#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

Gene Expression Regulation by Upstream Open Reading Frames and Human Disease


Upstream open reading frames (uORFs) are major gene expression regulatory elements. In many eukaryotic mRNAs, one or more uORFs precede the initiation codon of the main coding region. Indeed, several studies have revealed that almost half of human transcripts present uORFs. Very interesting examples have shown that these uORFs can impact gene expression of the downstream main ORF by triggering mRNA decay or by regulating translation. Also, evidence from recent genetic and bioinformatic studies implicates disturbed uORF-mediated translational control in the etiology of many human diseases, including malignancies, metabolic or neurologic disorders, and inherited syndromes. In this review, we will briefly present the mechanisms through which uORFs regulate gene expression and how they can impact on the organism's response to different cell stress conditions. Then, we will emphasize the importance of these structures by illustrating, with specific examples, how disturbed uORF-mediated translational control can be involved in the etiology of human diseases, giving special importance to genotype-phenotype correlations. Identifying and studying more cases of uORF-altering mutations will help us to understand and establish genotype-phenotype associations, leading to advancements in diagnosis, prognosis, and treatment of many human disorders.


Vyšlo v časopise: Gene Expression Regulation by Upstream Open Reading Frames and Human Disease. PLoS Genet 9(8): e32767. doi:10.1371/journal.pgen.1003529
Kategorie: Review
prolekare.web.journal.doi_sk: https://doi.org/10.1371/journal.pgen.1003529

Souhrn

Upstream open reading frames (uORFs) are major gene expression regulatory elements. In many eukaryotic mRNAs, one or more uORFs precede the initiation codon of the main coding region. Indeed, several studies have revealed that almost half of human transcripts present uORFs. Very interesting examples have shown that these uORFs can impact gene expression of the downstream main ORF by triggering mRNA decay or by regulating translation. Also, evidence from recent genetic and bioinformatic studies implicates disturbed uORF-mediated translational control in the etiology of many human diseases, including malignancies, metabolic or neurologic disorders, and inherited syndromes. In this review, we will briefly present the mechanisms through which uORFs regulate gene expression and how they can impact on the organism's response to different cell stress conditions. Then, we will emphasize the importance of these structures by illustrating, with specific examples, how disturbed uORF-mediated translational control can be involved in the etiology of human diseases, giving special importance to genotype-phenotype correlations. Identifying and studying more cases of uORF-altering mutations will help us to understand and establish genotype-phenotype associations, leading to advancements in diagnosis, prognosis, and treatment of many human disorders.


Zdroje

1. MorrisDR, GeballeAP (2000) Upstream open reading frames as regulators of mRNA translation. Mol Cell Biol 20: 8635–8642.

2. CalvoSE, PagliariniDJ, MoothaVK (2009) Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans. Proc Natl Acad Sci U S A 106: 7507–7512.

3. MendellJT, SharifiNA, MeyersJL, Martinez-MurilloF, DietzHC (2004) Nonsense surveillance regulates expression of diverse classes of mammalian transcripts and mutes genomic noise. Nat Genet 36: 1073–1078.

4. WittmannJ, HolEM, JäckH-M (2006) hUPF2 silencing identifies physiologic substrates of mammalian nonsense-mediated mRNA decay. Mol Cell Biol 26: 1272–1287.

5. YepiskoposyanH, AeschimannF, NilssonD, OkoniewskiM, MuhlemannO (2011) Autoregulation of the nonsense-mediated mRNA decay pathway in human cells. RNA 17: 2108–2118.

6. SpriggsKA, BushellM, WillisAE (2010) Translational regulation of gene expression during conditions of cell stress. Mol Cell 40: 228–237.

7. RogersGWJr, EdelmanGM, MauroVP (2004) Differential utilization of upstream AUGs in the beta-secretase mRNA suggests that a shunting mechanism regulates translation. Proc Natl Acad Sci U S A 101: 2794–2799.

8. LammichS, SchöbelS, ZimmerAK, LichtenthalerSF, HaassC (2004) Expression of the Alzheimer protease BACE1 is suppressed via its 5′-untranslated region. EMBO Rep 5: 620–625.

