#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

Limiting of the Innate Immune Response by SF3A-Dependent Control of MyD88 Alternative mRNA Splicing


Controlling infectious disease without inducing unwanted inflammatory disease requires proper regulation of the innate immune response. Thus, innate immunity needs to be activated when needed during an infection, but must be limited to prevent damage. To accomplish this, negative regulators of innate immunity limit the response. Here we investigate one such negative regulator encoded by an alternative splice form of MyD88. MyD88 mRNA exists in two alternative splice forms: MyD88L, a long form that encodes a protein that activates innate immunity by transducing Toll-like receptor (TLR) signals; and a short form that encodes a different protein, MyD88S, that inhibits the response. We find that MyD88S levels regulate the extent of inflammatory cytokine production in murine macrophages. MyD88S mRNA levels are regulated by the SF3A and SF3B mRNA splicing complexes, and these mRNA splicing complexes function with TLR signaling to regulate MyD88S production. Thus, the SF3A mRNA splicing complex controls production of a negative regulator of TLR signaling that limits the extent of innate immune activation.


Vyšlo v časopise: Limiting of the Innate Immune Response by SF3A-Dependent Control of MyD88 Alternative mRNA Splicing. PLoS Genet 9(10): e32767. doi:10.1371/journal.pgen.1003855
Kategorie: Research Article
prolekare.web.journal.doi_sk: https://doi.org/10.1371/journal.pgen.1003855

Souhrn

Controlling infectious disease without inducing unwanted inflammatory disease requires proper regulation of the innate immune response. Thus, innate immunity needs to be activated when needed during an infection, but must be limited to prevent damage. To accomplish this, negative regulators of innate immunity limit the response. Here we investigate one such negative regulator encoded by an alternative splice form of MyD88. MyD88 mRNA exists in two alternative splice forms: MyD88L, a long form that encodes a protein that activates innate immunity by transducing Toll-like receptor (TLR) signals; and a short form that encodes a different protein, MyD88S, that inhibits the response. We find that MyD88S levels regulate the extent of inflammatory cytokine production in murine macrophages. MyD88S mRNA levels are regulated by the SF3A and SF3B mRNA splicing complexes, and these mRNA splicing complexes function with TLR signaling to regulate MyD88S production. Thus, the SF3A mRNA splicing complex controls production of a negative regulator of TLR signaling that limits the extent of innate immune activation.


Zdroje

1. Kaufmann SHE, Medzhitov, Ruslan, Gordon, Siamon (2004) The innate immune response to infection. Washington D.C.: ASM Press.

2. ChaudhuriN, DowerSK, WhyteMK, SabroeI (2005) Toll-like receptors and chronic lung disease. Clin Sci (Lond) 109: 125–133.

3. CookDN, PisetskyDS, SchwartzDA (2004) Toll-like receptors in the pathogenesis of human disease. Nat Immunol 5: 975–979.

4. GrivennikovSI, GretenFR, KarinM (2010) Immunity, inflammation, and cancer. Cell 140: 883–899.

5. TakedaK, AkiraS (2005) Toll-like receptors in innate immunity. Int Immunol 17: 1–14.

6. KawaiT, AkiraS (2010) The role of pattern-recognition receptors in innate immunity: update on Toll-like receptors. Nat Immunol 11: 373–384.

7. TakeuchiO, AkiraS (2010) Pattern recognition receptors and inflammation. Cell 140: 805–820.

8. KondoT, KawaiT, AkiraS (2012) Dissecting negative regulation of Toll-like receptor signaling. Trends Immunol 33: 449–458.

9. LangT, MansellA (2007) The negative regulation of Toll-like receptor and associated pathways. Immunol Cell Biol 85: 425–434.

10. LiewFY, XuD, BrintEK, O'NeillLA (2005) Negative regulation of toll-like receptor-mediated immune responses. Nature Reviews Immunology 5: 446–458.

11. LuYC, YehWC, OhashiPS (2008) LPS/TLR4 signal transduction pathway. Cytokine 42: 145–151.

12. MurrayPJ, SmaleST (2012) Restraint of inflammatory signaling by interdependent strata of negative regulatory pathways. Nat Immunol 13: 916–924.

