Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

The roles of RNA processing in translating genotype to phenotype

Subjects

Key Points

  • Gene expression involves a cascade of chromosomal, transcriptional and post-transcriptional events. Genomic variants produce differences between individuals in terms of transcriptional and post-transcriptional events at equal frequency.

  • Common genetic variation between individuals contributes to phenotypic diversity. Single-nucleotide variants can have functional consequences by affecting RNA processing, including effects on pre-mRNA splicing, 3′ end formation, and RNA stability, localization, structure and translation efficiency.

  • Genomic variants affect RNA processing by altering the binding sites of sequence-specific RNA-binding proteins, by disrupting RNA structures required for processing or by introducing RNA structures that prevent RNA processing.

  • Comparisons of genomic and transcriptomic sequences from a large number of individuals from diverse backgrounds will not only identify functional variants that are relevant to human health, but also catalogue the cis-acting elements that are required for basal and regulated RNA processing.

  • Analysis of the effects of genomic variants on post-transcriptional regulation of gene expression is made possible by the recent rapid expansion of high-throughput experimental approaches and associated computational analyses.

Abstract

A goal of human genetics studies is to determine the mechanisms by which genetic variation produces phenotypic differences that affect human health. Efforts in this respect have previously focused on genetic variants that affect mRNA levels by altering epigenetic and transcriptional regulation. Recent studies show that genetic variants that affect RNA processing are at least equally as common as, and are largely independent from, those variants that affect transcription. We highlight the impact of genetic variation on pre-mRNA splicing and polyadenylation, and on the stability, translation and structure of mRNAs as mechanisms that produce phenotypic traits. These results emphasize the importance of including RNA processing signals in analyses to identify functional variants.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Genetic variation alters gene output by affecting RNA processing.
Figure 2: Genetic variants that affect RNA processing can have a range of effects on human health.
Figure 3: Cis-acting splicing elements are not restricted to exon–intron boundaries.
Figure 4: Genetic variants that create or abolish key RNA secondary structures affect multiple aspects of post-transcriptional regulation.

Similar content being viewed by others

References

  1. Edwards, S. L., Beesley, J., French, J. D. & Dunning, A. M. Beyond GWASs: illuminating the dark road from association to function. Am. J. Hum. Genet. 93, 779–797 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Kwan, T. et al. Genome-wide analysis of transcript isoform variation in humans. Nat. Genet. 40, 225–231 (2008). One of the early genome-wide comparisons showed that genomic variation between individuals correlated with differences in mRNA levels and structure.

    Article  CAS  PubMed  Google Scholar 

  3. Li, Y. I. et al. RNA splicing is a primary link between genetic variation and disease. Science 352, 600–604 (2016). A systematic analysis of LCLs from 68 individuals identified genomic variants that correlate with differences in up to eight molecular features of the gene regulatory cascade.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Lappalainen, T. et al. Transcriptome and genome sequencing uncovers functional variation in humans. Nature 501, 506–511 (2013). This study identified the genomic causes of variation in the transcriptome using RNA-seq and genomic sequence data from LCLs of 462 individuals in the 1000 Genomes Project.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Wang, E. T. et al. Alternative isoform regulation in human tissue transcriptomes. Nature 456, 470–476 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Tian, B., Hu, J., Zhang, H. & Lutz, C. S. A large-scale analysis of mRNA polyadenylation of human and mouse genes. Nucleic Acids Res. 33, 201–212 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Derti, A. et al. A quantitative atlas of polyadenylation in five mammals. Genome Res. 22, 1173–1183 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Halbeisen, R. E., Galgano, A., Scherrer, T. & Gerber, A. P. Post-transcriptional gene regulation: from genome-wide studies to principles. Cell. Mol. Life Sci. 65, 798–813 (2008).

    Article  CAS  PubMed  Google Scholar 

  9. Cooper, T. A., Wan, L. & Dreyfuss, G. RNA and disease. Cell 136, 777–793 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Scotti, M. M. & Swanson, M. S. RNA mis-splicing in disease. Nat. Rev. Genet. 17, 19–32 (2016).

    Article  CAS  PubMed  Google Scholar 

  11. Lim, L. P. & Burge, C. B. A computational analysis of sequence features involved in recognition of short introns. Proc. Natl Acad. Sci. USA 98, 11193–11198 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wang, G.-S. & Cooper, T. A. Splicing in disease: disruption of the splicing code and the decoding machinery. Nat. Rev. Genet. 8, 749–761 (2007).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Xiong, H. Y. et al. RNA splicing. The human splicing code reveals new insights into the genetic determinants of disease. Science 347, 1254806 (2015). A computational model was developed to predict splicing patterns based on genomic elements, which identified a large number of genetic variants among more than 650,000 SNVs that affect splicing and are associated with human disease.

