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.

  • Letter
  • Published:

Hepatitis-C-virus-like internal ribosome entry sites displace eIF3 to gain access to the 40S subunit

Abstract

Hepatitis C virus (HCV) and classical swine fever virus (CSFV) messenger RNAs contain related (HCV-like) internal ribosome entry sites (IRESs) that promote 5′-end independent initiation of translation, requiring only a subset of the eukaryotic initiation factors (eIFs) needed for canonical initiation on cellular mRNAs1. Initiation on HCV-like IRESs relies on their specific interaction with the 40S subunit2,3,4,5,6,7,8, which places the initiation codon into the P site, where it directly base-pairs with eIF2-bound initiator methionyl transfer RNA to form a 48S initiation complex. However, all HCV-like IRESs also specifically interact with eIF3 (refs 2, 5, 6, 7, 9, 10, 11, 12), but the role of this interaction in IRES-mediated initiation has remained unknown. During canonical initiation, eIF3 binds to the 40S subunit as a component of the 43S pre-initiation complex, and comparison of the ribosomal positions of eIF313 and the HCV IRES8 revealed that they overlap, so that their rearrangement would be required for formation of ribosomal complexes containing both components13. Here we present a cryo-electron microscopy reconstruction of a 40S ribosomal complex containing eIF3 and the CSFV IRES. Remarkably, although the position and interactions of the CSFV IRES with the 40S subunit in this complex are similar to those of the HCV IRES in the 40S–IRES binary complex8, eIF3 is completely displaced from its ribosomal position in the 43S complex, and instead interacts through its ribosome-binding surface exclusively with the apical region of domain III of the IRES. Our results suggest a role for the specific interaction of HCV-like IRESs with eIF3 in preventing ribosomal association of eIF3, which could serve two purposes: relieving the competition between the IRES and eIF3 for a common binding site on the 40S subunit, and reducing formation of 43S complexes, thereby favouring translation of viral mRNAs.

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: Cryo-electron microscopy structures of the CSFV ΔII-IRES–40S–DHX29 complex alone and bound to eIF3 compared to the structure of the DHX29-bound 43S preinitiation complex.
Figure 2: Different orientations of eIF3 and subdomain IIIb in the CSFV ΔII IRES–40S–DHX29 complex.
Figure 3: Structure and atomic model of the CSFV ΔII-IRES bound to the 40S subunit.
Figure 4: Binding of eIF3 to subdomain IIIb of the CSFV IRES and effects on translation of the eIF3–IRES interaction.

Similar content being viewed by others

Accession codes

Accessions

Protein Data Bank

Data deposits

The electron microscopy map has been deposited in the EMBL-European Bioinformatics Institute Electron Microscopy Data Bank under accession codes EMD-2450 and EMD-2451. Coordinates of electron microscopy-based model of the CSFV IRES have been deposited in the RCSB Protein Data Bank under accession number 4c4q.

References

  1. Jackson, R. J., Hellen, C. U. & Pestova, T. V. The mechanism of eukaryotic translation initiation and principles of its regulation. Nature Rev. Mol. Cell Biol. 11, 113–127 (2010)

    Article  CAS  Google Scholar 

  2. Pestova, T. V., Shatsky, I. N., Fletcher, S. P., Jackson, R. J. & Hellen, C. U. A prokaryotic-like mode of cytoplasmic eukaryotic ribosome binding to the initiation codon during internal translation initiation of hepatitis C and classical swine fever virus RNAs. Genes Dev. 12, 67–83 (1998)

    Article  CAS  Google Scholar 

  3. Kolupaeva, V. G., Pestova, T. V. & Hellen, C. U. Ribosomal binding to the internal ribosomal entry site of classical swine fever virus. RNA 6, 1791–1807 (2000)

    Article  CAS  Google Scholar 

  4. Kolupaeva, V. G., Pestova, T. V. & Hellen, C. U. An enzymatic footprinting analysis of the interaction of 40S ribosomal subunits with the internal ribosomal entry site of hepatitis C virus. J. Virol. 74, 6242–6250 (2000)

    Article  CAS  Google Scholar 

  5. Kieft, J. S., Zhou, K., Jubin, R. & Doudna, J. A. Mechanism of ribosome recruitment by hepatitis C IRES RNA. RNA 7, 194–206 (2001)

