625
Views
30
CrossRef citations to date
0
Altmetric
Review Article

SAMHD1: Recurring roles in cell cycle, viral restriction, cancer, and innate immunity

&
Pages 96-110 | Received 18 Sep 2017, Accepted 16 Mar 2018, Published online: 27 Mar 2018

References

  • Traut TW. Physiological concentrations of purines and pyrimidines. Mol Cell Biochem. 1994;140:1–22.
  • Reichard P. Interactions between deoxyribonucleotide and DNA synthesis. Annu Rev Biochem. 1988;57:349–374.
  • Mathews CK. Deoxyribonucleotides as genetic and metabolic regulators. FASEB J. 2014;28:3832–3840.
  • Mathews CK. DNA precursor metabolism and genomic stability. FASEB J. 2006;20:1300–1314.
  • Pai C, Kearsey S. A critical balance: dNTPs and the maintenance of genome stability. Genes (Basel). 2017;8:57.
  • Rampazzo C, Miazzi C, Franzolin E, et al. Regulation by degradation, a cellular defense against deoxyribonucleotide pool imbalances. Mutat Res Genet Toxicol Environ Mutagen. 2010;703:2–10.
  • Lane AN, Fan TWM. Regulation of mammalian nucleotide metabolism and biosynthesis. Nucleic Acids Res. 2015;43:2466–2485.
  • Hofer A, Crona M, Logan DT, et al. DNA building blocks: keeping control of manufacture. Crit Rev Biochem Mol Biol. 2012;47:50–63.
  • Walldén K, Nordlund P. Structural basis for the allosteric regulation and substrate recognition of human cytosolic 5′-nucleotidase II. J Mol Biol. 2011;408:684–696.
  • Johansson K, Ramaswamy S, Ljungcrantz C, et al. Structural basis for substrate specificities of cellular deoxyribonucleoside kinases. Nat Struct Biol. 2001;8:616–620.
  • Hunsucker SA, Mitchell BS, Spychala J. The 5′-nucleotidases as regulators of nucleotide and drug metabolism. Pharmacol Ther. 2005;107:1–30.
  • Bianchi V, Spychala J. Mammalian 5′-nucleotidases. J Biol Chem. 2003;278:46195–46198.
  • Nordlund P, Reichard P. Ribonucleotide reductases. Annu Rev Biochem. 2006;75:681–706.
  • Håkansson P, Hofer A, Thelander L. Regulation of mammalian ribonucleotide reduction and dNTP pools after DNA damage and in resting cells. J Biol Chem. 2006;281:7834–7841.
  • Pontarin G, Ferraro P, Bee L, et al. Mammalian ribonucleotide reductase subunit p53R2 is required for mitochondrial DNA replication and DNA repair in quiescent cells. Proc Natl Acad Sci USA. 2012;109:13302–13307.
  • Pontarin G, Ferraro P, Rampazzo C, et al. Deoxyribonucleotide metabolism in cycling and resting human fibroblasts with a missense mutation in p53R2, a subunit of ribonucleotide reductase. J Biol Chem. 2011;286:11132–11140.
  • Pontarin G, Fijolek A, Pizzo P, et al. Ribonucleotide reduction is a cytosolic process in mammalian cells independently of DNA damage. Proc Natl Acad Sci. 2008;105:17801–17806.
  • Choi JS, Berdis AJ. Nucleoside transporters: biological insights and therapeutic applications. Future Med Chem. 2012;4:1461–1478.
  • Nicander B, Reichard P. Evidence for the involvement of substrate cycles in the regulation of deoxyribonucleoside triphosphate pools in 3T6 cells. Pediatr Res. 1985;260:9216–9222.
  • Bianchi V, Pontis E, Reichard P. Dynamics of the dATP pool in cultured mammalian cells. Exp Cell Res. 1992;199:120–128.
  • Spyrou G, Reichard P. Dynamcis of the thymidine triphosphate pool during the cell cycle of synchronized 3T3 mouse fibroblasts. Mutat Res Fundam Mol Mech Mutagen. 1988;200:37–43.
  • Leanza L, Ferraro P, Reichard P, et al. Metabolic interrelations within guanine deoxynucleotide pools for mitochondrial and nuclear DNA maintenance. J Biol Chem. 2008;283:16437–16445.
  • Rampazzo C, Ferraro P, Pontarin G, et al. Mitochondrial deoxyribonucleotides, pool sizes, synthesis, and regulation. J Biol Chem. 2004;279:17019–17026.
  • Gandhi VV, Samuels DC. A review comparing deoxyribonucleoside triphosphate (dNTP) concentrations in the mitochondrial and cytoplasmic compartments of normal and transformed cells. Nucleosides Nucleotides Nucleic Acids. 2011;30:317–339.
  • Meuth M. The molecular basis of mutations induced by deoxyribonucleoside triphosphate pool imbalances in mammalian cells. Exp Cell Res. 1989;181:305–316.
  • Kumar D, Abdulovic AL, Viberg J, et al. Mechanisms of mutagenesis in vivo due to imbalanced dNTP pools. Nucleic Acids Res. 2011;39:1360–1371.
  • Kumar D, Viberg J, Nilsson AK, et al. Highly mutagenic and severely imbalanced dNTP pools can escape detection by the S-phase checkpoint. Nucleic Acids Res. 2010;38:3975–3983.
  • Buckland RJ, Watt DL, Chittoor B, et al. Increased and imbalanced dNTP pools symmetrically promote both leading and lagging strand replication infidelity. PLoS Genet. 2014;10:e1004846.