    Article  CAS  PubMed  Google Scholar 

  15. Busch, A. & Hertel, K. J. Splicing predictions reliably classify different types of alternative splicing. RNA 21, 813–823 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. 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).

    Article  CAS  PubMed  Google Scholar 

  17. Takata, A., Ionita-Laza, I., Gogos, J. A., Xu, B. & Karayiorgou, M. De novo synonymous mutations in regulatory elements contribute to the genetic etiology of autism and schizophrenia. Neuron 89, 940–947 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Bali, V. & Bebok, Z. Decoding mechanisms by which silent codon changes influence protein biogenesis and function. Int. J. Biochem. Cell Biol. 64, 58–74 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Lim, K. H., Ferraris, L., Filloux, M. E., Raphael, B. J. & Fairbrother, W. G. Using positional distribution to identify splicing elements and predict pre-mRNA processing defects in human genes. Proc. Natl Acad. Sci. USA 108, 11093–11098 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Cartegni, L., Chew, S. L. & Krainer, A. R. Listening to silence and understanding nonsense: exonic mutations that affect splicing. Nat. Rev. Genet. 3, 285–298 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Zemojtel, T. et al. Effective diagnosis of genetic disease by computational phenotype analysis of the disease-associated genome. Sci. Transl Med. 6, 252ra123 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Li, M.-X. et al. Predicting Mendelian disease-causing non-synonymous single nucleotide variants in exome sequencing studies. PLoS Genet. 9, e1003143 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Cargill, M. et al. Characterization of single-nucleotide polymorphisms in coding regions of human genes. Nat. Genet. 22, 231–238 (1999).

    Article  CAS  PubMed  Google Scholar 

  24. Wu, X. & Hurst, L. D. Determinants of the usage of splice-associated cis-motifs predict the distribution of human pathogenic SNPs. Mol. Biol. Evol. 33, 518–529 (2016).

    Article  CAS  PubMed  Google Scholar 

  25. Zhang, X. et al. Identification of common genetic variants controlling transcript isoform variation in human whole blood. Nat. Genet. 47, 345–352 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ulirsch, J. C. et al. Systematic functional dissection of common genetic variation affecting red blood cell traits. Cell 165, 1530–1545 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Fraser, H. B. & Xie, X. Common polymorphic transcript variation in human disease. Genome Res. 19, 567–575 (2009).

    Article  CAS  PubMed  Google Scholar 

  28. Zhang, W. et al. Identification of common genetic variants that account for transcript isoform variation between human populations. Hum. Genet. 125, 81–93 (2009).

    Article  CAS  PubMed  Google Scholar 

  29. ElSharawy, A. et al. Systematic evaluation of the effect of common SNPs on pre-mRNA splicing. Hum. Mutat. 30, 625–632 (2009).

    Article  CAS  PubMed  Google Scholar 

  30. Rivas, M. A. et al. Human genomics. Effect of predicted protein-truncating genetic variants on the human transcriptome. Science 348, 666–669 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Hsiao, Y.-H. E. et al. Alternative splicing modulated by genetic variants demonstrates accelerated evolution regulated by highly conserved proteins. Genome Res. 26, 440–450 (2016). The authors identified more than 600 exons for which the splicing is affected by SNVs using a computational approach to measure allele-specific gene expression, including splicing differences, in ENCODE RNA-seq data sets. The results demonstrate the prominence of variant-driven splicing differences and present a powerful approach to analyse allele-specific expression.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Matesanz, F. et al. A functional variant that affects exon-skipping and protein expression of SP140 as genetic mechanism predisposing to multiple sclerosis. Hum. Mol. Genet. 24, 5619–5627 (2015).

    Article  CAS  PubMed  Google Scholar 

  33. Paraboschi, E. M. et al. Functional variations modulating PRKCA expression and alternative splicing predispose to multiple sclerosis. Hum. Mol. Genet. 23, 6746–6761 (2014).

    Article  CAS  PubMed  Google Scholar 

  34. Bojesen, S. E. et al. Multiple independent variants at the TERT locus are associated with telomere length and risks of breast and ovarian cancer. Nat. Genet. 45, 371–384 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Di Giacomo, D. et al. Functional analysis of a large set of BRCA2 exon 7 variants highlights the predictive value of hexamer scores in detecting alterations of exonic splicing regulatory elements. Hum. Mutat. 34, 1547–1557 (2013).

    Article  CAS  PubMed  Google Scholar 

  36. Monlong, J., Calvo, M., Ferreira, P. G. & Guigó, R. Identification of genetic variants associated with alternative splicing using sQTLseekeR. Nat. Commun. 5, 4698 (2014).

