• 検索結果がありません。

参考⽂献

ドキュメント内 Kyushu University Institutional Repository (ページ 92-105)

1. Lander, E. S. et al. Initial sequencing and analysis of the human genome. Nature 409, 860–

921 (2001).

2. Venter, J. C. et al. The sequence of the human genome. Science 291, 1304–1351 (2001).

3. The ENCODE Project Consortium. The ENCODE (ENCyclopedia Of DNA Elements) Project. Science 306, 636–640 (2004).

4. Bernstein, B. E., Meissner, A. & Lander, E. S. The mammalian epigenome. Cell 128, 669–

681 (2007).

5. Kundaje, A. et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015).

6. Kawai, J. et al. Functional annotation of a full-length mouse cDNA collection. Nature 409, 685–690 (2001).

7. Park, P. J. ChIP–seq: advantages and challenges of a maturing technology. Nat. Rev. Genet.

10, 669–680 (2009).

8. Song, L. & Crawford, G. E. DNase-seq: a high-resolution technique for mapping active gene regulatory elements across the genome from mammalian cells. Cold Spring Harb.

Protoc. 2010, pdb.prot5384 (2010).

9. John, S. et al. Genome-scale mapping of DNase I hypersensitivity. Curr. Protoc. Mol. Biol.

103, 21.27.1–21.27.20 (2013).

10. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

11. Kanamori-Katayama, M. et al. Unamplified cap analysis of gene expression on a single-molecule sequencer. Genome Res. 21, 1150–1159 (2011).

12. Forrest, A. R. R. et al. A promoter-level mammalian expression atlas. Nature 507, 462

16. Albrecht, F., List, M., Bock, C. & Lengauer, T. DeepBlue epigenomic data server:

programmatic data retrieval and analysis of epigenome region sets. Nucleic Acids Res. 44, W581–W586 (2016).

17. Oki, S. et al. Atlas: a data-mining suite powered by full integration of public ChIP-seq data. EMBO Rep. 19, e46255 (2018).

18. Chèneby, J. et al. ReMap 2020: a database of regulatory regions from an integrative analysis of Human and Arabidopsis DNA-binding sequencing experiments. Nucleic Acids Res. 48, D180–D188 (2019).

19. Lieberman-Aiden, E. et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–293 (2009).

20. Fullwood, M. J. & Ruan, Y. ChIP-based methods for the identification of long-range chromatin interactions. J. Cell. Biochem. 107, 30–39 (2009).

21. Fullwood, M. J. et al. An oestrogen-receptor-alpha-bound human chromatin interactome.

Nature 462, 58–64 (2009).

22. Dixon, J. R. et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380 (2012).

23. Rao, S. S. P. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

24. Pomerantz, M. M. et al. The 8q24 cancer risk variant rs6983267 shows long-range interaction with MYC in colorectal cancer. Nat. Genet. 41, 882–884 (2009).

25. Tuupanen, S. et al. The common colorectal cancer predisposition SNP rs6983267 at chromosome 8q24 confers potential to enhanced Wnt signaling. Nat. Genet. 41, 885–890 (2009).

26. Musunuru, K. et al. From noncoding variant to phenotype via SORT1 at the 1p13 cholesterol locus. Nature 466, 714–719 (2010).

27. Wright, J. B., Brown, S. J. & Cole, M. D. Upregulation of c-MYC in cis through a large chromatin loop linked to a cancer risk-associated single-nucleotide polymorphism in colorectal cancer cells. Mol. Cell. Biol. 30, 1411–1420 (2010).

28. Harismendy, O. et al. 9p21 DNA variants associated with coronary artery disease impair interferon-γ signalling response. Nature 470, 264–268 (2011).

29. Visser, M., Kayser, M. & Palstra, R.-J. HERC2 rs12913832 modulates human pigmentation by attenuating chromatin-loop formation between a long-range enhancer and the OCA2 promoter. Genome Res. 22, 446–455 (2012).

30. Praetorius, C. et al. A polymorphism in IRF4 affects human pigmentation through a tyrosinase-dependent MITF/TFAP2A pathway. Cell 155, 1022–1033 (2013).

31. Guenther, C. A., Tasic, B., Luo, L., Bedell, M. A. & Kingsley, D. M. A molecular basis for classic blond hair color in Europeans. Nat. Genet. 46, 748–752 (2014).