13. WangJ, HuY, DengWW, SunB (2009) Negative regulation of Toll-like receptor signaling pathway. Microbes Infect 11: 321–327.

14. SunSC (2008) Deubiquitylation and regulation of the immune response. Nat Rev Immunol 8: 501–511.

15. AlamMM, O'NeillLA (2011) MicroRNAs and the resolution phase of inflammation in macrophages. Eur J Immunol 41: 2482–2485.

16. GrayP, MichelsenKS, SiroisCM, LoweE, ShimadaK, et al. (2010) Identification of a novel human MD-2 splice variant that negatively regulates Lipopolysaccharide-induced TLR4 signaling. J Immunol 184: 6359–6366.

17. HardyMP, O'NeillLA (2004) The murine IRAK2 gene encodes four alternatively spliced isoforms, two of which are inhibitory. J Biol Chem 279: 27699–27708.

18. IwamiKI, MatsuguchiT, MasudaA, KikuchiT, MusikacharoenT, et al. (2000) Cutting edge: naturally occurring soluble form of mouse Toll-like receptor 4 inhibits lipopolysaccharide signaling. J Immunol 165: 6682–6686.

19. LeemanJR, GilmoreTD (2008) Alternative splicing in the NF-kappaB signaling pathway. Gene 423: 97–107.

20. LynchKW (2004) Consequences of regulated pre-mRNA splicing in the immune system. Nat Rev Immunol 4: 931–940.

21. OhtaS, BahrunU, TanakaM, KimotoM (2004) Identification of a novel isoform of MD-2 that downregulates lipopolysaccharide signaling. Biochem Biophys Res Commun 323: 1103–1108.

22. WellsCA, ChalkAM, ForrestA, TaylorD, WaddellN, et al. (2006) Alternate transcription of the Toll-like receptor signaling cascade. Genome Biol 7: R10.

23. BurnsK, JanssensS, BrissoniB, OlivosN, BeyaertR, et al. (2003) Inhibition of interleukin 1 receptor/Toll-like receptor signaling through the alternatively spliced, short form of MyD88 is due to its failure to recruit IRAK-4. Journal of Experimental Medicine 197: 263–268.

24. JanssensS, BurnsK, TschoppJ, BeyaertR (2002) Regulation of interleukin-1- and lipopolysaccharide-induced NF-kappaB activation by alternative splicing of MyD88. Curr Biol 12: 467–471.

25. JanssensS, BurnsK, VercammenE, TschoppJ, BeyaertR (2003) MyD88S, a splice variant of MyD88, differentially modulates NF-kappaB- and AP-1-dependent gene expression. FEBS Lett 548: 103–107.

26. Mendoza-BarberaE, Corral-RodriguezMA, Soares-SchanoskiA, VelardeM, MacieiraS, et al. (2009) Contribution of globular death domains and unstructured linkers to MyD88.IRAK-4 heterodimer formation: an explanation for the antagonistic activity of MyD88s. Biochem Biophys Res Commun 380: 183–187.

27. Adib-ConquyM, AdrieC, FittingC, GattolliatO, BeyaertR, et al. (2006) Up-regulation of MyD88s and SIGIRR, molecules inhibiting Toll-like receptor signaling, in monocytes from septic patients. Crit Care Med 34: 2377–2385.

28. De ArrasL, SengA, LackfordB, KeikhaeeM, BowermanB, et al. (2013) An evolutionarily conserved innate immunity protein interaction network. J Biol Chem 288: 1967–1978.

29. KramerA, FerfogliaF, HuangCJ, MulhauptF, NesicD, et al. (2005) Structure-function analysis of the U2 snRNP-associated splicing factor SF3a. Biochem Soc Trans 33: 439–442.

30. HodgesPE, BeggsJD (1994) RNA splicing. U2 fulfils a commitment. Curr Biol 4: 264–267.

31. SperlingJ, AzubelM, SperlingR (2008) Structure and function of the Pre-mRNA splicing machine. Structure 16: 1605–1615.

