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.

  • Analysis
  • Published:

The structure, function and evolution of proteins that bind DNA and RNA

Key Points

  • DNA- and RNA-binding proteins (DRBPs) constitute a significant fraction of cellular proteins and have important roles in cells. Their functions include control of transcription and translation, DNA repair, splicing, apoptosis and mediating stress responses.

  • Orthogonal binding of DNA and RNA provides an opportunity for competitive regulation of transcription by decoy RNAs, which can be mRNAs, tRNAs and long non-coding RNAs (lncRNAs). Simultaneous binding of DNA and RNA facilitates the assembly of RNA-tethered transcriptional complexes, allowing the recruitment of RNA-containing complexes to specific DNA loci.

  • Binding to both DNA and RNA enables DRBPs to integrate multiple signals into cellular signalling networks and allows improved gene targeting, finer control of gene expression and incorporation of metabolic states or stresses to modulate protein activity.

  • The structural features of DRBPs have been remarkably well conserved during evolution, indicating that dual nucleic acid binding confers selective advantages. DRBPs may have less stringent criteria for interacting with nucleic acids or may adjust their structure when DNA and RNA compete for the same DRBP interaction surface.

  • Nucleic acid sequence evolution has an important role in the development of DRBP function. Emerging studies of rapidly evolving lncRNAs suggest that RNA binding by proteins that were previously thought of as DNA-specific may be a widespread phenomenon.

Abstract

Proteins that bind both DNA and RNA typify the ability of a single gene product to perform multiple functions. Such DNA- and RNA-binding proteins (DRBPs) have unique functional characteristics that stem from their specific structural features; these developed early in evolution and are widely conserved. Proteins that bind RNA have typically been considered as functionally distinct from proteins that bind DNA and studied independently. This practice is becoming outdated, in partly owing to the discovery of long non-coding RNAs (lncRNAs) that target DNA-binding proteins. Consequently, DRBPs were found to regulate many cellular processes, including transcription, translation, gene silencing, microRNA biogenesis and telomere maintenance.

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: Defining human DRBPs.
Figure 2: Functional and structural properties of DRBPs.
Figure 3: Three archetypes of DRBP function.
Figure 4: The structural basis for dual DNA and RNA recognition by TDP43 and by the NF-κB subunit p50.
Figure 5: DNA methyltransferases target both DNA and RNA.

Similar content being viewed by others

References

  1. Kino, T., Hurt, D. E., Ichijo, T., Nader, N. & Chrousos, G. P. Noncoding RNA GAS5 is a growth arrest- and starvation-associated repressor of the glucocorticoid receptor. Sci. Signal. 3, ra8 (2010). This paper shows that a widely expressed lncRNA was able to accumulate during stress and act as a decoy RNA to prevent steroid receptors from binding to their target DNA.

    PubMed  PubMed Central  Google Scholar 

  2. Ishmael, F. T. et al. The human glucocorticoid receptor as an RNA-binding protein: global analysis of glucocorticoid receptor-associated transcripts and identification of a target RNA motif. J. Immunol. 186, 1189–1198 (2011). This study shows that the GCR binds to mRNAs involved in inflammation to accelerate their degradation.

    CAS  PubMed  Google Scholar 

  3. Davis, B. N., Hilyard, A. C., Nguyen, P. H., Lagna, G. & Hata, A. SMAD proteins bind a conserved RNA sequence to promote microRNA maturation by Drosha. Mol. Cell 39, 373–384 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Shi, L. et al. Dynamic binding of Ku80, Ku70 and NF90 to the IL-2 promoter in vivo in activated T-cells. Nucleic Acids Res. 35, 2302–2310 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Shi, L., Godfrey, W. R., Lin, J., Zhao, G. & Kao, P. N. NF90 regulates inducible IL-2 gene expression in T cells. J. Exp. Med. 204, 971–977 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Xu, B. et al. Dax-1 and steroid receptor RNA activator (SRA) function as transcriptional coactivators for steroidogenic factor 1 in steroidogenesis. Mol. Cell. Biol. 29, 1719–1734 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Lanz, R. B. et al. A steroid receptor coactivator, SRA, functions as an RNA and is present in an SRC-1 complex. Cell 97, 17–27 (1999). This paper details the discovery of a lncRNA involved in coordinating steroid receptor-driven gene activation.

    CAS  PubMed  Google Scholar 

  8. Hung, T. et al. Extensive and coordinated transcription of noncoding RNAs within cell-cycle promoters. Nature Genet. 43, 621–629 (2011). This study describes a high-throughput approach to identify lncRNAs that interfere with transcription factors to repress gene expression, suggesting potentially widespread roles for promoter lncRNAs in cell-growth control.

    CAS  PubMed  Google Scholar 

  9. Rapicavoli, N. A. et al. A mammalian pseudogene lncRNA at the interface of inflammation and anti-inflammatory therapeutics. eLife 2, e00762 (2013). This study shows that the expression of pseudogene lncRNAs is actively regulated and that these lncRNAs may interact directly with proteins to modulate cell signalling.