9. SuzukiY, IshiharaD, SasakiM, NakagawaH, HataH, et al. (2000) Statistical analysis of the 5′ untranslated region of human mRNA using “Oligo-Capped” cDNA libraries. Genomics 64: 286–297.

10. IaconoM, MignoneF, PesoleG (2005) uAUG and uORFs in human and rodent 5′ untranslated mRNAs. Gene 349: 97–105.

11. KochetovAV, AhmadS, IvanisenkoV, VolkovaOA, KolchanovNA, et al. (2008) uORFs, reinitiation and alternative translation start sites in human mRNAs. FEBS Lett 582: 1293–1297.

12. SathirapongsasutiJF, SathiraN, SuzukiY, HuttenhowerC, SuganoS (2011) Ultraconserved cDNA segments in the human transcriptome exhibit resistance to folding and implicate function in translation and alternative splicing. Nucleic Acids Res 39: 1967–1979.

13. KozakM (1987) An analysis of 5′-noncoding sequences from 699 vertebrate messenger RNAs. Nucleic Acids Res 15: 8125–4128.

14. KozakM (1991) An analysis of vertebrate mRNA sequences: intimations of translational control. J Cell Biol 115: 887–903.

15. MorrisDR (1995) Growth control of translation in mammalian cells. Prog Nucleic Acid Res Mol Biol 51: 339–363.

16. MatsuiM, YachieN, OkadaY, SaitoR, TomitaM (2007) Bioinformatic analysis of post-transcriptional regulation by uORF in human and mouse. FEBS Lett 581: 4184–4188.

17. LivingstoneM, AtasE, MellerA, SonenbergN (2010) Mechanisms governing the control of mRNA translation. Phys Biol 7: 021001.

18. SonenbergN, HinnebuschAG (2009) Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136: 731–745.

19. KozakM (1999) Initiation of translation in prokaryotes and eukaryotes. Gene 234: 187–208.

20. GebauerF, HentzeMW (2004) Molecular mechanisms of translational control. Nat Rev Mol Cell Biol 5: 827–835.

21. SachsMS, GeballeAP (2006) Downstream control of upstream open reading frames. Genes Dev 20: 915–921.

22. KozakM (1986) Point mutations define a sequence flanking the AUG initiator codon that modulates translation by eukaryotic ribosomes. Cell 44: 283–292.

23. KozakM (2002) Pushing the limits of the scanning mechanism for initiation of translation. Gene 299: 1–34.

24. MeijerHA, ThomasAAM (2002) Control of eukaryotic protein synthesis by upstream open reading frames in the 5′-untranslated region of an mRNA. Biochem J 367: 1–11.

25. PoyryTAA (2004) What determines whether mammalian ribosomes resume scanning after translation of a short upstream open reading frame? Genes Dev 18: 62–75.

26. ChildSJ, MillerMK, GeballeAP (1999) Translational control by an upstream open reading frame in the HER-2/neu transcript. J Biol Chem 274: 24335–24341.

27. RoyB, VaughnJN, KimB-H, ZhouF, GilchristMA, et al. (2010) The h subunit of eIF3 promotes reinitiation competence during translation of mRNAs harboring upstream open reading frames. RNA 16: 748–761.

28. KozakM (2005) Regulation of translation via mRNA structure in prokaryotes and eukaryotes. Gene 361: 13–37.

29. MunzarováV, PánekJ, GunišováS, DányiI, SzameczB, et al. (2011) Translation reinitiation relies on the interaction between eIF3a/TIF32 and progressively folded cis-acting mRNA elements preceding short uORFs. PLoS Genet 7: e1002137 doi:10.1371/journal.pgen.1002137

30. LovettPS, RogersEJ (1996) Ribosome regulation by the nascent peptide. Microbiol Mol Biol Rev 60: 366–385.

31. GeballeAP, MorrisDR (1994) Initiation codons within 5′-leaders of mRNAs as regulators of translation. Trends Biochem Sci 19: 159–164.

32. KaragyozovL, GodfreyR, BöhmerSA, PetermannA, HöltersS, et al. (2008) The structure of the 5′-end of the protein-tyrosine phosphatase PTPRJ mRNA reveals a novel mechanism for translation attenuation. Nucleic Acids Res 36: 4443–4453.