  • Watt DL, Buckland RJ, Lujan SA, et al. Genome-wide analysis of the specificity and mechanisms of replication infidelity driven by imbalanced dNTP pools. Nucleic Acids Res. 2015;44:1669–1680.
  • Chabes A, Stillman B. Constitutively high dNTP concentration inhibits cell cycle progression and the DNA damage checkpoint in yeast Saccharomyces cerevisiae. Proc Natl Acad Sci USA. 2007;104:1183–1188.
  • Davidson MB, Katou Y, Keszthelyi A, et al. Endogenous DNA replication stress results in expansion of dNTP pools and a mutator phenotype. EMBO J. 2012;31:895–907.
  • Gemble S, Buhagiar-Labarchède G, Onclercq-Delic R, et al. A balanced pyrimidine pool is required for optimal Chk1 activation to prevent ultrafine anaphase bridge formation. J Cell Sci. 2016;129:3167–3177.
  • Sabouri N, Viberg J, Goyal DK, et al. Evidence for lesion bypass by yeast replicative DNA polymerases during DNA damage. Nucleic Acids Res. 2008;36:5660–5667.
  • Ke PY, Kuo YY, Hu CM, et al. Control of dTTP pool size by anaphase promoting complex/cyclosome is essential for the maintenance of genetic stability. Genes Dev. 2005;19:1920–1933.
  • Hastak K, Paul RK, Agarwal MK, et al. DNA synthesis from unbalanced nucleotide pools causes limited DNA damage that triggers ATR-CHK1-dependent p53 activation. Proc Natl Acad Sci USA. 2008;105:6314–6319.
  • Anglana M, Apiou F, Bensimon A, et al. Dynamics of DNA replication in mammalian somatic cells: nucleotide pool modulates origin choice and interorigin spacing. Cell. 2003;114:385–394.
  • Poli J, Tsaponina O, Crabbé L, et al. dNTP pools determine fork progression and origin usage under replication stress. EMBO J. 2012;31:883–894.
  • Papadopoulou C, Guilbaud G, Schiavone D, et al. Nucleotide pool depletion induces G-quadruplex-dependent perturbation of gene expression. Cell Rep. 2015;13:2491–2503.
  • Jasencakova Z, Scharf AND, Ask K, et al. Replication stress interferes with histone recycling and predeposition marking of new histones. Mol Cell. 2010;37:736–743.
  • Mathews CK. Deoxyribonucleotide metabolism, mutagenesis and cancer. Nat Rev Cancer. 2015;15:528–539.
  • Boss GR, Seegmiller JE. Genetic defects in human purine and pyrimidine metabolism. Annu Rev Genet. 1982;16:297–328.
  • El-Hattab AW, Craigen WJ, Scaglia F. Mitochondrial DNA maintenance defects. Biochim Biophys Acta. 2017;1863:1539–1555.
  • Chabosseau P, Buhagiar-Labarchède G, Onclercq-Delic R, et al. Pyrimidine pool imbalance induced by BLM helicase deficiency contributes to genetic instability in Bloom syndrome. Nat Comms. 2011;2:368.
  • Kimura T, Takeda S, Sagiya Y, et al. Impaired function of p53R2 in Rrm2b-null mice causes severe renal failure through attenuation of dNTP pools. Nat Genet. 2003;34:440–445.
  • Aird KM, Zhang G, Li H, et al. Suppression of nucleotide metabolism underlies the establishment and maintenance of oncogene-induced senescence. Cell Rep. 2013;3:1252–1265.
  • Aird KM, Zhang R. Nucleotide metabolism, oncogene-induced senescence and cancer. Cancer Lett. 2015;356:204–210.
  • Chang L, Guo R, Huang Q, et al. Chromosomal instability triggered by Rrm2b loss leads to IL-6 secretion and plasmacytic neoplasms. Cell Rep. 2013;3:1389–1397.
  • Oliver FJ, Collins MKL, López-Rivas A. dNTP pools imbalance as a signal to initiate apoptosis. Experientia. 1996;52:995–1000.
  • James SJ, Basnakian AG, Miller BJ. In vitro folate deficiency induces deoxynucleotide pool imbalance, apoptosis, and mutagenesis in Chinese hamster ovary cells. Cancer Res. 1994;54:5075–5080.
  • Chandra D, Bratton SB, Person MD, et al. Intracellular nucleotides act as critical prosurvival factors by binding to cytochrome C and inhibiting apoptosome. Cell. 2006;125:1333–1346.
  • Li N, Zhang W, Cao X. Identification of human homologue of mouse IFN-γ induced protein from human dendritic cells. Immunol Lett. 2000;74:221–224.
  • Rice GI, Kasher PR, Forte GMA, et al. Mutations in ADAR1 cause Aicardi-Goutières syndrome associated with a type I interferon signature. Nat Genet. 2012;44:1243–1248.
  • Powell RD, Holland PJ, Hollis T, et al. Aicardi-goutieres syndrome gene and HIV-1 restriction factor SAMHD1 is a dGTP-regulated deoxynucleotide triphosphohydrolase. J Biol Chem. 2011;286:43596–43600.
  • Goldstone DC, Ennis-Adeniran V, Hedden JJ, et al. HIV-1 restriction factor SAMHD1 is a deoxynucleoside triphosphate triphosphohydrolase. Nature. 2011;480:379–382.
  • Hansen EC, Seamon KJ, Cravens SL, et al. GTP activator and dNTP substrates of HIV-1 restriction factor SAMHD1 generate a long-lived activated state. Proc Natl Acad Sci. 2014;111:E1843–E1851.
  • Amie SM, Bambara RA, Kim B. GTP Is the primary activator of the anti-HIV restriction factor SAMHD1. J Biol Chem. 2013;288:25001–25006.