    Article  CAS  Google Scholar 

  6. de Breyne, S., Yu, Y., Pestova, T. V. & Hellen, C. U. Factor requirements for translation initiation on the Simian picornavirus internal ribosomal entry site. RNA 14, 367–380 (2008)

    Article  CAS  Google Scholar 

  7. Pisarev, A. V. et al. Functional and structural similarities between the internal ribosome entry sites of hepatitis C virus and porcine teschovirus, a picornavirus. J. Virol. 78, 4487–4497 (2004)

    Article  CAS  Google Scholar 

  8. Spahn, C. M. et al. Hepatitis C virus IRES RNA-induced changes in the conformation of the 40S ribosomal subunit. Science 291, 1959–1962 (2001)

    Article  ADS  CAS  Google Scholar 

  9. Sizova, D. V., Kolupaeva, V. G., Pestova, T. V., Shatsky, I. N. & Hellen, C. U. Specific interaction of eukaryotic translation initiation factor 3 with the 5′ nontranslated regions of hepatitis C virus and classical swine fever virus RNAs. J. Virol. 72, 4775–4782 (1998)

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Buratti, E., Tisminetzky, S., Zotti, M. & Baralle, F. E. Functional analysis of the interaction between HCV 5′UTR and putative subunits of eukaryotic translation initiation factor eIF3. Nucleic Acids Res. 26, 3179–3187 (1998)

    Article  CAS  Google Scholar 

  11. Sun, C. et al. Two RNA-binding motifs in eIF3 direct HCV IRES-dependent translation. Nucleic Acids Res. 41, 7512–7521 (2013)

    Article  CAS  Google Scholar 

  12. Siridechadilok, B., Fraser, C. S., Hall, R. J., Doudna, J. A. & Nogales, E. Structural roles for human translation factor eIF3 in initiation of protein synthesis. Science 310, 1513–1515 (2005)

    Article  ADS  CAS  Google Scholar 

  13. Hashem, Y. et al. Structure of the mammalian ribosomal 43S preinitiation complex bound to the scanning factor DHX29. Cell 153, 1108–1119 (2013)

    Article  CAS  Google Scholar 

  14. Pisareva, V. P., Pisarev, A. V., Komar, A. A., Hellen, C. U. & Pestova, T. V. Translation initiation on mammalian mRNAs with structured 5′UTRs requires DExH-box protein DHX29. Cell 135, 1237–1250 (2008)

    Article  CAS  Google Scholar 

  15. Abaeva, I. S., Marintchev, A., Pisareva, V. P., Hellen, C. U. & Pestova, T. V. Bypassing of stems versus linear base-by-base inspection of mammalian mRNAs during ribosomal scanning. EMBO J. 30, 115–129 (2011)

    Article  CAS  Google Scholar 

  16. Lukavsky, P. J., Otto, G. A., Lancaster, A. M., Sarnow, P. & Puglisi, J. D. Structures of two RNA domains essential for hepatitis C virus internal ribosome entry site function. Nature Struct. Biol. 7, 1105–1110 (2000)

    Article  CAS  Google Scholar 

  17. Locker, N., Easton, L. E. & Lukavsky, P. J. HCV and CSFV IRES domain II mediate eIF2 release during 80S ribosome assembly. EMBO J. 26, 795–805 (2007)

    Article  CAS  Google Scholar 

  18. Pestova, T. V., de Breyne, S., Pisarev, A. V., Abaeva, I. S. & Hellen, C. U. eIF2-dependent and eIF2-independent modes of initiation on the CSFV IRES: a common role of domain II. EMBO J. 27, 1060–1072 (2008)

    Article  CAS  Google Scholar 

  19. Ji, H., Fraser, C. S., Yu, Y., Leary, J. & Doudna, J. A. Coordinated assembly of human translation initiation complexes by the hepatitis C virus internal ribosome entry site RNA. Proc. Natl Acad. Sci. USA 101, 16990–16995 (2004)

    Article  ADS  CAS  Google Scholar 

  20. Fletcher, S. P. & Jackson, R. J. Pestivirus internal ribosome entry site (IRES) structure and function: elements in the 5′ untranslated region important for IRES function. J. Virol. 76, 5024–5033 (2002)

    Article  CAS  Google Scholar 

  21. Friis, M. B., Rasmussen, T. B. & Belsham, G. J. Modulation of translation initiation efficiency in classical swine fever virus. J. Virol. 86, 8681–8692 (2012)