33. RaneyA, LawGL, MizeGJ, MorrisDR (2002) Regulated translation termination at the upstream open reading frame in S-adenosylmethionine decarboxylase mRNA. J Biol Chem 277: 5988–5994.

34. HaydenC, JorgensenR (2007) Identification of novel conserved peptide uORF homology groups in Arabidopsis and rice reveals ancient eukaryotic origin of select groups and preferential association with transcription factor-encoding genes. BMC Biol 5: 32.

35. MaquatLE, KinniburghAJ, RachmilewitzEA, RossJ (1981) Unstable beta-globin mRNA in mRNA-deficient beta° thalassemia. Cell 27: 543–553.

36. RehwinkelJ, RaesJ, IzaurraldeE (2006) Nonsense-mediated mRNA decay: target genes and functional diversification of effectors. Trends Biochem Sci 31: 639–646.

37. HoshinoS, ImaiM, KobayashiT, UchidaN, KatadaT (1999) The eukaryotic polypeptide chain releasing factor (eRF3/GSPT) carrying the translation termination signal to the 3′-Poly(A) tail of mRNA. Direct association of erf3/GSPT with polyadenylate-binding protein. J Biol Chem 274: 16677–16680.

38. AmraniN, GanesanR, KervestinS, MangusDA, GhoshS, et al. (2004) A faux 3′-UTR promotes aberrant termination and triggers nonsense-mediated mRNA decay. Nature 432: 112–118.

39. Behm-AnsmantI, IzaurraldeE (2006) Quality control of gene expression: a stepwise assembly pathway for the surveillance complex that triggers nonsense-mediated mRNA decay. Genes Dev 20: 391–398.

40. SinghG, RebbapragadaI, Lykke-AndersenJ (2008) A competition between stimulators and antagonists of Upf complex recruitment governs human nonsense-mediated mRNA decay. PLoS Biol 6: e111 doi:10.1371/journal.pbio.0060111

41. MühlemannO (2008) Recognition of nonsense mRNA: towards a unified model. Biochem Soc Trans 36: 497–501.

42. ShyuA-B, WilkinsonMF, van HoofA (2008) Messenger RNA regulation: to translate or to degrade. EMBO J 27: 471–481.

43. SilvaAL, RomãoL (2009) The mammalian nonsense-mediated mRNA decay pathway: to decay or not to decay! Which players make the decision? FEBS Lett 583: 499–505.

44. NagyE, MaquatLE (1998) A rule for termination-codon position within intron-containing genes: when nonsense affects RNA abundance. Trends Biochem Sci 23: 198–199.

45. Le HirH, IzaurraldeE, MaquatLE, MooreMJ (2000) The spliceosome deposits multiple proteins 20–24 nucleotides upstream of mRNA exon-exon junctions. EMBO J 19: 6860–6869.

46. StalderL, MühlemannO (2008) The meaning of nonsense. Trends Cell Biol 18: 315–321.

47. RomãoL, InácioA, SantosS, AvilaM, FaustinoP, et al. (2000) Nonsense mutations in the human beta-globin gene lead to unexpected levels of cytoplasmic mRNA accumulation. Blood 96: 2895–2901.

48. InácioA, SilvaAL, PintoJ, JiX, MorgadoA, et al. (2004) Nonsense mutations in close proximity to the initiation codon fail to trigger full nonsense-mediated mRNA decay. J Biol Chem 279: 32170–32180.

49. SilvaAL, PereiraFJC, MorgadoA, KongJ, MartinsR, et al. (2006) The canonical UPF1-dependent nonsense-mediated mRNA decay is inhibited in transcripts carrying a short open reading frame independent of sequence context. RNA 12: 2160–2170.

50. SilvaAL, RibeiroP, InácioA, LiebhaberSA, RomãoL (2008) Proximity of the poly(A)-binding protein to a premature termination codon inhibits mammalian nonsense-mediated mRNA decay. RNA 14: 563–576.

