#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

The HSV-1 Exonuclease, UL12, Stimulates Recombination by a Single Strand Annealing Mechanism


Production of concatemeric DNA is an essential step during HSV infection, as the packaging machinery must recognize longer-than-unit-length concatemers; however, the mechanism by which they are formed is poorly understood. Although it has been proposed that the viral genome circularizes and rolling circle replication leads to the formation of concatemers, several lines of evidence suggest that HSV DNA replication involves recombination-dependent replication reminiscent of bacteriophages λ and T4. Similar to λ, HSV-1 encodes a 5′-to-3′ exonuclease (UL12) and a single strand annealing protein [SSAP (ICP8)] that interact with each other and can perform strand exchange in vitro. By analogy with λ phage, HSV may utilize viral and/or cellular recombination proteins during DNA replication. At least four double strand break repair pathways are present in eukaryotic cells, and HSV-1 is known to manipulate several components of these pathways. Chromosomally integrated reporter assays were used to measure the repair of double strand breaks in HSV-infected cells. Single strand annealing (SSA) was increased in HSV-infected cells, while homologous recombination (HR), non-homologous end joining (NHEJ) and alternative non-homologous end joining (A-NHEJ) were decreased. The increase in SSA was abolished when cells were infected with a viral mutant lacking UL12. Moreover, expression of UL12 alone caused an increase in SSA, which was completely eliminated when a UL12 mutant lacking exonuclease activity was expressed. UL12-mediated stimulation of SSA was decreased in cells lacking the cellular SSAP, Rad52, and could be restored by coexpressing the viral SSAP, ICP8, indicating that an SSAP is also required. These results demonstrate that UL12 can specifically stimulate SSA and that either ICP8 or Rad52 can function as an SSAP. We suggest that SSA is the homology-mediated repair pathway utilized during HSV infection.


Vyšlo v časopise: The HSV-1 Exonuclease, UL12, Stimulates Recombination by a Single Strand Annealing Mechanism. PLoS Pathog 8(8): e32767. doi:10.1371/journal.ppat.1002862
Kategorie: Research Article
prolekare.web.journal.doi_sk: https://doi.org/10.1371/journal.ppat.1002862

Souhrn

Production of concatemeric DNA is an essential step during HSV infection, as the packaging machinery must recognize longer-than-unit-length concatemers; however, the mechanism by which they are formed is poorly understood. Although it has been proposed that the viral genome circularizes and rolling circle replication leads to the formation of concatemers, several lines of evidence suggest that HSV DNA replication involves recombination-dependent replication reminiscent of bacteriophages λ and T4. Similar to λ, HSV-1 encodes a 5′-to-3′ exonuclease (UL12) and a single strand annealing protein [SSAP (ICP8)] that interact with each other and can perform strand exchange in vitro. By analogy with λ phage, HSV may utilize viral and/or cellular recombination proteins during DNA replication. At least four double strand break repair pathways are present in eukaryotic cells, and HSV-1 is known to manipulate several components of these pathways. Chromosomally integrated reporter assays were used to measure the repair of double strand breaks in HSV-infected cells. Single strand annealing (SSA) was increased in HSV-infected cells, while homologous recombination (HR), non-homologous end joining (NHEJ) and alternative non-homologous end joining (A-NHEJ) were decreased. The increase in SSA was abolished when cells were infected with a viral mutant lacking UL12. Moreover, expression of UL12 alone caused an increase in SSA, which was completely eliminated when a UL12 mutant lacking exonuclease activity was expressed. UL12-mediated stimulation of SSA was decreased in cells lacking the cellular SSAP, Rad52, and could be restored by coexpressing the viral SSAP, ICP8, indicating that an SSAP is also required. These results demonstrate that UL12 can specifically stimulate SSA and that either ICP8 or Rad52 can function as an SSAP. We suggest that SSA is the homology-mediated repair pathway utilized during HSV infection.


Zdroje

1. BrownSM, RitchieDA, Subak-SharpeJH (1973) 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: 329–346.

2. HaywardGS, JacobRJ, WadsworthSC, RoizmanB (1975) 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 U S A 72: 4243–4247.

3. SchafferPA, TevethiaMJ, Benyesh-MelnickM (1974) Recombination between temperature-sensitive mutants of herpes simplex virus type 1. Virology 58: 219–228.