    Article  CAS  Google Scholar 

  22. Rijnbrand, R. et al. Almost the entire 5′ non-translated region of hepatitis C virus is required for cap-independent translation. FEBS Lett. 365, 115–119 (1995)

    Article  CAS  Google Scholar 

  23. Boehringer, D., Thermann, R., Ostareck-Lederer, A., Lewis, J. D. & Stark, H. Structure of the hepatitis C virus IRES bound to the human 80S ribosome: remodeling of the HCV IRES. Structure 13, 1695–1706 (2005)

    Article  CAS  Google Scholar 

  24. Querol-Audi, J. et al. Architecture of human translation initiation factor 3. Structure 21, 920–928 (2013)

    Article  CAS  Google Scholar 

  25. Melcher, S. E., Wilson, T. J. & Lilley, D. M. The dynamic nature of the four-way junction of the hepatitis C virus IRES. RNA 9, 809–820 (2003)

    Article  CAS  Google Scholar 

  26. Rijnbrand, R., van der Straaten, T., van Rijn, P. A., Spaan, W. J. & Bredenbeek, P. J. Internal entry of ribosomes is directed by the 5′ noncoding region of classical swine fever virus and is dependent on the presence of an RNA pseudoknot upstream of the initiation codon. J. Virol. 71, 451–457 (1997)

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Berry, K. E., Waghray, S., Mortimer, S. A., Bai, Y. & Doudna, J. A. Crystal structure of the HCV IRES central domain reveals strategy for start-codon positioning. Structure 19, 1456–1466 (2011)

    Article  CAS  Google Scholar 

  28. Hellen, C. U. & de Breyne, S. A distinct group of hepacivirus/pestivirus-like internal ribosomal entry sites in members of diverse picornavirus genera: evidence for modular exchange of functional noncoding RNA elements by recombination. J. Virol. 81, 5850–5863 (2007)

    Article  CAS  Google Scholar 

  29. Babaylova, E. et al. Positioning of subdomain IIId and apical loop of domain II of the hepatitis C IRES on the human 40S ribosome. Nucleic Acids Res. 37, 1141–1151 (2009)

    Article  CAS  Google Scholar 

  30. Malygin, A. A., Shatsky, I. N. & Karpova, G. G. Proteins of the human 40S ribosomal subunit involved in hepatitis C IRES binding as revealed from fluorescent labeling. Biochemistry (Moscow) 78, 53–59 (2013)

    Article  CAS  Google Scholar 

  31. Skabkin, M. A. et al. Activities of ligatin and MCT-1/DENR in eukaryotic translation initiation and ribosomal recycling. Genes Dev. 24, 1787–1801 (2010)

    Article  CAS  Google Scholar 

  32. Pisarev, A. V., Unbehaun, A., Hellen, C. U. & Pestova, T. V. Assembly and analysis of eukaryotic translation initiation complexes. Methods Enzymol. 430, 147–177 (2007)

    Article  CAS  Google Scholar 

  33. Lomakin, I. B., Shirokikh, N. E., Yusupov, M. M., Hellen, C. U. & Pestova, T. V. The fidelity of translation initiation: reciprocal activities of eIF1, IF3 and YciH. EMBO J. 25, 196–210 (2006)

    Article  CAS  Google Scholar 

  34. Pestova, T. V. & Hellen, C. U. Preparation and activity of synthetic unmodified mammalian tRNAiMet in initiation of translation in vitro. RNA 7, 1496–1505 (2001)

    Article  CAS  Google Scholar 

  35. Grassucci, R. A., Taylor, D. J. & Frank, J. Preparation of macromolecular complexes for cryo-electron microscopy. Nature Protocols 2, 3239–3246 (2007)

    Article  CAS  Google Scholar 

  36. Dubochet, J. et al. Cryo-electron microscopy of vitrified specimens. Q. Rev. Biophys. 21, 129–228 (1988)

    Article  CAS  Google Scholar 

  37. Wagenknecht, T., Frank, J., Boublik, M., Nurse, K. & Ofengand, J. Direct localization of the tRNA-anticodon interaction site on the Escherichia coli 30 S ribosomal subunit by electron microscopy and computerized image averaging. J. Mol. Biol. 203, 753–760 (1988)

    Article  CAS  Google Scholar 

  38. Suloway, C. et al. Automated molecular microscopy: the new Leginon system. J. Struct. Biol. 151, 41–60 (2005)

    Article  CAS  Google Scholar 

  39. Frank, J. et al. SPIDER and WEB: processing and visualization of images in 3D electron microscopy and related fields. J. Struct. Biol. 116, 190–199 (1996)