    Article  CAS  PubMed  Google Scholar 

  37. Tejedor, J. R., Tilgner, H., Iannone, C., Guigó, R. & Valcárcel, J. Role of six single nucleotide polymorphisms, risk factors in coronary disease, in OLR1 alternative splicing. RNA 21, 1187–1202 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Gregory, S. G. et al. Interleukin 7 receptor α chain (IL7R) shows allelic and functional association with multiple sclerosis. Nat. Genet. 39, 1083–1091 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Gretarsdottir, S. et al. A splice region variant in LDLR lowers non-high density lipoprotein cholesterol and protects against coronary artery disease. PLoS Genet. 11, e1005379 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Shi, Y. & Manley, J. L. The end of the message: multiple protein-RNA interactions define the mRNA polyadenylation site. Genes Dev. 29, 889–897 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Mu, X. J., Lu, Z. J., Kong, Y., Lam, H. Y. K. & Gerstein, M. B. Analysis of genomic variation in non-coding elements using population-scale sequencing data from the 1000 Genomes Project. Nucleic Acids Res. 39, 7058–7076 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Lu, J. & Clark, A. G. Impact of microRNA regulation on variation in human gene expression. Genome Res. 22, 1243–1254 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Sheets, M. D., Ogg, S. C. & Wickens, M. P. Point mutations in AAUAAA and the poly (A) addition site: effects on the accuracy and efficiency of cleavage and polyadenylation in vitro. Nucleic Acids Res. 18, 5799–5805 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Higgs, D. R. et al. α-Thalassaemia caused by a polyadenylation signal mutation. Nature 306, 398–400 (1983).

    Article  CAS  PubMed  Google Scholar 

  45. Yoon, O. K., Hsu, T. Y., Im, J. H. & Brem, R. B. Genetics and regulatory impact of alternative polyadenylation in human B-lymphoblastoid cells. PLoS Genet. 8, e1002882 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Zhernakova, D. V. et al. DeepSAGE reveals genetic variants associated with alternative polyadenylation and expression of coding and non-coding transcripts. PLoS Genet. 9, e1003594 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Graham, R. R. et al. Three functional variants of IFN regulatory factor 5 (IRF5) define risk and protective haplotypes for human lupus. Proc. Natl Acad. Sci. USA 104, 6758–6763 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Cunninghame Graham, D. S. et al. Association of IRF5 in UK SLE families identifies a variant involved in polyadenylation. Hum. Mol. Genet. 16, 579–591 (2007).

    Article  CAS  PubMed  Google Scholar 

  49. Hellquist, A. et al. The human GIMAP5 gene has a common polyadenylation polymorphism increasing risk to systemic lupus erythematosus. J. Med. Genet. 44, 314–321 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Fahiminiya, S. et al. A polyadenylation site variant causes transcript-specific BMP1 deficiency and frequent fractures in children. Hum. Mol. Genet. 24, 516–524 (2015).

    Article  CAS  PubMed  Google Scholar 

  51. Garin, I. et al. Recessive mutations in the INS gene result in neonatal diabetes through reduced insulin biosynthesis. Proc. Natl Acad. Sci. USA 107, 3105–3110 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Stacey, S. N. et al. A germline variant in the TP53 polyadenylation signal confers cancer susceptibility. Nat. Genet. 43, 1098–1103 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Uitte de Willige, S., Rietveld, I. M., De Visser, M. C. H., Vos, H. L. & Bertina, R. M. Polymorphism 10034C>T is located in a region regulating polyadenylation of FGG transcripts and influences the fibrinogen γ′/γA mRNA ratio. J. Thromb. Haemost. 5, 1243–1249 (2007).

    Article  CAS  PubMed  Google Scholar 

  54. Friedman, R. C., Farh, K. K.-H., Burge, C. B. & Bartel, D. P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 19, 92–105 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. John, B. et al. Human microRNA targets. PLoS Biol. 2, e363 (2004).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Lewis, B. P., Shih, I.-H., Jones-Rhoades, M. W., Bartel, D. P. & Burge, C. B. Prediction of mammalian microRNA targets. Cell 115, 787–798 (2003).

    Article  CAS  PubMed  Google Scholar 

  57. Lewis, B. P., Burge, C. B. & Bartel, D. P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 120, 15–20 (2005).

    Article  CAS  PubMed  Google Scholar 

  58. Huan, T. et al. Genome-wide identification of microRNA expression quantitative trait loci. Nat. Commun. 6, 6601 (2015).