4. SheldrickP, BerthelotN (1975) Inverted repetitions in the chromosome of herpes simplex virus. Cold Spring Harb Symp Quant Biol 39 Pt 2: 667–678.

5. ZhangX, EfstathiouS, SimmonsA (1994) Identification of novel herpes simplex virus replicative intermediates by field inversion gel electrophoresis: implications for viral DNA amplification strategies. Virology 202: 530–539.

6. DeliusH, ClementsJB (1976) A partial denaturation map of herpes simplex virus type 1 DNA: evidence for inversions of the unique DNA regions. J Gen Virol 33: 125–133.

7. DutchRE, BianchiV, LehmanIR (1995) Herpes simplex virus type 1 DNA replication is specifically required for high-frequency homologous recombination between repeated sequences. J Virol 69: 3084–3089.

8. FuX, WangH, ZhangX (2002) High-frequency intermolecular homologous recombination during herpes simplex virus-mediated plasmid DNA replication. J Virol 76: 5866–5874.

9. ReuvenNB, StaireAE, MyersRS, WellerSK (2003) The herpes simplex virus type 1 alkaline nuclease and single-stranded DNA binding protein mediate strand exchange in vitro. J Virol 77: 7425–7433.

10. ReuvenNB, WillcoxS, GriffithJD, WellerSK (2004) 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: 57–71.

11. EllisHM, YuD, DiTizioT, CourtDL (2001) High efficiency mutagenesis, repair, and engineering of chromosomal DNA using single-stranded oligonucleotides. Proc Natl Acad Sci U S A 98: 6742–6746.

12. MurphyKC (1998) Use of bacteriophage lambda recombination functions to promote gene replacement in Escherichia coli. J Bacteriol 180: 2063–2071.

13. CourtDL, SawitzkeJA, ThomasonLC (2002) Genetic engineering using homologous recombination. Annu Rev Genet 36: 361–388.

14. MuyrersJP, ZhangY, TestaG, StewartAF (1999) Rapid modification of bacterial artificial chromosomes by ET-recombination. Nucleic Acids Res 27: 1555–1557.

15. ZhangY, BuchholzF, MuyrersJP, StewartAF (1998) A new logic for DNA engineering using recombination in Escherichia coli. Nat Genet 20: 123–128.

16. YuD, EllisHM, LeeEC, JenkinsNA, CopelandNG, et al. (2000) An efficient recombination system for chromosome engineering in Escherichia coli. Proc Natl Acad Sci U S A 97: 5978–5983.

17. SzczepanskaAK (2009) Bacteriophage-encoded functions engaged in initiation of homologous recombination events. Crit Rev Microbiol 35: 197–220.

18. Lo PianoA, Martinez-JimenezMI, ZecchiL, AyoraS (2011) Recombination-dependent concatemeric viral DNA replication. Virus Res 160: 1–14.

19. KuzminovA (1999) Recombinational repair of DNA damage in Escherichia coli and bacteriophage lambda. Microbiol Mol Biol Rev 63: 751–813 table of contents.

20. WilkinsonDE, WellerSK (2003) The role of DNA recombination in herpes simplex virus DNA replication. IUBMB Life 55: 451–458.

21. KassEM, JasinM (2010) Collaboration and competition between DNA double-strand break repair pathways. FEBS Lett 584: 3703–3708.

22. WymanC, KanaarR (2006) DNA double-strand break repair: all's well that ends well. Annu Rev Genet 40: 363–383.

23. IyerLM, KooninEV, AravindL (2002) Classification and evolutionary history of the single-strand annealing proteins, RecT, Redbeta, ERF and RAD52. BMC Genomics 3: 8.

24. KawabataM, KawabataT, NishiboriM (2005) Role of recA/RAD51 family proteins in mammals. Acta Med Okayama 59: 1–9.

25. SingletonMR, WentzellLM, LiuY, WestSC, WigleyDB (2002) Structure of the single-strand annealing domain of human RAD52 protein. Proc Natl Acad Sci U S A 99: 13492–13497.

26. StarkJM, PierceAJ, OhJ, PastinkA, JasinM (2004) Genetic steps of mammalian homologous repair with distinct mutagenic consequences. Mol Cell Biol 24: 9305–9316.