    Article  CAS  Google Scholar 

  40. Scheres, S. H. RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012)

    Article  CAS  Google Scholar 

  41. Henderson, R. et al. Outcome of the first electron microscopy validation task force meeting. Structure 20, 205–214 (2012)

    Article  CAS  Google Scholar 

  42. Rabl, J., Leibundgut, M., Ataide, S. F., Haag, A. & Ban, N. Crystal structure of the eukaryotic 40S ribosomal subunit in complex with initiation factor 1. Science 331, 730–736 (2011)

    Article  ADS  CAS  Google Scholar 

  43. Pettersen, E. F. et al. UCSF Chimera–a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004)

    Article  CAS  Google Scholar 

  44. Penczek, P. A., Yang, C., Frank, J. & Spahn, C. M. Estimation of variance in single-particle reconstruction using the bootstrap technique. J. Struct. Biol. 154, 168–183 (2006)

    Article  CAS  Google Scholar 

  45. Zhang, W., Kimmel, M., Spahn, C. M. & Penczek, P. A. Heterogeneity of large macromolecular complexes revealed by 3D cryo-EM variance analysis. Structure 16, 1770–1776 (2008)

    Article  CAS  Google Scholar 

  46. Simonetti, A. et al. Structure of the 30S translation initiation complex. Nature 455, 416–420 (2008)

    Article  ADS  CAS  Google Scholar 

  47. Liao, H. Y. & Frank, J. Classification by bootstrapping in single particle methods. Proc. IEEE Int. Symp. Biom. Imaging 169–172 (2010)

  48. Pintilie, G. D., Zhang, J., Goddard, T. D., Chiu, W. & Gossard, D. C. Quantitative analysis of cryo-EM density map segmentation by watershed and scale-space filtering, and fitting of structures by alignment to regions. J. Struct. Biol. 170, 427–438 (2010)

    Article  CAS  Google Scholar 

  49. Jossinet, F. & Westhof, E. Sequence to Structure (S2S): display, manipulate and interconnect RNA data from sequence to structure. Bioinformatics 21, 3320–3321 (2005)

    Article  CAS  Google Scholar 

  50. Jossinet, F., Ludwig, T. E. & Westhof, E. Assemble: an interactive graphical tool to analyze and build RNA architectures at the 2D and 3D levels. Bioinformatics 26, 2057–2059 (2010)

    Article  CAS  Google Scholar 

  51. Trabuco, L. G., Villa, E., Mitra, K., Frank, J. & Schulten, K. Flexible fitting of atomic structures into electron microscopy maps using molecular dynamics. Structure 16, 673–683 (2008)

    Article  CAS  Google Scholar 

  52. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. Model. 14, 33–38 (1996)

    Article  CAS  Google Scholar 

  53. Phillips, J. C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005)

    Article  CAS  Google Scholar 

  54. Brooks, B. R. et al. CHARMM: a program for macromolecular energy, minimization, and dynamics calculations. J. Comput. Chem. 4, 187–217 (1983)

    Article  CAS  Google Scholar 

  55. MacKerell, A. D., Jr et al. CHARMM: the energy function and its parameterization with an overview of the program. Enc. Comput. Chem. 1, 271–277 (1998)

    Google Scholar 

Download references

Acknowledgements

We thank M. Thomas for assistance in the preparation of figures and A. Jobe for critical reading of the manuscript. This work was supported by the HHMI and National Institutes of Health (NIH) grants R01 GM29169 (to J.F.), NIH R01 AI51340 (to C.U.T.H.) and NIH R01 GM59660 (to T.V.P.).

Author information

Authors and Affiliations

Authors

Contributions

Y.H., A.d.G., V.D, T.V.P., C.U.T.H. and J.F. interpreted the data and wrote the manuscript. V.D. and T.V.P. prepared the sample. Y.H., A.d.G. and R.A.G. performed the cryo-electron microscopy experiments. H.Y.L. performed the three-dimensional variance estimation. Y.H. and R.L. performed the cryo-electron microscopy data processing. Y.H. modelled the CSFV IRES. T.V.P., C.U.T.H. and J.F. directed research.

