Translational regulation of rotavirus gene expression Free

Abstract

Rotavirus mRNAs are transcribed from 11 genomic dsRNA segments within a subviral particle. The mRNAs are extruded into the cytoplasm where they serve as mRNA for protein synthesis and as templates for packaging and replication into dsRNA. The molecular steps in the replication pathway that regulate the levels of viral gene expression are not well defined. We have investigated potential mechanisms of regulation of rotavirus gene expression by functional evaluation of two differentially expressed viral mRNAs. NSP1 (gene 5) and VP6 (gene 6) are expressed early in infection, and VP6 is expressed in excess over NSP1. We formulated the hypothesis that the amounts of NSP1 and VP6 were regulated by the translational efficiencies of the respective mRNAs. We measured the levels of gene 5 and gene 6 mRNA and showed that they were not significantly different, and protein analysis indicated no difference in stability of NSP1 compared with VP6. Polyribosome analysis showed that the majority of gene 6 mRNA was present on large polysomes. In contrast, sedimentation of more than half of the gene 5 mRNA was subpolysomal. The change in distribution of gene 5 mRNA in polyribosome gradients in response to treatment with low concentrations of cycloheximide suggested that gene 5 is a poor translation initiation template compared with gene 6 mRNA. These data define a regulatory mechanism for the difference in amounts of VP6 and NSP1 and provide evidence for post-transcriptional control of rotavirus gene expression mediated by the translational efficiency of individual viral mRNAs.

Loading

Article metrics loading...

/content/journal/jgv/10.1099/vir.0.18558-0
2003-02-01
2024-03-29
Loading full text...

Full text loading...

/deliver/fulltext/jgv/84/2/vir840383.html?itemId=/content/journal/jgv/10.1099/vir.0.18558-0&mimeType=html&fmt=ahah