27. BranzeiD, FoianiM (2008) Regulation of DNA repair throughout the cell cycle. Nat Rev Mol Cell Biol 9: 297–308.

28. ShrivastavM, De HaroLP, NickoloffJA (2008) Regulation of DNA double-strand break repair pathway choice. Cell Res 18: 134–147.

29. HarperJW, ElledgeSJ (2007) The DNA damage response: ten years after. Mol Cell 28: 739–745.

30. LavinMF, KozlovS (2007) DNA damage-induced signalling in ataxia-telangiectasia and related syndromes. Radiother Oncol 83: 231–237.

31. DeFazioLG, StanselRM, GriffithJD, ChuG (2002) Synapsis of DNA ends by DNA-dependent protein kinase. EMBO J 21: 3192–3200.

32. SpagnoloL, Rivera-CalzadaA, PearlLH, LlorcaO (2006) Three-dimensional structure of the human DNA-PKcs/Ku70/Ku80 complex assembled on DNA and its implications for DNA DSB repair. Mol Cell 22: 511–519.

33. ValerieK, PovirkLF (2003) Regulation and mechanisms of mammalian double-strand break repair. Oncogene 22: 5792–5812.

34. MohniKN, LivingstonCM, CortezD, WellerSK (2010) ATR and ATRIP are recruited to herpes simplex virus type 1 replication compartments even though ATR signaling is disabled. J Virol 84: 12152–12164.

35. MohniKN, MastrocolaAS, BaiP, WellerSK, HeinenCD (2011) DNA mismatch repair proteins are required for efficient herpes simplex virus 1 replication. J Virol 85: 12241–12253.

36. GregoryDA, BachenheimerSL (2008) Characterization of mre11 loss following HSV-1 infection. Virology 373: 124–136.

37. Lees-MillerSP, LongMC, KilvertMA, LamV, RiceSA, et al. (1996) Attenuation of DNA-dependent protein kinase activity and its catalytic subunit by the herpes simplex virus type 1 transactivator ICP0. J Virol 70: 7471–7477.

38. LilleyCE, CarsonCT, MuotriAR, GageFH, WeitzmanMD (2005) DNA repair proteins affect the lifecycle of herpes simplex virus 1. Proc Natl Acad Sci U S A 102: 5844–5849.

39. TaylorTJ, KnipeDM (2004) Proteomics of herpes simplex virus replication compartments: association of cellular DNA replication, repair, recombination, and chromatin remodeling proteins with ICP8. J Virol 78: 5856–5866.

40. WilcockD, LaneDP (1991) Localization of p53, retinoblastoma and host replication proteins at sites of viral replication in herpes-infected cells. Nature 349: 429–431.

41. WilkinsonDE, WellerSK (2004) 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: 4783–4796.

42. LilleyCE, ChaurushiyaMS, BoutellC, EverettRD, WeitzmanMD (2011) The intrinsic antiviral defense to incoming HSV-1 genomes includes specific DNA repair proteins and is counteracted by the viral protein ICP0. PLoS Pathog 7: e1002084.

43. LilleyCE, ChaurushiyaMS, BoutellC, LandryS, SuhJ, et al. (2010) A viral E3 ligase targets RNF8 and RNF168 to control histone ubiquitination and DNA damage responses. EMBO J 29: 943–955.

44. ParkinsonJ, Lees-MillerSP, EverettRD (1999) 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: 650–657.

45. ShirataN, KudohA, DaikokuT, TatsumiY, FujitaM, et al. (2005) Activation of ataxia telangiectasia-mutated DNA damage checkpoint signal transduction elicited by herpes simplex virus infection. J Biol Chem 280: 30336–30341.

46. MuylaertI, EliasP (2007) Knockdown of DNA ligase IV/XRCC4 by RNA interference inhibits herpes simplex virus type I DNA replication. J Biol Chem 282: 10865–10872.

47. PierceAJ, JohnsonRD, ThompsonLH, JasinM (1999) XRCC3 promotes homology-directed repair of DNA damage in mammalian cells. Genes Dev 13: 2633–2638.