32. Smemo, S. et al. Obesity-associated variants within FTO form long-range functional connections with IRX3. Nature 507, 371–375 (2014).

33. Zhang, Y. et al. Integrative genomic analysis predicts causative cis-regulatory mechanisms of the breast cancer-associated genetic variant rs4415084. Cancer Res. 78, 1579–1591 (2018).

34. Wang, Y. et al. The 3D Genome Browser: a web-based browser for visualizing 3D genome organization and long-range chromatin interactions. Genome Biol. 19, 151 (2018).

35. Tam, V. et al. Benefits and limitations of genome-wide association studies. Nat. Rev. Genet.

20, 467–484 (2019).

36. Gibbs, R. A. et al. The international HapMap project. Nature 426, 789–796 (2003).

37. Ozaki, K. et al. Functional SNPs in the lymphotoxin-α gene that are associated with susceptibility to myocardial infarction. Nat. Genet. 32, 650–654 (2002).

38. MacArthur, J. et al. The new NHGRI-EBI Catalog of published genome-wide association studies (GWAS Catalog). Nucleic Acids Res. 45, D896–D901 (2017).

39. Meenan, F. O. C. A note on the history of psoriasis. Ir. J. Med. Sci. 30, 141–142 (1955).

40. Mueller, W. & Herrmann, B. Cyclosporin A for psoriasis. N. Engl. J. Med. 301, 555–556 (1979).

41. Gottlieb, S. L. et al. Response of psoriasis to a lymphocyte-selective toxin (DAB389IL-2) suggests a primary immune, but not keratinocyte, pathogenic basis. Nat. Med. 1, 442–447 (1995).

42. Sommer, D. M., Jenisch, S., Suchan, M., Christophers, E. & Weichenthal, M. Increased prevalence of the metabolic syndrome in patients with moderate to severe psoriasis. Arch.

46. Mantovani, A., Gisondi, P., Lonardo, A. & Targher, G. Relationship between non-alcoholic fatty liver disease and psoriasis: a novel hepato-dermal axis? Int. J. Mol. Sci. 17, 217 (2016).

47. Cao, H. Adipocytokines in obesity and metabolic disease. J. Endocrinol. 220, T47–T59 (2014).

48. Anstee, Q. M., Targher, G. & Day, C. P. Progression of NAFLD to diabetes mellitus, cardiovascular disease or cirrhosis. Nat. Rev. Gastroenterol. Hepatol. 10, 330–44 (2013).

49. Stefan, N. & Häring, H.-U. The role of hepatokines in metabolism. Nat. Rev. Endocrinol. 9, 144–152 (2013).

50. Byrne, C. D. & Targher, G. Ectopic fat, insulin resistance, and nonalcoholic fatty liver disease. Arterioscler. Thromb. Vasc. Biol. 34, 1155–1161 (2014).

51. Byrne, C. D. & Targher, G. NAFLD: A multisystem disease. J. Hepatol. 62, S47–S64 (2015).

52. Chandran, V. & Raychaudhuri, S. P. Geoepidemiology and environmental factors of psoriasis and psoriatic arthritis. J. Autoimmun. 34, J314–J321 (2010).

53. Parisi, R., Symmons, D. P. M., Griffiths, C. E. M., Ashcroft, D. M. & team, on behalf of the Identification and Management of Psoriasis and Associated ComorbidiTy (IMPACT) project. Global epidemiology of psoriasis: a systematic review of incidence and prevalence.

J. Invest. Dermatol. 133, 377–385 (2013).

54. Capon, F. et al. Identification of ZNF313/RNF114 as a novel psoriasis susceptibility gene.

Hum. Mol. Genet. 17, 1938–1945 (2008).

55. Liu, Y. et al. A genome-wide association study of psoriasis and psoriatic arthritis identifies new disease loci. PLoS Genet. 4, e1000041 (2008).

56. Zhang, X. J. et al. Psoriasis genome-wide association study identifies susceptibility variants within LCE gene cluster at 1q21. Nat. Genet. 41, 205–210 (2009).

57. Nair, R. P. et al. Genome-wide scan reveals association of psoriasis with IL-23 and NF-κB pathways. Nat. Genet. 41, 199–204 (2009).