    Article  CAS  PubMed  Google Scholar 

  59. Wagschal, A. et al. Genome-wide identification of microRNAs regulating cholesterol and triglyceride homeostasis. Nat. Med. 21, 1290–1297 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Duan, J. et al. A rare functional noncoding variant at the GWAS-implicated MIR137/MIR2682 locus might confer risk to schizophrenia and bipolar disorder. Am. J. Hum. Genet. 95, 744–753 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Gamazon, E. R. et al. Genetic architecture of microRNA expression: implications for the transcriptome and complex traits. Am. J. Hum. Genet. 90, 1046–1063 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Ghanbari, M. et al. Genetic variations in microRNA-binding sites affect microRNA-mediated regulation of several genes associated with cardio-metabolic phenotypes. Circ. Cardiovasc. Genet. 8, 473–486 (2015).

    Article  CAS  PubMed  Google Scholar 

  63. Thomas, L. F. & Sætrom, P. Single nucleotide polymorphisms can create alternative polyadenylation signals and affect gene expression through loss of microRNA-regulation. PLoS Comput. Biol. 8, e1002621 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Arnold, M., Ellwanger, D. C., Hartsperger, M. L., Pfeufer, A. & Stümpflen, V. Cis-acting polymorphisms affect complex traits through modifications of microRNA regulation pathways. PLoS ONE 7, e36694 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Houlston, R. S. et al. Meta-analysis of three genome-wide association studies identifies susceptibility loci for colorectal cancer at 1q41, 3q26.2, 12q13.13 and 20q13.33. Nat. Genet. 42, 973–977 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Parkes, M. et al. Sequence variants in the autophagy gene IRGM and multiple other replicating loci contribute to Crohn's disease susceptibility. Nat. Genet. 39, 830–832 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Brest, P. et al. A synonymous variant in IRGM alters a binding site for miR-196 and causes deregulation of IRGM-dependent xenophagy in Crohn's disease. Nat. Genet. 43, 242–245 (2011).

    Article  CAS  PubMed  Google Scholar 

  68. Kulkarni, S. et al. Differential microRNA regulation of HLA-C expression and its association with HIV control. Nature 472, 495–498 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Caussy, C. et al. An APOA5 3′ UTR variant associated with plasma triglycerides triggers APOA5 downregulation by creating a functional miR-485-5p binding site. Am. J. Hum. Genet. 94, 129–134 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Gartner, J. J. et al. Whole-genome sequencing identifies a recurrent functional synonymous mutation in melanoma. Proc. Natl Acad. Sci. USA 110, 13481–13486 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Wang, G. et al. Variation in the miRNA-433 binding site of FGF20 confers risk for Parkinson disease by overexpression of α-synuclein. Am. J. Hum. Genet. 82, 283–289 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Haghnejad, L. et al. Variation in the miRNA-433 binding site of FGF20 is a risk factor for Parkinson's disease in Iranian population. J. Neurol. Sci. 355, 72–74 (2015).

    Article  CAS  PubMed  Google Scholar 

  73. Pai, A. A. et al. The contribution of RNA decay quantitative trait loci to inter-individual variation in steady-state gene expression levels. PLoS Genet. 8, e1003000 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Capon, F. et al. A synonymous SNP of the corneodesmosin gene leads to increased mRNA stability and demonstrates association with psoriasis across diverse ethnic groups. Hum. Mol. Genet. 13, 2361–2368 (2004).

    Article  CAS  PubMed  Google Scholar 

  75. van Hoof, A. & Wagner, E. J. A brief survey of mRNA surveillance. Trends Biochem. Sci. 36, 585–592 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Jacobs, E., Mills, J. D. & Janitz, M. The role of RNA structure in posttranscriptional regulation of gene expression. J. Genet. Genomics 39, 535–543 (2012).

    Article  CAS  PubMed  Google Scholar 

  77. Lu, Z. et al. RNA duplex map in living cells reveals higher-order transcriptome structure. Cell 165, 1267–1279 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Sharma, E., Sterne-Weiler, T., O'Hanlon, D. & Blencowe, B. J. Global mapping of human RNA–RNA interactions. Mol. Cell 62, 618–626 (2016).

    Article  CAS  PubMed  Google Scholar 

  79. Spitale, R. C. et al. Structural imprints in vivo decode RNA regulatory mechanisms. Nature 519, 486–490 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Wan, Y. et al. Landscape and variation of RNA secondary structure across the human transcriptome. Nature 505, 706–709 (2014). This study identified more than 1,900 SNVs that affected local RNA structure using high-throughput analysis of RNA structure in LCLs from a family trio.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Halvorsen, M., Martin, J. S., Broadaway, S. & Laederach, A. Disease-associated mutations that alter the RNA structural ensemble. PLoS Genet. 6, e1001074 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  82. Rogler, L. E. et al. Small RNAs derived from lncRNA RNase MRP have gene-silencing activity relevant to human cartilage–hair hypoplasia. Hum. Mol. Genet. 23, 368–382 (2014).