48. BennardoN, ChengA, HuangN, StarkJM (2008) Alternative-NHEJ is a mechanistically distinct pathway of mammalian chromosome break repair. PLoS Genet 4: e1000110.

49. SeveriniA, MorganAR, TovellDR, TyrrellDL (1994) Study of the structure of replicative intermediates of HSV-1 DNA by pulsed-field gel electrophoresis. Virology 200: 428–435.

50. GoldsteinDJ, WellerSK (1988) lacZ insertional mutagen is used to demonstrate that the UL52 gene of herpes simplex virus type 1 is required for virus growth and DNA synthesis. J Virol 62: 2970–2977.

51. LambertiC, WellerSK (1998) The herpes simplex virus type 1 cleavage/packaging protein, UL32, is involved in efficient localization of capsids to replication compartments. J Virol 72: 2463–2473.

52. LimSI, MinBE, JungGY (2008) Lagging strand-biased initiation of red recombination by linear double-stranded DNAs. J Mol Biol 384: 1098–1105.

53. PoteeteAR (2008) Involvement of DNA replication in phage lambda Red-mediated homologous recombination. Mol Microbiol 68: 66–74.

54. StahlFW, McMilinKD, StahlMM, CrasemannJM, LamS (1974) The distribution of crossovers along unreplicated lambda bacteriophage chromosomes. Genetics 77: 395–408.

55. LiZ, KarakousisG, ChiuSK, ReddyG, RaddingCM (1998) The beta protein of phage lambda promotes strand exchange. J Mol Biol 276: 733–744.

56. MuniyappaK, RaddingCM (1986) The homologous recombination system of phage lambda. Pairing activities of beta protein. J Biol Chem 261: 7472–7478.

57. StahlMM, ThomasonL, PoteeteAR, TarkowskiT, KuzminovA, et al. (1997) Annealing vs. invasion in phage lambda recombination. Genetics 147: 961–977.

58. GaoM, KnipeDM (1989) Genetic evidence for multiple nuclear functions of the herpes simplex virus ICP8 DNA-binding protein. J Virol 63: 5258–5267.

59. WellerSK, SeghatoleslamiMR, ShaoL, RowseD, CarmichaelEP (1990) The herpes simplex virus type 1 alkaline nuclease is not essential for viral DNA synthesis: isolation and characterization of a lacZ insertion mutant. J Gen Virol 71 (Pt 12) 2941–2952.

60. NimonkarAV, BoehmerPE (2002) In vitro strand exchange promoted by the herpes simplex virus type-1 single strand DNA-binding protein (ICP8) and DNA helicase-primase. J Biol Chem 277: 15182–15189.

61. NimonkarAV, BoehmerPE (2003) On the mechanism of strand assimilation by the herpes simplex virus type-1 single-strand DNA-binding protein (ICP8). Nucleic Acids Res 31: 5275–5281.

62. BalasubramanianN, BaiP, BuchekG, KorzaG, WellerSK (2010) Physical interaction between the herpes simplex virus type 1 exonuclease, UL12, and the DNA double-strand break-sensing MRN complex. J Virol 84: 12504–12514.

63. ThomasMS, GaoM, KnipeDM, PowellKL (1992) Association between the herpes simplex virus major DNA-binding protein and alkaline nuclease. J Virol 66: 1152–1161.

64. AntrobusR, GrantK, GangadharanB, ChittendenD, EverettRD, et al. (2009) Proteomic analysis of cells in the early stages of herpes simplex virus type-1 infection reveals widespread changes in the host cell proteome. Proteomics 9: 3913–3927.

65. StarkJM, HuP, PierceAJ, MoynahanME, EllisN, et al. (2002) ATP hydrolysis by mammalian RAD51 has a key role during homology-directed DNA repair. J Biol Chem 277: 20185–20194.

66. GoldsteinJN, WellerSK (1998) The exonuclease activity of HSV-1 UL12 is required for in vivo function. Virology 244: 442–457.

67. ShirasawaS, FuruseM, YokoyamaN, SasazukiT (1993) Altered growth of human colon cancer cell lines disrupted at activated Ki-ras. Science 260: 85–88.

