We use cookies to improve your experience. By continuing to browse this site, you accept our cookie policy.×
Skip main navigation
Aging Health
Bioelectronics in Medicine
Biomarkers in Medicine
Breast Cancer Management
CNS Oncology
Colorectal Cancer
Concussion
Epigenomics
Future Cardiology
Future Medicine AI
Future Microbiology
Future Neurology
Future Oncology
Future Rare Diseases
Future Virology
Hepatic Oncology
HIV Therapy
Immunotherapy
International Journal of Endocrine Oncology
International Journal of Hematologic Oncology
Journal of 3D Printing in Medicine
Lung Cancer Management
Melanoma Management
Nanomedicine
Neurodegenerative Disease Management
Pain Management
Pediatric Health
Personalized Medicine
Pharmacogenomics
Regenerative Medicine

HSV-I and the cellular DNA damage response

    Samantha Smith

    Department of Molecular Biology & Biophysics, University of Connecticut Health Center, Farmington, CT 06030, USA

    &
    Sandra K Weller

    *Author for correspondence:

    E-mail Address: weller@uchc.edu

    Department of Molecular Biology & Biophysics, University of Connecticut Health Center, Farmington, CT 06030, USA

    Published Online:https://doi.org/10.2217/fvl.15.18

    ABSTRACT 

    Peter Wildy first observed genetic recombination between strains of HSV in 1955. At the time, knowledge of DNA repair mechanisms was limited, and it has only been in the last decade that particular DNA damage response (DDR) pathways have been examined in the context of viral infections. One of the first reports addressing the interaction between a cellular DDR protein and HSV-1 was the observation by Lees-Miller et al. that DNA-dependent protein kinase catalytic subunit levels were depleted in an ICP0-dependent manner during HSV-1 infection. Since then, there have been numerous reports describing the interactions between HSV infection and cellular DDR pathways. Due to space limitations, this review will focus predominantly on the most recent observations regarding how HSV navigates a potentially hostile environment to replicate its genome.

    Papers of special note have been highlighted as: • of interest; •• of considerable interest