Corresponding authors

Correspondence to Christopher U. T. Hellen or Joachim Frank.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Comparison of HCV and CSFV IRES-bound ribosomal complexes.

a, Secondary structures of (left) the HCV IRES and (right) the CSFV IRES. Domain II of each IRES is indicated by a red dashed oval; elements of the pseudoknot and subdomains IIIa–IIIe are colour-coded as in Extended Data Fig. 6. b, Cryo-electron microscopy reconstructions of the HCV IRES bound to the rabbit 40S subunit at 20 Å resolution8 (left), the HCV IRES bound to the 40S subunit of cycloheximide-stalled human 80S ribosomes at 15 Å resolution23 (middle) (accession code EMD-1138) and the CSFV ΔII-IRES bound to the rabbit 40S subunits at 8.5 Å resolution (right) (this study). In all panels, the IRES-40S subunit is viewed from the solvent side; the 40S subunit is displayed in yellow and the IRES in cyan. The red dashed circles in left and middle panels show a discontinuity in the density of domain II in the HCV IRES bound to the 40S subunit compared to the HCV IRES bound to 80S ribosomes. The dashed circle in the right hand panel highlights CSFV IRES subdomain IIId2, which has no counterpart in the HCV IRES.

Extended Data Figure 2 Analysis of 40S–ΔII-IRES–eIF3–DHX29 complexes.

40S–ΔII-IRES–eIF3–DHX29 complexes were assembled in vitro using CSFV ΔII-IRES mRNA, native eIF2, eIF3 and 40S subunits purified from rabbit reticulocyte lysate and recombinant DHX29, and assayed by toeprinting. Lanes C, T, A and G show the cDNA sequence corresponding to CSFV ΔII-IRES mRNA. The position of the initiation codon is indicated on the left. This analysis revealed (lane 2) that deletion of domain II of the IRES or the presence of DHX29 did not influence IRES’s contacts with either 40S subunit (the toeprint stops at UUU387–389, G345 and C334) or eIF3 (the toeprint stops at A250) that have been previously observed2,18. Moreover, upon addition of the eIF2-TC, 40S–ΔII-IRES–eIF3–DHX29 complexes were quantitatively converted into 48S complexes on the authentic initiation codon AUG373 (lane 3). The low-efficiency 48S complex formation on the preceding AUG366 was also observed before and was not related to the presence of DHX2918. The gel reported in the figure is representative of results obtained from three technical replicates.

Extended Data Figure 3 Unsupervised three-dimensional classification of IRES-bound ribosomal complexes.

Unsupervised three-dimensional classification of IRES-bound ribosomal complexes identified 423,000 particles inconsistent with the known structure of the 40S subunit (rejects) and six well-populated classes containing complexes of the 40S subunit in a binary complex with the ΔII-IRES (class 1), of the 40S subunit bound to the ΔII-IRES and DHX29 (class 2), of the 40S subunit bound to DHX29 and eIF3 (class 3) and of the 40S subunit bound to the ΔII-IRES, DHX29 and eIF3 in orientation 1 (class 4), in orientation 2 (class 5) and in orientation 3 (class 6), viewed from (left) the back, (centre) the intersubunit side and (right) the solvent side.

Extended Data Figure 4 Measured resolution and reference-free two-dimensional classification of IRES-bound ribosomal complexes.

a, Gold-standard Fourier shell correlation (FSC) curves of the cryo-electron microscopy reconstruction of classes 2 (red line) and 4 (blue line) (also see Extended Data Fig. 3) indicating their estimated resolution. b, Right column on each side, two-dimensional classes obtained by reference-free classification of particles corresponding to 40S–eIF3–DHX29–ΔII-IRES complexes (class 4 in Extended Data Fig. 3). Middle column on each side, projection views of the class 4 cryo-electron microscopy map corresponding to the two-dimensional classes. Right column on each side, corresponding views of the segmented three-dimensional map coloured as in Fig. 1.

Extended Data Figure 5 Correspondence between individual subunits and anthropomorphic features of the eIF3 core complex and three-dimensional variance of the 40S–DHX29–ΔII-IRES–eIF3 map.

a, b, Front (upper panels) and back views (lower panels) of cryo-electron microscopy reconstructions of eIF3 as it appears in class 4 of the CSFV ΔII-IRES–40S–DHX29–eIF3 complex bound to the CSFV ΔII-IRES (a) and alone13 (b), labelled to show anthropomorphic terms12 and the localization of individual subunits in the core complex11,24. c, three-dimensional variance of the class 4 cryo-electron microscopy map, filtered to 20 Å, and coloured according to the computed three-dimensional variance (see Methods), from dark blue for the lowest variance to red for the highest variance. The map is filtered to the resolution at which the three-dimensional variance was estimated (20 Å).

