [Google Scholar] [CrossRef] [PubMed]
- Black, A.C.; Luo, J.; Chun, S.; Bakker, A.; Fraser, J.K.; Rosenblatt, J.D. Specific binding of polypyrimidine tract binding protein and hnRNP A1 to HIV-1 CRS elements. Virus Genes 1996, 12, 275–285. [Google Scholar] [CrossRef] [PubMed]
- Zolotukhin, A.S.; Michalowski, D.; Bear, J.; Smulevitch, S.V.; Traish, A.M.; Peng, R.; Patton, J.; Shatsky, I.N.; Felber, B.K. PSF acts through the human immunodeficiency virus type 1 mRNA instability elements to regulate virus expression. Mol. Cell. Biol. 2003, 23, 6618–6630. [Google Scholar] [CrossRef] [PubMed]
- Yandrapally, S.; Sarkar, S.; Banerjee, S. HIV-1 Tat commandeers nuclear export of Rev-viral RNA complex by controlling hnRNPA2-mediated splicing. J. Virol. 2023, 97, e0104423. [Google Scholar] [CrossRef] [PubMed]
- Cazalla, D.; Zhu, J.; Manche, L.; Huber, E.; Krainer, A.R.; Caceres, J.F. Nuclear export and retention signals in the RS domain of SR proteins. Mol. Cell. Biol. 2002, 22, 6871–6882. [Google Scholar] [CrossRef] [PubMed]
- Jacks, T.; Varmus, H.E. Expression of the Rous sarcoma virus pol gene by ribosomal frameshifting. Science 1985, 230, 1237–1242. [Google Scholar] [CrossRef] [PubMed]
- Namy, O.; Moran, S.J.; Stuart, D.I.; Gilbert, R.J.; Brierley, I. A mechanical explanation of RNA pseudoknot function in programmed ribosomal frameshifting. Nature 2006, 441, 244–247. [Google Scholar] [CrossRef] [PubMed]
- Jacks, T.; Power, M.D.; Masiarz, F.R.; Luciw, P.A.; Barr, P.J.; Varmus, H.E. Characterization of ribosomal frameshifting in HIV-1 gag-pol expression. Nature 1988, 331, 280–283. [Google Scholar] [CrossRef] [PubMed]
- Nam, S.H.; Kidokoro, M.; Shida, H.; Hatanaka, M. Processing of gag precursor polyprotein of human T-cell leukemia virus type I by virus-encoded protease. J. Virol. 1988, 62, 3718–3728. [Google Scholar] [CrossRef] [PubMed]
- Marra, M.A.; Jones, S.J.; Astell, C.R.; Holt, R.A.; Brooks-Wilson, A.; Butterfield, Y.S.; Khattra, J.; Asano, J.K.; Barber, S.A.; Chan, S.Y.; et al. The Genome sequence of the SARS-associated coronavirus. Science 2003, 300, 1399–1404. [Google Scholar] [CrossRef] [PubMed]
- Bhatt, P.R.; Scaiola, A.; Loughran, G.; Leibundgut, M.; Kratzel, A.; Meurs, R.; Dreos, R.; O‘Connor, K.M.; McMillan, A.; Bode, J.W.; et al. Structural basis of ribosomal frameshifting during translation of the SARS-CoV-2 RNA genome. Science 2021, 372, 1306–1313. [Google Scholar] [CrossRef] [PubMed]
- Belew, A.T.; Meskauskas, A.; Musalgaonkar, S.; Advani, V.M.; Sulima, S.O.; Kasprzak, W.K.; Shapiro, B.A.; Dinman, J.D. Ribosomal frameshifting in the CCR5 mRNA is regulated by miRNAs and the NMD pathway. Nature 2014, 512, 265–269. [Google Scholar] [CrossRef] [PubMed]
- Kang, H. Direct structural evidence for formation of a stem-loop structure involved in ribosomal frameshifting in human immunodeficiency virus type 1. Biochim. Biophys. Acta 1998, 1397, 73–78. [Google Scholar] [CrossRef]
- Mikl, M.; Pilpel, Y.; Segal, E. High-throughput interrogation of programmed ribosomal frameshifting in human cells. Nat. Commun. 2020, 11, 3061. [Google Scholar] [CrossRef] [PubMed]