68. KoiM, UmarA, ChauhanDP, CherianSP, CarethersJM, et al. (1994) Human chromosome 3 corrects mismatch repair deficiency and microsatellite instability and reduces N-methyl-N′-nitro-N-nitrosoguanidine tolerance in colon tumor cells with homozygous hMLH1 mutation. Cancer Res 54: 4308–4312.

69. GunnA, BennardoN, ChengA, StarkJM (2011) Correct End Use during End Joining of Multiple Chromosomal Double Strand Breaks Is Influenced by Repair Protein RAD50, DNA-dependent Protein Kinase DNA-PKcs, and Transcription Context. J Biol Chem 286: 42470–42482.

70. BrooksK, ClarkAJ (1967) Behavior of lambda bacteriophage in a recombination deficienct strain of Escherichia coli. J Virol 1: 283–293.

71. EcholasH, GingeryR (1968) Mutants of bacteriophage lambda defective in vegetative genetic recombination. J Mol Biol 34: 239–249.

72. FranklinNC (1967) Deletions and functions of the center of the phi80 -lambda phage genome. Evidence for a phage function promoting genetic recombination. Genetics 57: 301–318.

73. ShulmanMJ, HallickLM, EcholsH, SignerER (1970) Properties of recombination-deficient mutants of bacteriophage lambda. J Mol Biol 52: 501–520.

74. SignerER, WeilJ (1968) Recombination in bacteriophage lambda. I. Mutants deficient in general recombination. J Mol Biol 34: 261–271.

75. van de PutteP, ZwenkH, RorschA (1966) Properties of four mutants of Escherichia coli defective in genetic recombination. Mutat Res 3: 381–392.

76. KulkarniAS, FortunatoEA (2011) Stimulation of homology-directed repair at I-SceI-induced DNA breaks during the permissive life cycle of human cytomegalovirus. J Virol 85: 6049–6054.

77. MimitouEP, SymingtonLS (2009) DNA end resection: many nucleases make light work. DNA Repair (Amst) 8: 983–995.

78. NimonkarAV, GenschelJ, KinoshitaE, PolaczekP, CampbellJL, et al. (2011) BLM-DNA2-RPA-MRN and EXO1-BLM-RPA-MRN constitute two DNA end resection machineries for human DNA break repair. Genes Dev 25: 350–362.

79. SartoriAA, LukasC, CoatesJ, MistrikM, FuS, et al. (2007) Human CtIP promotes DNA end resection. Nature 450: 509–514.

80. de WindN, DekkerM, BernsA, RadmanM, te RieleH (1995) Inactivation of the mouse Msh2 gene results in mismatch repair deficiency, methylation tolerance, hyperrecombination, and predisposition to cancer. Cell 82: 321–330.

81. LyndakerAM, AlaniE (2009) A tale of tails: insights into the coordination of 3′ end processing during homologous recombination. Bioessays 31: 315–321.

82. SugawaraN, PaquesF, ColaiacovoM, HaberJE (1997) Role of Saccharomyces cerevisiae Msh2 and Msh3 repair proteins in double-strand break-induced recombination. Proc Natl Acad Sci U S A 94: 9214–9219.

83. JacobRJ, RoizmanB (1977) Anatomy of herpes simplex virus DNA VIII. Properties of the replicating DNA. J Virol 23: 394–411.

84. RWHY, OakesJE, KudlerL (1977) In vitro repair of the preexisting nicks and gaps in herpes simplex virus DNA. Virology 76: 286–294.

85. BatailleD, EpsteinA (1994) Herpes simplex virus replicative concatemers contain L components in inverted orientation. Virology 203: 384–388.

86. BatailleD, EpsteinAL (1997) Equimolar generation of the four possible arrangements of adjacent L components in herpes simplex virus type 1 replicative intermediates. J Virol 71: 7736–7743.

87. SariskyRT, WeberPC (1994) Requirement for double-strand breaks but not for specific DNA sequences in herpes simplex virus type 1 genome isomerization events. J Virol 68: 34–47.

88. RybalchenkoN, GolubEI, BiB, RaddingCM (2004) Strand invasion promoted by recombination protein beta of coliphage lambda. Proc Natl Acad Sci U S A 101: 17056–17060.

89. MarescaM, ErlerA, FuJ, FriedrichA, ZhangY, et al. (2010) Single-stranded heteroduplex intermediates in lambda Red homologous recombination. BMC Mol Biol 11: 54.