51. PeixeiroI, InácioÂ, BarbosaC, SilvaAL, LiebhaberSA, et al. (2012) Interaction of PABPC1 with the translation initiation complex is critical to the NMD resistance of AUG-proximal nonsense mutations. Nucleic Acids Res 40: 1160–1173.

52. McGlincyNJ, TanL-Y, PaulN, ZavolanM, LilleyKS, et al. (2010) Expression proteomics of UPF1 knockdown in HeLa cells reveals autoregulation of hnRNP A2/B1 mediated by alternative splicing resulting in nonsense-mediated mRNA decay. BMC Genomics 11: 565.

53. ZhaoC, DattaS, MandalP, XuS, HamiltonT (2010) Stress-sensitive regulation of IFRD1 mRNA decay is mediated by an upstream open reading frame. J Biol Chem 285: 8552–8562.

54. DaviesWL, VandenbergJI, SayeedRA, TreziseAE (2004) Post-transcriptional regulation of the cystic fibrosis gene in cardiac development and hypertrophy. Biochem Biophys Res Comm 319: 410–418.

55. NyikóT, SonkolyB, MéraiZ, BenkovicsAH, SilhavyD (2009) Plant upstream ORFs can trigger nonsense-mediated mRNA decay in a size-dependent manner. Plant Mol Biol 71: 367–378.

56. GardnerLB (2008) Hypoxic inhibition of nonsense-mediated RNA decay regulates gene expression and the integrated stress response. Mol Cell Biol 28: 3729–3741.

57. Hinnebusch AG, Dever TE, Asano K (2007) Mechanisms of translation initiation in the yeast Saccharomyces cerevisiae. In: Mathews M, Sonenberg N, Hershey JWB, editors. Translational Control in Biology and Medicine. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. pp 225–268.

58. Raught B, Gingras AC (2007) Signaling to translation initiation. In: Mathews M, Sonenberg N, Hershey JWB, editors. Translational Control in Biology and Medicine. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. pp 369–400.

59. MarcotrigianoJ, GingrasAC, SonenbergN, BurleySK (1999) Cap-dependent translation initiation in eukaryotes is regulated by a molecular mimic of eIF4G. Mol Cell 3: 707–716.

60. HoodHM, NeafseyDE, GalaganJ, SachsMS (2009) Evolutionary roles of upstream open reading frames in mediating gene regulation in fungi. Annu Rev Microbiol 63: 85–409.

61. MuellerPP, HinnebuschAG (1986) Multiple upstream AUG codons mediate translational control of GCN4. Cell 45: 201–207.

62. HinnebuschAG (2005) Translational regulation of GCN4 and the general amino acid control of yeast. Annu Rev Microbiol 59: 407–450.

63. LewerenzJ, SatoH, AlbrechtP, HenkeN, NoackR, et al. (2011) Mutation of ATF4 mediates resistance of neuronal cell lines against oxidative stress by inducing xCT expression. Cell Death Differ 19: 847–858.

64. BlaisJD, FilipenkoV, BiM, HardingHP, RonD, et al. (2004) Activating transcription factor 4 is translationally regulated by hypoxic stress. Mol Cell Biol 24: 7469–7482.

65. Ron D, Harding HP (2007) eIF2α phosphorylation in cellular stress responses and disease. In: Mathews M, Sonenberg N, Hershey JWB, editors. Translational Control in Biology and Medicine. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. pp 345–368.

66. PaliiSS, KaysCE, DevalC, BruhatA, FafournouxP, et al. (2009) Specificity of amino acid regulated gene expression: analysis of genes subjected to either complete or single amino acid deprivation. Amino Acids 37: 79–88.

67. WatataniY, IchikawaK, NakanishiN, FujimotoM, TakedaH, et al. (2008) Stress-induced translation of ATF5 mRNA is regulated by the 5′-untranslated region. J Biol Chem 283: 2543–2553.

68. HansenMB, MitchelmoreC, KjaerulffKM, RasmussenTE, PedersenKM, et al. (2002) Mouse Atf5: molecular cloning of two novel mRNAs, genomic organization, and odorant sensory neuron localization. Genomics 80: 344–350.

