1932

Abstract

Evolutionary conservation has been an accurate predictor of functional elements across the first decade of metazoan genomics. More recently, there has been a move to define functional elements instead from biochemical annotations. Evolutionary methods are, however, more comprehensive than biochemical approaches can be and can assess quantitatively, especially for subtle effects, how biologically important—how injurious after mutation—different types of elements are. Evolutionary methods are thus critical for understanding the large fraction (up to 10%) of the human genome that does not encode proteins and yet might convey function. These methods can also capture the ephemeral nature of much noncoding functional sequence, with large numbers of functional elements having been gained and lost rapidly along each mammalian lineage. Here, we review how different strengths of purifying selection have impacted on protein-coding and non-protein-coding loci and on transcription factor binding sites in mammalian and fruit fly genomes.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-genom-090413-025621
2014-08-31
2024-05-05
Loading full text...

Full text loading...

/deliver/fulltext/genom/15/1/annurev-genom-090413-025621.html?itemId=/content/journals/10.1146/annurev-genom-090413-025621&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 1000 Genomes Proj. Consort 2010. A map of human genome variation from population-scale sequencing. Nature 467:1061–73 [Google Scholar]
  2. 2. 1000 Genomes Proj. Consort 2012. An integrated map of genetic variation from 1,092 human genomes. Nature 491:56–65 [Google Scholar]
  3. Ahituv N, Zhu Y, Visel A, Holt A, Afzal V. 3.  et al. 2007. Deletion of ultraconserved elements yields viable mice. PLoS Biol. 5:e234 [Google Scholar]
  4. Andolfatto P. 4.  2008. Controlling type-I error of the McDonald–Kreitman test in genomewide scans for selection on noncoding DNA. Genetics 180:1767–71 [Google Scholar]
  5. Arbiza L, Gronau I, Aksoy BA, Hubisz MJ, Gulko B. 5.  et al. 2013. Genome-wide inference of natural selection on human transcription factor binding sites. Nat. Genet. 45:723–29 [Google Scholar]
  6. Asthana S, Noble WS, Kryukov G, Grant CE, Sunyaev S, Stamatoyannopoulos JA. 6.  2007. Widely distributed noncoding purifying selection in the human genome. Proc. Natl. Acad. Sci. USA 104:12410–15 [Google Scholar]
  7. Barreiro LB, Laval G, Quach H, Patin E, Quintana-Murci L. 7.  2008. Natural selection has driven population differentiation in modern humans. Nat. Genet. 40:340–45 [Google Scholar]
  8. Bartel DP. 8.  2004. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116:281–97 [Google Scholar]
  9. Blow MJ, McCulley DJ, Li Z, Zhang T, Akiyama JA. 9.  et al. 2010. ChIP-Seq identification of weakly conserved heart enhancers. Nat. Genet. 42:806–10 [Google Scholar]
  10. Boffelli D, Nobrega MA, Rubin EM. 10.  2004. Comparative genomics at the vertebrate extremes. Nat. Rev. Genet. 5:456–65 [Google Scholar]
  11. Boyko AR, Williamson SH, Indap AR, Degenhardt JD, Hernandez RD. 11.  et al. 2008. Assessing the evolutionary impact of amino acid mutations in the human genome. PLoS Genet. 4:e1000083 [Google Scholar]
  12. Cabili MN, Trapnell C, Goff L, Koziol M, Tazon-Vega B. 12.  et al. 2011. Integrative annotation of human large intergenic noncoding RNAs reveals global properties and specific subclasses. Genes Dev. 25:1915–27 [Google Scholar]
  13. Calin GA, Liu CG, Ferracin M, Hyslop T, Spizzo R. 13.  et al. 2007. Ultraconserved regions encoding ncRNAs are altered in human leukemias and carcinomas. Cancer Cell 12:215–29 [Google Scholar]