    PubMed  PubMed Central  Google Scholar 

  10. Lebruska, L. L. & Maher, L. J. 3rd. Selection and characterization of an RNA decoy for transcription factor NF-κB. Biochemistry 38, 3168–3174 (1999).

    CAS  PubMed  Google Scholar 

  11. Duss, O. et al. Structural basis of the non-coding RNA RsmZ acting as a protein sponge. Nature 509, 588–592 (2014). This paper shows how a bacterial lncRNA acts as a 'sponge' to sequester and release proteins involved in translation.

    CAS  PubMed  Google Scholar 

  12. Binns, D. et al. QuickGO: a web-based tool for gene ontology searching. Bioinformatics 25, 3045–3046 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Hu, S. et al. Profiling the human protein-DNA interactome reveals ERK2 as a transcriptional repressor of interferon signaling. Cell 139, 610–622 (2009). This paper details the use of a combined bioinformatics- and protein microarray-based strategy to systematically characterize the human protein–dsDNA interactome.

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Castello, A. et al. Insights into RNA biology from an atlas of mammalian mRNA-binding proteins. Cell 149, 1393–1406 (2012). This study identifies the mRNA interactome by using high-throughput covalent capture techniques coupled with proteomic-based detection strategies.

    CAS  Google Scholar 

  15. Kadmiel, M. & Cidlowski, J. A. Glucocorticoid receptor signaling in health and disease. Trends Pharmacol. Sci. 34, 518–530 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Reddy, T. E. et al. Genomic determination of the glucocorticoid response reveals unexpected mechanisms of gene regulation. Genome Res. 19, 2163–2171 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Clark, A. R. & Belvisi, M. G. Maps and legends: The quest for dissociated ligands of the glucocorticoid receptor. Pharmacol. Ther. 134, 54–67 (2012).

    CAS  PubMed  Google Scholar 

  18. Kanai, A., Tani, H., Torimura, M. & Akimitsu, N. The RNA degradation pathway regulates the function of GAS5 a non-coding RNA in mammalian cells. PLoS ONE 8, e55684 (2013).

    Google Scholar 

  19. Fukunaga, J. et al. The Runt domain of AML1 (RUNX1) binds a sequence-conserved RNA motif that mimics a DNA element. RNA 19, 927–936 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Gros, C. et al. DNA methylation inhibitors in cancer: recent and future approaches. Biochimie 94, 2280–2296 (2012).

    CAS  PubMed  Google Scholar 

  21. Holz-Schietinger, C. & Reich, N. O. RNA modulation of the human DNA methyltransferase 3A. Nucleic Acids Res. 40, 8550–8557 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Di Ruscio, A. et al. DNMT1-interacting RNAs block gene-specific DNA methylation. Nature 503, 371–376 (2013). This paper describes how RNA transcripts are able to regulate local DNA methylation by interacting with DNA methyltransferases.

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Cattaneo, A., Biocca, S., Corvaja, N. & Calissano, P. Nuclear localization of a lactic dehydrogenase with single-stranded DNA-binding properties. Exp. Cell Res. 161, 130–140 (1985).

    CAS  PubMed  Google Scholar 

  24. Calissano, P., Volontè, C., Biocca, S. & Cattaneo, A. Synthesis and content of a DNA-binding protein with lactic dehydrogenase activity are reduced by nerve growth factor in the neoplastic cell line PC12. Exp. Cell Res. 161, 117–129 (1985).

    CAS  PubMed  Google Scholar 

  25. Pioli, P. A., Hamilton, B. J., Connolly, J. E., Brewer, G. & Rigby, W. F. Lactate dehydrogenase is an AU-rich element-binding protein that directly interacts with AUF1. J. Biol. Chem. 277, 35738–35745 (2002).

    CAS  PubMed  Google Scholar 

  26. Demarse, N. A. et al. Direct binding of glyceraldehyde 3-phosphate dehydrogenase to telomeric DNA protects telomeres against chemotherapy-induced rapid degradation. J. Mol. Biol. 394, 789–803 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Singh, R. & Green, M. Sequence-specific binding of transfer RNA by glyceraldehyde-3-phosphate dehydrogenase. Science 259, 365–368 (1993). This study identifies GAPDH as a tRNA-binding protein, demonstrating in principle how tRNA levels may control metabolism and how metabolites may influence protein synthesis.

    CAS  PubMed  Google Scholar 

  28. Nagy, E. & Rigby, W. F. C. Glyceraldehyde-3-phosphate dehydrogenase selectively binds AU-rich RNA in the NAD+-binding region (Rossmann fold). J. Biol. Chem. 270, 2755–2763 (1995).

    CAS  PubMed  Google Scholar 

  29. Ray, R. & Miller, D. M. Cloning and characterization of a human c-Myc promoter-binding protein. Mol. Cell. Biol. 11, 2154–2161 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Hernandez-Perez, L. et al. α-enolase binds to RNA. Biochimie 93, 1520–1528 (2011).

    CAS  PubMed  Google Scholar 

  31. Fletcher, L., Rider, C. C. & Taylor, C. B. Enolase isoenzymes. III. Chromatographic and immunological characteristics of rat brain enolase. Biochim. Biophys. Acta 452, 245–252 (1976).