69. LohseI, ReillyP, ZauggK (2011) The CPT1C 5′UTR contains a repressing upstream open reading frame that is regulated by cellular energy availability and AMPK. PLoS ONE 6: e21486 doi:10.1371/journal.pone.0021486

70. PalamLR, BairdTD, WekRC (2011) Phosphorylation of eIF2 facilitates ribosomal bypass of an inhibitory upstream ORF to enhance CHOP translation. J Biol Chem 286: 10939–10949.

71. ChenY-J, TanBC-M, ChengY-Y, ChenJ-S, LeeS-C (2010) Differential regulation of CHOP translation by phosphorylated eIF4E under stress conditions. Nucleic Acids Res 38: 764–777.

72. LeeY-Y, CevallosRC, JanE (2008) An upstream open reading frame regulates translation of GADD34 during cellular stresses that induce eIF2 phosphorylation. J Biol Chem 284: 6661–6673.

73. O'ConnorT, SadleirKR, MausE, VelliquetteRA, ZhaoJ, et al. (2008) Phosphorylation of the translation initiation factor eIF2α increases BACE1 levels and promotes amyloidogenesis. Neuron 60: 988–1009.

74. Mouton-LigerF, PaquetC, DumurgierJ, BourasC, PradierL, et al. (2012) Oxidative stress increases BACE1 protein levels through activation of the PKR-eIF2α pathway. Biochim Biophys Acta 1822: 885–896.

75. Raveh-AmitH, MaisselA, PollerJ, MaromL, Elroy-SteinO, et al. (2009) Translational control of protein kinase C by two upstream open reading frames. Mol Cell Biol 29: 6140–6148.

76. van den BeuckenT, MagagninMG, SavelkoulsK, LambinP, KoritzinskyM, et al. (2007) Regulation of Cited2 expression provides a functional link between translational and transcriptional responses during hypoxia. Radiother Oncol 83: 346–352.

77. BastideA, KaraaZ, BornesS, HieblotC, LacazetteE, et al. (2008) An upstream open reading frame within an IRES controls expression of a specific VEGF-A isoform. Nucleic Acids Res 3: 2434.

78. GopfertU, KullmannM, HengstL (2003) Cell cycle-dependent translation of p27 involves a responsive element in its 5′-UTR that overlaps with a uORF. Hum Mol Genet 12: 1767–1779.

79. ParkE-H, LeeJM, BlaisJD, BellJC, PelletierJ (2005) Internal translation initiation mediated by the angiogenic factor Tie2. J Biol Chem 280: 20945–20953.

80. GrobeK, EskoJD (2002) Regulated translation of heparan sulfate N-acetylglucosamine N-deacetylase/N-sulfotransferase isozymes by structured 5′-untranslated regions and internal ribosome entry sites. J Biol Chem 277: 30699–30706.

81. FernandezJ, YamanI, SarnowP, SniderMD, HatzoglouM (2002) Regulation of internal ribosomal entry site-mediated translation by phosphorylation of the translation initiation factor eIF2alpha. J Biol Chem 277: 19198–19205.

82. YamanI, FernandezJ, LiuH, CapraraM, KomarAA, et al. (2003) The zipper model of translational control: a small upstream ORF is the switch that controls structural remodeling of an mRNA leader. Cell 113: 519–531.

83. ÖrdT, ÖrdD, KõivomägiM, JuhkamK, ÖrdT (2009) Human TRB3 is upregulated in stressed cells by the induction of translationally efficient mRNA containing a truncated 5′-UTR. Gene 444: 24–32.

84. KoschmiederS, D'AloF, RadomskaH, SchoneichC, ChangJS, et al. (2007) CDDO induces granulocytic differentiation of myeloid leukemic blasts through translational up-regulation of p42 CCAAT enhancer binding protein alpha. Blood 110: 3695–3705.

85. ReA, HamelN, FuC, BushH, McCaffreyT, et al. (2001) A link between diabetes and atherosclerosis: glucose regulates expression of CD36 at the level of translation. Nat Med 7: 840–846.

86. MehtaA (2006) Derepression of the Her-2 uORF is mediated by a novel post-transcriptional control mechanism in cancer cells. Genes Dev 20: 939–953.