    Article  CAS  PubMed  Google Scholar 

  83. Martin, J. S. et al. Structural effects of linkage disequilibrium on the transcriptome. RNA 18, 77–87 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Kutchko, K. M. et al. Multiple conformations are a conserved and regulatory feature of the RB1 5′ UTR. RNA 21, 1274–1285 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Duan, J. et al. Synonymous mutations in the human dopamine receptor D2 (DRD2) affect mRNA stability and synthesis of the receptor. Hum. Mol. Genet. 12, 205–216 (2003).

    Article  CAS  PubMed  Google Scholar 

  86. Akdeli, N. et al. A 3′UTR polymorphism modulates mRNA stability of the oncogene and drug target polo-like kinase 1. Mol. Cancer 13, 87 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  87. Niu, T. et al. Identification of a novel FGFRL1 microRNA target site polymorphism for bone mineral density in meta-analyses of genome-wide association studies. Hum. Mol. Genet. 24, 4710–4727 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Sabarinathan, R. et al. Transcriptome-wide analysis of UTRs in non-small cell lung cancer reveals cancer-related genes with SNV-induced changes on RNA secondary structure and miRNA target sites. PLoS ONE 9, e82699 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Wu, L. et al. Variation and genetic control of protein abundance in humans. Nature 499, 79–82 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Battle, A. et al. Genomic variation. Impact of regulatory variation from RNA to protein. Science 347, 664–667 (2015).

    Article  CAS  PubMed  Google Scholar 

  91. Suhl, J. A. et al. A 3′ untranslated region variant in FMR1 eliminates neuronal activity-dependent translation of FMRP by disrupting binding of the RNA-binding protein HuR. Proc. Natl Acad. Sci. USA 112, E6553–E6561 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Somers, J. et al. A common polymorphism in the 5′ UTR of ERCC5 creates an upstream ORF that confers resistance to platinum-based chemotherapy. Genes Dev. 29, 1891–1896 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Raj, A. et al. Thousands of novel translated open reading frames in humans inferred by ribosome footprint profiling. eLife 5, e13328 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  94. Cenik, C. et al. Integrative analysis of RNA, translation, and protein levels reveals distinct regulatory variation across humans. Genome Res. 25, 1610–1621 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Zhang, F., Saha, S., Shabalina, S. A. & Kashina, A. Differential arginylation of actin isoforms is regulated by coding sequence-dependent degradation. Science 329, 1534–1537 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Mortimer, S. A., Kidwell, M. A. & Doudna, J. A. Insights into RNA structure and function from genome-wide studies. Nat. Rev. Genet. 15, 469–479 (2014).

    Article  CAS  PubMed  Google Scholar 

  97. Wen, J.-D. et al. Following translation by single ribosomes one codon at a time. Nature 452, 598–603 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Shabalina, S. A., Ogurtsov, A. Y. & Spiridonov, N. A. A periodic pattern of mRNA secondary structure created by the genetic code. Nucleic Acids Res. 34, 2428–2437 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Yu, C.-H. et al. Codon usage influences the local rate of translation elongation to regulate co-translational protein folding. Mol. Cell 59, 744–754 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Riordan, J. R. et al. Identification of the cystic fibrosis gene: cloning and characterization of complementary DNA. Science 245, 1066–1073 (1989).

    Article  CAS  PubMed  Google Scholar 

  101. Ward, C. L. & Kopito, R. R. Intracellular turnover of cystic fibrosis transmembrane conductance regulator. Inefficient processing and rapid degradation of wild-type and mutant proteins. J. Biol. Chem. 269, 25710–25718 (1994).

    CAS  PubMed  Google Scholar 

  102. Bartoszewski, R. A. et al. A synonymous single nucleotide polymorphism in ΔF508 CFTR alters the secondary structure of the mRNA and the expression of the mutant protein. J. Biol. Chem. 285, 28741–28748 (2010). This study proposes that the effects on CFTR mRNA secondary structure induced by the ΔF508 mutation, rather than the deletion of Phe, destabilize the protein, leading to loss of its function.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Nackley, A. G. et al. Human catechol-O- methyltransferase haplotypes modulate protein expression by altering mRNA secondary structure. Science 314, 1930–1933 (2006).

    Article  CAS  PubMed  Google Scholar 

  104. Wang, T., Zhou, T., Saadat, L. & Garcia, J. G. N. A MYLK variant regulates asthmatic inflammation via alterations in mRNA secondary structure. Eur. J. Hum. Genet. 23, 874–876 (2015).

    Article  PubMed  CAS  Google Scholar 

  105. Chiaruttini, C. et al. Dendritic trafficking of BDNF mRNA is mediated by translin and blocked by the G196A (Val66Met) mutation. Proc. Natl Acad. Sci. USA 106, 16481–16486 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Bergalet, J. & Lécuyer, E. The functions and regulatory principles of mRNA intracellular trafficking. Adv. Exp. Med. Biol. 825, 57–96 (2014).