58. Stuart, P. E. et al. Genome-wide association analysis identifies three psoriasis susceptibility loci. Nat. Genet. 42, 1000–1004 (2010).

59. Ellinghaus, E. et al. Genome-wide association study identifies a psoriasis susceptibility locus at TRAF3IP2. Nat. Genet. 42, 991–995 (2010).

60. Strange, A. et al. A genome-wide association study identifies new psoriasis susceptibility loci and an interaction between HLA-C and ERAP1. Nat. Genet. 42, 985 (2010).

61. Tsoi, L. C. et al. Identification of 15 new psoriasis susceptibility loci highlights the role of innate immunity. Nat. Genet. 44, 1341–1348 (2012).

62. Baurecht, H. et al. Genome-wide comparative analysis of atopic dermatitis and psoriasis gives insight into opposing genetic mechanisms. Am. J. Hum. Genet. 96, 104–120 (2015).

63. Yin, X. et al. Genome-wide meta-analysis identifies multiple novel associations and ethnic heterogeneity of psoriasis susceptibility. Nat. Commun. 6, 6916 (2015).

64. Tsoi, L. C. et al. Enhanced meta-analysis and replication studies identify five new psoriasis susceptibility loci. Nat. Commun. 6, 7001 (2015).

65. Stuart, P. E. et al. Genome-wide association analysis of psoriatic arthritis and cutaneous psoriasis reveals differences in their genetic architecture. Am. J. Hum. Genet. 97, 816–836 (2015).

66. Tsoi, L. C. et al. Large scale meta-analysis characterizes genetic architecture for common psoriasis associated variants. Nat. Commun. 8, 15382 (2017).

67. Hirata, J. et al. Variants at HLA-A, HLA-C, and HLA-DQB1 confer risk of psoriasis vulgaris in Japanese. J. Invest. Dermatol. 138, 542–548 (2018).

68. Farh, K. K. H. et al. Genetic and epigenetic fine mapping of causal autoimmune disease variants. Nature 518, 337–343 (2015).

69. Machiela, M. J. & Chanock, S. J. LDlink: a web-based application for exploring population-specific haplotype structure and linking correlated alleles of possible functional variants.

Bioinformatics 31, 3555–3557 (2015).

70. Auton, A. et al. A global reference for human genetic variation. Nature 526, 68 (2015).

71. Hill, W. G. & Robertson, A. Linkage disequilibrium in finite populations. TAG Theor. Appl.

Genet. Theor. Angew. Genet. 38, 226–231 (1968).

72. Carlson, C. S. et al. Selecting a maximally informative set of single-nucleotide polymorphisms for association analyses using linkage disequilibrium. Am. J. Hum. Genet.

74, 106–120 (2004).

77. Aguet, F. et al. Genetic effects on gene expression across human tissues. Nature 550, 204–

213 (2017).

78. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

79. Kumar, S., Ambrosini, G. & Bucher, P. SNP2TFBS – a database of regulatory SNPs affecting predicted transcription factor binding site affinity. Nucleic Acids Res. 45, D139–

D144 (2017).

80. Zhou, Y. et al. Metascape provides a biologist-oriented resource for the analysis of systems-level datasets. Nat. Commun. 10, 1523 (2019).

81. Keenan, A. B. et al. ChEA3: transcription factor enrichment analysis by orthogonal omics integration. Nucleic Acids Res. 47, W212–W224 (2019).

82. Lachmann, A. et al. Massive mining of publicly available RNA-seq data from human and mouse. Nat. Commun. 9, 1366 (2018).

83. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B Methodol. 57, 289–300 (1995).

84. Haeussler, M. et al. The UCSC Genome Browser database: 2019 update. Nucleic Acids Res.

47, D853–D858 (2019).

85. Ernst, J. & Kellis, M. ChromHMM: automating chromatin-state discovery and characterization. Nat. Methods 9, 215–216 (2012).

86. Ernst, J. & Kellis, M. Large-scale imputation of epigenomic datasets for systematic annotation of diverse human tissues. Nat. Biotechnol. 33, 364–376 (2015).

87. Ernst, J. & Kellis, M. Chromatin-state discovery and genome annotation with ChromHMM.

Nat. Protoc. 12, 2478–2492 (2017).