87. SherryST, WardMH, KholodovM, BakerJ, PhanL, et al. (2001) dbSNP: the NCBI database of genetic variation. Nucleic Acids Res 29: 308–311.

88. BersanoA, BallabioE, BresolinN, CandeliseL (2008) Genetic polymorphisms for the study of multifactorial stroke. Hum Mutat 29: 776–795.

89. BachJ, EndlerG, WinkelmannBR, BoehmBO, MaerzW, et al. (2008) Coagulation factor XII (FXII) activity, activated FXII, distribution of FXII C46T gene polymorphism and coronary risk. J Thromb Haemost 6: 291–296.

90. KanajiT, OkamuraT, OsakiK, KuroiwaM, ShimodaK, et al. (1998) A common genetic polymorphism (46 C to T substitution) in the 5′-untranslated region of the coagulation factor XII gene is associated with low translation efficiency and decrease in plasma factor XII level. Blood 91: 2010–2014.

91. OnerR, AgarwalS, DimovskiAJ, EfremovGD, PetkovGH, et al. (1991) The G to A mutation at position +22 3′ to the Cap site of the beta-globin gene as a possible cause for a beta-thalassemia. Hemoglobin 15: 67–76.

92. KondoS, SchutteBC, RichardsonRJ, BjorkBC, KnightAS, et al. (2002) Mutations in IRF6 cause Van der Woude and popliteal pterygium syndromes. Nat Genet 32: 285–289.

93. PoulatF, DesclozeauxM, TufferyS, JayP, BoizetB, et al. (1998) Mutation in the 5′ noncoding region of the SRY gene in an XY sex-reversed patient. Hum Mutat Suppl 1: S192–194.

94. WittH, LuckW, HenniesHC, ClassenM, KageA, et al. (2000) Mutations in the gene encoding the serine protease inhibitor, Kazal type 1 are associated with chronic pancreatitis. Nat Genet 25: 213–216.

95. LiuL, DilworthD, GaoL, MonzonJ, SummersA, et al. (1999) Mutation of the CDKN2A 5′ UTR creates an aberrant initiation codon and predisposes to melanoma. Nat Genet 21: 128–132.

96. BisioA, NastiS, JordanJJ, GargiuloS, PastorinoL, et al. (2010) Functional analysis of CDKN2A/p16INK4a 5′-UTR variants predisposing to melanoma. Hum Mol Genet 19: 1479–1491.

97. SözenMM, WhittallR, OnerC, TokatliA, KalkanoğluHS, et al. (2005) The molecular basis of familial hypercholesterolaemia in Turkish patients. Atherosclerosis 180: 63–71.

98. LukowskiSW, BombieriC, TreziseAEO (2011) Disrupted posttranscriptional regulation of the cystic fibrosis transmembrane conductance regulator (CFTR) by a 5′UTR mutation is associated with a CFTR-related disease. Hum Mutat 32: e2266–e2282.

99. HuopioH, JääskeläinenJ, KomulainenJ, MiettinenR, KärkkäinenP, et al. (2002) Acute insulin response tests for the differential diagnosis of congenital hyperinsulinism. J Clin Endocrinol Metab 87: 4502–4507.

100. BravermanN, ChenL, LinP, ObieC, SteelG, et al. (2002) Mutation analysis of PEX7 in 60 probands with rhizomelic chondrodysplasia punctata and functional correlations of genotype with phenotype. Hum Mutat 20: 284–297.

101. KrudeH, BiebermannH, LuckW, HornR, BrabantG, et al. (1998) Severe early-onset obesity, adrenal insufficiency and red hair pigmentation caused by POMC mutations in humans. Nat Genet 19: 155–157.

102. TassinJ, DürrA, BonnetAM, GilR, VidailhetM, et al. (2000) Levodopa-responsive dystonia. GTP cyclohydrolase I or parkin mutations? Brain 123: 1112–1121.

103. RideauA, MangeatB, MatthesT, TronoD, BerisP (2007) Molecular mechanism of hepcidin deficiency in a patient with juvenile hemochromatosis. Haematologica 92: 127–128.