90. MosbergJA, LajoieMJ, ChurchGM (2010) Lambda red recombineering in Escherichia coli occurs through a fully single-stranded intermediate. Genetics 186: 791–799.

91. BujnickiJM, RychlewskiL (2001) The herpesvirus alkaline exonuclease belongs to the restriction endonuclease PD-(D/E)XK superfamily: insight from molecular modeling and phylogenetic analysis. Virus Genes 22: 219–230.

92. DatsenkoKA, WannerBL (2000) One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A 97: 6640–6645.

93. MurphyKC, CampelloneKG, PoteeteAR (2000) PCR-mediated gene replacement in Escherichia coli. Gene 246: 321–330.

94. MuyrersJP, ZhangY, StewartAF (2001) Techniques: Recombinogenic engineering–new options for cloning and manipulating DNA. Trends Biochem Sci 26: 325–331.

95. CampbellA (1994) Comparative molecular biology of lambdoid phages. Annu Rev Microbiol 48: 193–222.

96. BowdenR, SakaokaH, DonnellyP, WardR (2004) High recombination rate in herpes simplex virus type 1 natural populations suggests significant co-infection. Infect Genet Evol 4: 115–123.

97. MuylkensB, FarnirF, MeurensF, SchyntsF, VanderplasschenA, et al. (2009) Coinfection with two closely related alphaherpesviruses results in a highly diversified recombination mosaic displaying negative genetic interference. J Virol 83: 3127–3137.

98. ThiryE, MeurensF, MuylkensB, McVoyM, GogevS, et al. (2005) Recombination in alphaherpesviruses. Rev Med Virol 15: 89–103.

99. NorbergP, KasubiMJ, HaarrL, BergstromT, LiljeqvistJA (2007) Divergence and recombination of clinical herpes simplex virus type 2 isolates. J Virol 81: 13158–13167.

100. TopalogluO, HurleyPJ, YildirimO, CivinCI, BunzF (2005) Improved methods for the generation of human gene knockout and knockin cell lines. Nucleic Acids Res 33: e158.

101. GoldsteinJN, WellerSK (1998) In vitro processing of herpes simplex virus type 1 DNA replication intermediates by the viral alkaline nuclease, UL12. J Virol 72: 8772–8781.

102. ReuvenNB, AntokuS, WellerSK (2004) The UL12.5 gene product of herpes simplex virus type 1 exhibits nuclease and strand exchange activities but does not localize to the nucleus. J Virol 78: 4599–4608.

103. RichardsonC, MoynahanME, JasinM (1998) Double-strand break repair by interchromosomal recombination: suppression of chromosomal translocations. Genes Dev 12: 3831–3842.

104. FattahF, LeeEH, WeisenselN, WangY, LichterN, et al. (2010) Ku regulates the non-homologous end joining pathway choice of DNA double-strand break repair in human somatic cells. PLoS Genet 6: e1000855.

105. HeilbronnR, zur HausenH (1989) A subset of herpes simplex virus replication genes induces DNA amplification within the host cell genome. J Virol 63: 3683–3692.

106. LivingstonCM, DeLucaNA, WilkinsonDE, WellerSK (2008) Oligomerization of ICP4 and rearrangement of heat shock proteins may be important for herpes simplex virus type 1 prereplicative site formation. J Virol 82: 6324–6336.

Štítky
Hygiena a epidemiológia Infekčné lekárstvo Laboratórium

Článok vyšiel v časopise

PLOS Pathogens


2012 Číslo 8
Najčítanejšie tento týždeň
Najčítanejšie v tomto čísle
Kurzy

Zvýšte si kvalifikáciu online z pohodlia domova

Získaná hemofilie - Povědomí o nemoci a její diagnostika
nový kurz

Eozinofilní granulomatóza s polyangiitidou
Autori: doc. MUDr. Martina Doubková, Ph.D.

Všetky kurzy
Prihlásenie
Zabudnuté heslo

Zadajte e-mailovú adresu, s ktorou ste vytvárali účet. Budú Vám na ňu zasielané informácie k nastaveniu nového hesla.

Prihlásenie

Nemáte účet?  Registrujte sa

#ADS_BOTTOM_SCRIPTS#