88. Khan, A. et al. JASPAR 2018: update of the open-access database of transcription factor binding profiles and its web framework. Nucleic Acids Res. 46, D260–D266 (2018).

89. Crooks, G. E., Hon, G., Chandonia, J.-M. & Brenner, S. E. WebLogo: A sequence logo generator. Genome Res. 14, 1188–1190 (2004).

90. Yang, D. et al. 3DIV: A 3D-genome Interaction Viewer and database. Nucleic Acids Res.

46, D52–57 (2018).

91. Zhang, Y. et al. Model-based Analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

92. Torre, D., Lachmann, A. & Ma’ayan, A. BioJupies: Automated generation of interactive notebooks for RNA-Seq data analysis in the cloud. Cell Syst. 7, 556–561.e3 (2018).

93. Bray, N. L., Pimentel, H., Melsted, P. & Pachter, L. Near-optimal probabilistic RNA-seq quantification. Nat. Biotechnol. 34, 525–527 (2016).

94. Clark, N. R. et al. The characteristic direction: a geometrical approach to identify differentially expressed genes. BMC Bioinformatics 15, 79 (2014).

95. Berge, T. et al. Quantitative proteomic analyses of CD4+ and CD8+ T cells reveal differentially expressed proteins in multiple sclerosis patients and healthy controls. Clin.

Proteomics 16, 19 (2019).

96. Zhu, K. J., Zhu, C. Y., Shi, G. & Fan, Y. M. Association of IL23R polymorphisms with psoriasis and psoriatic arthritis: a meta-analysis. Inflamm. Res. Off. J. Eur. Histamine Res.

Soc. Al 61, 1149–1154 (2012).

97. Wain, L. V. et al. Genome-wide association analyses for lung function and chronic obstructive pulmonary disease identify new loci and potential druggable targets. Nat. Genet.

49, 416–425 (2017).

98. Bowes, J. et al. Dense genotyping of immune-related susceptibility loci reveals new insights into the genetics of psoriatic arthritis. Nat. Commun. 6, 6046 (2015).

99. Nititham, J. et al. Psoriasis risk SNPs and their association with HIV-1 control. Hum.

Immunol. 78, 179–184 (2016).

100. Weidinger, S. et al. A genome-wide association study of atopic dermatitis identifies loci with overlapping effects on asthma and psoriasis. Hum. Mol. Genet. 22, 4841–4856 (2013).

101. Enerbäck, C. et al. The psoriasis-protective TYK2 I684S variant impairs IL-12 stimulated pSTAT4 response in skin-homing CD4+ and CD8+ memory T-cells. Sci. Rep. 8, 7043 (2018).

102. Parmar, A. S. et al. Association study of FUT2 (rs601338) with celiac disease and inflammatory bowel disease in the Finnish population. Tissue Antigens 80, 488–493 (2012).

103. Azad, M. B., Wade, K. H. & Timpson, N. J. FUT2 secretor genotype and susceptibility to infections and chronic conditions in the ALSPAC cohort. Wellcome Open Res. 3, 65 (2018).

106. Banaganapalli, B. et al. Comprehensive computational analysis of GWAS loci identifies CCR2 as a candidate gene for celiac disease pathogenesis. J. Cell. Biochem. 118, 2193–

2207 (2017).

107. Neville, M. D. C., Choi, J., Lieberman, J. & Duan, Q. L. Identification of deleterious and regulatory genomic variations in known asthma loci. Respir. Res. 19, 248 (2018).

108. Lykke-Andersen, S. & Jensen, T. H. Nonsense-mediated mRNA decay: an intricate machinery that shapes transcriptomes. Nat. Rev. Mol. Cell Biol. 16, 665–677 (2015).

109. Andrés, A. M. et al. Balancing selection maintains a form of ERAP2 that undergoes nonsense-mediated decay and affects antigen presentation. PLoS Genet. 6, e1001157 (2010).

110. Maurano, M. T. et al. Large-scale identification of sequence variants influencing human transcription factor occupancy in vivo. Nat. Genet. 47, 1393–1401 (2015).

111. Di Meglio, P. et al. Activation of the aryl hydrocarbon receptor dampens the severity of inflammatory skin conditions. Immunity 40, 989–1001 (2014).