    References

    • 1 Wildy P. Recombination with herpes simplex virus. J. Gen. Microbiol. 13(2), 346–360 (1955).
    • 2 Lees-Miller SP, Long MC, Kilvert MA, Lam V, Rice SA, Spencer CA. Attenuation of DNA-dependent protein kinase activity and its catalytic subunit by the herpes simplex virus type 1 transactivator ICP0. J. Virol. 70(11), 7471–7477 (1996).
    • 3 Wilkinson DE, Weller SK. The role of DNA recombination in herpes simplex virus DNA replication. IUBMB Life 55(8), 451–458 (2003).
    • 4 Everett RD. Interactions between DNA viruses, ND10 and the DNA damage response. Cell. Microbiol. 8(3), 365–374 (2006).
    • 5 Lilley CE, Schwartz RA, Weitzman MD. Using or abusing: viruses and the cellular DNA damage response. Trends Microbiol. 15(3), 119–126 (2007).
    • 6 Weitzman MD, Lilley CE, Chaurushiya MS. Genomes in conflict: maintaining genome integrity during virus infection. Annu. Rev. Microbiol. 64, 61–81 (2010).
    • 7 Weller SK. Herpes simplex virus reorganizes the cellular DNA repair and protein quality control machinery. PLoS Pathog. 6(11), e1001105 (2010).
    • 8 Everett RD. DNA viruses and viral proteins that interact with PML nuclear bodies. Oncogene 20(49), 7266–7273 (2001).
    • 9 Everett RD, Meredith M, Orr A. The ability of herpes simplex virus type 1 immediate-early protein Vmw110 to bind to a ubiquitin-specific protease contributes to its roles in the activation of gene expression and stimulation of virus replication. J. Virol. 73(1), 417–426 (1999).
    • 10 Blümel J, Matz B. Thermosensitive UL9 gene function is required for early stages of herpes simplex virus type 1 DNA synthesis. J. Gen. Virol. 76(Pt 12), 3119–3124 (1995).
    • 11 Schildgen O, Graper S, Blumel J, Matz B. Genome replication and progeny virion production of herpes simplex virus type 1 mutants with temperature-sensitive lesions in the origin-binding protein. J. Virol. 79(11), 7273–7278 (2005).
    • 12 Weller SK, Coen DM. Herpes simplex viruses: mechanisms of DNA replication. Cold Spring Harb. Perspect. Biol. 4(9), a013011 (2012).• Reviews current models for HSV-1 replication.
    • 13 Delius H, Clements JB. A partial denaturation map of herpes simplex virus type 1 DNA: evidence for inversions of the unique DNA regions. J. Gen. Virol. 33(1), 125–133 (1976).
    • 14 Hayward GS, Jacob RJ, Wadsworth SC, Roizman B. Anatomy of herpes simplex virus DNA: evidence for four populations of molecules that differ in the relative orientations of their long and short components. Proc. Natl Acad. Sci. USA 72(11), 4243–4247 (1975).
    • 15 Zhang X, Efstathiou S, Simmons A. Identification of novel herpes simplex virus replicative intermediates by field inversion gel electrophoresis: implications for viral DNA amplification strategies. Virology 202(2), 530–539 (1994).
    • 16 Brown SM, Ritchie DA, Subak-Sharpe JH. Genetic studies with herpes simplex virus type 1. The isolation of temperature-sensitive mutants, their arrangement into complementation groups and recombination analysis leading to a linkage map. J. Gen. Virol. 18(3), 329–346 (1973).
    • 17 Sheldrick P, Berthelot N. Inverted repetitions in the chromosome of herpes simplex virus. Cold Spring Harb. Symp. Quant. Biol. 39(Pt 2), 667–678 (1975).
    • 18 Dutch RE, Bianchi V, Lehman IR. Herpes simplex virus type 1 DNA replication is specifically required for high-frequency homologous recombination between repeated sequences. J. Virol. 69(5), 3084–3089 (1995).
    • 19 Fu X, Wang H, Zhang X. High-frequency intermolecular homologous recombination during herpes simplex virus-mediated plasmid DNA replication. J. Virol. 76(12), 5866–5874 (2002).