- Cen, S.; Niu, M.; Saadatmand, J.; Guo, F.; Huang, Y.; Nabel, G.J.; Kleiman, L. Incorporation of pol into human immunodeficiency virus type 1 Gag virus-like particles occurs independently of the upstream Gag domain in Gag-pol. J. Virol. 2004, 78, 1042–1049. [Google Scholar] [CrossRef] [PubMed]
- Shehu-Xhilaga, M.; Crowe, S.M.; Mak, J. Maintenance of the Gag/Gag-Pol ratio is important for human immunodeficiency virus type 1 RNA dimerization and viral infectivity. J. Virol. 2001, 75, 1834–1841. [Google Scholar] [CrossRef] [PubMed]
- Napthine, S.; Hill, C.H.; Nugent, H.C.M.; Brierley, I. Modulation of Viral Programmed Ribosomal Frameshifting and Stop Codon Readthrough by the Host Restriction Factor Shiftless. Viruses 2021, 13, 1230. [Google Scholar] [CrossRef] [PubMed]
- Zimmer, M.M.; Kibe, A.; Rand, U.; Pekarek, L.; Ye, L.; Buck, S.; Smyth, R.P.; Cicin-Sain, L.; Caliskan, N. The short isoform of the host antiviral protein ZAP acts as an inhibitor of SARS-CoV-2 programmed ribosomal frameshifting. Nat. Commun. 2021, 12, 7193. [Google Scholar] [CrossRef] [PubMed]
- Wang, X.; Xuan, Y.; Han, Y.; Ding, X.; Ye, K.; Yang, F.; Gao, P.; Goff, S.P.; Gao, G. Regulation of HIV-1 Gag-Pol Expression by Shiftless, an Inhibitor of Programmed -1 Ribosomal Frameshifting. Cell 2019, 176, 625–635.e14. [Google Scholar] [CrossRef] [PubMed]
- Jager, N.; Ayyub, S.A.; Korniy, N.; Peske, F.; Hoffmann, M.; Rodnina, M.V.; Pohlmann, S. Mutagenic Analysis of the HIV Restriction Factor Shiftless. Viruses 2022, 14, 1454. [Google Scholar] [CrossRef] [PubMed]
- Yu, D.; Zhao, Y.; Pan, J.; Yang, X.; Liang, Z.; Xie, S.; Cao, R. C19orf66 Inhibits Japanese Encephalitis Virus Replication by Targeting -1 PRF and the NS3 Protein. Virol. Sin. 2021, 36, 1443–1455. [Google Scholar] [CrossRef] [PubMed]
- Suzuki, Y.; Chin, W.X.; Han, Q.; Ichiyama, K.; Lee, C.H.; Eyo, Z.W.; Ebina, H.; Takahashi, H.; Takahashi, C.; Tan, B.H.; et al. Characterization of RyDEN (C19orf66) as an Interferon-Stimulated Cellular Inhibitor against Dengue Virus Replication. PLoS Pathog. 2016, 12, e1005357. [Google Scholar] [CrossRef] [PubMed]
- Wu, Y.; Yang, X.; Yao, Z.; Dong, X.; Zhang, D.; Hu, Y.; Zhang, S.; Lin, J.; Chen, J.; An, S.; et al. C19orf66 interrupts Zika virus replication by inducing lysosomal degradation of viral NS3. PLoS Neglected Trop. Dis. 2020, 14, e0008083. [Google Scholar] [CrossRef] [PubMed]
- Kinast, V.; Plociennikowska, A.; Anggakusuma; Bracht, T.; Todt, D.; Brown, R.J.P.; Boldanova, T.; Zhang, Y.; Bruggemann, Y.; Friesland, M.; et al. C19orf66 is an interferon-induced inhibitor of HCV replication that restricts formation of the viral replication organelle. J. Hepatol. 2020, 73, 549–558. [Google Scholar] [CrossRef] [PubMed]
- Rodriguez, W.; Mehrmann, T.; Hatfield, D.; Muller, M. Shiftless Restricts Viral Gene Expression and Influences RNA Granule Formation during Kaposi’s Sarcoma-Associated Herpesvirus Lytic Replication. J. Virol. 2022, 96, e0146922. [Google Scholar] [CrossRef] [PubMed]
- Gao, G.; Guo, X.; Goff, S.P. Inhibition of retroviral RNA production by ZAP, a CCCH-type zinc finger protein. Science 2002, 297, 1703–1706. [Google Scholar] [CrossRef] [PubMed]
- Müller, S.; Möller, P.; Bick, M.J.; Wurr, S.; Becker, S.; Günther, S.; Kümmerer, B.M. Inhibition of filovirus replication by the zinc finger antiviral protein. J. Virol. 2007, 81, 2391–2400. [Google Scholar] [CrossRef] [PubMed]
- Tang, Q.; Wang, X.; Gao, G. The Short Form of the Zinc Finger Antiviral Protein Inhibits Influenza A Virus Protein Expression and Is Antagonized by the Virus-Encoded NS1. J. Virol. 2017, 91. [Google Scholar] [CrossRef]
- Chiu, H.P.; Chiu, H.; Yang, C.F.; Lee, Y.L.; Chiu, F.L.; Kuo, H.C.; Lin, R.J.; Lin, Y.L. Inhibition of Japanese encephalitis virus infection by the host zinc-finger antiviral protein. PLoS Pathog. 2018, 14, e1007166. [Google Scholar] [CrossRef] [PubMed]
- Zhu, Y.; Chen, G.; Lv, F.; Wang, X.; Ji, X.; Xu, Y.; Sun, J.; Wu, L.; Zheng, Y.T.; Gao, G. Zinc-finger antiviral protein inhibits HIV-1 infection by selectively targeting multiply spliced viral mRNAs for degradation. Proc. Natl. Acad. Sci. USA 2011, 108, 15834–15839. [Google Scholar] [CrossRef] [PubMed]
- Trono, D.; Van Lint, C.; Rouzioux, C.; Verdin, E.; Barre-Sinoussi, F.; Chun, T.W.; Chomont, N. HIV persistence and the prospect of long-term drug-free remissions for HIV-infected individuals. Science 2010, 329, 174–180. [Google Scholar] [CrossRef] [PubMed]
- Chun, T.W.; Moir, S.; Fauci, A.S. HIV reservoirs as obstacles and opportunities for an HIV cure. Nat. Immunol. 2015, 16, 584–589. [Google Scholar] [CrossRef] [PubMed]
- Pennings, P.S. HIV Drug Resistance: Problems and Perspectives. Infect. Dis. Rep. 2013, 5, e5. [Google Scholar] [CrossRef] [PubMed]
- Saayman, S.M.; Lazar, D.C.; Scott, T.A.; Hart, J.R.; Takahashi, M.; Burnett, J.C.; Planelles, V.; Morris, K.V.; Weinberg, M.S. Potent and Targeted Activation of Latent HIV-1 Using the CRISPR/dCas9 Activator Complex. Mol. Ther. J. Am. Soc. Gene Ther. 2016, 24, 488–498. [Google Scholar] [CrossRef] [PubMed]
- Soret, J.; Bakkour, N.; Maire, S.; Durand, S.; Zekri, L.; Gabut, M.; Fic, W.; Divita, G.; Rivalle, C.; Dauzonne, D.; et al. Selective modification of alternative splicing by indole derivatives that target serine-arginine-rich protein splicing factors. Proc. Natl. Acad. Sci. USA 2005, 102, 8764–8769. [Google Scholar] [CrossRef] [PubMed]
- Bakkour, N.; Lin, Y.L.; Maire, S.; Ayadi, L.; Mahuteau-Betzer, F.; Nguyen, C.H.; Mettling, C.; Portales, P.; Grierson, D.; Chabot, B.; et al. Small-molecule inhibition of HIV pre-mRNA splicing as a novel antiretroviral therapy to overcome drug resistance. PLoS Pathog. 2007, 3, 1530–1539. [Google Scholar] [CrossRef] [PubMed]
- Soret, J.; Gabut, M.; Tazi, J. SR proteins as potential targets for therapy. Prog. Mol. Subcell. Biol. 2006, 44, 65–87. [Google Scholar] [CrossRef] [PubMed]
- Cheung, P.K.; Horhant, D.; Bandy, L.E.; Zamiri, M.; Rabea, S.M.; Karagiosov, S.K.; Matloobi, M.; McArthur, S.; Harrigan, P.R.; Chabot, B.; et al. A Parallel Synthesis Approach to the Identification of Novel Diheteroarylamide-Based Compounds Blocking HIV Replication: Potential Inhibitors of HIV-1 Pre-mRNA Alternative Splicing. J. Med. Chem. 2016, 59, 1869–1879. [Google Scholar] [CrossRef] [PubMed]
- Shkreta, L.; Blanchette, M.; Toutant, J.; Wilhelm, E.; Bell, B.; Story, B.A.; Balachandran, A.; Cochrane, A.; Cheung, P.K.; Harrigan, P.R.; et al. Modulation of the splicing regulatory function of SRSF10 by a novel compound that impairs HIV-1 replication. Nucleic Acids Res. 2017, 45, 4051–4067. [Google Scholar] [CrossRef] [PubMed]