  14. Carninci P, Sandelin A, Lenhard B, Katayama S, Shimokawa K. 14.  et al. 2006. Genome-wide analysis of mammalian promoter architecture and evolution. Nat. Genet. 38:626–35 [Google Scholar]
  15. Casillas S, Barbadilla A, Bergman CM. 15.  2007. Purifying selection maintains highly conserved noncoding sequences in Drosophila. Mol. Biol. Evol. 24:2222–34 [Google Scholar]
  16. Cesana M, Cacchiarelli D, Legnini I, Santini T, Sthandier O. 16.  et al. 2011. A long noncoding RNA controls muscle differentiation by functioning as a competing endogenous RNA. Cell 147:358–69 [Google Scholar]
  17. Chamary JV, Hurst LD. 17.  2004. Similar rates but different modes of sequence evolution in introns and at exonic silent sites in rodents: evidence for selectively driven codon usage. Mol. Biol. Evol. 21:1014–23 [Google Scholar]
  18. Chamary JV, Parmley JL, Hurst LD. 18.  2006. Hearing silence: non-neutral evolution at synonymous sites in mammals. Nat. Rev. Genet. 7:98–108 [Google Scholar]
  19. Charlesworth B, Morgan MT, Charlesworth D. 19.  1993. The effect of deleterious mutations on neutral molecular variation. Genetics 134:1289–303 [Google Scholar]
  20. Chen K, Rajewsky N. 20.  2006. Natural selection on human microRNA binding sites inferred from SNP data. Nat. Genet. 38:1452–56 [Google Scholar]
  21. Clark AG, Eisen MB, Smith DR, Bergman CM, Oliver B. 21.  et al. 2007. Evolution of genes and genomes on the Drosophila phylogeny. Nature 450:203–18 [Google Scholar]
  22. Clemente F, Vogl C. 22.  2012. Evidence for complex selection on four-fold degenerate sites in Drosophila melanogaster. J. Evol. Biol. 25:2582–95 [Google Scholar]
  23. Cretekos CJ, Wang Y, Green ED, Martin JF, Rasweiler JJ IV, Behringer RR. 23.  2008. Regulatory divergence modifies limb length between mammals. Genes Dev. 22:141–51 [Google Scholar]
  24. Derrien T, Johnson R, Bussotti G, Tanzer A, Djebali S. 24.  et al. 2012. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res. 22:1775–89 [Google Scholar]
  25. Dobzhansky T. 25.  1964. Biology, molecular and organismic. Am. Zool. 4:443–52 [Google Scholar]
  26. Drake JA, Bird C, Nemesh J, Thomas DJ, Newton-Cheh C. 26.  et al. 2006. Conserved noncoding sequences are selectively constrained and not mutation cold spots. Nat. Genet. 38:223–27 [Google Scholar]
  27. Ebisuya M, Yamamoto T, Nakajima M, Nishida E. 27.  2008. Ripples from neighbouring transcription. Nat. Cell Biol. 10:1106–13 [Google Scholar]
  28. Eddy SR. 28.  2005. A model of the statistical power of comparative genome sequence analysis. PLoS Biol. 3:e10 [Google Scholar]
  29. Gutschner T, Hämmerle M, Günther S, Caudron-Herger M. 29.  Eißmann, et al. 2012. Loss of the abundant nuclear non-coding RNA MALAT1 is compatible with life and development. RNA Biol. 9:1076–87 [Google Scholar]
  30. Elgar G, Vavouri T. 30.  2008. Tuning in to the signals: noncoding sequence conservation in vertebrate genomes. Trends Genet. 24:344–52 [Google Scholar]
  31. Ellegren H, Smith NG, Webster MT. 31.  2003. Mutation rate variation in the mammalian genome. Curr. Opin. Genet. Dev. 13:562–68 [Google Scholar]
  32. 32. ENCODE Proj. Consort 2012. An integrated encyclopedia of DNA elements in the human genome. Nature 489:57–74 [Google Scholar]
  33. Eory L, Halligan DL, Keightley PD. 33.  2010. Distributions of selectively constrained sites and deleterious mutation rates in the hominid and murid genomes. Mol. Biol. Evol. 27:177–92 [Google Scholar]
  34. Eyre-Walker A, Keightley PD. 34.  2009. Estimating the rate of adaptive molecular evolution in the presence of slightly deleterious mutations and population size change. Mol. Biol. Evol. 26:2097–108 [Google Scholar]