    • 20 Jackson SA, DeLuca NA. Relationship of herpes simplex virus genome configuration to productive and persistent infections. Proc. Natl Acad. Sci. USA 100(13), 7871–7876 (2003).• Demonstrates that circularization of the HSV-1 genome does not occur in cells infected with wild-type virus; however, circularization can be detected in cells infected with an ICP0 mutant virus. This result suggests that ICP0 inhibits circularization of the HSV-1 genome.
    • 21 Strang BL, Stow ND. Circularization of the herpes simplex virus type 1 genome upon lytic infection. J. Virol. 79(19), 12487–12494 (2005).
    • 22 Falkenberg M, Lehman IR, Elias P. Leading and lagging strand DNA synthesis in vitro by a reconstituted herpes simplex virus type 1 replisome. Proc. Natl Acad. Sci. USA 97(8), 3896–3900 (2000).
    • 23 Graves-Woodward KL, Gottlieb J, Challberg MD, Weller SK. Biochemical analyses of mutations in the HSV-1 helicase-primase that alter ATP hydrolysis, DNA unwinding, and coupling between hydrolysis and unwinding. J. Biol. Chem. 272(7), 4623–4630 (1997).
    • 24 Skaliter R, Lehman IR. Rolling circle DNA replication in vitro by a complex of herpes simplex virus type 1-encoded enzymes. Proc. Natl Acad. Sci. USA 91(22), 10665–10669 (1994).
    • 25 Skaliter R, Makhov AM, Griffith JD, Lehman IR. Rolling circle DNA replication by extracts of herpes simplex virus type 1-infected human cells. J. Virol. 70(2), 1132–1136 (1996).
    • 26 Lamberti C, Weller SK. The herpes simplex virus type 1 UL6 protein is essential for cleavage and packaging but not for genomic inversion. Virology 226(2), 403–407 (1996).
    • 27 Severini A, Scraba DG, Tyrrell DL. Branched structures in the intracellular DNA of herpes simplex virus type 1. J. Virol. 70(5), 3169–3175 (1996).
    • 28 Jacob RJ, Roizman B. Anatomy of herpes simplex virus DNA VIII. Properties of the replicating DNA. J. Virol. 23(2), 394–411 (1977).
    • 29 Weller SK, Sawitzke JA. Recombination promoted by DNA viruses: phage lambda to herpes simplex virus. Annu. Rev. Microbiol. 68, 237–258 (2014).•• Explores recombination-dependent replication employed by HSV-1 and the phage λ.
    • 30 Blumel J, Graper S, Matz B. Structure of simian virus 40 DNA replicated by herpes simplex virus type 1. Virology 276(2), 445–454 (2000).
    • 31 Matz B. Herpes simplex virus infection generates large tandemly reiterated simian virus 40 DNA molecules in a transformed hamster cell line. J. Virol. 61(5), 1427–1434 (1987).
    • 32 Nicolas A, Alazard-Dany N, Biollay C et al. Identification of rep-associated factors in herpes simplex virus type 1-induced adeno-associated virus type 2 replication compartments. J. Virol. 84(17), 8871–8887 (2010).
    • 33 Stahl MM, Thomason L, Poteete AR, Tarkowski T, Kuzminov A, Stahl FW. Annealing vs. invasion in phage lambda recombination. Genetics 147(3), 961–977 (1997).
    • 34 Reuven NB, Willcox S, Griffith JD, Weller SK. Catalysis of strand exchange by the HSV-1 UL12 and ICP8 proteins: potent ICP8 recombinase activity is revealed upon resection of dsDNA substrate by nuclease. J. Mol. Biol. 342(1), 57–71 (2004).• Demonstrates that UL12 and ICP8 form a two-subunit recombinase that can perform strand exchange in vitro.
    • 35 Goldstein JN, Weller SK. The exonuclease activity of HSV-1 UL12 is required for in vivo function. Virology 244(2), 442–457 (1998).
    • 36 Schumacher AJ, Mohni KN, Kan Y, Hendrickson EA, Stark JM, Weller SK. The HSV-1 exonuclease, UL12, stimulates recombination by a single strand annealing mechanism. PLoS Pathog. 8(8), e1002862 (2012).•• Demonstrates that HSV-1 stimulates single-strand annealing and inhibits homologous recombination and nonhomologous end joining.
    • 37 Lilley CE, Carson CT, Muotri AR, Gage FH, Weitzman MD. DNA repair proteins affect the lifecycle of herpes simplex virus 1. Proc. Natl Acad. Sci. USA 102(16), 5844–5849 (2005).