32. WahlMC, WillCL, LuhrmannR (2009) The spliceosome: design principles of a dynamic RNP machine. Cell 136: 701–718.

33. RinoJ, Carmo-FonsecaM (2009) The spliceosome: a self-organized macromolecular machine in the nucleus? Trends Cell Biol 19: 375–384.

34. CollinsLJ, KurlandCG, BiggsP, PennyD (2009) The modern RNP world of eukaryotes. J Hered 100: 597–604.

35. BrosiR, GroningK, BehrensSE, LuhrmannR, KramerA (1993) Interaction of mammalian splicing factor SF3a with U2 snRNP and relation of its 60-kD subunit to yeast PRP9. Science 262: 102–105.

36. NesicD, KramerA (2001) Domains in human splicing factors SF3a60 and SF3a66 required for binding to SF3a120, assembly of the 17S U2 snRNP, and prespliceosome formation. Mol Cell Biol 21: 6406–6417.

37. RubySW, ChangTH, AbelsonJ (1993) Four yeast spliceosomal proteins (PRP5, PRP9, PRP11, and PRP21) interact to promote U2 snRNP binding to pre-mRNA. Genes Dev 7: 1909–1925.

38. TanackovicG, KramerA (2005) Human splicing factor SF3a, but not SF1, is essential for pre-mRNA splicing in vivo. Mol Biol Cell 16: 1366–1377.

39. WiestDK, O'DayCL, AbelsonJ (1996) In vitro studies of the Prp9.Prp11.Prp21 complex indicate a pathway for U2 small nuclear ribonucleoprotein activation. J Biol Chem 271: 33268–33276.

40. DasBK, XiaL, PalandjianL, GozaniO, ChyungY, et al. (1999) Characterization of a protein complex containing spliceosomal proteins SAPs 49, 130, 145, and 155. Mol Cell Biol 19: 6796–6802.

41. KramerA, GruterP, GroningK, KastnerB (1999) Combined biochemical and electron microscopic analyses reveal the architecture of the mammalian U2 snRNP. J Cell Biol 145: 1355–1368.

42. WillCL, SchneiderC, ReedR, LuhrmannR (1999) Identification of both shared and distinct proteins in the major and minor spliceosomes. Science 284: 2003–2005.

43. AlperS, LawsR, LackfordB, BoydWA, DunlapP, et al. (2008) Identification of innate immunity genes and pathways using a comparative genomics approach. ProcNatl Acad Sci USA 105: 7016–7021.

44. CorrioneroA, MinanaB, ValcarcelJ (2011) Reduced fidelity of branch point recognition and alternative splicing induced by the anti-tumor drug spliceostatin A. Genes Dev 25: 445–459.

45. KaidaD, MotoyoshiH, TashiroE, NojimaT, HagiwaraM, et al. (2007) Spliceostatin A targets SF3b and inhibits both splicing and nuclear retention of pre-mRNA. Nat Chem Biol 3: 576–583.

46. RoybalGA, JuricaMS (2010) Spliceostatin A inhibits spliceosome assembly subsequent to prespliceosome formation. Nucleic Acids Res 38: 6664–6672.

47. AliprantisAO, YangRB, MarkMR, SuggettS, DevauxB, et al. (1999) Cell activation and apoptosis by bacterial lipoproteins through toll-like receptor-2. Science 285: 736–739.

48. MercurioF, ZhuH, MurrayBW, ShevchenkoA, BennettBL, et al. (1997) IKK-1 and IKK-2: cytokine-activated IkappaB kinases essential for NF-kappaB activation. Science 278: 860–866.

49. HäckerH, RedeckeV, BlagoevB, KratchmarovaI, HsuLC, et al. (2006) Specificity in Toll-like receptor signalling through distinct effector functions of TRAF3 and TRAF6. Nature 439: 204–207.

50. MaruyamaK, SuganoS (1994) Oligo-capping: a simple method to replace the cap structure of eukaryotic mRNAs with oligoribonucleotides. Gene 138: 171–174.

51. SuzukiY, Yoshitomo-NakagawaK, MaruyamaK, SuyamaA, SuganoS (1997) Construction and characterization of a full length-enriched and a 5′-end-enriched cDNA library. Gene 200: 149–156.