104. WenY, LiuY, XuY, ZhaoY, HuaR, et al. (2009) Loss-of-function mutations of an inhibitory upstream ORF in the human hairless transcript cause Marie Unna hereditary hypotrichosis. Nat Genet 41: 228–233.

105. BaekIC, KimJK, ChoK-H, ChaD-S, ChoJ-W, et al. (2009) A novel mutation in Hr causes abnormal hair follicle morphogenesis in hairpoor mouse, an animal model for Marie Unna Hereditary Hypotrichosis. Mamm Genome 20: 350–358.

106. CazzolaM, SkodaRC (2000) Translational pathophysiology: a novel molecular mechanism of human disease. Blood 95: 3280–3288.

107. KikuchiM, TayamaT, HayakawaH, TakahashiI, HoshinoH, et al. (1995) Familial thrombocytosis. Br J Haematol 89: 900–902.

108. GhilardiN, SkodaRC (1999) A single-base deletion in the thrombopoietin (TPO) gene causes familial essential thrombocythemia through a mechanism of more efficient translation of TPO mRNA. Blood 94: 1480–1482.

109. GhilardiN, WiestnerA, KikuchiM, OhsakaA, SkodaRC (1999) Hereditary thrombocythaemia in a Japanese family is caused by a novel point mutation in the thrombopoietin gene. Br J Haematol 107: 310–316.

110. WiestnerA, SchlemperRJ, van der MaasAP, SkodaRC (1998) An activating splice donor mutation in the thrombopoietin gene causes hereditary thrombocythaemia. Nat Genet 18: 49–52.

111. KondoT, OkabeM, SanadaM, KurosawaM, SuzukiS, et al. (1998) Familial essential thrombocythemia associated with one-base deletion in the 5′-untranslated region of the thrombopoietin gene. Blood 92: 1091–1096.

112. SivagnanasundaramS, MorrisAG, GaitondeEJ, McKennaPJ, MollonJD, et al. (2000) A cluster of single nucleotide polymorphisms in the 5′-leader of the human dopamine D3 receptor gene (DRD3) and its relationship to schizophrenia. Neurosci Lett 279: 13–16.

113. PasajeCF, BaeJS, ParkBL, CheongHS, KimJH, et al. (2012) HWDR46 is a genetic risk factor for aspirin-exacerbated respiratory disease in a Korean population. Allergy Asthma Immunol Res 4: 199–205.

114. BeffagnaG, OcchiG, NavaA, VitielloL, DitadiA, et al. (2005) Regulatory mutations in transforming growth factor-beta3 gene cause arrhythmogenic right ventricular cardiomyopathy type 1. Cardiovasc Res 65: 366–373.

115. NieslerB, FlohrT, NöthenMM, FischerC, RietschelM, et al. (2001) Association between the 5′ UTR variant C178T of the serotonin receptor gene HTR3A and bipolar affective disorder. Pharmacogenetics 11: 471–475.

116. WethmarK, SminkJJ, LeutzA (2010) Upstream open reading frames: Molecular switches in (patho)physiology. Bioessays 32: 885–893.

117. WethmarK, BegayV, SminkJJ, ZaragozaK, WiesenthalV, et al. (2010) C/EBP uORF mice: a genetic model for uORF-mediated translational control in mammals. Genes Dev 24: 15–20.

118. OlinerJD, KinzlerKW, MeltzerPS, GeorgeDL, VogelsteinB (1992) Amplification of a gene encoding a p53-associated protein in human sarcomas. Nature 358: 80–83.

119. BrownCY, MizeGJ, PinedaM, GeorgeDL, MorrisDR (1999) Role of two upstream open reading frames in the translational control of oncogene mdm2. Oncogene 18: 5631–5637.

120. ZhouW, SongW (2006) Leaky scanning and reinitiation regulate BACE1 gene expression. Mol Cell Biol 26: 3353–3364.

121. MihailovichM, ThermannR, GrohovazF, HentzeMW, ZacchettiD (2007) Complex translational regulation of BACE1 involves upstream AUGs and stimulatory elements within the 5′ untranslated region. Nucleic Acids Res 35: 2975–2985.

Štítky
Genetika Reprodukčná medicína

Článok vyšiel v časopise

PLOS Genetics


2013 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#