    • 38 Shirata N, Kudoh A, Daikoku T et al. Activation of ataxia telangiectasia-mutated DNA damage checkpoint signal transduction elicited by herpes simplex virus infection. J. Biol. Chem. 280(34), 30336–30341 (2005).
    • 39 Taylor TJ, Knipe DM. Proteomics of herpes simplex virus replication compartments: association of cellular DNA replication, repair, recombination, and chromatin remodeling proteins with ICP8. J. Virol. 78(11), 5856–5866 (2004).
    • 40 Wilkinson DE, Weller SK. Recruitment of cellular recombination and repair proteins to sites of herpes simplex virus type 1 DNA replication is dependent on the composition of viral proteins within prereplicative sites and correlates with the induction of the DNA damage response. J. Virol. 78(9), 4783–4796 (2004).
    • 41 Gregory DA, Bachenheimer SL. Characterization of mre11 loss following HSV-1 infection. Virology 373(1), 124–136 (2008).
    • 42 Balasubramanian N, Bai P, Buchek G, Korza G, Weller SK. Physical interaction between the herpes simplex virus type 1 exonuclease, UL12, and the DNA double-strand break-sensing MRN complex. J. Virol. 84(24), 12504–12514 (2010).
    • 43 Mohni KN, Mastrocola AS, Bai P, Weller SK, Heinen CD. DNA mismatch repair proteins are required for efficient herpes simplex virus 1 replication. J. Virol. 85(23), 12241–12253 (2011).
    • 44 Karttunen H, Savas JN, McKinney C et al. Co-opting the Fanconi anemia genomic stability pathway enables herpesvirus DNA synthesis and productive growth. Mol. Cell 55(1), 111–122 (2014).•• Demonstrates the novel discovery that HSV-1 interacts with and utilizes components of the Fanconi anemia pathway for efficient replication.
    • 45 Dueva R, Iliakis G. Alternative pathways of non-homologous end joining (NHEJ) in genomic instability and cancer. Transl. Cancer Res. 2, 163–177 (2013).
    • 46 Neal JA, Meek K. Choosing the right path: does DNA-PK help make the decision? Mutat. Res. 711(1–2), 73–86 (2011).
    • 47 Ahnesorg P, Smith P, Jackson SP. XLF interacts with the XRCC4-DNA ligase IV complex to promote DNA nonhomologous end-joining. Cell 124(2), 301–313 (2006).
    • 48 Smith S, Reuven N, Mohni KN, Schumacher AJ, Weller SK. Structure of the herpes simplex virus 1 genome: manipulation of nicks and gaps can abrogate infectivity and alter the cellular DNA damage response. J. Virol. 88(17), 10146–10156 (2014).•• Shows that the viral genome can activate DNA-dependent protein kinase catalytic subunit in the absence of viral proteins.
    • 49 Parkinson J, Lees-Miller SP, Everett RD. Herpes simplex virus type 1 immediate-early protein vmw110 induces the proteasome-dependent degradation of the catalytic subunit of DNA-dependent protein kinase. J. Virol. 73(1), 650–657 (1999).
    • 50 Rogakou EP, Pilch DR, Orr AH, Ivanova VS, Bonner WM. DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J. Biol. Chem. 273(10), 5858–5868 (1998).
    • 51 Harper JW, Elledge SJ. The DNA damage response: ten years after. Mol. Cell 28(5), 739–745 (2007).
    • 52 Fernandez-Capetillo O, Celeste A, Nussenzweig A. Focusing on foci: H2AX and the recruitment of DNA-damage response factors. Cell Cycle 2(5), 426–427 (2003).
    • 53 Jackson SP, Durocher D. Regulation of DNA damage responses by ubiquitin and SUMO. Mol. Cell 49(5), 795–807 (2013).
    • 54 Huen MS, Grant R, Manke I et al. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell 131(5), 901–914 (2007).
    • 55 Mailand N, Bekker-Jensen S, Faustrup H et al. RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins. Cell 131(5), 887–900 (2007).
    • 56 Stewart GS, Panier S, Townsend K et al. The RIDDLE syndrome protein mediates a ubiquitin-dependent signaling cascade at sites of DNA damage. 136(3), 420–434 (2009).