References

  1. Allen A. M., Desselberger U. 1985; Reassortment of human rotaviruses carrying rearranged genomes with bovine rotavirus. J Gen Virol 66:2703–2714
    [Google Scholar]
  2. Bass D. M., Baylor M. R., Chen C., Mackow E. M., Bremont M., Greenberg H. B. 1992; Liposome-mediated transfection of intact viral particles reveals that plasma membrane penetration determines permissivity of tissue culture cells to rotavirus. J Clin Invest 90:2313–2320
    [Google Scholar]
  3. Biryahwaho B., Hundley F., Desselberger U. 1987; Bovine rotavirus with rearranged genome reassorts with human rotavirus. Brief report. Arch Virol 96:257–264
    [Google Scholar]
  4. Bremont M., Charpilienne A., Chabanne D., Cohen J. 1987; Nucleotide sequence and expression in Escherichia coli of the gene encoding the nonstructural protein NCVP2 of bovine rotavirus. Virology 161:138–144
    [Google Scholar]
  5. Brendler T., Godefroy-Colburn T., Carlill R. D., Thach R. E. 1981a; The role of mRNA competition in regulating translation. II. Development of a quantitative in vitro assay. J Biol Chem 256:11747–11754
    [Google Scholar]
  6. Brendler T., Godefroy-Colburn T., Yu S., Thach R. E. 1981b; The role of mRNA competition in regulating translation. III. Comparison of in vitro and in vivo results.. J Biol Chem 256:11755–11761
    [Google Scholar]
  7. Chizhikov V., Patton J. T. 2000; A four-nucleotide translation enhancer in the 3′-terminal consensus sequence of the nonpolyadenylated mRNAs of rotavirus. RNA 6:814–825
    [Google Scholar]
  8. Detjen B. M., Walden W. E., Thach R. E. 1982; Translational specificity in reovirus-infected mouse fibroblasts. J Biol Chem 257:9855–9860
    [Google Scholar]
  9. Ericson B. L., Graham D. Y., Mason B. B., Estes M. K. 1982; Identification, synthesis, and modifications of simian rotavirus SA11 polypeptides in infected cells. J Virol 42:825–839
    [Google Scholar]
  10. Estes M. K., Cohen J. 1989; Rotavirus gene structure and function. Microbiol Rev 53:410–449
    [Google Scholar]
  11. Estes M. K., Mason B. B., Crawford S., Cohen J. 1984; Cloning and nucleotide sequence of the simian rotavirus gene 6 that codes for the major inner capsid protein. Nucleic Acids Res 12:1875–1887
    [Google Scholar]
  12. Gallie D. R. 1996; Translational control of cellular and viral mRNAs. Plant Mol Biol 32:145–158
    [Google Scholar]
  13. Gallie D. R., Kobayashi M. 1994; The role of the 3′-untranslated region of non-polyadenylated plant viral mRNAs in regulating translational efficiency. Gene 142:159–165
    [Google Scholar]
  14. Geballe A. P., Sachs M. S. 2000; Translational control by upstream open reading frames. In Translational Control of Gene Expression pp  595–614 Edited by Sonenberg N., Hershey J. W., Mathews M. B. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory;
    [Google Scholar]
  15. Godefroy-Colburn T., Thach R. E. 1981; The role of mRNA competition in regulating translation. IV. Kinetic model. J Biol Chem 256:11762–11773
    [Google Scholar]
  16. Golini F., Thach S. S., Birge C. H., Safer B., Merrick W. C., Thach R. E. 1976; Competition between cellular and viral mRNAs in vitro is regulated by a messenger discriminatory initiation factor. Proc Natl Acad Sci U S A 73:3040–3044
    [Google Scholar]
  17. Hann L. E., Webb A. C., Cai J. M., Gehrke L. 1997; Identification of a competitive translation determinant in the 3′ untranslated region of alfalfa mosaic virus coat protein mRNA. Mol Cell Biol 17:2005–2013
    [Google Scholar]
  18. Hardy M. E., Woode G. N., Xu Z. C., Gorziglia M. 1991; Comparative amino acid sequence analysis of VP4 for VP7 serotype 6 bovine rotavirus strains NCDV, B641, and UK. J Virol 65:5535–5538
    [Google Scholar]
  19. Hardy M. E., Gorziglia M., Woode G. N. 1992; Amino acid sequence analysis of bovine rotavirus B223 reveals a unique outer capsid protein VP4 and confirms a third bovine VP4 type. Virology 191:291–300
    [Google Scholar]
  20. Hua J., Patton J. T. 1994; The carboxyl-half of the rotavirus nonstructural protein NS53 (NSP1) is not required for virus replication. Virology 198:567–576
    [Google Scholar]
  21. Hua J., Mansell E. A., Patton J. T. 1993; Comparative analysis of the rotavirus NS53 gene: conservation of basic and cysteine-rich regions in the protein and possible stem–loop structures in the RNA. Virology 196:372–378
    [Google Scholar]
  22. Hundley F., Biryahwaho B., Gow M., Desselberger U. 1985; Genome rearrangements of bovine rotavirus after serial passage at high multiplicity of infection. Virology 143:88–103
    [Google Scholar]
  23. Hundley F., McIntyre M., Clark B., Beards G., Wood D., Chrystie I., Desselberger U. 1987; Heterogeneity of genome rearrangements in rotaviruses isolated from a chronically infected immunodeficient child. J Virol 61:3365–3372
    [Google Scholar]
  24. Imai M., Akatani K., Ikegami N., Furuichi Y. 1983; Capped and conserved terminal structures in human rotavirus genome double-stranded RNA segments. J Virol 47:125–136
    [Google Scholar]
  25. Johnson M. A., McCrae M. A. 1989; Molecular biology of rotaviruses. VIII. Quantitative analysis of regulation of gene expression during virus replication. J Virol 63:2048–2055
    [Google Scholar]
  26. Kaspar R. L., Kakegawa T., Cranston H., Morris D. R., White M. W. 1992; A regulatory cis element and a specific binding factor involved in the mitogenic control of murine ribosomal protein L32 translation. J Biol Chem 267:508–514
    [Google Scholar]
  27. Kojima K., Taniguchi K., Kobayashi N. 1996; Species-specific and interspecies relatedness of NSP1 sequences in human, porcine, bovine, feline, and equine rotavirus strains. Arch Virol 141:1–12
    [Google Scholar]