    CAS  PubMed  Google Scholar 

  32. Kato, K. et al. Immunoassay of human muscle enolase subunit in serum: a novel marker antigen for muscle diseases. Clin. Chim. Acta 131, 75–85 (1983).

    CAS  PubMed  Google Scholar 

  33. Anderson, S. L., Minard, K. I. & McAlister-Henn, L. Allosteric inhibition of NAD+-specific isocitrate dehydrogenase by a mitochondrial mRNA. Biochemistry 39, 5623–5629 (2000).

    CAS  PubMed  Google Scholar 

  34. Hentze, M. W. & Preiss, T. The REM phase of gene regulation. Trends Biochem. Sci. 35, 423–426 (2010).

    CAS  PubMed  Google Scholar 

  35. Auphan, N., DiDonato, J. A., Rosette, C., Helmberg, A. & Karin, M. Immunosuppression by glucocorticoids: inhibition of NF-κB activity through induction of IκB synthesis. Science 270, 286–290 (1995).

    CAS  PubMed  Google Scholar 

  36. Surjit, M. et al. Widespread negative response elements mediate direct repression by agonist-liganded glucocorticoid receptor. Cell 145, 224–241 (2011).

    CAS  PubMed  Google Scholar 

  37. Jonat, C. et al. Antitumor promotion and antiinflammation: down-modulation of AP-1 (Fos/Jun) activity by glucocorticoid hormone. Cell 62, 1189–1204 (1990).

    CAS  PubMed  Google Scholar 

  38. Dhawan, L., Liu, B., Blaxall, B. C. & Taubman, M. B. A novel role for the glucocorticoid receptor in the regulation of monocyte chemoattractant protein-1 mRNA stability. J. Biol. Chem. 282, 10146–10152 (2007).

    CAS  PubMed  Google Scholar 

  39. Yang-Yen, H.-F. et al. Transcriptional interference between c-Jun and the glucocorticoid receptor: mutual inhibition of DNA binding due to direct protein–protein interaction. Cell 62, 1205–1215 (1990).

    CAS  PubMed  Google Scholar 

  40. Ray, A. & Prefontaine, K. E. Physical association and functional antagonism between the p65 subunit of transcription factor NF-κB and the glucocorticoid receptor. Proc. Natl Acad. Sci. USA 91, 752–756 (1994).

    CAS  PubMed  Google Scholar 

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

    Google Scholar 

  42. Kim, V. N., Han, J. & Siomi, M. C. Biogenesis of small RNAs in animals. Nature Rev. Mol. Cell Biol. 10, 126–139 (2009).

    CAS  Google Scholar 

  43. Massagué, J. & Wotton, D. Transcriptional control by the TGF-β/SMAD signaling system. EMBO J. 19, 1745–1754 (2000).

    PubMed  PubMed Central  Google Scholar 

  44. Davis, B. N., Hilyard, A. C., Lagna, G. & Hata, A. SMAD proteins control DROSHA-mediated microRNA maturation. Nature 454, 56–61 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Kumarswamy, R., Volkmann, I. & Thum, T. Regulation and function of miRNA-21 in health and disease. RNA Biol. 8, 706–713 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Kao, P. N. et al. Cloning and expression of cyclosporin A- and FK506-sensitive nuclear factor of activated T-cells: NF45 and NF90. J. Biol. Chem. 269, 20691–20699 (1994).

    CAS  PubMed  Google Scholar 

  47. Kuwano, Y. et al. NF90 selectively represses the translation of target mRNAs bearing an AU-rich signature motif. Nucleic Acids Res. 38, 225–238 (2009).

    PubMed  PubMed Central  Google Scholar 

  48. Shim, J., Lim, H., R. Yates Iii, J. & Karin, M. Nuclear export of NF90 is required for interleukin-2 mRNA stabilization. Mol. Cell 10, 1331–1344 (2002).

    CAS  PubMed  Google Scholar 

  49. Sakamoto, S. et al. The NF90–NF45 complex functions as a negative regulator in the microRNA processing pathway. Mol. Cell. Biol. 29, 3754–3769 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Malek, T. R. & Castro, I. Interleukin-2 receptor signaling: at the interface between tolerance and immunity. Immunity 33, 153–165 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Dienz, O. & Rincon, M. The effects of IL-6 on CD4 T cell responses. Clim. Immunol. 130, 27–33 (2009).

    CAS  Google Scholar 

  52. Iliopoulos, D., Hirsch, H. A. & Struhl, K. An epigenetic switch involving NF-κB, lin28, let-7 microRNA, and IL6 links inflammation to cell transformation. Cell 139, 693–706 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Takamizawa, J. et al. Reduced expression of the let-7 microRNAs in human lung cancers in association with shortened postoperative survival. Cancer Res. 64, 3753–3756 (2004).

    CAS  PubMed  Google Scholar 

  54. Guo, N. L. et al. Confirmation of gene expression-based prediction of survival in non-small cell lung cancer. Clin. Cancer Res. 14, 8213–8220 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Ng, S.-Y., Bogu, G. K., Soh, B. S. & Stanton, L. W. The long noncoding RNA RMST interacts with SOX2 to regulate neurogenesis. Mol. Cell 51, 349–359 (2013).