  • Ji X, Tang C, Zhao Q, et al. Structural basis of cellular dNTP regulation by SAMHD1. Proc Natl Acad Sci USA. 2014;111:E4305–E4314.
  • Beauchamp BB, Richardson CC. A unique deoxyguanosine triphosphatase is responsible for the optA1 phenotype of Escherichia coli. Proc Natl Acad Sci USA. 1988;85:2563–2567.
  • Vorontsov II, Minasov G, Kiryukhina O, et al. Characterization of the deoxynucleotide triphosphate triphosphohydrolase (dNTPase) activity of the EF1143 protein from Enterococcus faecalis and crystal structure of the activator-substrate complex. J Biol Chem. 2011;286:33158–33166.
  • Oganesyan V, Adams PD, Jancarik J, et al. Structure of O67745_AQUAE, a hypothetical protein from Aquifex aeolicus. Acta Crystallogr Sect F Struct Biol Cryst Commun. 2007;63:369–374.
  • Kondo N, Kuramitsu S, Masui R. Biochemical characterization of TT1383 from Thermus thermophilus identifies a novel dNTP triphosphohydrolase activity stimulated by dATP and dTTP. J Biochem. 2004;136:221–231.
  • Kondo N, Nakagawa N, Ebihara A, et al. Structure of dNTP-inducible dNTP triphosphohydrolase: insight into broad specificity for dNTPs and triphosphohydrolase-type hydrolysis. Acta Crystallogr Sect D Biol Crystallogr. 2007;63:230–239.
  • Mega R, Kondo N, Nakagawa N, et al. Two dNTP triphosphohydrolases from Pseudomonas aeruginosa possess diverse substrate specificities. FEBS J. 2009;276:3211–3221.
  • Zimmerman MD, Proudfoot M, Yakunin A, et al. Structural insight into the mechanism of substrate specificity and catalytic activity of an HD-domain phosphohydrolase: the 5’-deoxyribonucleotidase YfbR from Escherichia coli. J Mol Biol. 2008;378:215–226.
  • Qiao F, Bowie JU. The many faces of SAM. Sci Signal. 2005;2005:re7–re7.
  • Brandariz-Nuñez A, Valle-Casuso J, White TE, et al. Role of SAMHD1 nuclear localization in restriction of HIV-1 and SIVmac. Retrovirology. 2012;9:49.
  • Hofmann H, Logue EC, Bloch N, et al. The Vpx lentiviral accessory protein targets SAMHD1 for degradation in the nucleus. J Virol. 2012;86:12552–12560.
  • Aravind L, Koonin EV. The HD domain defines a new superfamily of metal-dependent phosphohydrolases. Trends Biochem Sci. 1998;23:469–472.
  • DeLucia M, Mehrens J, Wu Y, et al. HIV-2 and SIVmac accessory virulence factor Vpx down-regulates SAMHD1 enzyme catalysis prior to proteasome-dependent degradation. J Biol Chem. 2013;288:19116–19126.
  • Zhu C, Gao W, Zhao K, et al. Structural insight into dGTP-dependent activation of tetrameric SAMHD1 deoxynucleoside triphosphate triphosphohydrolase. Nat Commun. 2013;4:2722.
  • Ji X, Wu Y, Yan J, et al. Mechanism of allosteric activation of SAMHD1 by dGTP. Nat Struct Mol Biol. 2013;20:1304–1309.
  • Yan J, Kaur S, DeLucia M, et al. Tetramerization of SAMHD1 is required for biological activity and inhibition of HIV infection. J Biol Chem. 2013;288:10406–10417.
  • Miazzi C, Ferraro P, Pontarin G, et al. Allosteric regulation of the human and mouse deoxyribonucleotide triphosphohydrolase sterile α-motif/histidine-aspartate domain-containing protein 1 (SAMHD1). J Biol Chem. 2014;289:18339–18346.
  • Zhu CF, Wei W, Peng X, et al. The mechanism of substrate-controlled allosteric regulation of SAMHD1 activated by GTP. Acta Crystallogr D Biol Crystallogr. 2015;71:516–524.
  • Li Y, Kong J, Peng X, et al. Structural insights into the high-efficiency catalytic mechanism of the sterile α-motif/histidine-aspartate domain-containing protein. J Biol Chem. 2015;290:29428–29437.
  • Koharudin LMI, Wu Y, DeLucia M, et al. Structural basis of allosteric activation of sterile motif and histidine-aspartate domain-containing protein 1 (SAMHD1) by nucleoside triphosphates. J Biol Chem. 2014;289:32617–32627.
  • Wang Z, Bhattacharya A, Villacorta J, et al. Allosteric activation of SAMHD1 protein by deoxynucleotide triphosphate (dNTP)-dependent tetramerization requires dNTP concentrations that are similar to dNTP concentrations observed in cycling T cells. J Biol Chem. 2016;291:21407–21413.
  • Jang S, Zhou X, Ahn J. Substrate specificity of SAMHD1 triphosphohydrolase activity is controlled by deoxyribonucleoside triphosphates and phosphorylation at thr592. Biochemistry. 2016;55:5635–5646.
  • Amie SM, Daly MB, Noble E, et al. Anti-HIV host factor SAMHD1 regulates viral sensitivity to nucleoside reverse transcriptase inhibitors via modulation of cellular deoxyribonucleoside triphosphate (dNTP) levels. J Biol Chem. 2013;288:20683–20691.
  • Seamon KJ, Hansen EC, Kadina AP, et al. Small molecule inhibition of SAMHD1 dNTPase by tetramer destabilization. J Am Chem Soc. 2014;136:9822–9825.