112. Gupta, R. et al. Landscape of long noncoding RNAs in psoriatic and healthy skin. J. Invest.

Dermatol. 136, 603–609 (2016).

113. Eriksen, K. W. et al. Increased sensitivity to interferon-α in psoriatic T cells. J. Invest.

Dermatol. 125, 936–944 (2005).

114. van der Fits, L., van der Wel, L. I., Laman, J. D., Prens, E. P. & Verschuren, M. C. M. In psoriasis lesional skin the type I interferon signaling pathway is activated, whereas interferon-α sensitivity is unaltered. J. Invest. Dermatol. 122, 51–60 (2004).

115. Andrés, R. M., Hald, A., Johansen, C., Kragballe, K. & Iversen, L. Studies of Jak/STAT3 expression and signalling in psoriasis identifies STAT3-Ser727 phosphorylation as a modulator of transcriptional activity. Exp. Dermatol. 22, 323–328 (2013).

116. Hald, A. et al. STAT1 expression and activation is increased in lesional psoriatic skin. Br.

J. Dermatol. 168, 302–310 (2013).

117. Johansen, C. et al. STAT2 is involved in the pathogenesis of psoriasis by promoting CXCL11 and CCL5 production by keratinocytes. PLOS ONE 12, e0176994 (2017).

118. Langrish, C. L. et al. IL-23 drives a pathogenic T cell population that induces autoimmune inflammation. J. Exp. Med. 201, 233–240 (2005).

119. Bettelli, E. et al. Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature 441, 235–238 (2006).

120. Tsui, H. S. et al. Human COQ10A and COQ10B are distinct lipid-binding START domain proteins required for coenzyme Q function. J. Lipid Res. 60, 1293–1310 (2019).

121. Lee, S.-Y. et al. Coenzyme Q10 inhibits Th17 and STAT3 signaling pathways to ameliorate colitis in mice. J. Med. Food 20, 821–829 (2017).

122. Folkers, K., Morita, M. & Mcree, J. The activities of coenzyme Q10 and vitamin B6 for immune responses. Biochem. Biophys. Res. Commun. 193, 88–92 (1993).

123. Farough, S. et al. Coenzyme Q10 and immunity: A case report and new implications for treatment of recurrent infections in metabolic diseases. Clin. Immunol. 155, 209–212 (2014).

124. Wootton, J. C. Non-globular domains in protein sequences: Automated segmentation using complexity measures. Comput. Chem. 18, 269–285 (1994).

125. Martin, E. W. & Mittag, T. Relationship of sequence and phase separation in protein low-complexity regions. Biochemistry 57, 2478–2487 (2018).

126. Murthy, A. C. et al. Molecular interactions underlying liquid−liquid phase separation of the FUS low-complexity domain. Nat. Struct. Mol. Biol. 26, 637–648 (2019).

127. Xue, S. et al. Low-complexity domain of U1-70K modulates phase separation and aggregation through distinctive basic-acidic motifs. Sci. Adv. 5, eaax5349 (2019).

128. Fujimura, T. et al. Increased expression of tripartite motif (TRIM) 47 is a negative prognostic predictor in human prostate cancer. Clin. Genitourin. Cancer 14, 298–303 (2016).

129. Han, Y., Tian, H., Chen, P. & Lin, Q. TRIM47 overexpression is a poor prognostic factor and contributes to carcinogenesis in non-small cell lung carcinoma. Oncotarget 8, 22730–

22740 (2017).

130. Liang, Q. et al. TRIM47 is up-regulated in colorectal cancer, promoting ubiquitination and degradation of SMAD4. J. Exp. Clin. Cancer Res. 38, 159 (2019).

131. Ozato, K., Shin, D. M., Chang, T. H. & III, H. C. M. TRIM family proteins and their emerging roles in innate immunity. Nat. Rev. Immunol. 8, 849–860 (2008).

132. Saveanu, L. et al. Concerted peptide trimming by human ERAP1 and ERAP2 aminopeptidase complexes in the endoplasmic reticulum. Nat. Immunol. 6, 689–697 (2005).

136. Jia, X. et al. Imputing amino acid polymorphisms in human leukocyte antigens. PloS One 8, e64683 (2013).

137. Raychaudhuri, S. et al. Five amino acids in three HLA proteins explain most of the association between MHC and seropositive rheumatoid arthritis. Nat. Genet. 44, 291–296 (2012).