    CAS  PubMed  Google Scholar 

  56. Sarkar, A. & Hochedlinger, K. The Sox family of transcription factors: versatile regulators of stem and progenitor cell fate. Cell Stem Cell 12, 15–30 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Ng, S.-Y., Johnson, R. & Stanton, L. W. Human long non-coding RNAs promote pluripotency and neuronal differentiation by association with chromatin modifiers and transcription factors. EMBO J. 31, 522–533 (2012). References 55 and 57 describe how the lncRNA RMST physically interacts with the transcription factor SOX2 to co-regulate a large pool of downstream genes implicated in neurogenesis.

    CAS  PubMed  Google Scholar 

  58. Feng, J. et al. The Evf-2 noncoding RNA is transcribed from the Dlx-5/6 ultraconserved region and functions as a Dlx-2 transcriptional coactivator. Genes Dev. 20, 1470–1484 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Bond, A. M. et al. Balanced gene regulation by an embryonic brain ncRNA is critical for adult hippocampal GABA circuitry. Nature Neurosci. 12, 1020–1027 (2009).

    CAS  PubMed  Google Scholar 

  60. Jeffery, L. & Nakielny, S. Components of the DNA methylation system of chromatin control are RNA-binding proteins. J. Biol. Chem. 279, 49479–49487 (2004).

    CAS  PubMed  Google Scholar 

  61. Wassenegger, M., Heimes, S., Riedel, L. & Sänger, H. L. RNA-directed de novo methylation of genomic sequences in plants. Cell 76, 567–576 (1994).

    CAS  PubMed  Google Scholar 

  62. Schmitz, K. M., Mayer, C., Postepska, A. & Grummt, I. Interaction of noncoding RNA with the rDNA promoter mediates recruitment of DNMT3b and silencing of rRNA genes. Genes Dev. 24, 2264–2269 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Morris, K. V., Chan, S. W., Jacobsen, S. E. & Looney, D. J. Small interfering RNA-induced transcriptional gene silencing in human cells. Science 305, 1289–1292 (2004).

    CAS  PubMed  Google Scholar 

  64. Sun, B. K., Deaton, A. M. & Lee, J. T. A transient heterochromatic state in Xist preempts X inactivation choice without RNA stabilization. Mol. Cell 21, 617–628 (2006).

    CAS  PubMed  Google Scholar 

  65. Zazopoulos, E., Lalli, E., Stocco, D. M. & Sassone-Corsi, P. DNA binding and transcriptional repression by DAX-1 blocks steroidogenesis. Nature 390, 311–315 (1997).

    CAS  PubMed  Google Scholar 

  66. Kabe, Y. et al. The role of human MBF1 as a transcriptional coactivator. J. Biol. Chem. 274, 34196–34202 (1999).

    CAS  PubMed  Google Scholar 

  67. Xu, B. & Koenig, R. J. An RNA-binding domain in the thyroid hormone receptor enhances transcriptional activation. J. Biol. Chem. 279, 33051–33056 (2004).

    CAS  PubMed  Google Scholar 

  68. Zhao, X. et al. Regulation of nuclear receptor activity by a pseudouridine synthase through posttranscriptional modification of steroid receptor RNA activator. Mol. Cell 15, 549–558 (2004).

    CAS  PubMed  Google Scholar 

  69. Watanabe, M. et al. A subfamily of RNA-binding DEAD-box proteins acts as an estrogen receptor α coactivator through the N-terminal activation domain (AF-1) with an RNA coactivator, SRA. EMBO J. 20, 1341–1352 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Deblois, G. & Giguere, V. Ligand-independent coactivation of ERα AF-1 by steroid receptor RNA activator (SRA) via MAPK activation. J. Steroid Biochem. Mol. Biol. 85, 123–131 (2003).

    CAS  PubMed  Google Scholar 

  71. Poon, M. M. & Chen, L. Retinoic acid-gated sequence-specific translational control by RARα. Proc. Natl Acad. Sci. USA 105, 20303–20308 (2008).

    CAS  PubMed  Google Scholar 

  72. van Steensel, B., Smogorzewska, A. & de Lange, T. TRF2 protects human telomeres from end-to-end fusions. Cell 92, 401–413 (1998).

    CAS  PubMed  Google Scholar 

  73. Court, R., Chapman, L., Fairall, L. & Rhodes, D. How the human telomeric proteins TRF1 and TRF2 recognize telomeric DNA: a view from high-resolution crystal structures. EMBO Rep. 6, 39–45 (2005).

    CAS  PubMed  Google Scholar 

  74. Chesnokov, I. N. in International Review of Cytolology. Vol. 256 (ed. Kwang, W. J.) 69–109 (Academic Press, 2007).

    Google Scholar 

  75. Deng, Z., Dheekollu, J., Broccoli, D., Dutta, A. & Lieberman, P. M. The origin recognition complex localizes to telomere repeats and prevents telomere-circle formation. Curr. Biol. 17, 1989–1995 (2007).