  • Arnold LH, Groom HCT, Kunzelmann S, et al. Phospho-dependent regulation of SAMHD1 oligomerisation couples catalysis and restriction. PLoS Pathog. 2015;11:e1005194.
  • Mauney CH, Rogers LC, Harris RS, et al. The SAMHD1 dNTP triphosphohydrolase is controlled by a redox switch. Antioxid Redox Signal. 2017;27:1317–1331.
  • Seamon KJ, Sun Z, Shlyakhtenko LS, et al. SAMHD1 is a single-stranded nucleic acid binding protein with no active site-associated nuclease activity. Nucleic Acids Res. 2015;43:6486–6499.
  • Goncalves A, Karayel E, Rice GI, et al. SAMHD1 is a nucleic-acid binding protein that is mislocalized due to Aicardi-Goutières syndrome-associated mutations. Hum Mutat. 2012;33:1116–1122.
  • White TE, Brandariz-Nuñez A, Valle-Casuso JC, et al. Contribution of SAM and HD domains to retroviral restriction mediated by human SAMHD1. Virology. 2013;436:81–90.
  • Beloglazova N, Flick R, Tchigvintsev A, et al. Nuclease activity of the human SAMHD1 protein implicated in the Aicardi-Goutières syndrome and HIV-1 restriction. J Biol Chem. 2013;288:8101–8110.
  • Tüngler V, Staroske W, Kind B, et al. Single-stranded nucleic acids promote SAMHD1 complex formation. J Mol Med. 2013;91:759–770.
  • Seamon KJ, Bumpus NN, Stivers JT. Single-stranded nucleic acids bind to the tetramer interface of SAMHD1 and prevent formation of the catalytic homotetramer. Biochemistry. 2016;55:6087–6099.
  • Choi J, Ryoo J, Oh C, et al. SAMHD1 specifically restricts retroviruses through its RNase activity. Retrovirology. 2015;12:46.
  • Ryoo J, Choi J, Oh C, et al. The ribonuclease activity of SAMHD1 is required for HIV-1 restriction. Nat Med. 2014;20:936–941.
  • Ryoo J, Hwang SY, Choi J, et al. SAMHD1, the Aicardi-Goutières syndrome gene and retroviral restriction factor, is a phosphorolytic ribonuclease rather than a hydrolytic ribonuclease. Biochem Biophys Res Commun. 2016;477:977–981.
  • Antonucci JM, St. Gelais C, de Silva S, et al. SAMHD1-mediated HIV-1 restriction in cells does not involve ribonuclease activity. Nat Med. 2016;22:1072–1074.
  • Wittmann S, Behrendt R, Eissmann K, et al. Phosphorylation of murine SAMHD1 regulates its antiretroviral activity. Retrovirology. 2015;12:103.
  • Daddacha W, Koyen AE, Bastien AJ, et al. SAMHD1 promotes DNA end resection to facilitate DNA repair by homologous recombination. Cell Rep. 2017;20:1921–1935.
  • Liao W, Bao Z, Cheng C, et al. Dendritic cell-derived interferon-γ-induced protein mediates tumor necrosis factor-α stimulation of human lung fibroblasts. Proteomics. 2008;8:2640–2650.
  • Schmidt S, Schenkova K, Adam T, et al. SAMHD1’s protein expression profile in humans. J Leukoc Biol. 2015;98:5–14.
  • Franzolin E, Pontarin G, Rampazzo C, et al. The deoxynucleotide triphosphohydrolase SAMHD1 is a major regulator of DNA precursor pools in mammalian cells. Proc Natl Acad Sci. 2013;110:14272–14277.
  • Cribier A, Descours B, Valadão A, et al. Phosphorylation of SAMHD1 by cyclin A2/CDK1 regulates its restriction activity toward HIV-1. Cell Rep. 2013;3:1036–1043.
  • Yan J, Hao C, DeLucia M, et al. CyclinA2-cyclin-dependent kinase regulates SAMHD1 protein phosphohydrolase domain. J Biol Chem. 2015;290:13279–13292.
  • St. Gelais C, de Silva S, Hach JC, et al. Identification of cellular proteins interacting with the retroviral restriction factor SAMHD1. J Virol. 2014;88:7689–7689.
  • Baldauf HM, Pan X, Erikson E, et al. SAMHD1 restricts HIV-1 infection in resting CD4(+) T cells. Nat Med. 2012;18:1682–1689.
  • St. Gelais C, de Silva S, Amie SM, et al. SAMHD1 restricts HIV-1 infection in dendritic cells (DCs) by dNTP depletion, but its expression in DCs and primary CD4+ T-lymphocytes cannot be upregulated by interferons. Retrovirology. 2012;9:105.
  • Diamond TL, Roshal M, Jamburuthugoda VK, et al. Macrophage tropism of HIV-1 depends on efficient cellular dNTP utilization by reverse transcriptase. J Biol Chem. 2004;279:51545–51553.
  • De Silva S, Wang F, Hake TS, et al. Downregulation of SAMHD1 expression correlates with promoter DNA methylation in sézary syndrome patients. J Invest Dermatol. 2013;134:562–565.
  • De Silva S, Hoy H, Hake TS, et al. Promoter methylation regulates SAMHD1 gene expression in human CD4 + T cells. J Biol Chem. 2013;288:9284–9292.
  • Wang J, Lu F, Shen XY, et al. SAMHD1 is down regulated in lung cancer by methylation and inhibits tumor cell proliferation. Biochem Biophys Res Commun. 2014;455:229–233.
  • Welbourn S, Miyagi E, White TE, et al. Identification and characterization of naturally occurring splice variants of SAMHD1. Retrovirology. 2012;9:86.