138. Woolf, E. et al. Runx3 and Runx1 are required for CD8 T cell development during thymopoiesis. Proc. Natl. Acad. Sci. 100, 7731–7736 (2003).

139. Lin, J. X. & Leonard, W. J. The role of Stat5a and Stat5b in signaling by IL-2 family cytokines. Oncogene 19, 2566–2576 (2000).

140. Wu, L. et al. A novel IL-25 signaling pathway through STAT5. J. Immunol. 194, 4528–

4534 (2015).

141. Kulski, J. K. & Dawkins, R. L. The P5 multicopy gene family in the MHC is related in sequence to human endogenous retroviruses HERV-L and HERV-16. Immunogenetics 49, 404–412 (1999).

142. Kulski, J. K. Long noncoding RNA HCP5, a hybrid HLA class I endogenous retroviral gene: structure, expression, and disease associations. Cells 8, 480 (2019).

143. Zhang, X. et al. Inhibition of the EGF receptor by binding of MIG6 to an activating kinase domain interface. Nature 450, 741–744 (2007).

144. Hopkins, S. et al. Mig6 is a sensor of EGF receptor inactivation that directly activates c-Abl to induce apoptosis during epithelial homeostasis. Dev. Cell 23, 547–559 (2012).

145. Park, E. et al. Structure and mechanism of activity-based inhibition of the EGF receptor by Mig6. Nat. Struct. Mol. Biol. 22, 703–711 (2015).

146. Ferby, I. et al. Mig6 is a negative regulator of EGF receptor–mediated skin morphogenesis and tumor formation. Nat. Med. 12, 568–573 (2006).

147. Ku, B. J. et al. Mig-6 plays a critical role in the regulation of cholesterol homeostasis and bile acid synthesis. PLoS ONE 7, e42915 (2012).

148. Park, B. et al. Fatty liver and insulin resistance in the liver-specific knockout mice of mitogen inducible gene-6. J. Diabetes Res. 2016, 1632061 (2016).

149. Yoo, J. et al. Role of Mig-6 in hepatic glucose metabolism. J. Diabetes 8, 86–97 (2016).

150. Jin, N. et al. Mig-6 is required for appropriate lung development and to ensure normal adult lung homeostasis. Development 136, 3347–3356 (2009).

151. Yoo, J. Y. et al. MIG-6 suppresses endometrial epithelial cell proliferation by inhibiting phospho-AKT. BMC Cancer 18, 605 (2018).

152. Li, Z. et al. Downregulation of Mig-6 in nonsmall-cell lung cancer is associated with EGFR signaling. Mol. Carcinog. 51, 522–534 (2012).

153. Maity, T. K. et al. Loss of MIG6 accelerates initiation and progression of mutant epidermal growth factor receptor–driven lung adenocarcinoma. Cancer Discov. 5, 534–549 (2015).

154. Liu, J. et al. Mig-6 deficiency cooperates with oncogenic Kras to promote mouse lung tumorigenesis. Lung Cancer 112, 47–56 (2017).

155. Duncan, C. G. et al. Integrated genomic analyses identify ERRFI1 and TACC3 as glioblastoma-targeted genes. Oncotarget 1, 265–277 (2010).

156. Ying, H. et al. Mig-6 controls EGFR trafficking and suppresses gliomagenesis. Proc. Natl.

Acad. Sci. 107, 6912–6917 (2010).

157. Reschke, M. et al. Mitogen-inducible gene-6 is a negative regulator of epidermal growth factor receptor signaling in hepatocytes and human hepatocellular carcinoma. Hepatology 51, 1383–1390 (2010).

158. Li, Z. et al. Mig-6 is down-regulated in HCC and inhibits the proliferation of HCC cells via the P-ERK/Cyclin D1 pathway. Exp. Mol. Pathol. 102, 492–499 (2017).

159. Lee, J. et al. Mig-6 gene knockout induces neointimal hyperplasia in the vascular smooth muscle cell. Dis. Markers 2014, 549054 (2014).

160. Chang, X. et al. The relative expression of Mig6 and EGFR is associated with resistance to EGFR kinase inhibitors. PLoS ONE 8, e68966 (2013).