  35. Eyre-Walker A, Keightley PD, Smith NG, Gaffney D. 35.  2002. Quantifying the slightly deleterious mutation model of molecular evolution. Mol. Biol. Evol. 19:2142–49 [Google Scholar]
  36. Eyre-Walker A, Woolfit M, Phelps T. 36.  2006. The distribution of fitness effects of new deleterious amino acid mutations in humans. Genetics 173:891–900 [Google Scholar]
  37. Fairbrother WG, Holste D, Burge CB, Sharp PA. 37.  2004. Single nucleotide polymorphism-based validation of exonic splicing enhancers. PLoS Biol. 2:E268 [Google Scholar]
  38. Fisher S, Grice EA, Vinton RM, Bessling SL, McCallion AS. 38.  2006. Conservation of RET regulatory function from human to zebrafish without sequence similarity. Science 312:276–79 [Google Scholar]
  39. Fisher WW, Li JJ, Hammonds AS, Brown JB, Pfeiffer BD. 39.  et al. 2012. DNA regions bound at low occupancy by transcription factors do not drive patterned reporter gene expression in Drosophila. Proc. Natl. Acad. Sci. USA 109:21330–35 [Google Scholar]
  40. Friedman RC, Farh KKH, Burge CB, Bartel DP. 40.  2009. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 19:92–105 [Google Scholar]
  41. Fu W, Akey JM. 41.  2013. Selection and adaptation in the human genome. Annu. Rev. Genomics Hum. Genet. 14:467–89 [Google Scholar]
  42. Fu W, O'Connor TD, Jun G, Kang HM, Abecasis G. 42.  et al. 2013. Analysis of 6,515 exomes reveals the recent origin of most human protein-coding variants. Nature 493:216–20 [Google Scholar]
  43. Fu YX, Li WH. 43.  1993. Statistical tests of neutrality of mutations. Genetics 133:693–709 [Google Scholar]
  44. Galtier N, Duret L, Glémin S, Ranwez V. 44.  2009. GC-biased gene conversion promotes the fixation of deleterious amino acid changes in primates. Trends Genet. 25:1–5 [Google Scholar]
  45. Gazave E, Marqués-Bonet T, Fernando O, Charlesworth B, Navarro A. 45.  2007. Patterns and rates of intron divergence between humans and chimpanzees. Genome Biol. 8:R21 [Google Scholar]
  46. Haddrill PR, Charlesworth B, Halligan DL, Andolfatto P. 46.  2005. Patterns of intron sequence evolution in Drosophila are dependent upon length and GC content. Genome Biol. 6:R67 [Google Scholar]
  47. Haerty W, Jagadeeshan S, Kulathinal RJ, Wong A, Ravi Ram K. 47.  et al. 2007. Evolution in the fast lane: rapidly evolving sex-related genes in Drosophila. Genetics 177:1321–35 [Google Scholar]
  48. Haerty W, Ponting CP. 48.  2013. Mutations within lncRNAs are effectively selected against in fruitfly but not in human. Genome Biol. 14:R49 [Google Scholar]
  49. Halligan DL, Oliver F, Eyre-Walker A, Harr B, Keightley PD. 49.  2010. Evidence for pervasive adaptive protein evolution in wild mice. PLoS Genet. 6:e1000825 [Google Scholar]
  50. Halligan DL, Oliver F, Guthrie J, Stemshorn KC, Harr B, Keightley PD. 50.  2011. Positive and negative selection in murine ultraconserved noncoding elements. Mol. Biol. Evol. 28:2651–60 [Google Scholar]
  51. Hangauer MJ, Vaughn IW, McManus MT. 51.  2013. Pervasive transcription of the human genome produces thousands of previously unidentified long intergenic noncoding RNAs. PLoS Genet. 9:e1003569 [Google Scholar]
  52. Hedges SB. 52.  2002. The origin and evolution of model organisms. Nat. Rev. Genet. 3:838–49 [Google Scholar]
  53. Heger A, Ponting CP. 53.  2007. Variable strength of translational selection among 12 Drosophila species. Genetics 177:1337–48 [Google Scholar]
  54. Hellmann I, Zöllner S, Enard W, Ebersberger I, Nickel B, Pääbo S. 54.  2003. Selection on human genes as revealed by comparisons to chimpanzee cDNA. Genome Res. 13:831–37 [Google Scholar]