    Article  CAS  PubMed  Google Scholar 

  107. Chabanon, H., Mickleburgh, I., Burtle, B., Pedder, C. & Hesketh, J. An AU-rich stem-loop structure is a critical feature of the perinuclear localization signal of c-myc mRNA. Biochem. J. 392, 475–483 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Chen, Z.-Y. et al. Genetic variant BDNF (Val66Met) polymorphism alters anxiety-related behavior. Science 314, 140–143 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Egan, M. F. et al. The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function. Cell 112, 257–269 (2003).

    Article  CAS  PubMed  Google Scholar 

  110. Notaras, M., Hill, R. & van den Buuse, M. The BDNF gene Val66Met polymorphism as a modifier of psychiatric disorder susceptibility: progress and controversy. Mol. Psychiatry 20, 916–930 (2015).

    Article  CAS  PubMed  Google Scholar 

  111. Mallei, A. et al. Expression and dendritic trafficking of BDNF-6 splice variant are impaired in knock-in mice carrying human BDNF Val66Met polymorphism. Int. J. Neuropsychopharmacol. 18, pyv069 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  112. Li, G. et al. Identification of allele-specific alternative mRNA processing via transcriptome sequencing. Nucleic Acids Res. 40, e104 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Maniatis, T. & Reed, R. An extensive network of coupling among gene expression machines. Nature 416, 499–506 (2002).

    Article  CAS  PubMed  Google Scholar 

  114. Braunschweig, U., Gueroussov, S., Plocik, A. M., Graveley, B. R. & Blencowe, B. J. Dynamic integration of splicing within gene regulatory pathways. Cell 152, 1252–1269 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Naftelberg, S., Schor, I. E., Ast, G. & Kornblihtt, A. R. Regulation of alternative splicing through coupling with transcription and chromatin structure. Annu. Rev. Biochem. 84, 165–198 (2015).

    Article  CAS  PubMed  Google Scholar 

  116. Mayr, C. & Bartel, D. P. Widespread shortening of 3′ UTRs by alternative cleavage and polyadenylation activates oncogenes in cancer cells. Cell 138, 673–684 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Taliaferro, J. M. et al. Distal alternative last exons localize mRNAs to neural projections. Mol. Cell 61, 821–833 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Sterne-Weiler, T. et al. Frac-seq reveals isoform-specific recruitment to polyribosomes. Genome Res. 23, 1615–1623 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Simms, C. L., Thomas, E. N. & Zaher, H. S. Ribosome-based quality control of mRNA and nascent peptides. Wiley Interdiscip. Rev. RNA http://dx.doi.org/10.1002/wrna.1366 (2016).

  120. Van Nostrand, E. L. et al. Robust transcriptome-wide discovery of RNA-binding protein binding sites with enhanced CLIP (eCLIP). Nat. Methods 13, 508–514 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. de Klerk, E. & 't Hoen, P. A. C. Alternative mRNA transcription, processing, and translation: insights from RNA sequencing. Trends Genet. 31, 128–139 (2015).

    Article  CAS  PubMed  Google Scholar 

  122. Weng, L., Li, Y., Xie, X. & Shi, Y. Poly(A) code analyses reveal key determinants for tissue-specific mRNA alternative polyadenylation. RNA 22, 813–821 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Brar, G. A. & Weissman, J. S. Ribosome profiling reveals the what, when, where and how of protein synthesis. Nat. Rev. Mol. Cell Biol. 16, 651–664 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. van der Maarel, S. M., Tawil, R. & Tapscott, S. J. Facioscapulohumeral muscular dystrophy and DUX4: breaking the silence. Trends Mol. Med. 17, 252–258 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Lemmers, R. J. L. F. et al. Digenic inheritance of an SMCHD1 mutation and an FSHD-permissive D4Z4 allele causes facioscapulohumeral muscular dystrophy type 2. Nat. Genet. 44, 1370–1374 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Lemmers, R. J. L. F. et al. A unifying genetic model for facioscapulohumeral muscular dystrophy. Science 329, 1650–1653 (2010). The authors demonstrate the role of allelic differences in a polyadenylation signal that allows expression of a toxic protein in muscle tissue of patients with FSHD. This followed a detailed genetic analysis to identify permissive and non-permissive alleles within complex genomic regions.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Spear, B. B., Heath-Chiozzi, M. & Huff, J. Clinical application of pharmacogenetics. Trends Mol. Med. 7, 201–204 (2001).