- Tazi, J.; Bakkour, N.; Marchand, V.; Ayadi, L.; Aboufirassi, A.; Branlant, C. Alternative splicing: Regulation of HIV-1 multiplication as a target for therapeutic action. FEBS J. 2010, 277, 867–876. [Google Scholar] [CrossRef] [PubMed]
- Wong, R.W.; Balachandran, A.; Ostrowski, M.A.; Cochrane, A. Digoxin suppresses HIV-1 replication by altering viral RNA processing. PLoS Pathog. 2013, 9, e1003241. [Google Scholar] [CrossRef] [PubMed]
- Wong, R.W.; Balachandran, A.; Haaland, M.; Stoilov, P.; Cochrane, A. Characterization of novel inhibitors of HIV-1 replication that function via alteration of viral RNA processing and rev function. Nucleic Acids Res. 2013, 41, 9471–9483. [Google Scholar] [CrossRef] [PubMed]
- Campos, N.; Myburgh, R.; Garcel, A.; Vautrin, A.; Lapasset, L.; Nadal, E.S.; Mahuteau-Betzer, F.; Najman, R.; Fornarelli, P.; Tantale, K.; et al. Long lasting control of viral rebound with a new drug ABX464 targeting Rev—Mediated viral RNA biogenesis. Retrovirology 2015, 12, 30. [Google Scholar] [CrossRef] [PubMed]
- Hammond, S.M.; Caudy, A.A.; Hannon, G.J. Post-transcriptional gene silencing by double-stranded RNA. Nat. Rev. Genet. 2001, 2, 110–119. [Google Scholar] [CrossRef] [PubMed]
- Meister, G.; Tuschl, T. Mechanisms of gene silencing by double-stranded RNA. Nature 2004, 431, 343–349. [Google Scholar] [CrossRef] [PubMed]
- Asparuhova, M.B.; Marti, G.; Liu, S.; Serhan, F.; Trono, D.; Schumperli, D. Inhibition of HIV-1 multiplication by a modified U7 snRNA inducing Tat and Rev exon skipping. J. Gene Med. 2007, 9, 323–334. [Google Scholar] [CrossRef] [PubMed]
- Asparuhova, M.B.; Barde, I.; Trono, D.; Schranz, K.; Schumperli, D. Development and characterization of a triple combination gene therapy vector inhibiting HIV-1 multiplication. J. Gene Med. 2008, 10, 1059–1070. [Google Scholar] [CrossRef] [PubMed]
- Chen, M.J.; Gatignol, A.; Scarborough, R.J. The discovery and development of RNA-based therapies for treatment of HIV-1 infection. Expert Opin. Drug Discov. 2023, 18, 163–179. [Google Scholar] [CrossRef] [PubMed]
- Mandal, D.; Feng, Z.; Stoltzfus, C.M. Excessive RNA splicing and inhibition of HIV-1 replication induced by modified U1 small nuclear RNAs. J. Virol. 2010, 84, 12790–12800. [Google Scholar] [CrossRef] [PubMed]
- Del Corpo, O.; Goguen, R.P.; Malard, C.M.G.; Daher, A.; Colby-Germinario, S.; Scarborough, R.J.; Gatignol, A. A U1i RNA that Enhances HIV-1 RNA Splicing with an Elongated Recognition Domain Is an Optimal Candidate for Combination HIV-1 Gene Therapy. Mol. Ther. Nucleic Acids 2019, 18, 815–830. [Google Scholar] [CrossRef] [PubMed]
- Grunweller, A.; Hartmann, R.K. Locked nucleic acid oligonucleotides: The next generation of antisense agents? BioDrugs Clin. Immunother. Biopharm. Gene Ther. 2007, 21, 235–243. [Google Scholar]
- Lundin, K.E.; Hojland, T.; Hansen, B.R.; Persson, R.; Bramsen, J.B.; Kjems, J.; Koch, T.; Wengel, J.; Smith, C.I. Biological activity and biotechnological aspects of locked nucleic acids. Adv. Genet. 2013, 82, 47–107. [Google Scholar] [CrossRef] [PubMed]