  55. Hillenmeyer ME, Fung E, Wildenhain J, Pierce SE, Hoon S. 55.  et al. 2008. The chemical genomic portrait of yeast: uncovering a phenotype for all genes. Science 320:362–65 [Google Scholar]
  56. Hurst LD. 56.  2002. The Ka/Ks ratio: diagnosing the form of sequence evolution. Trends Genet. 18:486 [Google Scholar]
  57. Kaessmann H. 57.  2010. Origins, evolution, and phenotypic impact of new genes. Genome Res. 20:1313–26 [Google Scholar]
  58. Katzman S, Kern AD, Bejerano G, Fewell G, Fulton L. 58.  et al. 2007. Human genome ultraconserved elements are ultraselected. Science 317:915 [Google Scholar]
  59. Keightley PD, Eyre-Walker A. 59.  2007. Joint inference of the distribution of fitness effects of deleterious mutations and population demography based on nucleotide polymorphism frequencies. Genetics 177:2251–61 [Google Scholar]
  60. Keightley PD, Kryukov GV, Sunyaev S, Halligan DL, Gaffney DJ. 60.  2005. Evolutionary constraints in conserved nongenic sequences of mammals. Genome Res. 15:1373–78 [Google Scholar]
  61. Keightley PD, Lercher MJ, Eyre-Walker A. 61.  2005. Evidence for widespread degradation of gene control regions in hominid genomes. PLoS Biol. 3:e42 [Google Scholar]
  62. Kiezun A, Pulit SL, Francioli LC, van Dijk F, Swertz M. 62.  et al. 2013. Deleterious alleles in the human genome are on average younger than neutral alleles of the same frequency. PLoS Genet. 9:e1003301 [Google Scholar]
  63. Kim SY, Pritchard JK. 63.  2007. Adaptive evolution of conserved noncoding elements in mammals. PLoS Genet. 3:1572–86 [Google Scholar]
  64. Kimura M. 64.  1983. The Neutral Theory of Molecular Evolution Cambridge, UK: Cambridge Univ. Press
  65. Klopfstein S, Currat M, Excoffier L. 65.  2006. The fate of mutations surfing on the wave of a range expansion. Mol. Biol. Evol. 23:482–90 [Google Scholar]
  66. Knowles DG, McLysaght A. 66.  2009. Recent de novo origin of human protein-coding genes. Genome Res. 19:1752–59 [Google Scholar]
  67. Kondrashov AS, Sunyaev S, Kondrashov FA. 67.  2002. Dobzhansky-Muller incompatibilities in protein evolution. Proc. Natl. Acad. Sci. USA 99:14878–83 [Google Scholar]
  68. Kousathanas A, Keightley PD. 68.  2013. A comparison of models to infer the distribution of fitness effects of new mutations. Genetics 193:1197–208 [Google Scholar]
  69. Kousathanas A, Oliver F, Halligan DL, Keightley PD. 69.  2011. Positive and negative selection on noncoding DNA close to protein-coding genes in wild house mice. Mol. Biol. Evol. 28:1183–91 [Google Scholar]
  70. Kryukov GV, Schmidt S, Sunyaev S. 70.  2005. Small fitness effect of mutations in highly conserved non-coding regions. Hum. Mol. Genet. 14:2221–29 [Google Scholar]
  71. Künstner A, Nabholz B, Ellegren H. 71.  2011. Significant selective constraint at 4-fold degenerate sites in the avian genome and its consequence for detection of positive selection. Genome Biol. Evol. 3:1381–89 [Google Scholar]
  72. Kutter C, Watt S, Stefflova K, Wilson MD, Goncalves A. 72.  et al. 2012. Rapid turnover of long noncoding RNAs and the evolution of gene expression. PLoS Genet. 8:e1002841 [Google Scholar]
  73. Lachance J, Vernot B, Elbers CC, Ferwerda B, Froment A. 73.  et al. 2012. Evolutionary history and adaptation from high-coverage whole-genome sequences of diverse African hunter-gatherers. Cell 150:457–69 [Google Scholar]
  74. Lawrie DS, Messer PW, Hershberg R, Petrov DA. 74.  2013. Strong purifying selection at synonymous sites in D. melanogaster. PLoS Genet. 9:e1003527 [Google Scholar]
  75. Lewejohann L, Skryabin BV, Sachser N, Prehn C, Heiduschka P. 75.  et al. 2004. Role of a neuronal small non-messenger RNA: behavioural alterations in BC1 RNA-deleted mice. Behav. Brain Res. 154:273–89 [Google Scholar]