  28. Kojima K., Taniguchi K., Kawagishi-Kobayashi M., Matsuno S., Urasawa S. 2000; Rearrangement generated in double genes, NSP1 and NSP3, of viable progenies from a human rotavirus strain. Virus Res 67:163–171
    [Google Scholar]
  29. Kozak M. 1984; Compilation and analysis of sequences upstream from the translational start site in eukaryotic mRNAs. Nucleic Acids Res 12:857–872
    [Google Scholar]
  30. Kozak M. 1987; An analysis of 5′-noncoding sequences from 699 vertebrate messenger RNAs. Nucleic Acids Res 15:8125–8148
    [Google Scholar]
  31. Kozak M. 1988; Leader length and secondary structure modulate mRNA function under conditions of stress. Mol Cell Biol 8:2737–2744
    [Google Scholar]
  32. Lawson T. G., Cladaras M. H., Ray B. K., Lee K. A., Abramson R. D., Merrick W. C., Thach R. E. 1988; Discriminatory interaction of purified eukaryotic initiation factors 4F plus 4A with the 5′ ends of reovirus messenger RNAs. J Biol Chem 263:7266–7276
    [Google Scholar]
  33. Lawton J. A., Estes M. K., Prasad B. V. 1997; Three-dimensional visualization of mRNA release from actively transcribing rotavirus particles [letter]. Nature Struct Biol 4:118–121
    [Google Scholar]
  34. Leathers V., Tanguay R., Kobayashi M., Gallie D. R. 1993; A phylogenetically conserved sequence within viral 3′ untranslated RNA pseudoknots regulates translation. Mol Cell Biol 13:5331–5347
    [Google Scholar]
  35. Lodish H. F. 1971; Alpha and beta globin messenger ribonucleic acid. Different amounts and rates of initiation of translation. J Biol Chem 246:7131–7138
    [Google Scholar]
  36. Mattion N. M., Mitchell D. B., Both G. W., Estes M. K. 1991; Expression of rotavirus proteins encoded by alternative open reading frames of genome segment 11. Virology 181:295–304
    [Google Scholar]
  37. McCrae M. A., McCorquodale J. G. 1983; Molecular biology of rotaviruses. V. Terminal structure of viral RNA species. Virology 126:204–212
    [Google Scholar]
  38. Patton J. T. 1995; Structure and function of the rotavirus RNA-binding proteins. J Gen Virol 76:2633–2644
    [Google Scholar]
  39. Patton J. T., Taraporewala Z., Chen D. 8 other authors 2001; Effect of intragenic rearrangement and changes in the 3′ consensus sequence on NSP1 expression and rotavirus replication. J Virol 75:2076–2086
    [Google Scholar]
  40. Pedley S., Hundley F., Chrystie I., McCrae M. A., Desselberger U. 1984; The genomes of rotaviruses isolated from chronically infected immunodeficient children. J Gen Virol 65:1141–1150
    [Google Scholar]
  41. Piron M., Vende P., Cohen J., Poncet D. 1998; Rotavirus RNA-binding protein NSP3 interacts with eIF4GI and evicts the poly(A) binding protein from eIF4F. EMBO J 17:5811–5821
    [Google Scholar]
  42. Poncet D., Aponte C., Cohen J. 1993; Rotavirus protein NSP3 (NS34) is bound to the 3′ end consensus sequence of viral mRNAs in infected cells. J Virol 67:3159–3165
    [Google Scholar]
  43. Poncet D., Laurent S., Cohen J. 1994; Four nucleotides are the minimal requirement for RNA recognition by rotavirus non-structural protein NSP3. EMBO J 13:4165–4173
    [Google Scholar]
  44. Prasad B. V., Chiu W. 1994; Structure of rotavirus. Curr Top Microbiol Immunol 185:9–29
    [Google Scholar]
  45. Ray A., Walden W. E., Brendler T., Zenger V. E., Thach R. E. 1985; Effect of medium hypertonicity on reovirus translation rates. An application of kinetic modeling in vivo. Biochemistry 24:7525–7532
    [Google Scholar]
  46. Sambrook J., Russell D. W. 2001 Molecular Cloning: a Laboratory Manual Cold Spring Harbor, NY: Cold Spring Harbor Laboratory;
    [Google Scholar]
  47. Tanguay R. L., Gallie D. R. 1996; Translational efficiency is regulated by the length of the 3′ untranslated region. Mol Cell Biol 16:146–156
    [Google Scholar]
  48. Taniguchi K., Kojima K., Urasawa S. 1996; Nondefective rotavirus mutants with an NSP1 gene which has a deletion of 500 nucleotides, including a cysteine-rich zinc finger motif-encoding region (nucleotides 156 to 248), or which has a nonsense codon at nucleotides 153–155. J Virol 70:4125–4130
    [Google Scholar]
  49. Vende P., Piron M., Castagne N., Poncet D. 2000; Efficient translation of rotavirus mRNA requires simultaneous interaction of NSP3 with the eukaryotic translation initiation factor eIF4G and the mRNA 3′ end. J Virol 74:7064–7071
    [Google Scholar]
  50. Walden W. E., Thach R. E. 1986; Translational control of gene expression in a normal fibroblast. Characterization of a subclass of mRNAs with unusual kinetic properties. Biochemistry 25:2033–2041
    [Google Scholar]
  51. Walden W. E., Godefroy-Colburn T., Thach R. E. 1981; The role of mRNA competition in regulating translation. I. Demonstration of competition in vivo.. J Biol Chem 256:11739–11746
    [Google Scholar]
  52. White M. W., Kameji T., Pegg A. E., Morris D. R. 1987; Increased efficiency of translation of ornithine decarboxylase mRNA in mitogen-activated lymphocytes. Eur J Biochem 170:87–92
    [Google Scholar]
  53. White M. W., Degnin C., Hill J., Morris D. R. 1990; Specific regulation by endogenous polyamines of translational initiation of S-adenosylmethionine decarboxylase mRNA in Swiss 3T3 fibroblasts. Biochem J 268:657–660
    [Google Scholar]
  54. Woode G. N., Kelso N. E., Simpson T. F., Gaul S. K., Evans L. E., Babiuk L. 1983; Antigenic relationships among some bovine rotaviruses: serum neutralization and cross-protection in gnotobiotic calves. J Clin Microbiol 18:358–364
    [Google Scholar]
  55. Zeyenko V. V., Ryabova L. A., Gallie D. R., Spirin A. S. 1994; Enhancing effect of the 3′-untranslated region of tobacco mosaic virus RNA on protein synthesis in vitro. FEBS Lett 354:271–273
    [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/jgv/10.1099/vir.0.18558-0
Loading
/content/journal/jgv/10.1099/vir.0.18558-0
Loading

Data & Media loading...

Most cited Most Cited RSS feed