  • Bloch N, Gläsker Sitaram S, Hofmann PH, et al. A highly active isoform of lentivirus restriction Factor SAMHD1 in mouse. J Biol Chem. 2016;292:1068–1080.
  • Pauls E, Jimenez E, Ruiz A, et al. Restriction of HIV-1 replication in primary macrophages by IL-12 and IL-18 through the upregulation of SAMHD1. J Immunol. 2013;190:4736–4741.
  • Yang S, Zhan Y, Zhou Y, et al. Interferon regulatory factor 3 is a key regulation factor for inducing the expression of SAMHD1 in antiviral innate immunity. Sci Rep. 2016;6:1–16.
  • Goujon C, Schaller T, Galão RP, et al. Evidence for IFNα-induced, SAMHD1-independent inhibitors of early HIV-1 infection. Retrovirology. 2013;10:23.
  • Riess M, Fuchs NV, Idica A, et al. Interferons induce expression of SAMHD1 in monocytes through down-regulation of miR-181a and miR-30a. J Biol Chem. 2017;292:264–277.
  • Kyei GB, Cheng X, Ramani R, et al. Cyclin L2 is a critical HIV dependency factor in macrophages that controls samhd1 abundance. Cell Host Microbe. 2015;17:98–106.
  • Morrissey C, Schwefel D, Ennis-Adeniran V, et al. The eukaryotic elongation factor eEF1A1 interacts with SAMHD1. Biochem J. 2015;466:69–76.
  • Rocha-Perugini V, Suárez H, Álvarez S, et al. CD81 association with SAMHD1 enhances HIV-1 reverse transcription by increasing dNTP levels. Nat Microbiol. 2017;2:1513–1522.
  • Welbourn S, Dutta SM, Semmes OJ, et al. Restriction of virus infection but not catalytic dNTPase activity is regulated by phosphorylation of SAMHD1. J Virol. 2013;87:11516–11524.
  • White TE, Brandariz-Nuñez A, Valle-Casuso JC, et al. The retroviral restriction ability of SAMHD1, but not its deoxynucleotide triphosphohydrolase activity, is regulated by phosphorylation. Cell Host Microbe. 2013;13:441–451.
  • St. Gelais C, Kim SH, Ding L, et al. A putative cyclin-binding motif in human SAMHD1 contributes to protein phosphorylation, localization, and stability. J Biol Chem. 2016;291:26332–26342.
  • Coiras M, Bermejo M, Descours B, et al. IL-7 induces SAMHD1 phosphorylation in CD4+ T lymphocytes, improving early steps of HIV-1 life cycle. Cell Rep. 2016;14:2100–2107.
  • Pauls E, Ruiz A, Badia R, et al. Cell cycle control and HIV-1 susceptibility are linked by CDK6-dependent CDK2 phosphorylation of SAMHD1 in myeloid and lymphoid cells. J Immunol. 2014;193:1988–1997.
  • Pauls E, Badia R, Torres-Torronteras J, et al. Palbociclib, a selective inhibitor of cyclin-dependent kinase4/6, blocks HIV-1 reverse transcription through the control of sterile alpha motif and HD domain-containing protein-1 (SAMHD1) activity. AIDS. 2014;28:2213–2222.
  • Mlcochova P, Sutherland KA, Watters SA, et al. A G1-like state allows HIV-1 to bypass SAMHD1 restriction in macrophages. EMBO J. 2017;36:604–616.
  • Mandal M, Bandyopadhyay D, Goepfert TM, et al. Interferon-induces expression of cyclin-dependent kinase-inhibitors p21WAF1 and p27Kip1 that prevent activation of cyclin-dependent kinase by CDK-activating kinase (CAK). Oncogene. 1998;16:217–225.
  • Pauls E, Ruiz A, Riveira-Muñoz E, et al. p21 regulates the HIV-1 restriction factor SAMHD1. Proc Natl Acad Sci USA. 2014;111:E1322–E1324.
  • Valle-Casuso JC, Allouch A, David A, et al. p21 restricts HIV-1 in monocyte-derived dendritic cells through the reduction of deoxynucleoside triphosphate biosynthesis and regulation of SAMHD1 antiviral activity. J Virol. 2017;91:e01324-17.
  • Mlcochova P, Caswell SJ, Taylor IA, et al. DNA damage induced by topoisomerase inhibitors activates SAMHD1 and blocks HIV-1 infection of macrophages. EMBO J. 2018;37:50–62.
  • Tang C, Ji X, Wu L, et al. Impaired dNTPase activity of SAMHD1 by phosphomimetic mutation of Thr-592. J Biol Chem. 2015;290:26352–26359.
  • Ruiz A, Pauls E, Badia R, et al. Cyclin D3-dependent control of the dNTP pool and HIV-1 replication in human macrophages. Cell Cycle. 2015;14:1657–1665.
  • Bhattacharya A, Wang Z, White T, et al. Effects of T592 phosphomimetic mutations on tetramer stability and dNTPase activity of SAMHD1 cannot explain the retroviral restriction defect. Sci Rep. 2016;6:31353.
  • Poole LB. The basics of thiols and cysteines in redox biology and chemistry. Free Radic Biol Med. 2015;80:148–157.
  • Go YM, Jones DP. The redox proteome. J Biol Chem. 2013;288:26512–26520.
  • Lee EJ, Seo JH, Park JH, et al. SAMHD1 acetylation enhances its deoxynucleotide triphosphohydrolase activity and promotes cancer cell proliferation. Oncotarget. 2017;8:68517–68529.
  • Bonifati S, Daly MB, St. Gelais C, et al. SAMHD1 controls cell cycle status, apoptosis and HIV-1 infection in monocytic THP-1 cells. Virology. 2016;495:92–100.