52. IsonoK, Mizutani-KosekiY, KomoriT, Schmidt-ZachmannMS, KosekiH (2005) Mammalian polycomb-mediated repression of Hox genes requires the essential spliceosomal protein Sf3b1. Genes Dev 19: 536–541.

53. KramerA, MulhauserF, WersigC, GroningK, BilbeG (1995) Mammalian splicing factor SF3a120 represents a new member of the SURP family of proteins and is homologous to the essential splicing factor PRP21p of Saccharomyces cerevisiae. RNA 1: 260–272.

54. AnM, HenionPD (2012) The zebrafish sf3b1b460 mutant reveals differential requirements for the sf3b1 pre-mRNA processing gene during neural crest development. Int J Dev Biol 56: 223–237.

55. FanL, LagisettiC, EdwardsCC, WebbTR, PotterPM (2011) Sudemycins, novel small molecule analogues of FR901464, induce alternative gene splicing. ACS Chem Biol 6: 582–589.

56. VisconteV, RogersHJ, SinghJ, BarnardJ, BupathiM, et al. (2012) SF3B1 haploinsufficiency leads to formation of ring sideroblasts in myelodysplastic syndromes. Blood 120: 3173–3186.

57. DammF, KosmiderO, Gelsi-BoyerV, RennevilleA, CarbucciaN, et al. (2012) Mutations affecting mRNA splicing define distinct clinical phenotypes and correlate with patient outcome in myelodysplastic syndromes. Blood 119: 3211–3218.

58. DammF, TholF, KosmiderO, KadeS, LoffeldP, et al. (2012) SF3B1 mutations in myelodysplastic syndromes: clinical associations and prognostic implications. Leukemia 26: 1137–1140.

59. EllisMJ, DingL, ShenD, LuoJ, SumanVJ, et al. (2012) Whole-genome analysis informs breast cancer response to aromatase inhibition. Nature 486: 353–360.

60. MakishimaH, VisconteV, SakaguchiH, JankowskaAM, Abu KarS, et al. (2012) Mutations in the spliceosome machinery, a novel and ubiquitous pathway in leukemogenesis. Blood 119: 3203–3210.

61. MalcovatiL, PapaemmanuilE, BowenDT, BoultwoodJ, Della PortaMG, et al. (2011) Clinical significance of SF3B1 mutations in myelodysplastic syndromes and myelodysplastic/myeloproliferative neoplasms. Blood 118: 6239–6246.

62. PapaemmanuilE, CazzolaM, BoultwoodJ, MalcovatiL, VyasP, et al. (2011) Somatic SF3B1 mutation in myelodysplasia with ring sideroblasts. N Engl J Med 365: 1384–1395.

63. PatnaikMM, LashoTL, HodnefieldJM, KnudsonRA, KetterlingRP, et al. (2012) SF3B1 mutations are prevalent in myelodysplastic syndromes with ring sideroblasts but do not hold independent prognostic value. Blood 119: 569–572.

64. QuesadaV, CondeL, VillamorN, OrdonezGR, JaresP, et al. (2012) Exome sequencing identifies recurrent mutations of the splicing factor SF3B1 gene in chronic lymphocytic leukemia. Nat Genet 44: 47–52.

65. RossiD, BruscagginA, SpinaV, RasiS, KhiabanianH, et al. (2011) Mutations of the SF3B1 splicing factor in chronic lymphocytic leukemia: association with progression and fludarabine-refractoriness. Blood 118: 6904–6908.

66. VisconteV, MakishimaH, JankowskaA, SzpurkaH, TrainaF, et al. (2012) SF3B1, a splicing factor is frequently mutated in refractory anemia with ring sideroblasts. Leukemia 26: 542–545.

67. WangL, LawrenceMS, WanY, StojanovP, SougnezC, et al. (2011) SF3B1 and other novel cancer genes in chronic lymphocytic leukemia. N Engl J Med 365: 2497–2506.