Extended Data Figure 6 Comparison between the CSFV and the HCV pseudoknots.

Views of the structures of the HCV pseudoknot, from the 3.6 Å resolution crystal structure, with an additional crystallization module extending from helix III127 (a) (PDB: 3T4B) and the CSFV pseudoknot in the context of the 40S-subunit-bound ΔII-IRES (b) (this study) are shown in ribbon representation and coloured according to the scheme of the respective secondary structure diagrams (Extended Data Fig. 1a). c, d, Close-up views of HCV and CSFV pseudoknots, showing the ‘main’ helix, formed by helix III1 and pseudoknot (pk) stem 1A (in HCV) or helix III1, pk stem 1a and pk stem 1b (in CSFV), and the ‘sidecar’ helix, which contains subdomain IIIe, pk stem 2 and the two-base-pair helical segment of subdomain IIIf (see Extended Data Fig. 5a, b).

Extended Data Figure 7 Molecular interactions of the CSFV ΔII-IRES with the 40S subunit and interactions of eIF3 with the HCV and CSFV IRESs.

a, Secondary structure diagram of the CSFV ΔII-IRES, with nucleotides shown in different degrees of bold to show qualitatively their flexibility in the cryo-electron microscopy map (the more flexible, the bolder). Circled nucleotides interact with the indicated components of the 40S subunit. Ribosomal protein names and residue numbers are indicated according to the Tetrahymena thermophila 40S subunit42. bd, Secondary structure diagram of the apical region of domain III of the CSFV IRES (b, c) and the HCV IRES (d). b, Contacts of eIF3 with the IRES in the cryoEM map of the 40S–ΔII-IRES–eIF3 complex. c, d, Sites of strong protection of CSFV and HCV IRESs by native eIF3 from enzymatic cleavage and chemical modification, of protection of the HCV IRES by a 10-subunit form of eIF3 from 1M7 modification, or of interference with binding of eIF3 to the IRES by modification, as indicated in the keys5,9,11. Abbreviations: dimethyl sulphate (DMS), 1-cyclohexyl-3-(2-morpholinoethyl)carbodiimide metho-p-toluene sulphonate (CMCT), diethylpyrocarbonate (DEPC), 1-methyl-7-nitroisatoic anhydride (1M7). The inset panels show CSFV(c) and HCV IRESs (d), with helix III4 and subdomains IIIa, IIIb and IIIc in bold.

Extended Data Figure 8 Formation of elongation-competent 80S ribosomes on the HCV IRES depending on the presence of eIF3.

Toe-printing analysis of 48S initiation and 80S pre-termination complexes (pre-TC) assembled on the wild-type and ΔIIIb HCV (MSTN-STOP) mRNAs with translation components as indicated. The positions of the initiation and stop codons are shown on the left. Lanes C, T, A and G depict the cDNA sequence corresponding to the wild-type HCV (MSTN-STOP) mRNA. The gel reported in the figure is representative of results obtained from three technical replicates.

Extended Data Figure 9 Unsupervised three-dimensional classification protocol.

Details of the unsupervised three-dimensional classification. The classification included 6 rounds. For each round, the number of the particles included is indicated, as well as their percentages calculated over the full data set. The classes of rejected particles are crossed out in red and their percentages are indicated, also in red, as calculated over the full data set. Lines and brackets are drawn in different colours for clarity. Classes generated in rounds 3 to 6 are displayed and coloured by radial distance in Chimera UCSF43 in order to help in the visual discrimination of differences in features among the classes.

Supplementary information

Range of different orientations of eIF3 within the 40S•DHX29•∆II-IRES•eIF3 complex

This video displaying a morphing of density maps from classes 4 through 6 (Fig. S3), displaying eIF3 in three distinct orientations when bound on the IRES•40S complex. The complex is viewed from the top and the morphing illustrates the continuum of orientations sampled by eIF3 due to the flexibility of domain IIIb of the CSFV IRES. The video was made using Chimera UCSF47. (MP4 974 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hashem, Y., des Georges, A., Dhote, V. et al. Hepatitis-C-virus-like internal ribosome entry sites displace eIF3 to gain access to the 40S subunit. Nature 503, 539–543 (2013). https://doi.org/10.1038/nature12658

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature12658

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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