  76. Li H, Durbin R. 76.  2011. Inference of human population history from individual whole-genome sequences. Nature 475:493–96 [Google Scholar]
  77. Lickwar CR, Mueller F, Lieb JD. 77.  2013. Genome-wide measurement of protein-DNA binding dynamics using competition ChIP. Nat Protoc. 8:1337–53 [Google Scholar]
  78. Lindblad-Toh K, Garber M, Zuk O, Lin MF, Parker BJ. 78.  et al. 2011. A high-resolution map of human evolutionary constraint using 29 mammals. Nature 478:476–82 [Google Scholar]
  79. Lunter G, Ponting CP, Hein J. 79.  2006. Genome-wide identification of human functional DNA using a neutral indel model. PLoS Comput. Biol. 2:e5 [Google Scholar]
  80. Main BJ, Smith AD, Jang H, Nuzhdin SV. 80.  2013. Transcription start site evolution in Drosophila. Mol. Biol. Evol. 30:1966–74 [Google Scholar]
  81. Majewski J, Ott J. 81.  2002. Distribution and characterization of regulatory elements in the human genome. Genome Res. 12:1827–36 [Google Scholar]
  82. Marques AC, Ponting CP. 82.  2009. Catalogues of mammalian long noncoding RNAs: modest conservation and incompleteness. Genome Biol. 10:R124 [Google Scholar]
  83. McGaughey DM, Vinton RM, Huynh J, Al-Saif A, Beer MA, McCallion AS. 83.  2008. Metrics of sequence constraint overlook regulatory sequences in an exhaustive analysis at phox2b. Genome Res. 18:252–60 [Google Scholar]
  84. Meader S, Ponting CP, Lunter G. 84.  2010. Massive turnover of functional sequence in human and other mammalian genomes. Genome Res. 20:1335–43 [Google Scholar]
  85. Messer PW, Petrov DA. 85.  2013. Frequent adaptation and the McDonald–Kreitman test. Proc. Natl. Acad. Sci. USA 110:8615–20 [Google Scholar]
  86. Roy S, Ernst J, Kharchenko PV, Kheradpour P. 86.  modENCODE Consort. et al. 2010. Identification of functional elements and regulatory circuits by Drosophila modENCODE. Science 330:1787–97 [Google Scholar]
  87. 87. Mouse Genome Seq. Consort 2002. Initial sequencing and comparative analysis of the mouse genome. Nature 420:520–62 [Google Scholar]
  88. Mu XJ, Lu ZJ, Kong Y, Lam HYK, Gerstein MB. 88.  2011. Analysis of genomic variation in non-coding elements using population-scale sequencing data from the 1000 Genomes Project. Nucleic Acids Res. 39:7058–76 [Google Scholar]
  89. Nakagawa S, Ip JY, Shioi G, Tripathi V, Zong X. 89.  et al. 2012. Malat1 is not an essential component of nuclear speckles in mice. RNA 18:1487–99 [Google Scholar]
  90. Nakagawa S, Naganuma T, Shioi G, Hirose T. 90.  2011. Paraspeckles are subpopulation-specific nuclear bodies that are not essential in mice. J. Cell Biol. 193:31–39 [Google Scholar]
  91. Odom DT, Dowell RD, Jacobsen ES, Gordon W, Danford TW. 91.  et al. 2007. Tissue-specific transcriptional regulation has diverged significantly between human and mouse. Nat. Genet. 39:730–32 [Google Scholar]
  92. Ohta T. 92.  1973. Slightly deleterious mutant substitutions in evolution. Nature 246:96–98 [Google Scholar]
  93. Ohta T, Gillespie JH. 93.  1996. Development of neutral and nearly neutral theories. Theor. Popul. Biol. 49:128–42 [Google Scholar]
  94. Osada N, Hirata M, Tanuma R, Kusuda J, Hida M. 94.  et al. 2005. Substitution rate and structural divergence of 5′UTR evolution: comparative analysis between human and cynomolgus monkey cDNAs. Mol. Biol. Evol. 22:1976–82 [Google Scholar]
  95. Parmley JL, Urrutia AO, Potrzebowski L, Kaessmann H, Hurst LD. 95.  2007. Splicing and the evolution of proteins in mammals. PLoS Biol. 5:e14 [Google Scholar]