  • Stillman B. Deoxynucleoside triphosphate (dNTP) synthesis and destruction regulate the replication of both cell and virus genomes. Proc Natl Acad Sci USA. 2013;110:14120–14121.
  • Kim B, Nguyen LA, Daddacha W, et al. Tight interplay among SAMHD1 protein level, cellular dNTP levels, and HIV-1 proviral DNA synthesis kinetics in human primary monocyte-derived macrophages. J Biol Chem. 2012;287:21570–21574.
  • Hollenbaugh JA, Tao S, Lenzi GM, et al. dNTP pool modulation dynamics by SAMHD1 protein in monocyte-derived macrophages. Retrovirology. 2014;11:63.
  • Kretschmer S, Wolf C, Konig N, et al. SAMHD1 prevents autoimmunity by maintaining genome stability. Ann Rheum Dis. 2015;74:e17–e17.
  • Dragin L, Munir-Matloob S, Froehlich J, et al. Evidence that HIV-1 restriction factor SAMHD1 facilitates differentiation of myeloid THP-1 cells. Virol J. 2015;12:201.
  • Kodigepalli KM, Li M, Liu SL, et al. Exogenous expression of SAMHD1 inhibits proliferation and induces apoptosis in cutaneous T-cell lymphoma-derived HuT78 cells. Cell Cycle. 2017;16:179–188.
  • Antonucci JM, St. Gelais C, Wu L. The dynamic interplay between HIV-1, SAMHD1, and the innate antiviral response. Front Immunol. 2017;8:1–9.
  • Srivastava S, Swanson SK, Manel N, et al. Lentiviral Vpx accessory factor targets VprBP/DCAF1 substrate adaptor for cullin 4 E3 ubiquitin ligase to enable macrophage infection. PLoS Pathog. 2008;4:e1000059.
  • Pertel T, Reinhard C, Luban J. Vpx rescues HIV-1 transduction of dendritic cells from the antiviral state established by type 1 interferon. Retrovirology. 2011;8:49.
  • Amie SM, Noble E, Kim B. Intracellular nucleotide levels and the control of retroviral infections. Virology. 2013;436:247–254.
  • Kennedy EM, Gavegnano C, Nguyen L, et al. Ribonucleoside triphosphates as substrate of human immunodeficiency virus type 1 reverse transcriptase in human macrophages. J Biol Chem. 2010;285:39380–39391.
  • Descours B, Cribier A, Chable-Bessia C, et al. SAMHD1 restricts HIV-1 reverse transcription in quiescent CD4(+) T-cells. Retrovirology. 2012;9:87.
  • Hrecka K, Hao C, Gierszewska M, et al. Vpx relieves inhibition of HIV-1 infection of macrophages mediated by the SAMHD1 protein. Nature. 2011;474:658–661.
  • Laguette N, Sobhian B, Casartelli N, et al. SAMHD1 is the dendritic- and myeloid-cell-specific HIV-1 restriction factor counteracted by Vpx. Nature. 2011;474:654–657.
  • Berger A, Sommer AFR, Zwarg J, et al. SAMHD1-deficient CD14+ cells from individuals with Aicardi-Goutières syndrome are highly susceptible to HIV-1 infection. PLoS Pathog. 2011;7:1–12.
  • Sunseri N, O’Brien M, Bhardwaj N, et al. Human immunodeficiency virus type 1 modified to package Simian immunodeficiency virus Vpx efficiently infects macrophages and dendritic cells. J Virol. 2011;85:6263–6274.
  • Lahouassa H, Daddacha W, Hofmann H, et al. SAMHD1 restricts the replication of human immunodeficiency virus type 1 by depleting the intracellular pool of deoxynucleoside triphosphates. Nat Immunol. 2012;13:223–228.
  • Berger G, Turpin J, Cordeil S, et al. Functional analysis of the relationship between Vpx and the restriction factor SAMHD1. J Biol Chem. 2012;287:41210–41217.
  • Schwefel D, Groom HCT, Boucherit VC, et al. Structural basis of lentiviral subversion of a cellular protein degradation pathway. Nature. 2014;505:234–238.
  • Ahn J, Hao C, Yan J, et al. HIV/simian immunodeficiency virus (SIV) accessory virulence factor Vpx loads the host cell restriction factor SAMHD1 onto the E3 ubiquitin ligase complex CRL4 DCAF1. J Biol Chem. 2012;287:12550–12558.
  • Ayinde D, Bruel T, Cardinaud S, et al. SAMHD1 limits HIV-1 antigen presentation by monocyte-derived dendritic cells. J Virol. 2015;89:6994–7006.
  • Brégnard C, Benkirane M, Laguette N. DNA damage repair machinery and HIV escape from innate immune sensing. Front Microbiol. 2014;5:1–10.
  • Manel N, Hogstad B, Wang Y, et al. A cryptic sensor for HIV-1 activates antiviral innate immunity in dendritic cells. Nature. 2010;467:214–217.
  • Maelfait J, Bridgeman A, Benlahrech A, et al. Restriction by SAMHD1 limits cGAS/STING-dependent innate and adaptive immune responses to HIV-1. Cell Rep. 2016;16:1492–1501.
  • van Montfoort N, Olagnier D, Hiscott J. Unmasking immune sensing of retroviruses: interplay between innate sensors and host effectors. Cytokine Growth Factor Rev. 2014;25:657–668.
  • Kim ET, White TE, Brandariz-Núñez A, et al. SAMHD1 restricts herpes simplex virus 1 in macrophages by limiting DNA replication. J Virol. 2013;87:12949–12956.