    Article  CAS  PubMed  Google Scholar 

  128. Tate, S. K. et al. Genetic predictors of the maximum doses patients receive during clinical use of the anti-epileptic drugs carbamazepine and phenytoin. Proc. Natl Acad. Sci. USA 102, 5507–5512 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Burkhardt, R. et al. Common SNPs in HMGCR in Micronesians and whites associated with LDL-cholesterol levels affect alternative splicing of exon13. Arterioscler. Thromb. Vasc. Biol. 28, 2078–2084 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Medina, M. W., Gao, F., Ruan, W., Rotter, J. I. & Krauss, R. M. Alternative splicing of 3-hydroxy-3-methylglutaryl coenzyme A reductase is associated with plasma low-density lipoprotein cholesterol response to simvastatin. Circulation 118, 355–362 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Yu, C.-Y. et al. HNRNPA1 regulates HMGCR alternative splicing and modulates cellular cholesterol metabolism. Hum. Mol. Genet. 23, 319–332 (2014).

    Article  CAS  PubMed  Google Scholar 

  132. Corrado, L. et al. A novel synonymous mutation in the MPZ gene causing an aberrant splicing pattern and Charcot–Marie–Tooth disease type 1b. Neuromuscul. Disord. 26, 516–520 (2016).

    Article  CAS  PubMed  Google Scholar 

  133. Llewellyn, D. H. et al. Acute intermittent porphyria caused by defective splicing of porphobilinogen deaminase RNA: a synonymous codon mutation at -22 bp from the 5′ splice site causes skipping of exon 3. J. Med. Genet. 33, 437–438 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Thanopoulou, E. et al. The single nucleotide polymorphism g.1548A>G (K469E) of the ICAM-1 gene is associated with worse prognosis in non-small cell lung cancer. Tumour Biol. 33, 1429–1436 (2012).

    Article  CAS  PubMed  Google Scholar 

  135. Boni, V. et al. Role of primary miRNA polymorphic variants in metastatic colon cancer patients treated with 5-fluorouracil and irinotecan. Pharmacogenomics J. 11, 429–436 (2010).

    Article  PubMed  CAS  Google Scholar 

  136. Xu, J. et al. A heroin addiction severity-associated intronic single nucleotide polymorphism modulates alternative pre-mRNA splicing of the μ opioid receptor gene OPRM1 via hnRNPH Interactions. J. Neurosci. 34, 11048–11066 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  137. Chiba-Falek, O. et al. The molecular basis of disease variability among cystic fibrosis patients carrying the 3849 + 10 kb C>T mutation. Genomics 53, 276–283 (1998).

    Article  CAS  PubMed  Google Scholar 

  138. Hinzpeter, A. et al. Alternative splicing at a NAGNAG acceptor site as a novel phenotype modifier. PLoS Genet. 6, e1001153 (2010). This study demonstrates that selection of the upstream AG in the NAGNAG sequence at the 3′ splice site introduces a termination codon in the CFTR mRNA, thereby modifying disease severity.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  139. Morini, E. et al. The human rs1050286 polymorphism alters LOX-1 expression through modifying miR-24 binding. J. Cell. Mol. Med. 20, 181–187 (2016).

    Article  CAS  PubMed  Google Scholar 

  140. Zhou, B. et al. Identification of a splicing variant that regulates type 2 diabetes risk factor CDKAL1 level by a coding-independent mechanism in human. Hum. Mol. Genet. 23, 4639–4650 (2014).

    Article  CAS  PubMed  Google Scholar 

  141. Ray, D. et al. A compendium of RNA-binding motifs for decoding gene regulation. Nature 499, 172–177 (2013). This study identified the preferred binding motifs of hundreds of RNA-binding proteins with broad implications regarding roles in post-transcriptional regulation. A high-throughput approach was used to interrogate more than 200 RNA-binding protein motifs from vertebrate and invertebrate organisms using a randomized pool of 30–41-nucleotide RNAs containing all possible 9-nucleotide combinations.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Stephens, R. M. & Schneider, T. D. Features of spliceosome evolution and function inferred from an analysis of the information at human splice sites. J. Mol. Biol. 228, 1124–1136 (1992).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank members of the Cooper laboratory as well as the reviewers for valuable input. The work of the authors is supported by the US National Institutes of Health (NIH) (R01AR045653, R01AR060733 and R01HL045565 to T.A.C.), the Muscular Dystrophy Association (MDA276796 to T.A.C.), NIH training grant T32 GM008231 (to K.S.M.) and Baylor Research Advocates for Student Scientists (to K.S.M.).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas A. Cooper.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Related links

Related links

FURTHER INFORMATION

ENCODE

GEUVADIS

GTEx

Glossary

Genome-wide association studies

(GWAS). Large studies across many individuals to determine whether the presence of a specific genotype is correlated with the manifestation of natural and pathological phenotypic traits or diseases.

mRNA 3′ ends

The 3′ end of an mRNA is formed by endonucleolytic cleavage. The vast majority of protein-coding mRNAs have approximately 250 non-templated adenosines added to the last templated nucleotide at the site of endonucleolytic cleavage.