  96. Parsch J, Novozhilov S, Saminadin-Peter SS, Wong KM, Andolfatto P. 96.  2010. On the utility of short intron sequences as a reference for the detection of positive and negative selection in Drosophila. Mol. Biol. Evol. 27:1226–34 [Google Scholar]
  97. Pennacchio LA, Ahituv N, Moses AM, Prabhakar S, Nobrega MA. 97.  et al. 2006. In vivo enhancer analysis of human conserved non-coding sequences. Nature 444:499–502 [Google Scholar]
  98. Piskol R, Stephan W. 98.  2008. Analyzing the evolution of RNA secondary structures in vertebrate introns using Kimura's model of compensatory fitness interactions. Mol. Biol. Evol. 25:2483–92 [Google Scholar]
  99. Pollard KS, Salama SR, Lambert N, Lambot MA, Coppens S. 99.  et al. 2006. An RNA gene expressed during cortical development evolved rapidly in humans. Nature 443:167–72 [Google Scholar]
  100. Ponjavic J, Ponting CP, Lunter G. 100.  2007. Functionality or transcriptional noise? Evidence for selection within long noncoding RNAs. Genome Res. 17:556–65 [Google Scholar]
  101. Ponting CP, Hardison RC. 101.  2011. What fraction of the human genome is functional?. Genome Res. 21:1769–76 [Google Scholar]
  102. Ponting CP, Lunter G. 102.  2006. Signatures of adaptive evolution within human non-coding sequence. Hum. Mol. Genet. 15:Suppl. 2R170–75 [Google Scholar]
  103. Ponting CP, Nellåker C, Meader S. 103.  2011. Rapid turnover of functional sequence in human and other genomes. Annu. Rev. Genomics Hum. Genet. 12:275–99 [Google Scholar]
  104. Ponting CP, Oliver PL, Reik W. 104.  2009. Evolution and functions of long noncoding RNAs. Cell 136:629–41 [Google Scholar]
  105. Ramani AK, Chuluunbaatar T, Verster AJ, Na H, Vu V. 105.  et al. 2012. The majority of animal genes are required for wild-type fitness. Cell 148:792–802 [Google Scholar]
  106. Rands CM, Meader S, Ponting CP, Lunter G. 106.  2014. 8.2% of the human genome is constrained: variation in rates of turnover across functional element classes in the human lineage. PLoS Genet. In press
  107. Roy SW, Penny D, Neafsey DE. 107.  2007. Evolutionary conservation of UTR intron boundaries in Cryptococcus. Mol. Biol. Evol. 24:1140–48 [Google Scholar]
  108. Schorderet P, Duboule D. 108.  2011. Structural and functional differences in the long non-coding RNA Hotair in mouse and human. PLoS Genet. 7:e1002071 [Google Scholar]
  109. Siepel A, Bejerano G, Pedersen JS, Hinrichs AS, Hou M. 109.  et al. 2005. Evolutionarily conserved elements in vertebrate, insect, worm, and yeast genomes. Genome Res. 15:1034–50 [Google Scholar]
  110. Singh ND, Larracuente AM, Clark AG. 110.  2008. Contrasting the efficacy of selection on the X and autosomes in Drosophila. Mol. Biol. Evol. 25:454–67 [Google Scholar]
  111. Spivakov M, Akhtar J, Kheradpour P, Beal K, Girardot C. 111.  et al. 2012. Analysis of variation at transcription factor binding sites in Drosophila and humans. Genome Biol. 13:R49 [Google Scholar]
  112. Stark A, Brennecke J, Bushati N, Russell RB, Cohen SM. 112.  2005. Animal microRNAs confer robustness to gene expression and have a significant impact on 3′UTR evolution. Cell 123:1133–46 [Google Scholar]
  113. Stone EA, Cooper GM, Sidow A. 113.  2005. Trade-offs in detecting evolutionarily constrained sequence by comparative genomics. Annu. Rev. Genomics Hum. Genet. 6:143–64 [Google Scholar]
  114. Struhl K. 114.  2007. Transcriptional noise and the fidelity of initiation by RNA polymerase II. Nat. Struct. Mol. Biol. 14:103–5 [Google Scholar]
  115. Taylor MS, Kai C, Kawai J, Carninci P, Hayashizaki Y, Semple CAM. 115.  2006. Heterotachy in mammalian promoter evolution. PLoS Genet. 2:e30 [Google Scholar]