    CAS  PubMed  Google Scholar 

  76. Deng, Z., Norseen, J., Wiedmer, A., Riethman, H. & Lieberman, P. M. TERRA RNA binding to TRF2 facilitates heterochromatin formation and ORC recruitment at telomeres. Mol. Cell 35, 403–413 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Huang da, W., Sherman, B. T. & Lempicki, R. A. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nature Protoc. 4, 44–57 (2009).

    Google Scholar 

  78. Huang da, W., Sherman, B. T. & Lempicki, R. A. Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res. 37, 1–13 (2009).

    PubMed  Google Scholar 

  79. Maris, C., Dominguez, C. & Allain, F. H. T. The RNA recognition motif, a plastic RNA-binding platform to regulate post-transcriptional gene expression. FEBS J. 272, 2118–2131 (2005).

    CAS  PubMed  Google Scholar 

  80. Fiset, S. & Chabot, B. hnRNP A1 may interact simultaneously with telomeric DNA and the human telomerase RNA in vitro. Nucleic Acids Res. 29, 2268–2275 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Bahadur, R. P., Zacharias, M. & Janin, J. Dissecting protein–RNA recognition sites. Nucleic Acids Res. 36, 2705–2716 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Allers, J. & Shamoo, Y. Structure-based analysis of protein–RNA interactions using the program ENTANGLE. J. Mol. Biol. 311, 75–86 (2001).

    CAS  PubMed  Google Scholar 

  83. Buratti, E. et al. Nuclear factor TDP43 and SR proteins promote in vitro and in vivo CFTR exon 9 skipping. EMBO J. 20, 1774–1784 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Kawahara, Y. & Mieda-Sato, A. TDP43 promotes microRNA biogenesis as a component of the Drosha and Dicer complexes. Proc. Natl Acad. Sci. USA 109, 3347–3352 (2012).

    CAS  PubMed  Google Scholar 

  85. Lukavsky, P. J. et al. Molecular basis of UG-rich RNA recognition by the human splicing factor TDP43. Nature Struct. Mol. Biol. 20, 1443–1449 (2013).

    CAS  Google Scholar 

  86. Kuo, P. H., Doudeva, L. G., Wang, Y. T., Shen, C. K. J. & Yuan, H. S. Structural insights into TDP43 in nucleic-acid binding and domain interactions. Nucleic Acids Res. 37, 1799–1808 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Deo, R. C., Bonanno, J. B., Sonenberg, N. & Burley, S. K. Recognition of polyadenylate RNA by the poly(A)-binding protein. Cell 98, 835–845 (1999).

    CAS  PubMed  Google Scholar 

  88. Hayden, M. S. & Ghosh, S. NF-κB, the first quarter-century: remarkable progress and outstanding questions. Genes Dev. 26, 203–234 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Wurster, S. E. & Maher, L. J. Selection and characterization of anti-NF-κB p65 RNA aptamers. RNA 14, 1037–1047 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Huang, D. B. et al. Crystal structure of NF-κB (p50)2 complexed to a high-affinity RNA aptamer. Proc. Natl Acad. Sci. USA 100, 9268–9273 (2003).

    CAS  PubMed  Google Scholar 

  91. Müller, C. W., Rey, F. A., Sodeoka, M., Verdine, G. L. & Harrison, S. C. Structure of the NF-κB p50 homodimer bound to DNA. Nature 373, 311–317 (1995).

    PubMed  Google Scholar 

  92. Lu, X.-J. & Olson, W. K. 3DNA: a versatile, integrated software system for the analysis, rebuilding and visualization of three-dimensional nucleic-acid structures. Nature Protoc. 3, 1213–1227 (2008).

    CAS  Google Scholar 

  93. Barrick, J. E., Sudarsan, N., Weinberg, Z., Ruzzo, W. L. & Breaker, R. R. 6S RNA is a widespread regulator of eubacterial RNA polymerase that resembles an open promoter. RNA 11, 774–784 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Schwartz, T., Rould, M. A., Lowenhaupt, K., Herbert, A. & Rich, A. Crystal structure of the Zα domain of the human editing enzyme ADAR1 bound to left-handed Z-DNA. Science 284, 1841–1845 (1999).

    CAS  PubMed  Google Scholar 

  95. Placido, D., Brown, B. A., Lowenhaupt, K., Rich, A. & Athanasiadis, A. A left-handed RNA double helix bound by the Zα domain of the RNA-editing enzyme ADAR1. Structure 15, 395–404 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Goll, M. G. et al. Methylation of tRNAAsp by the DNA methyltransferase homolog Dnmt2. Science 311, 395–398 (2006). This study shows that the DNA methylase DNMT2 evolved the ability to methylate tRNA.

    CAS  PubMed  Google Scholar 

  97. Didier, D. K., Schiffenbauer, J., Woulfe, S. L., Zacheis, M. & Schwartz, B. D. Characterization of the cDNA encoding a protein binding to the major histocompatibility complex class II Y box. Proc. Natl Acad. Sci. USA 85, 7322–7326 (1988).