    • 57 Williams RS, Williams JS, Tainer JA. Mre11-Rad50-Nbs1 is a keystone complex connecting DNA repair machinery, double-strand break signaling, and the chromatin template. Biochem. Cell Biol. 85(4), 509–520 (2007).
    • 58 Bolderson E, Tomimatsu N, Richard DJ et al. Phosphorylation of Exo1 modulates homologous recombination repair of DNA double-strand breaks. Nucl. Acids Res. 38(6), 1821–1831 (2010).
    • 59 Lilley CE, Chaurushiya MS, Boutell C et al. A viral E3 ligase targets RNF8 and RNF168 to control histone ubiquitination and DNA damage responses. EMBO J. 29(5), 943–955 (2010).• Identifies RNF8 and RNF168 as targets for degradation by ICP0 during HSV-1 infection.
    • 60 Jazayeri A, Falck J, Lukas C et al. ATM- and cell cycle-dependent regulation of ATR in response to DNA double-strand breaks. Nat. Cell Biol. 8(1), 37–45 (2006).
    • 61 Sartori AA, Lukas C, Coates J et al. Human CtIP promotes DNA end resection. Nature 450(7169), 509–514 (2007).
    • 62 Mohni KN, Livingston CM, Cortez D, Weller SK. ATR and ATRIP are recruited to herpes simplex virus type 1 replication compartments even though ATR signaling is disabled. J. Virol. 84(23), 12152–12164 (2010).
    • 63 Gordin M, Olshevsky U, Rosenkranz HS, Becker Y. Studies on herpes simplex virus DNA: denaturation properties. Virology 55(1), 280–284 (1973).
    • 64 Peasland A, Wang LZ, Rowling E et al. Identification and evaluation of a potent novel ATR inhibitor, NU6027, in breast and ovarian cancer cell lines. Br. J. Cancer 105(3), 372–381 (2011).
    • 65 Mohni KN, Dee AR, Smith S, Schumacher AJ, Weller SK. Efficient herpes simplex virus 1 replication requires cellular ATR pathway proteins. J. Virol. 87(1), 531–542 (2013).•• Identifies certain components of the ATR pathway that contribute to efficient HSV-1 replication. Also identifies the mechanism by which HSV-1 inhibits ATR-CHK1 signaling during infection.
    • 66 Shigechi T, Tomida J, Sato K et al. ATR-ATRIP kinase complex triggers activation of the Fanconi anemia DNA repair pathway. Cancer Res. 72(5), 1149–1156 (2012).
    • 67 Tomida J, Itaya A, Shigechi T et al. A novel interplay between the Fanconi anemia core complex and ATR-ATRIP kinase during DNA cross-link repair. Nucl. Acids Res. 41(14), 6930–6941 (2013).
    • 68 Nakanishi K, Yang YG, Pierce AJ et al. Human Fanconi anemia monoubiquitination pathway promotes homologous DNA repair. Proc. Natl Acad. Sci. USA 102(4), 1110–1115 (2005).
    • 69 Sale JE, Lehmann AR, Woodgate R. Y-family DNA polymerases and their role in tolerance of cellular DNA damage. Nat. Rev. Mol. Cell Biol. 13(3), 141–152 (2012).
    • 70 Zhi G, Wilson JB, Chen X et al. Fanconi anemia complementation group FANCD2 protein serine 331 phosphorylation is important for fanconi anemia pathway function and BRCA2 interaction. Cancer Res. 69(22), 8775–8783 (2009).
    • 71 Peng M, Xie J, Ucher A, Stavnezer J, Cantor SB. Crosstalk between BRCA-Fanconi anemia and mismatch repair pathways prevents MSH2-dependent aberrant DNA damage responses. EMBO J. 33(15), 1698–1712 (2014).
    • 72 Adamo A, Collis SJ, Adelman CA et al. Preventing nonhomologous end joining suppresses DNA repair defects of Fanconi anemia. Mol. Cell 39(1), 25–35 (2010).
    • 73 Andreassen PR, D'Andrea AD, Taniguchi T. ATR couples FANCD2 monoubiquitination to the DNA-damage response. Genes Dev. 18(16), 1958–1963 (2004).
    • 74 Ho GP, Margossian S, Taniguchi T, D'Andrea AD. Phosphorylation of FANCD2 on two novel sites is required for mitomycin C resistance. Mol. Cell Biol. 26(18), 7005–7015 (2006).
    • 75 Singleton MR, Wentzell LM, Liu Y, West SC, Wigley DB. Structure of the single-strand annealing domain of human RAD52 protein. Proc. Natl Acad. Sci. USA 99(21), 13492–13497 (2002).
    • 76 Stark JM, Pierce AJ, Oh J, Pastink A, Jasin M. Genetic steps of mammalian homologous repair with distinct mutagenic consequences. Mol. Cell Biol. 24(21), 9305–9316 (2004).