  • Hollenbaugh JA, Gee P, Baker J, et al. Host factor SAMHD1 restricts DNA viruses in non-dividing myeloid cells. PLoS Pathog. 2013;9:e1003481.
  • Behrendt R, Schumann T, Gerbaulet A, et al. Mouse SAMHD1 has antiretroviral activity and suppresses a spontaneous cell-intrinsic antiviral response. Cell Rep. 2013;4:689–696.
  • Sze A, Belgnaoui SM, Olagnier D, et al. Host restriction factor SAMHD1 limits human T cell leukemia virus type 1 infection of monocytes via STING-mediated apoptosis. Cell Host Microbe. 2013;14:422–434.
  • Sommer AFR, Rivière L, Qu B, et al. Restrictive influence of SAMHD1 on Hepatitis B virus life cycle. Sci Rep. 2016;6:26616.
  • Chen Z, Zhu M, Pan X, et al. Inhibition of Hepatitis B virus replication by SAMHD1. Biochem Biophys Res Commun. 2014;450:1462–1468.
  • Gramberg T, Kahle T, Bloch N, et al. Restriction of diverse retroviruses by SAMHD1. Retrovirology. 2013;10:26.
  • Huber HE, Beauchamp BB, Richardson CC. Escherichia coli dGTP triphosphohydrolase is inhibited by gene 1.2 protein of bacteriophage T7. J Biol Chem. 1988;263:13549–13556.
  • Pfister SX, Markkanen E, Jiang Y, et al. Inhibiting WEE1 selectively kills histone H3K36me3-deficient cancers by dNTP starvation. Cancer Cell. 2015;28:557–568.
  • Aye Y, Li M, Long MJC, et al. Ribonucleotide reductase and cancer: biological mechanisms and targeted therapies. Oncogene. 2015;34:2011–2021.
  • Welbourn S, Strebel K. Low dNTP levels are necessary but may not be sufficient for lentiviral restriction by SAMHD1. Virology. 2016;488:271–277.
  • Brandariz-Nuñez A, Valle-Casuso JC, White TE, et al. Contribution of oligomerization to the anti-HIV-1 properties of SAMHD1. Retrovirology. 2013;10:131.
  • Crow YJ, Rehwinkel J. Aicardi-goutieres syndrome and related phenotypes: linking nucleic acid metabolism with autoimmunity. Hum Mol Genet. 2009;18:R130–R136.
  • Crow YJ, Manel N. Aicardi-Goutières syndrome and the type I interferonopathies. Nat Rev Immunol. 2015;15:429–440.
  • Plander M, Kalman B. Rare autoimmune disorders with Mendelian inheritance. Autoimmunity. 2016;49:285–297.
  • Goutières F, Aicardi J, Barth PG, et al. Aicardi-Goutières syndrome: an update and results of interferon-alpha studies. Ann Neurol. 1998;44:900–907.
  • Ravenscroft JC, Suri M, Rice GI, et al. Autosomal dominant inheritance of a heterozygous mutation in SAMHD1 causing familial chilblain lupus. Am J Med Genet. 2011;155:235–237.
  • Crow MK, Kirou KA, Wohlgemuth J. Microarray analysis of interferon-regulated genes in SLE. Autoimmunity. 2003;36:481–490.
  • Crow YJ, Chase DS, Schmidt J, et al. Characterization of human disease phenotypes associated with mutations in TREX1, RNASEH2A, RNASEH2B, RNASEH2C, SAMHD1, ADAR, and IFIH1. Am J Med Genet Part A. 2015;167A:296–312.
  • Rice GI, Bond J, Asipu A, et al. Mutations involved in Aicardi-Goutières syndrome implicate SAMHD1 as regulator of the innate immune response. Nat Genet. 2009;41:829–832.
  • Oda H, Nakagawa K, Abe J, et al. Aicardi-Goutières syndrome is caused by IFIH1 mutations. Am J Hum Genet. 2014;95:121–125.
  • Crow YJ, Leitch A, Hayward BE, et al. Mutations in genes encoding ribonuclease H2 subunits cause Aicardi-Goutières syndrome and mimic congenital viral brain infection. Nat Genet. 2006;38:910–916.
  • Rice G, Newman WG, Dean J, et al. Heterozygous mutations in TREX1 cause familial chilblain lupus and dominant Aicardi-Goutieres syndrome. Am J Hum Genet. 2007;80:811–815.
  • Crow YJ, Hayward BE, Parmar R, et al. Mutations in the gene encoding the 3′-5′ DNA exonuclease TREX1 cause Aicardi-Goutières syndrome at the AGS1 locus. Nat Genet. 2006;38:917–920.
  • Hu S, Li J, Xu F, et al. SAMHD1 inhibits LINE-1 retrotransposition by promoting stress granule formation. PLoS Genet. 2015;11:e1005367.
  • Zhao K, Du J, Han X, et al. Modulation of LINE-1 and Alu/SVA retrotransposition by Aicardi-Goutières syndrome-related SAMHD1. Cell Rep. 2013;4:1108–1115.
  • Crow MK. Long interspersed nuclear elements (LINE-1): potential triggers of systemic autoimmune disease. Autoimmunity. 2010;43:7–16.
  • McKinnon PJ. Maintaining genome stability in the nervous system. Nat Neurosci. 2013;16:1523–1529.
  • Xin B, Jones S, Puffenberger EG, et al. Homozygous mutation in SAMHD1 gene causes cerebral vasculopathy and early onset stroke. Proc Natl Acad Sci USA. 2011;108:5372–5377.
  • Ramesh V, Bernardi B, Stafa A, et al. Intracerebral large artery disease in Aicardi-Goutières syndrome implicates SAMHD1 in vascular homeostasis. Dev Med Child Neurol. 2010;52:725–732.