    CAS  PubMed  Google Scholar 

  98. Stickeler, E. et al. The RNA binding protein YB-1 binds A/C-rich exon enhancers and stimulates splicing of the CD44 alternative exon v4. EMBO J. 20, 3821–3830 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Skabkina, O. V., Lyabin, D. N., Skabkin, M. A. & Ovchinnikov, L. P. YB-1 autoregulates translation of its own mRNA at or prior to the step of 40S ribosomal subunit joining. Mol. Cell. Biol. 25, 3317–3323 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Evdokimova, V. et al. The major mRNA-associated protein YB-1 is a potent 5′ cap-dependent mRNA stabilizer. EMBO J. 20, 5491–5502 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Marenstein, D. R. et al. Stimulation of human endonuclease III by Y box-binding protein 1 (DNA-binding protein B). Interaction between a base excision repair enzyme and a transcription factor. J. Biol. Chem. 276, 21242–21249 (2001).

    CAS  PubMed  Google Scholar 

  102. de Souza-Pinto, N. C. et al. Novel DNA mismatch-repair activity involving YB-1 in human mitochondria. DNA Repair 8, 704–719 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Ise, T. et al. Transcription factor Y-box binding protein 1 binds preferentially to cisplatin-modified DNA and interacts with proliferating cell nuclear antigen. Cancer Res. 59, 342–346 (1999).

    CAS  PubMed  Google Scholar 

  104. Stein, U. et al. Hyperthermia-induced nuclear translocation of transcription factor YB-1 leads to enhanced expression of multidrug resistance-related ABC transporters. J. Biol. Chem. 276, 28562–28569 (2001).

    CAS  PubMed  Google Scholar 

  105. Koike, K. et al. Nuclear translocation of the Y-box binding protein by ultraviolet irradiation. FEBS Lett. 417, 390–394 (1997).

    CAS  PubMed  Google Scholar 

  106. Yamanaka, K., Fang, L. & Inouye, M. The CspA family in Escherichia coli: multiple gene duplication for stress adaptation. Mol. Microbiol. 27, 247–255 (1998).

    CAS  PubMed  Google Scholar 

  107. Etchegaray, J. P., Jones, P. G. & Inouye, M. Differential thermoregulation of two highly homologous cold-shock genes, cspA and cspB, of Escherichia coli. Genes Cells 1, 171–178 (1996).

    CAS  PubMed  Google Scholar 

  108. Wang, N., Yamanaka, K. & Inouye, M. CspI, the ninth member of the CspA family of Escherichia coli, is induced upon cold shock. J. Bacteriol. 181, 1603–1609 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Nakashima, K., Kanamaru, K., Mizuno, T. & Horikoshi, K. A novel member of the cspA family of genes that is induced by cold shock in Escherichia coli. J. Bacteriol. 178, 2994–2997 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Goldstein, J., Pollitt, N. S. & Inouye, M. Major cold shock protein of Escherichia coli. Proc. Natl Acad. Sci. USA 87, 283–287 (1990).

    CAS  PubMed  Google Scholar 

  111. Xia, B., Ke, H. & Inouye, M. Acquirement of cold sensitivity by quadruple deletion of the cspA family and its suppression by PNPase S1 domain in Escherichia coli. Mol. Microbiol. 40, 179–188 (2001).

    CAS  PubMed  Google Scholar 

  112. Yamanaka, K. & Inouye, M. Growth-phase-dependent expression of cspD, encoding a member of the CspA family in Escherichia coli. J. Bacteriol. 179, 5126–5130 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Phadtare, S. & Inouye, M. Role of CspC and CspE in regulation of expression of RpoS and UspA, the stress response proteins in Escherichia coli. J. Bacteriol. 183, 1205–1214 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Jiang, W., Hou, Y. & Inouye, M. CspA, the major cold-shock protein of Escherichia coli, is an RNA chaperone. J. Biol. Chem. 272, 196–202 (1997).

    CAS  PubMed  Google Scholar 

  115. Phadtare, S. & Inouye, M. Sequence-selective interactions with RNA by CspB, CspC and CspE, members of the CspA family of Escherichia coli. Mol. Endocrinol. 33, 1004–1014 (1999).

    CAS  Google Scholar 

  116. Hanna, M. M. & Liu, K. Nascent RNA in transcription complexes interacts with CspE, a small protein in E. coli implicated in chromatin condensation. J. Mol. Biol. 282, 227–239 (1998).

    CAS  PubMed  Google Scholar 

  117. Bae, W., Xia, B., Inouye, M. & Severinov, K. Escherichia coli CspA-family RNA chaperones are transcription antiterminators. Proc. Natl Acad. Sci. USA 97, 7784–7789 (2000).

    CAS  PubMed  Google Scholar 

  118. Yamanaka, K., Zheng, W., Crooke, E., Wang, Y. H. & Inouye, M. CspD, a novel DNA replication inhibitor induced during the stationary phase in Escherichia coli. Mol. Microbiol. 39, 1572–1584 (2001).

    CAS  PubMed  Google Scholar 

  119. Chen, C. S. et al. A proteome ChIP approach reveals new DNA damage recognition activities in Escherichia coli. Nature Methods 5, 69–74 (2008).