- Stein, C.A.; Hansen, J.B.; Lai, J.; Wu, S.; Voskresenskiy, A.; Hog, A.; Worm, J.; Hedtjarn, M.; Souleimanian, N.; Miller, P.; et al. Efficient gene silencing by delivery of locked nucleic acid antisense oligonucleotides, unassisted by transfection reagents. Nucleic Acids Res. 2010, 38, e3. [Google Scholar] [CrossRef] [PubMed]
- Zhang, Y.; Qu, Z.; Kim, S.; Shi, V.; Liao, B.; Kraft, P.; Bandaru, R.; Wu, Y.; Greenberger, L.M.; Horak, I.D. Down-modulation of cancer targets using locked nucleic acid (LNA)-based antisense oligonucleotides without transfection. Gene Ther. 2011, 18, 326–333. [Google Scholar] [CrossRef] [PubMed]
- Hillebrand, F.; Ostermann, P.N.; Muller, L.; Degrandi, D.; Erkelenz, S.; Widera, M.; Pfeffer, K.; Schaal, H. Gymnotic Delivery of LNA Mixmers Targeting Viral SREs Induces HIV-1 mRNA Degradation. Int. J. Mol. Sci. 2019, 20, 1088. [Google Scholar] [CrossRef] [PubMed]
- Arzumanov, A.; Walsh, A.P.; Rajwanshi, V.K.; Kumar, R.; Wengel, J.; Gait, M.J. Inhibition of HIV-1 Tat-dependent trans activation by steric block chimeric 2′-O-methyl/LNA oligoribonucleotides. Biochemistry 2001, 40, 14645–14654. [Google Scholar] [CrossRef] [PubMed]
- Lebars, I.; Richard, T.; Di Primo, C.; Toulme, J.J. LNA derivatives of a kissing aptamer targeted to the trans-activating responsive RNA element of HIV-1. Blood Cells Mol. Dis. 2007, 38, 204–209. [Google Scholar] [CrossRef] [PubMed]
- Jakobsen, M.R.; Haasnoot, J.; Wengel, J.; Berkhout, B.; Kjems, J. Efficient inhibition of HIV-1 expression by LNA modified antisense oligonucleotides and DNAzymes targeted to functionally selected binding sites. Retrovirology 2007, 4, 29. [Google Scholar] [CrossRef] [PubMed]
- Holmes, S.C.; Arzumanov, A.A.; Gait, M.J. Steric inhibition of human immunodeficiency virus type-1 Tat-dependent trans-activation in vitro and in cells by oligonucleotides containing 2′-O-methyl G-clamp ribonucleoside analogues. Nucleic Acids Res. 2003, 31, 2759–2768. [Google Scholar] [CrossRef] [PubMed]
- Uleri, E.; Beltrami, S.; Gordon, J.; Dolei, A.; Sariyer, I.K. Extinction of Tumor Antigen Expression by SF2/ASF in JCV-Transformed Cells. Genes Cancer 2011, 2, 728–736. [Google Scholar] [CrossRef] [PubMed]
- Uleri, E.; Regan, P.; Dolei, A.; Sariyer, I.K. SF2/ASF binding region within JC virus NCCR limits early gene transcription in glial cells. Virol. J. 2013, 10, 147. [Google Scholar] [CrossRef] [PubMed]
- Sariyer, I.K.; Khalili, K. Regulation of human neurotropic JC virus replication by alternative splicing factor SF2/ASF in glial cells. PLoS ONE 2011, 6, e14630. [Google Scholar] [CrossRef] [PubMed]
- Sariyer, I.K.; Sariyer, R.; Otte, J.; Gordon, J. Pur-Alpha Induces JCV Gene Expression and Viral Replication by Suppressing SRSF1 in Glial Cells. PLoS ONE 2016, 11, e0156819. [Google Scholar] [CrossRef] [PubMed]
- McPhillips, M.G.; Veerapraditsin, T.; Cumming, S.A.; Karali, D.; Milligan, S.G.; Boner, W.; Morgan, I.M.; Graham, S.V. SF2/ASF binds the human papillomavirus type 16 late RNA control element and is regulated during differentiation of virus-infected epithelial cells. J. Virol. 2004, 78, 10598–10605. [Google Scholar] [CrossRef] [PubMed]