  • Thiele H, Du Moulin M, Barczyk K, et al. Cerebral arterial stenoses and stroke: novel features of Aicardi-Goutières syndrome caused by the Arg164X mutation in SAMHD1 are associated with altered cytokine expression. Hum Mutat. 2010;31:1836–1850.
  • Leshinsky-Silver E, Malinger G, Ben-Sira L, et al. A large homozygous deletion in the SAMHD1 gene causes atypical Aicardi-Goutiéres syndrome associated with mtDNA deletions. Eur J Hum Genet. 2011;19:287–292.
  • Rehwinkel J, Maelfait J, Bridgeman A, et al. SAMHD1-dependent retroviral control and escape in mice. EMBO J. 2013;32:2454–2462.
  • Kasher PR, Jenkinson EM, Briolat V, et al. Characterization of samhd1 morphant zebrafish recapitulates features of the human type I interferonopathy Aicardi-Goutières syndrome. J Immunol. 2015;194:2819–2825.
  • Kunz BA. Mutagenesis and deoxyribonucleotide pool imbalance. Mutat Res. 1988;200:133–147.
  • Weinberg G, Ullman B, Martin DW. Mutator phenotypes in mammalian cell mutants with distinct biochemical defects and abnormal deoxyribonucleoside triphosphate pools. Proc Natl Acad Sci USA. 1981;78:2447–2451.
  • Sohl CD, Ray S, Sweasy JB. Pools and pols: mechanism of a mutator phenotype. Proc Natl Acad Sci USA. 2015;112:201505169.
  • Williams LN, Marjavaara L, Knowels GM, et al. dNTP pool levels modulate mutator phenotypes of error-prone DNA polymerase ε variants. Proc Natl Acad Sci USA. 2015;112:E2457–E2466.
  • Mertz TM, Sharma S, Chabes A, et al. Colon cancer-associated mutator DNA polymerase δ variant causes expansion of dNTP pools increasing its own infidelity. Proc Natl Acad Sci USA. 2015;112:E2467–E2476.
  • Bester AC, Roniger M, Oren YS, et al. Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell. 2011;145:435–446.
  • Pajalunga D, Franzolin Stevanoni E, Zribi M, et al. A defective dNTP pool hinders DNA replication in cell cycle-reactivated terminally differentiated muscle cells. Cell Death Differ. 2017;24:774–784.
  • Rentoft M, Lindell K, Tran P, et al. Heterozygous colon cancer-associated mutations of SAMHD1 have functional significance. Proc Natl Acad Sci USA. 2016;113:4723–4728.
  • Clifford R, Louis T, Robbe P, et al. SAMHD1 is mutated recurrently in chronic lymphocytic leukemia and is involved in response to DNA damage. Blood. 2014;123:1021–1031.
  • Johansson P, Klein-Hitpass L, Bergmann AK, et al. SAMHD1 Is frequently involved in T-cell prolymphocytic leukemia (T-PLL) pathogenesis. Hematol Oncol. 2017;35:164–164.
  • Forbes SA, Beare D, Boutselakis H, et al. COSMIC: somatic cancer genetics at high-resolution. Nucleic Acids Res. 2017;45:D777–D783.
  • Kohnken R, Kodigepalli KM, Wu L. Regulation of deoxynucleotide metabolism in cancer: novel mechanisms and therapeutic implications. Mol Cancer. 2015;14:176.
  • Mathews CK. Oxidized deoxyribonucleotides, mutagenesis, and cancer. FASEB J. 2017;31:11–13.
  • Rudd SG, Valerie NCK, Helleday T. Pathways controlling dNTP pools to maintain genome stability. DNA Repair (Amst). 2016;44:193–204.
  • Parker WB. Enzymology of purine and pyrimidine antimetabolites used in the treatment of cancer. Chem Rev. 2009;109:2880–2893.
  • Ewald B, Sampath D, Plunkett W. Nucleoside analogs: molecular mechanisms signaling cell death. Oncogene. 2008;27:6522–6537.
  • Herold N, Rudd SG, Sanjiv K, et al. SAMHD1 protects cancer cells from various nucleoside-based antimetabolites. Cell Cycle. 2017;4101:1–10.
  • Schneider C, Oellerich T, Baldauf HM, et al. SAMHD1 is a biomarker for cytarabine response and a therapeutic target in acute myeloid leukemia. Nat Med. 2016;23:250–255.
  • Herold N, Rudd SG, Ljungblad L, et al. Targeting SAMHD1 with the Vpx protein to improve cytarabine therapy for hematological malignancies. Nat Med. 2017;23:256–263.
  • Herold N, Rudd SG, Sanjiv K, et al. With me or against me: tumor suppressor and drug resistance activities of SAMHD1. Exp Hematol. 2017;52:32–39.
  • Ballana E, Badia R, Terradas G, et al. SAMHD1 specifically affects the antiviral potency of thymidine analog HIV reverse transcriptase inhibitors. Antimicrob Agents Chemother. 2014;58:4804–4813.
  • Huber AD, Michailidis E, Schultz ML, et al. SAMHD1 has differential impact on the efficacies of HIV nucleoside reverse transcriptase inhibitors. Antimicrob Agents Chemother. 2014;58:4915–4919.
  • Arnold LH, Kunzelmann S, Webb MR, et al. A continuous enzyme-coupled assay for triphosphohydrolase activity of HIV-1 restriction factor SAMHD1. Antimicrob Agents Chemother. 2015;59:186–192.
  • Seamon KJ, Stivers JT. A high-throughput enzyme-coupled assay for SAMHD1 dNTPase. J Biomol Screen. 2015;20:801–809.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.