68. YoshidaK, SanadaM, ShiraishiY, NowakD, NagataY, et al. (2011) Frequent pathway mutations of splicing machinery in myelodysplasia. Nature 478: 64–69.

69. KarinM, GretenFR (2005) NF-kappaB: linking inflammation and immunity to cancer development and progression. Nat Rev Immunol 5: 749–759.

70. LiQ, WithoffS, VermaIM (2005) Inflammation-associated cancer: NF-kappaB is the lynchpin. Trends Immunol 26: 318–325.

71. OhshimaH, TatemichiM, SawaT (2003) Chemical basis of inflammation-induced carcinogenesis. Arch Biochem Biophys 417: 3–11.

72. StarczynowskiDT, KarsanA (2010) Innate immune signaling in the myelodysplastic syndromes. Hematol Oncol Clin North Am 24: 343–359.

73. TsanMF (2006) Toll-like receptors, inflammation and cancer. Semin Cancer Biol 16: 32–37.

74. SchusterTB, CostinaV, FindeisenP, NeumaierM, Ahmad-NejadP (2011) Identification and functional characterization of 14-3-3 in TLR2 signaling. J Proteome Res 10: 4661–4670.

75. LiS, ArmstrongCM, BertinN, GeH, MilsteinS, et al. (2004) A map of the interactome network of the metazoan C. elegans. Science 303: 540–543.

76. CartyM, GoodbodyR, SchroderM, StackJ, MoynaghPN, et al. (2006) The human adaptor SARM negatively regulates adaptor protein TRIF-dependent Toll-like receptor signaling. Nat Immunol 7: 1074–1081.

77. PengJ, YuanQ, LinB, PanneerselvamP, WangX, et al. (2010) SARM inhibits both TRIF- and MyD88-mediated AP-1 activation. Eur J Immunol 40: 1738–1747.

78. CouillaultC, PujolN, ReboulJ, SabatierL, GuichouJF, et al. (2004) TLR-independent control of innate immunity in Caenorhabditis elegans by the TIR domain adaptor protein TIR-1, an ortholog of human SARM. Nat Immunol 5: 488–494.

79. LiberatiNT, FitzgeraldKA, KimDH, FeinbaumR, GolenbockDT, et al. (2004) Requirement for a conserved Toll/interleukin-1 resistance domain protein in the Caenorhabditis elegans immune response. Proc Natl Acad Sci U S A 101: 6593–6598.

80. MuhammedM, FuchsBB, WuMP, BregerJ, ColemanJJ, et al. (2012) The role of mycelium production and a MAPK-mediated immune response in the C. elegans-Fusarium model system. Med Mycol 50: 488–496.

81. GuruharshaKG, RualJF, ZhaiB, MintserisJ, VaidyaP, et al. (2011) A protein complex network of Drosophila melanogaster. Cell 147: 690–703.

82. HuangCJ, FerfogliaF, RaleffF, KramerA (2011) Interaction domains and nuclear targeting signals in subunits of the U2 small nuclear ribonucleoprotein particle-associated splicing factor SF3a. J Biol Chem 286: 13106–13114.

83. NesicD, TanackovicG, KramerA (2004) A role for Cajal bodies in the final steps of U2 snRNP biogenesis. J Cell Sci 117: 4423–4433.

84. HunterP (2012) The inflammation theory of disease. The growing realization that chronic inflammation is crucial in many diseases opens new avenues for treatment. EMBO Rep 13: 968–970.

85. Fernandez-Botran R, Větvička V (2001) Methods in Cellular Immunology. Boca Raton: CRC Press. p. 8.

86. VickersTA, ZhangH, GrahamMJ, LemonidisKM, ZhaoC, et al. (2006) Modification of MyD88 mRNA splicing and inhibition of IL-1beta signaling in cell culture and in mice with a 2′-O-methoxyethyl-modified oligonucleotide. J Immunol 176: 3652–3661.

87. SchneiderCA, RasbandWS, EliceiriKW (2012) NIH Image to ImageJ: 25 years of image analysis. Nat Methods 9: 671–675.

Štítky
Genetika Reprodukčná medicína

Článok vyšiel v časopise

PLOS Genetics


2013 Číslo 10
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#