    CAS  PubMed  Google Scholar 

  120. Hu, K. H. et al. Overproduction of three genes leads to camphor resistance and chromosome condensation in Escherichia coli. Genetics 143, 1521–1532 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Karlson, D. & Imai, R. Conservation of the cold shock domain protein family in plants. Plant Physiol. 131, 12–15 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Karlson, D., Nakaminami, K., Toyomasu, T. & Imai, R. A cold-regulated nucleic acid-binding protein of winter wheat shares a domain with bacterial cold shock proteins. J. Biol. Chem. 277, 35248–35256 (2002).

    CAS  PubMed  Google Scholar 

  123. Nakaminami, K., Karlson, D. T. & Imai, R. Functional conservation of cold shock domains in bacteria and higher plants. Proc. Natl Acad. Sci. USA 103, 10122–10127 (2006).

    CAS  PubMed  Google Scholar 

  124. Imai, R., Kim, M. H., Sasaki, K., Sato, S. & Sonoda, Y. Plant and Microbe Adaptations to Cold in a Changing World (Springer, 2013).

    Google Scholar 

  125. Juntawong, P., Sorenson, R. & Bailey-Serres, J. Cold shock protein 1 chaperones mRNAs during translation in Arabidopsis thaliana. Plant J. 74, 1016–1028 (2013).

    CAS  PubMed  Google Scholar 

  126. Kim, J. S. et al. Cold shock domain proteins and glycine-rich RNA-binding proteins from Arabidopsis thaliana can promote the cold adaptation process in Escherichia coli. Nucleic Acids Res. 35, 506–516 (2007).

    CAS  PubMed  Google Scholar 

  127. Triqueneaux, G., Velten, M., Franzon, P., Dautry, F. & Jacquemin-Sablon, H. RNA binding specificity of Unr, a protein with five cold shock domains. Nucleic Acids Res. 27, 1926–1934 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Eliseeva, I. A., Kim, E. R., Guryanov, S. G., Ovchinnikov, L. P. & Lyabin, D. N. Y-box-binding protein 1 (YB-1) and its functions. Biochem. 76, 1402–1433 (2012).

    Google Scholar 

  129. Mihailovich, M., Militti, C., Gabaldón, T. & Gebauer, F. Eukaryotic cold shock domain proteins: highly versatile regulators of gene expression. BioEssays 32, 109–118 (2010).

    CAS  PubMed  Google Scholar 

  130. Hermann, A., Schmitt, S. & Jeltsch, A. The human Dnmt2 has residual DNA-(cytosine-C5) methyltransferase activity. J. Biol. Chem. 278, 31717–31721 (2003).

    CAS  PubMed  Google Scholar 

  131. Lyko, F., Jurkowski, T. P. & Jeltsch, A. On the evolutionary origin of eukaryotic DNA methyltransferases and Dnmt2. PLoS ONE 6, e28104 (2011).

    Google Scholar 

  132. Tuorto, F. et al. RNA cytosine methylation by Dnmt2 and NSun2 promotes tRNA stability and protein synthesis. Nature Struct. Mol. Biol. 19, 900–905 (2012).

    CAS  Google Scholar 

  133. Pavlopoulou, A. & Kossida, S. Phylogenetic analysis of the eukaryotic RNA (cytosine-5)-methyltransferases. Genomics 93, 350–357 (2009).

    CAS  PubMed  Google Scholar 

  134. Sakita-Suto, S. et al. Aurora-B regulates RNA methyltransferase NSUN2. Mol. Biol. Cell 18, 1107–1117 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Hussain, S. et al. NSun2-mediated cytosine-5 methylation of vault noncoding RNA determines its processing into regulatory small RNAs. Cell Rep. 4, 255–261 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Reiter, N. J., Maher, L. J. & Butcher, S. E. DNA mimicry by a high-affinity anti-NF-κB RNA aptamer. Nucleic Acids Res. 36, 1227–1236 (2008). This paper shows that the RNA aptamer developed in reference 10 mimics DNA to target the DNA-binding interface of NF-κB.

    CAS  PubMed  Google Scholar 

  137. Ray, D. et al. A compendium of RNA-binding motifs for decoding gene regulation. Nature 499, 172–177 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Nakagawa, S., Gisselbrecht, S. S., Rogers, J. M., Hartl, D. L. & Bulyk, M. L. DNA-binding specificity changes in the evolution of forkhead transcription factors. Proc. Natl Acad. Sci. USA 110, 12349–12354 (2013).

    CAS  PubMed  Google Scholar 

  139. Tompa, P. & Csermely, P. The role of structural disorder in the function of RNA and protein chaperones. FASEB J. 18, 1169–1175 (2004).

    CAS  PubMed  Google Scholar 

  140. Volders, P. J. et al. LNCipedia: a database for annotated human lncRNA transcript sequences and structures. Nucleic Acids Res. 41, D246–D251 (2012).

    PubMed  PubMed Central  Google Scholar 

  141. Ali, M. & Vedeckis, W. The glucocorticoid receptor protein binds to transfer RNA. Science 235, 467–470 (1987).

    CAS  PubMed  Google Scholar 

  142. Feldman, M., Kallos, J. & Hollander, V. P. RNA inhibits estrogen receptor binding to DNA. J. Biol. Chem. 256, 1145–1148 (1981).