    • 77 Iyama T, Wilson DM 3rd. DNA repair mechanisms in dividing and non-dividing cells. DNA Repair (Amsterdam) 12(8), 620–636 (2013).
    • 78 Bennardo N, Cheng A, Huang N, Stark JM. Alternative-NHEJ is a mechanistically distinct pathway of mammalian chromosome break repair. PLoS Genet. 4(6), e1000110 (2008).
    • 79 Wang M, Wu W, Wu W et al. PARP-1 and Ku compete for repair of DNA double strand breaks by distinct NHEJ pathways. Nucl. Acids Res. 34(21), 6170–6182 (2006).
    • 80 Audebert M, Salles B, Calsou P. Involvement of poly(ADP-ribose) polymerase-1 and XRCC1/DNA ligase III in an alternative route for DNA double-strand breaks rejoining. J. Biol. Chem. 279(53), 55117–55126 (2004).
    • 81 Cheng Q, Barboule N, Frit P et al. Ku counteracts mobilization of PARP1 and MRN in chromatin damaged with DNA double-strand breaks. Nucl. Acids Res. 39(22), 9605–9619 (2011).
    • 82 Rass E, Grabarz A, Plo I, Gautier J, Bertrand P, Lopez BS. Role of Mre11 in chromosomal nonhomologous end joining in mammalian cells. Nat. Struct. Mol. Biol. 16(8), 819–824 (2009).
    • 83 Xie A, Kwok A, Scully R. Role of mammalian Mre11 in classical and alternative nonhomologous end joining. Nat. Struct. Mol. Biol. 16(8), 814–818 (2009).
    • 84 Zhuang J, Jiang G, Willers H, Xia F. Exonuclease function of human Mre11 promotes deletional nonhomologous end joining. J. Biol. Chem. 284(44), 30565–30573 (2009).
    • 85 Truong LN, Li Y, Shi LZ et al. Microhomology-mediated end joining and homologous recombination share the initial end resection step to repair DNA double-strand breaks in mammalian cells. Proc. Natl Acad. Sci. USA 110(19), 7720–7725 (2013).
    • 86 Ottaviani D, LeCain M, Sheer D. The role of microhomology in genomic structural variation. Trends Genet. 30(3), 85–94 (2014).
    • 87 Maresca M, Erler A, Fu J, Friedrich A, Zhang Y, Stewart AF. Single-stranded heteroduplex intermediates in lambda red homologous recombination. BMC Mol. Biol. 11, 54 (2010).
    • 88 Kuzminov A. Recombinational repair of DNA damage in Escherichia coli and bacteriophage lambda. Microbiol. Mol. Biol. Rev. 63(4), 751–813, table of contents (1999).
    • 89 Huber A, Bai P, de Murcia JM, de Murcia G. PARP-1, PARP-2 and ATM in the DNA damage response: functional synergy in mouse development. DNA Repair (Amsterdam) 3(8–9), 1103–1108 (2004).
    • 90 Min W, Cortes U, Herceg Z, Tong WM, Wang ZQ. Deletion of the nuclear isoform of poly(ADP-ribose) glycohydrolase (PARG) reveals its function in DNA repair, genomic stability and tumorigenesis. Carcinogenesis 31(12), 2058–2065 (2010).
    • 91 Illuzzi G, Fouquerel E, Ame JC et al. PARG is dispensable for recovery from transient replicative stress but required to prevent detrimental accumulation of poly(ADP-ribose) upon prolonged replicative stress. Nucl. Acids Res. 42(12), 7776–7792 (2014).
    • 92 Vastag L, Koyuncu E, Grady SL, Shenk TE, Rabinowitz JD. Divergent effects of human cytomegalovirus and herpes simplex virus-1 on cellular metabolism. PLoS Pathog. 7(7), e1002124 (2011).
    • 93 Grady SL, Hwang J, Vastag L, Rabinowitz JD, Shenk T. Herpes simplex virus 1 infection activates poly(ADP-ribose) polymerase and triggers the degradation of poly(ADP-ribose) glycohydrolase. J. Virol. 86(15), 8259–8268 (2012).• Identifies PARG as a target for degradation by ICP0 during HSV-1 infection.
    • 94 Mohni KN, Smith S, Dee AR, Schumacher AJ, Weller SK. Herpes simplex virus type 1 single strand DNA binding protein and helicase/primase complex disable cellular ATR signaling. PLoS Pathog. 9(10), e1003652 (2013).
    • 95 Bieniasz PD. Intrinsic immunity: a front-line defense against viral attack. Nat. Immunol. 5(11), 1109–1115 (2004).