- Molin, M.; Akusjarvi, G. Overexpression of essential splicing factor ASF/SF2 blocks the temporal shift in adenovirus pre-mRNA splicing and reduces virus progeny formation. J. Virol. 2000, 74, 9002–9009. [Google Scholar] [CrossRef] [PubMed]
- Ofori, L.O.; Hilimire, T.A.; Bennett, R.P.; Brown, N.W., Jr.; Smith, H.C.; Miller, B.L. High-affinity recognition of HIV-1 frameshift-stimulating RNA alters frameshifting in vitro and interferes with HIV-1 infectivity. J. Med. Chem. 2014, 57, 723–732. [Google Scholar] [CrossRef] [PubMed]
- Anokhina, V.S.; Miller, B.L. Targeting Ribosomal Frameshifting as an Antiviral Strategy: From HIV-1 to SARS-CoV-2. Acc. Chem. Res. 2021, 54, 3349–3361. [Google Scholar] [CrossRef] [PubMed]
- Knizhnik, E.; Chumakov, S.; Svetlova, J.; Pavlova, I.; Khodarovich, Y.; Brylev, V.; Severov, V.; Alieva, R.; Kozlovskaya, L.; Andreev, D.; et al. Unwinding the SARS-CoV-2 Ribosomal Frameshifting Pseudoknot with LNA and G-Clamp-Modified Phosphorothioate Oligonucleotides Inhibits Viral Replication. Biomolecules 2023, 13, 1660. [Google Scholar] [CrossRef] [PubMed]
IFN-I-Regulated Genes | ||||
---|---|---|---|---|
Protein Levels | SRSF1 High | SRSF1 Low | hnRNP A0 High | hnRNP A0 Low |
transcription | LTR inhibition, Competes with Tat-TAR-binding and trans-activation | LTR activation, increased levels of total HIV-1 mRNAs | Tat-LTR inhibition | Tat-LTR activation |
alternative splicing | increase in vif, vpr, tat2, tat3 splicing efficiency; decreased exon3, tat1, and US HIV-1 mRNA efficiency | increase in US HIV-1 mRNA; decreased exon2, exon3, vif, vpr, tat2, and MS HIV-1 mRNA splicing efficiency | increase in exon2 and vif mRNA splicing | no impact on alternative splice site use |
mRNA trafficking | Facilitated export of unspliced 9 kb gag/pol mRNA Nuclear retention of 4 kb intron-containing env1 mRNA | Facilitated export of 4 kb intron-containing env1 mRNA Nuclear retainment of unspliced 9 kb gag/pol mRNA | Retains unspliced HIV-1 mRNA in the nucleus | Facilitated nuclear export of unspliced HIV-1 mRNA |
programmed ribosomal frameshifting | n.d. | n.d. | Reduced -1PRF efficiency | Slightly facilitated -1PRF efficiency |
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Roesmann, F.; Müller, L.; Klaassen, K.; Heß, S.; Widera, M. Interferon-Regulated Expression of Cellular Splicing Factors Modulates Multiple Levels of HIV-1 Gene Expression and Replication. Viruses 2024, 16, 938. https://doi.org/10.3390/v16060938
Roesmann F, Müller L, Klaassen K, Heß S, Widera M. Interferon-Regulated Expression of Cellular Splicing Factors Modulates Multiple Levels of HIV-1 Gene Expression and Replication. Viruses. 2024; 16(6):938. https://doi.org/10.3390/v16060938
Chicago/Turabian StyleRoesmann, Fabian, Lisa Müller, Katleen Klaassen, Stefanie Heß, and Marek Widera. 2024. "Interferon-Regulated Expression of Cellular Splicing Factors Modulates Multiple Levels of HIV-1 Gene Expression and Replication" Viruses 16, no. 6: 938. https://doi.org/10.3390/v16060938
APA StyleRoesmann, F., Müller, L., Klaassen, K., Heß, S., & Widera, M. (2024). Interferon-Regulated Expression of Cellular Splicing Factors Modulates Multiple Levels of HIV-1 Gene Expression and Replication. Viruses, 16(6), 938. https://doi.org/10.3390/v16060938