    CAS  PubMed  Google Scholar 

  143. Rossini, G. P. RNA-containing nuclear binding sites for glucocorticoid-receptor complexes. Biochem. Biophys. Res. Commun. 123, 78–83 (1984).

    CAS  PubMed  Google Scholar 

  144. Necsulea, A. et al. The evolution of lncRNA repertoires and expression patterns in tetrapods. Nature 505, 635–640 (2014). References 2, 10, 57, 143 and 144 reveal that RNA binding to transcription factors may be a widespread phenomenon.

    CAS  PubMed  Google Scholar 

  145. Flicek, P. et al. Ensembl 2014. Nucleic Acids Res. 42, D749–D755 (2013).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

W.H.H. is supported by American Heart Association predoctoral fellowship 13PRE16920012 and by a US National Institutes of Health (NIH) training grant to Emory University (5T32GM008602-14). Work in the laboratory of E.A.O. is supported by grant R01DK095750 from the NIH National Institute of Diabetes and Digestive and Kidney Diseases and by grant 14GRNT20460124 from the American Heart Association.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Eric A. Ortlund.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Supplementary information

Supplementary information S1 (table)

Human proteins demonstrated to bind DNA and RNA. (PDF 983 kb)

Related links

Related links

DATABASE

QuickGo

GO:0003723

GO:0003677

FURTHER INFORMATION

Unpublished structure of TDP43 RRMs

Glossary

Rossmann fold

A common protein-folding pattern that contains the topology of a β-α-β fold. It is found in many nucleotide- binding proteins.

Drosha

A nuclease of the RNase III family that, as part of a complex, cleaves primary microRNA (pri-miRNA) transcripts into precursor miRNAs (pre-miRNAs) in the nucleus.

Primary miRNA

(Pri-miRNA). The primary transcripts of microRNA-encoding genes, which are stem–loop structures processed by the Drosha complex into precursor miRNAs (pre-miRNAs). Typical pri-miRNAs are hundreds of base pairs long and can contain several pre-miRNAs.

Precursor miRNA

(Pre-miRNA). The product of Drosha-mediated cleavage of primary microRNAs (pri-miRNAs). Stem–loop-structured pre-miRNAs are exported to the cytoplasm and further processed by Dicer to fully mature miRNAs.

MAD homology 1 domain

An evolutionarily conserved domain found in SMAD proteins. It contains four α-helices, six short β-strands and five loops, and it recognizes specific DNA sequences.

Homeodomain

A protein structural domain that binds DNA and RNA, and that is found most commonly in transcription factors. It consists of a helix–turn–helix structure of 60 amino acids.

GAR domain

A Gly- and Arg-rich motif that adopts a repeated β-turn structure. It is found most commonly in proteins that bind RNA.

Zinc-finger domains

Structural motifs that are characterized by the coordination of one or more zinc ions. There are several different zinc-finger motifs, and each displays a different binding mode and structure.

WW domains

Small protein motifs of 40 amino acids that mediate specific protein–protein interactions with short Pro-rich or Pro-containing motifs.

Ftz-F1 domain

A protein domain first identified in the Fushi Tarazu factor 1 (FTZ-F1) nuclear receptor. It contains an evolutionarily conserved LXXLL motif that recognizes other LXXLL-related motifs.

K-homology domain

A conserved protein domain that interacts with both RNA and DNA through a binding cleft formed between two α-helices, two β-sheets and the GXXG loop.

Cold-shock domain

(CSD). A small ( 70 kDa) domain with high similarity to the ribonucleoprotein 1 RNA-binding motif. This domain is found in DNA- and RNA-binding proteins in bacteria, archaea and eukaryotes.

Interfaces

The solvent-accessible portions of a protein that are capable of binding a ligand — including DNA and RNA — in a competitive or non-competitive manner.

π-stacking interactions

Non-covalent interactions between two aromatic molecules owing to the attractive force originating from the opposing electrostatic potentials between two adjacent aromatic amino acid residues.

Aptamers

Single-stranded DNA or RNA molecules that selectively bind small molecules, proteins and peptides with high affinity. Aptamers have dynamic tertiary structures, which contribute to the diversity of their binding targets.

Z-DNA

DNA in the conformation of a left-handed double helix.

Z-RNA

RNA in the conformation of a left-handed double helix.

Y box

A DNA target sequence with the consensus CTGATTG, which is recognized and bound by certain proteins.

RNA homopolymers

Sequences of ribonucleotides consisting of a single base; for example, CCCCCCCCC.

Intrinsically disordered protein

A protein that does not have a well-ordered three-dimensional structure, such as proteins containing random coils or multiple domains connected with flexible linkers.

Systematic evolution of ligands by exponential enrichment

(SELEX). A technique to identify ligand-binding-sequence specificity that is based on sequential rounds of binding of an oligonucleotide library to the ligand, followed by PCR amplification of the bound sequences.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hudson, W., Ortlund, E. The structure, function and evolution of proteins that bind DNA and RNA. Nat Rev Mol Cell Biol 15, 749–760 (2014). https://doi.org/10.1038/nrm3884

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm3884

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