    • 96 Boutell C, Everett RD. Regulation of alphaherpesvirus infections by the ICP0 family of proteins. J. Gen. Virol. 94(Pt 3), 465–481 (2013).
    • 97 Tavalai N, Papior P, Rechter S, Leis M, Stamminger T. Evidence for a role of the cellular ND10 protein PML in mediating intrinsic immunity against human cytomegalovirus infections. J. Virol. 80(16), 8006–8018 (2006).
    • 98 Maldonado E, Shiekhattar R, Sheldon M et al. A human RNA polymerase II complex associated with SRB and DNA-repair proteins. Nature 381(6577), 86–89 (1996).
    • 99 Pankotai T, Bonhomme C, Chen D, Soutoglou E. DNAPKcs-dependent arrest of RNA polymerase II transcription in the presence of DNA breaks. Nat. Struct. Mol. Biol. 19(3), 276–282 (2012).
    • 100 Knipe DM, Cliffe A. Chromatin control of herpes simplex virus lytic and latent infection. Nat. Rev. Microbiol. 6(3), 211–221 (2008).
    • 101 Orzalli MH, Conwell SE, Berrios C, DeCaprio JA, Knipe DM. Nuclear interferon-inducible protein 16 promotes silencing of herpesviral and transfected DNA. Proc. Natl Acad. Sci. USA 110(47), E4492–E4501 (2013).
    • 102 Orzalli MH, Knipe DM. Cellular sensing of viral DNA and viral evasion mechanisms. Annu. Rev. Microbiol. 68, 477–492 (2014).•• Excellent review of the mechanisms by which mammalian cells sense foreign DNAs to trigger antimicrobial responses.
    • 103 Unterholzner L. The interferon response to intracellular DNA: why so many receptors? Immunobiology 218(11), 1312–1321 (2013).
    • 104 Ferguson BJ, Mansur DS, Peters NE, Ren H, Smith GL. DNA-PK is a DNA sensor for IRF-3-dependent innate immunity. Elife 1, e00047 (2012).• Identifies a novel role for DNA-dependent protein kinase catalytic subunit as a DNA sensor for the innate immune response.
    • 105 Unterholzner L, Keating SE, Baran M et al. IFI16 is an innate immune sensor for intracellular DNA. Nat. Immunol. 11(11), 997–1004 (2010).
    • 106 Orzalli MH, DeLuca NA, Knipe DM. Nuclear IFI16 induction of IRF-3 signaling during herpesviral infection and degradation of IFI16 by the viral ICP0 protein. Proc. Natl Acad. Sci. USA 109(44), E3008–3017 (2012).
    • 107 Li Z, Yamauchi Y, Kamakura M et al. Herpes simplex virus requires poly(ADP-ribose) polymerase activity for efficient replication and induces extracellular signal-related kinase-dependent phosphorylation and ICP0-dependent nuclear localization of tankyrase 1. J. Virol. 86(1), 492–503 (2012).
    • 108 Johnson KE, Chikoti L, Chandran B. Herpes simplex virus 1 infection induces activation and subsequent inhibition of the IFI16 and NLRP3 inflammasomes. J. Virol. 87(9), 5005–5018 (2013).
    • 109 Bacon TH, Levin MJ, Leary JJ, Sarisky RT, Sutton D. Herpes simplex virus resistance to acyclovir and penciclovir after two decades of antiviral therapy. Clin. Microbiol. Rev. 16(1), 114–128 (2003).
    • 110 van Velzen M, Missotten T, van Loenen FB et al. Acyclovir-resistant herpes simplex virus type 1 in intra-ocular fluid samples of herpetic uveitis patients. J. Clin. Virol. 57(3), 215–221 (2013).
    • 111 Yan Z, Bryant KF, Gregory SM et al. HIV integrase inhibitors block replication of alpha-, beta-, and gammaherpesviruses. MBio 5(4), e01318–e01314 (2014).
    • 112 Alekseev O, Donovan K, Azizkhan-Clifford J. Inhibition of ataxia telangiectasia mutated (ATM) kinase suppresses herpes simplex virus type 1 (HSV-1) keratitis. Invest. Ophthalmol. Vis. Sci. 55(2), 706–715 (2014).
    • 113 Myers RS, Rudd KE. Mining DNA sequences for molecular enzymology: the Red alpha superfamily defines a set of recombination nucleases. In: Proceedings of the 1998 Miami Nature Biotechnology Winter Symposium. Ahmed F (Ed.). Oxford, United Kingdom, 49–50 (1998).
    • 114 Lo Piano A, Martinez-Jimenez MI, Zecchi L, Ayora S. Recombination-dependent concatemeric viral DNA replication. Virus Res 160(1–2), 1–14 (2011).