  116. Tennessen JA, Bigham AW, O'Connor TD, Fu W, Kenny EE. 116.  et al. 2012. Evolution and functional impact of rare coding variation from deep sequencing of human exomes. Science 337:64–69 [Google Scholar]
  117. Thomas LF, Sætrom P. 117.  2012. Single nucleotide polymorphisms can create alternative polyadenylation signals and affect gene expression through loss of microRNA-regulation. PLoS Comput. Biol. 8:e1002621 [Google Scholar]
  118. Thurman RE, Rynes E, Humbert R, Vierstra J, Maurano MT. 118.  et al. 2012. The accessible chromatin landscape of the human genome. Nature 489:75–82 [Google Scholar]
  119. Torgerson DG, Boyko AR, Hernandez RD, Indap A, Hu X. 119.  et al. 2009. Evolutionary processes acting on candidate cis-regulatory regions in humans inferred from patterns of polymorphism and divergence. PLoS Genet. 5:e1000592 [Google Scholar]
  120. Ulitsky I, Bartel DP. 120.  2013. lincRNAs: genomics, evolution, and mechanisms. Cell 154:26–46 [Google Scholar]
  121. Vernot B, Stergachis AB, Maurano MT, Vierstra J, Neph S. 121.  et al. 2012. Personal and population genomics of human regulatory variation. Genome Res. 22:1689–97 [Google Scholar]
  122. Visel A, Prabhakar S, Akiyama JA, Shoukry M, Lewis KD. 122.  et al. 2008. Ultraconservation identifies a small subset of extremely constrained developmental enhancers. Nat. Genet. 40:158–60 [Google Scholar]
  123. Wakeley J, Aliacar N. 123.  2001. Gene genealogies in a metapopulation. Genetics 159:893–905 [Google Scholar]
  124. Ward LD, Kellis M. 124.  2012. Evidence of abundant purifying selection in humans for recently acquired regulatory functions. Science 337:1675–78 [Google Scholar]
  125. Warnecke T, Parmley JL, Hurst LD. 125.  2008. Finding exonic islands in a sea of non-coding sequence: Splicing related constraints on protein composition and evolution are common in intron-rich genomes. Genome Biol. 9:R29 [Google Scholar]
  126. Warnefors M, Pereira V, Eyre-Walker A. 126.  2010. Transposable elements: insertion pattern and impact on gene expression evolution in hominids. Mol. Biol. Evol. 27:1955–62 [Google Scholar]
  127. Williamson SH, Hernandez R, Fledel-Alon A, Zhu L, Nielsen R, Bustamante CD. 127.  2005. Simultaneous inference of selection and population growth from patterns of variation in the human genome. Proc. Natl. Acad. Sci. USA 102:7882–87 [Google Scholar]
  128. Woolfe A, Goodson M, Goode DK, Snell P, McEwen GK. 128.  et al. 2005. Highly conserved non-coding sequences are associated with vertebrate development. PLoS Biol. 3:e7 [Google Scholar]
  129. Xu J, Zhang R, Shen Y, Liu G, Lu X, Wu CI. 129.  2013. The evolution of evolvability of microRNA target sites in vertebrates. Genome Res. 23:1810–16 [Google Scholar]
  130. Yokoyama KD, Thorne JL, Wray GA. 130.  2011. Coordinated genome-wide modifications within proximal promoter cis-regulatory elements during vertebrate evolution. Genome Biol. Evol. 3:66–74 [Google Scholar]
  131. Young RS, Marques AC, Tibbit C, Haerty W, Bassett AR. 131.  et al. 2012. Identification and properties of 1,119 lincRNA loci in the Drosophila melanogaster genome. Genome Biol. Evol. 4:427–42 [Google Scholar]
  132. Zhang B, Arun G, Mao YS, Lazar Z, Hung G. 132.  et al. 2012. The lncRNA Malat1 is dispensable for mouse development but its transcription plays a cis-regulatory role in the adult. Cell Rep. 2:111–23 [Google Scholar]
  133. Zhen Y, Andolfatto P. 133.  2012. Methods to detect selection on noncoding DNA. Methods Mol. Biol. 856:141–59 [Google Scholar]
/content/journals/10.1146/annurev-genom-090413-025621
Loading
/content/journals/10.1146/annurev-genom-090413-025621
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error