Polyadenylation sites

Technically refers to the site of addition of the poly(A) tail at the last templated nucleotide of an mRNA. However, the term is sometimes used to refer to the AAUAAA hexanucleotide motif that is required for polyadenylation and that is typically located within 25 nucleotides of the last templated nucleotide.

5′ and 3′ untranslated regions

(5′ and 3′ UTRs). The open reading frames of protein-coding mRNAs can constitute less than half of their nucleotide sequence. The 5′ UTR is the mRNA region upstream of the translation start codon and the 3′ UTR is the region downstream of the translation stop codon.

MicroRNAs

(miRNAs). Small RNAs (about 22 nucleotides long) that base pair to their mRNA targets and reduce protein expression by mRNA destabilization or translational repression.

Branch site

A consensus mRNA splicing element typically located within 30 nucleotides of the 3′ splice site. It is recognized by sequential binding of branch point-binding protein followed by the U2 small nuclear ribonucleoprotein by base pairing.

Exonic splicing enhancer

(ESE). One of several auxiliary mRNA splicing elements that are bound by splicing factors to promote (ESE or intronic splicing enhancer (ISE)) or inhibit (exonic splicing silencer (ESS) or intronic splicing silencer (ISS)) splicing. Disruption of these motifs and the resulting effects on splicing indicate that genetic variants outside exon–intron junctions can affect gene output.

Synonymous and non-synonymous variants

Refers to variants in the coding region of a gene that either alter the encoded amino acid (non-synonymous) or do not affect the encoded amino acid (synonymous).

Splicing quantitative trait loci

(sQTLs). Genomic regions containing variants that affect the splicing patterns of the associated pre-mRNA.

Expression QTLs

(eQTLs). Genomic regions containing variants that affect the level of mRNA expression from one or more genes.

HapMap-derived lymphoblastoid cell lines

(HapMap-derived LCLs). Refers to the International HapMap Project, which derived LCLs from 270 individuals of European, African and Asian ancestry for analysis of genotype, gene expression and pharmacological response.

Single-nucleotide variants

(SNVs). We use the term SNV to refer to any genetic variant that changes one nucleotide, which may or may not have functional or pathological consequences. We largely focus on examples of common variants (also known as single-nucleotide polymorphisms (SNPs)) that are associated with either disease risk or disease severity.

GEUVADIS and GTEx

Genetic European Variation in Health and Disease (GEUVADIS) and Genotype Tissue Expression (GTEx) are consortia for the high-throughput sequencing of human genomes and transcriptomes.

Crosslinking immunoprecipitation

(CLIP). Used to identify the direct targets of an RNA-binding protein by covalently linking RNA and the protein in vivo by UV crosslinking, followed by immunoprecipitation and then high-throughput sequencing.

Linkage disequilibrium

The nonrandom association of two or more genetic variants, usually owing to close proximity on the same chromosome and reflecting shared ancestry of alleles at flanking loci.

Cis-acting miRNA QTLs

(Cis-mirQTLs). Genomic regions containing variants that affect the level of expression of individual microRNAs (miRNAs) located in cis with the variant.

1000 Genomes Project

This database was designed to identify genetic variants present in at least 1% of individuals across 26 populations, and contains DNA sequencing data for more than 2,500 individuals. This data set has been expanded by the addition of RNA sequencing data from the GEUVADIS consortium.

Nonsense-mediated decay

An RNA surveillance mechanism that recognizes and degrades RNA transcripts containing premature termination codons.

Nonstop decay

An RNA surveillance mechanism that recognizes and degrades RNA transcripts lacking a stop codon.

No-go decay

An RNA and translation quality control mechanism that recognizes and degrades RNA -transcripts containing stalled ribosomes.

RNA-induced silencing complex

(RISC). A ribonucleoprotein complex that mediates binding between a microRNA and its target mRNA.

Protein QTLs

Genomic regions containing variants that affect the expression level of protein from one or more genes.

Kozak sequence

A consensus sequence surrounding the start codon in eukaryotic mRNAs that promotes translation initiation.

Endoplasmic reticulum-associated degradation pathway

(ERAD pathway). A quality control system for misfolded proteins in the endoplasmic reticulum that marks them for degradation by the ubiquitin–proteasome system.

RNA zipcode motifs

Regulatory cis-acting RNA elements containing a specific sequence and secondary structure that are sufficient to confer mRNA subcellular localization.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Manning, K., Cooper, T. The roles of RNA processing in translating genotype to phenotype. Nat Rev Mol Cell Biol 18, 102–114 (2017). https://doi.org/10.1038/nrm.2016.139

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm.2016.139

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing