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.

  • Article
  • Published:

A metazoan ortholog of SpoT hydrolyzes ppGpp and functions in starvation responses

Abstract

In nutrient-starved bacteria, RelA and SpoT proteins have key roles in reducing cell growth and overcoming stresses. Here we identify functional SpoT orthologs in metazoa (named Mesh1, encoded by HDDC3 in human and Q9VAM9 in Drosophila melanogaster) and reveal their structures and functions. Like the bacterial enzyme, Mesh1 proteins contain an active site for ppGpp hydrolysis and a conserved His-Asp–box motif for Mn2+ binding. Consistent with these structural data, Mesh1 efficiently catalyzes hydrolysis of guanosine 3′,5′-diphosphate (ppGpp) both in vitro and in vivo. Mesh1 also suppresses SpoT-deficient lethality and RelA-induced delayed cell growth in bacteria. Notably, deletion of Mesh1 (Q9VAM9) in Drosophila induces retarded body growth and impaired starvation resistance. Microarray analyses reveal that the amino acid–starved Mesh1 null mutant has highly downregulated DNA and protein synthesis–related genes and upregulated stress-responsible genes. These data suggest that metazoan SpoT orthologs have an evolutionarily conserved function in starvation responses.

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: Identification of SpoT homologs in metazoa.
Figure 2: Mesh1 compared with bacterial ppGpp hydrolase shows a conserved structure and similar enzymatic activity.
Figure 3: Mesh1 hydrolyzes ppGpp in vivo.
Figure 4: Mesh1 null mutant Drosophila shows retarded body growth and impaired starvation resistance.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Cashel, M. & Gallant, J. Two compounds implicated in the function of the RC gene of Escherichia coli. Nature 221, 838–841 (1969).

    Google Scholar 

  2. Cashel, M., Gentry, D.R., Hernandez, V.J. & Vinella, D. The stringent response. in Escherichia coli and Salmonella typhimurium. Cellular and Molecular Biology 1458–1496: (ASM Press, Washington, DC, 1996).

  3. Potrykus, K. & Cashel, M. (p)ppGpp: still magical? Annu. Rev. Microbiol. 62, 35–51 (2008).

    Google Scholar 

  4. Haseltine, W.A. & Block, R. Synthesis of guanosine tetra- and pentaphosphate requires the presence of a codon-specific, uncharged transfer ribonucleic acid in the acceptor site of ribosomes. Proc. Natl. Acad. Sci. USA 70, 1564–1568 (1973).

    Google Scholar 

  5. Battesti, A. & Bouveret, E. Acyl carrier protein/SpoT interaction, the switch linking SpoT-dependent stress response to fatty acid metabolism. Mol. Microbiol. 62, 1048–1063 (2006).

    Google Scholar 

  6. Metzger, S. et al. Characterization of the relA1 mutation and a comparison of relA1 with new relA null alleles in Escherichia coli. J. Biol. Chem. 264, 21146–21152 (1989).

    Google Scholar 

  7. Xiao, H. et al. Residual guanosine 3′,5′-bispyrophosphate synthetic activity of relA null mutants can be eliminated by spoT null mutations. J. Biol. Chem. 266, 5980–5990 (1991).

    Google Scholar 

  8. Laffler, T. & Gallant, J.A. Stringent control of protein synthesis in E. coli. Cell 3, 47–49 (1974).

    Google Scholar 

  9. Schreiber, G. et al. Overexpression of the relA gene in Escherichia coli. J. Biol. Chem. 266, 3760–3767 (1991).

    Google Scholar 

  10. Kasai, K. et al. A RelA-SpoT homolog (Cr-RSH) identified in Chlamydomonas reinhardtii generates stringent factor in vivo and localizes to chloroplasts in vitro. Nucleic Acids Res. 30, 4985–4992 (2002).

    Google Scholar 

  11. Givens, R.M. et al. Inducible expression, enzymatic activity, and origin of higher plant homologues of bacterial RelA/SpoT stress proteins in Nicotiana tabacum. J. Biol. Chem. 279, 7495–7504 (2004).

    Google Scholar 

  12. van der Biezen, E.A., Sun, J., Coleman, M.J., Bibb, M.J. & Jones, J.D. Arabidopsis RelA/SpoT homologs implicate (p)ppGpp in plant signaling. Proc. Natl. Acad. Sci. USA 97, 3747–3752 (2000).

    Google Scholar 

  13. Masuda, S. et al. The bacterial stringent response, conserved in chloroplasts, controls plant fertilization. Plant Cell Physiol. 49, 135–141 (2008).

    Google Scholar 

  14. Mizusawa, K., Masuda, S. & Ohta, H. Expression profiling of four RelA/SpoT-like proteins, homologues of bacterial stringent factors, in Arabidopsis thaliana. Planta 228, 553–562 (2008).

    Google Scholar 

  15. Tozawa, Y. et al. Calcium-activated (p)ppGpp synthetase in chloroplasts of land plants. J. Biol. Chem. 282, 35536–35545 (2007).

    Google Scholar 

  16. Braeken, K. et al. New horizons for (p)ppGpp in bacterial and plant physiology. Trends Microbiol. 14, 45–54 (2006).

    Google Scholar 

  17. Silverman, R.H. & Atherly, A.G. The search for guanosine tetraphosphate (ppGpp) and other unusual nucleotides in eucaryotes. Microbiol. Rev. 43, 27–41 (1979).

    Google Scholar 

  18. Mittenhuber, G. Comparative genomics and evolution of genes encoding bacterial (p)ppGpp synthetases/hydrolases (the Rel, RelA and SpoT proteins). J. Mol. Microbiol. Biotechnol. 3, 585–600 (2001).

    Google Scholar 

  19. Aravind, L. & Koonin, E.V. The HD domain defines a new superfamily of metal-dependent phosphohydrolases. Trends Biochem. Sci. 23, 469–472 (1998).

    Google Scholar 

  20. Rhee, H.W. et al. Selective fluorescent chemosensor for the bacterial alarmone (p)ppGpp. J. Am. Chem. Soc. 130, 784–785 (2008).

    Google Scholar 

  21. Gallant, J., Margason, G. & Finch, B. On the turnover of ppGpp in Escherichia coli. J. Biol. Chem. 247, 6055–6058 (1972).

    Google Scholar 

  22. Takahashi, K., Kasai, K. & Ochi, K. Identification of the bacterial alarmone guanosine 5′-diphosphate 3′-diphosphate (ppGpp) in plants. Proc. Natl. Acad. Sci. USA 101, 4320–4324 (2004).

    Google Scholar 

  23. Guranowski, A. Metabolism of diadenosine tetraphosphate (Ap4A) and related nucleotides in plants; review with historical and general perspective. Front. Biosci. 9, 1398–1411 (2004).

    Google Scholar 

  24. Holm, L. & Sander, C. Protein structure comparison by alignment of distance matrices. J. Mol. Biol. 233, 123–138 (1993).

    Google Scholar 

  25. Zimmerman, M.D., Proudfoot, M., Yakunin, A. & Minor, W. Structural insight into the mechanism of substrate specificity and catalytic activity of an HD-domain phosphohydrolase: the 5′-deoxyribonucleotidase YfbR from Escherichia coli. J. Mol. Biol. 378, 215–226 (2008).

    Google Scholar 

  26. Xu, R.X. et al. Atomic structure of PDE4: insights into phosphodiesterase mechanism and specificity. Science 288, 1822–1825 (2000).

    Google Scholar 

  27. Mechold, U. et al. Functional analysis of a relA/spoT gene homolog from Streptococcus equisimilis. J. Bacteriol. 178, 1401–1411 (1996).

    Google Scholar 

  28. Hogg, T. et al. Conformational antagonism between opposing active sites in a bifunctional RelA/SpoT homolog modulates (p)ppGpp metabolism during the stringent response. Cell 117, 57–68 (2004).

    Google Scholar 

  29. Avarbock, D., Avarbock, A. & Rubin, H. Differential regulation of opposing RelMtb activities by the aminoacylation state of a tRNA.ribosome.mRNA.RelMtb complex. Biochemistry 39, 11640–11648 (2000).

    Google Scholar 

  30. Wehmeier, L. et al. A Corynebacterium glutamicum mutant with a defined deletion within the rplK gene is impaired in (p)ppGpp accumulation upon amino acid starvation. Microbiology 147, 691–700 (2001).

    Google Scholar 

  31. Song, M. et al. ppGpp-dependent stationary phase induction of genes on Salmonella pathogenicity island 1. J. Biol. Chem. 279, 34183–34190 (2004).

    Google Scholar 

  32. Artsimovitch, I. et al. Structural basis for transcription regulation by alarmone ppGpp. Cell 117, 299–310 (2004).

    Google Scholar 

  33. Paul, B.J., Ross, W., Gaal, T. & Gourse, R.L. rRNA transcription in Escherichia coli. Annu. Rev. Genet. 38, 749–770 (2004).

    Google Scholar 

  34. Magnusson, L.U., Farewell, A. & Nyström, T. ppGpp: a global regulator in Escherichia coli. Trends Microbiol. 13, 236–242 (2005).

    Google Scholar 

  35. Durfee, T. et al. Transcription profiling of the stringent response in Escherichia coli. J. Bacteriol. 190, 1084–1096 (2008).

    Google Scholar 

  36. Eymann, C., Homuth, G., Scharf, C. & Hecker, M. Bacillus subtilis functional genomics: global characterization of the stringent response by proteome and transcriptome analysis. J. Bacteriol. 184, 2500–2520 (2002).

    Google Scholar 

  37. Britton, J.S. & Edgar, B.A. Environmental control of the cell cycle in Drosophila: nutrition activates mitotic and endoreplicative cells by distinct mechanisms. Development 125, 2149–2158 (1998).

    Google Scholar 

  38. Beissbarth, T. & Speed, T.P. GOstat: find statistically overrepresented Gene Ontologies within a group of genes. Bioinformatics 20, 1464–1465 (2004).

    Google Scholar 

  39. Palanker, L., Tennessen, J.M., Lam, G. & Thummel, C.S. Drosophila HNF4 regulates lipid mobilization and beta-oxidation. Cell Metab. 9, 228–239 (2009).

    Google Scholar 

  40. Zinke, I. et al. Nutrient control of gene expression in Drosophila: microarray analysis of starvation and sugar-dependent response. EMBO J. 21, 6162–6173 (2002).

    Google Scholar 

  41. Jones, G.H. & Bibb, M.J. Guanosine pentaphosphate synthetase from Streptomyces antibioticus is also a polynucleotide phosphorylase. J. Bacteriol. 178, 4281–4288 (1996).

    Google Scholar 

  42. Hoyt, S. & Jones, G.H. relA is required for actinomycin production in Streptomyces antibioticus. J. Bacteriol. 181, 3824–3829 (1999).

    Google Scholar 

  43. Mukai, J., Hirashima, A. & Mikuniya, T. Nucleotide 2′,3′-cyclic monophosphokinase from actinomycetes. Nucleic Acids Symp. Ser. 22, 89–90 (1980).

    Google Scholar 

  44. Muta, S. et al. Streptomyces ATP nucleotide 3′-pyrophosphokinase and its gene. Nucleic Acids Symp. Ser. 27, 165–166 (1992).

    Google Scholar 

  45. Sawaya, M.R. et al. Crystal structures of human DNA polymerase beta complexed with gapped and nicked DNA: evidence for an induced fit mechanism. Biochemistry 36, 11205–11215 (1997).

    Google Scholar 

  46. Ochi, K., Kandala, J.C. & Freese, E. Initiation of Bacillus subtilis sporulation by the stringent response to partial amino acid deprivation. J. Biol. Chem. 256, 6866–6875 (1981).

    Google Scholar 

  47. Ochi, K. Streptomyces relC mutants with an altered ribosomal protein ST-L11 and genetic analysis of a Streptomyces griseus relC mutant. J. Bacteriol. 172, 4008–4016 (1990).

    Google Scholar 

  48. Rhaese, H.J. Studies on the control of development synthesis of regulatory nucleotides, HPN and MS, in mammalian cells in tissue cultures. FEBS Lett. 53, 113–118 (1975).

    Google Scholar 

  49. Thammana, P., Buerk, R.R. & Gordon, J. Absence of ppGpp production in synchronised Balb/C mouse 3T3 cells on isoleucine starvation. FEBS Lett. 68, 187–190 (1976).

    Google Scholar 

  50. Park, J. et al. Mitochondrial dysfunction in Drosophila PINK1 mutants is complemented by parkin. Nature 441, 1157–1161 (2006).

    Google Scholar 

  51. Gassner, N.C. & Matthews, B.W. Use of differentially substituted selenomethionine proteins in X-ray structure determination. Acta Crystallogr. D Biol. Crystallogr. 55, 1967–1970 (1999).

    Google Scholar 

  52. Ohmura, T., Ueda, T., Hashimoto, Y. & Imoto, T. Tolerance of point substitution of methionine for isoleucine in hen egg white lysozyme. Protein Eng. 14, 421–425 (2001).

    Google Scholar 

  53. Terwilliger, T.C. & Berendzen, J. Automated MAD and MIR structure solution. Acta Crystallogr. D Biol. Crystallogr. 55, 849–861 (1999).

    Google Scholar 

  54. Terwilliger, T.C. Maximum-likelihood density modification. Acta Crystallogr. D Biol. Crystallogr. 56, 965–972 (2000).

    Google Scholar 

  55. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).

    Google Scholar 

  56. Murshudov, G.N., Vagin, A.A. & Dodson, E.J. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr. D Biol. Crystallogr. 53, 240–255 (1997).

    Google Scholar 

  57. Brünger, A.T. et al. Crystallography & NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr. D Biol. Crystallogr. 54, 905–921 (1998).

    Google Scholar 

  58. Collaborative Computational Project, Number 4. The CCP4 suite: programs for protein crystallography. Acta Crystallogr. D Biol. Crystallogr. 50, 760–763 (1994).

  59. Laskowski, R.A., MacArthur, M.W., Moss, D.S. & Thornton, J.M. PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Cryst. 26, 283–291 (1993).

    Google Scholar 

Download references

Acknowledgements

This research was supported by grants from the Korean National Creative Research Initiatives Program (2010-0018291) to J.C. and from the NMR Research Program of Korea Basic Science Institute to Y.H.J. G.L. was supported by the Priority Research Centers Program through the National Research Foundation of Korea, funded by the Ministry of Education, Science and Technology (2008-005-J00203). Y.K. and B.Y.K. were supported by World Class Institute Program of the National Research Foundation of Korea, funded by the Ministry of Education, Science and Technology. Wild-type (CF1648) and ΔrelA ΔspoT (CF1693) E. coli were kindly provided by M. Cashel (US National Institutes of Health).

Author information

Authors and Affiliations

Authors

Contributions

D.S. determined the structures and performed HPLC and MS analyses. G.L. and J.H.L. performed Drosophila and bacteria experiments. H.-Y.K. and K.-J.K. helped to determine the structures. H.-W.R. and J.-I.H. prepared PyDPA analyses. S.-Y.P. analyzed nucleotides from larval extracts. Y.K. performed mammalian cell culture experiments. All authors, including B.Y.K., C.P., H.E.C. and J.H.K. discussed the experimental results and the text of the manuscript. D.S., G.L., J.H.L., Y.H.J. and J.C. prepared the manuscript. J.C. and Y.H.J. developed the original research idea and supervised the research project.

Corresponding authors

Correspondence to Young Ho Jeon or Jongkyeong Chung.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6, Supplementary Tables 1–2 and Supplementary Methods (PDF 827 kb)

Supplementary Table 3

Differentially expressed genes in response to amino acid starvation in wild-type and Mesh1 null mutant flies. (ac) GO stat analysis was performed for the downregulated and upregulated genes in response to amino acid starvation in dMesh1 null mutant (b,c) or wild-type flies (a). The affected GO categories and the P value of match are indicated. Gene name and the fold change of each gene included in each GO category are indicated. (XLS 71 kb)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Sun, D., Lee, G., Lee, J. et al. A metazoan ortholog of SpoT hydrolyzes ppGpp and functions in starvation responses. Nat Struct Mol Biol 17, 1188–1194 (2010). https://doi.org/10.1038/nsmb.1906

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.1906

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