161. Papp, K. et al. Phase 2 trial of selective tyrosine kinase 2 inhibition in psoriasis. N. Engl. J.

Med. 379, 1313–1321 (2018).

162. Huang, D., Petrykowska, H. M., Miller, B. F., Elnitski, L. & Ovcharenko, I. Identification of human silencers by correlating cross-tissue epigenetic profiles and gene expression.

Genome Res. 29, 657–667 (2019).

163. Pang, B. & Snyder, M. P. Systematic identification of silencers in human cells. Nat. Genet.

52, 254–263 (2020).

167. Trieu, T., Martinez-Fundichely, A. & Khurana, E. DeepMILO: a deep learning approach to predict the impact of non-coding sequence variants on 3D chromatin structure. Genome Biol.

21, 79 (2020).

168. Wang, Z. & Burge, C. B. Splicing regulation: from a parts list of regulatory elements to an integrated splicing code. RNA N. Y. N 14, 802–813 (2008).

169. Supek, F., Miñana, B., Valcárcel, J., Gabaldón, T. & Lehner, B. Synonymous mutations frequently act as driver mutations in human cancers. Cell 156, 1324–1335 (2014).

170. Ikemura, T. Correlation between the abundance of yeast transfer RNAs and the occurrence of the respective codons in protein genes: Differences in synonymous codon choice patterns of yeast and Escherichia coli with reference to the abundance of isoaccepting transfer RNAs.

J. Mol. Biol. 158, 573–597 (1982).

171. Dana, A. & Tuller, T. The effect of tRNA levels on decoding times of mRNA codons.

Nucleic Acids Res. 42, 9171–9181 (2014).

172. Presnyak, V. et al. Codon optimality is a major determinant of mRNA stability. Cell 160, 1111–24 (2015).

173. Gingold, H. et al. A dual program for translation regulation in cellular proliferation and differentiation. Cell 158, 1281–92 (2014).

174. Goodarzi, H. et al. Modulated expression of specific tRNAs drives gene expression and cancer progression. Cell 165, 1416–1427 (2016).

175. Stergachis, A. B. et al. Exonic transcription factor binding directs codon choice and affects protein evolution. Science 342, 1367–1372 (2013).

176. Tumpel, S., Cambronero, F., Sims, C., Krumlauf, R. & Wiedemann, L. M. A regulatory module embedded in the coding region of Hoxa2 controls expression in rhombomere 2.

Proc. Natl. Acad. Sci. 105, 20077–20082 (2008).

177. Birnbaum, R. Y. et al. Coding exons function as tissue-specific enhancers of nearby genes.

Genome Res. 22, 1059–1068 (2012).

178. Mercer, T. R. et al. DNase I–hypersensitive exons colocalize with promoters and distal regulatory elements. Nat. Genet. 45, 852–859 (2013).

179. Ahituv, N. Exonic enhancers: proceed with caution in exome and genome sequencing studies. Genome Med. 8, 14 (2016).

180. Wang, K., Li, M. & Hakonarson, H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 38, e164 (2010).

181. Bartel, D. P. MicroRNAs: Target recognition and regulatory functions. Cell 136, 215–233 (2009).

182. Filipowicz, W., Bhattacharyya, S. N. & Sonenberg, N. Mechanisms of post-transcriptional regulation by microRNAs: are the answers in sight? Nat. Rev. Genet. 9, 102–114 (2008).

183. Jackson, R. J. Alternative mechanisms of initiating translation of mammalian mRNAs.

Biochem. Soc. Trans. 33, 1231–1241 (2005).

184. Deveci, M., Çatalyürek, Ü. V. & Toland, A. E. mrSNP: Software to detect SNP effects on microRNA binding. BMC Bioinformatics 15, 73 (2014).

185. Bhattacharya, A. & Cui, Y. miR2GO: comparative functional analysis for microRNAs.

Bioinforma. Oxf. Engl. 31, 2403–2405 (2015).

186. Ryan, B. C., Werner, T. S., Howard, P. L. & Chow, R. L. ImiRP: a computational approach to microRNA target site mutation. BMC Bioinformatics 17, 190 (2016).

ドキュメント内 Kyushu University Institutional Repository (ページ 92-105)