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:

Functionally uncoupled transcription–translation in Bacillus subtilis

Abstract

Tight coupling of transcription and translation is considered a defining feature of bacterial gene expression1,2. The pioneering ribosome can both physically associate and kinetically coordinate with RNA polymerase (RNAP)3,4,5,6,7,8,9,10,11, forming a signal-integration hub for co-transcriptional regulation that includes translation-based attenuation12,13 and RNA quality control2. However, it remains unclear whether transcription–translation coupling—together with its broad functional consequences—is indeed a fundamental characteristic of bacteria other than Escherichia coli. Here we show that RNAPs outpace pioneering ribosomes in the Gram-positive model bacterium Bacillus subtilis, and that this ‘runaway transcription’ creates alternative rules for both global RNA surveillance and translational control of nascent RNA. In particular, uncoupled RNAPs in B. subtilis explain the diminished role of Rho-dependent transcription termination, as well as the prevalence of mRNA leaders that use riboswitches and RNA-binding proteins. More broadly, we identified widespread genomic signatures of runaway transcription in distinct phyla across the bacterial domain. Our results show that coupled RNAP–ribosome movement is not a general hallmark of bacteria. Instead, translation-coupled transcription and runaway transcription constitute two principal modes of gene expression that determine genome-specific regulatory mechanisms in prokaryotes.

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

Fig. 1: Fast RNAP movement results in runaway transcription.
Fig. 2: Lack of translational control on transcription.
Fig. 3: Signals of Rho-dependent termination.
Fig. 4: Phylogenomic distribution of uncoupling.

Similar content being viewed by others

Data availability

All data generated and analysed during this study are included in this published article (and its Supplementary Information). The high-throughput sequencing datasets analysed during the current study are available from the Gene Expression Omnibus repository with accession numbers: GSE53767, GSE95211 and GSE108295 (see ‘High-throughput expression datasets used’ for details). Uncropped gel source data for northern blots can be found in Supplementary Fig. 1Source data are provided with this paper.

Code availability

Scripts for terminator identification have been deposited to GitHub (https://github.com/jblalanne/intrinsic_trx_terminator_identifier). Core Rend-seq analysis scripts used can be found on Github (https://github.com/jblalanne/Rend_seq_core_scripts). Other custom scripts used for data analysis are available upon request.

References

  1. Adhya, S. & Gottesman, M. Control of transcription termination. Annu. Rev. Biochem. 47, 967–996 (1978).

    Article  CAS  PubMed  Google Scholar 

  2. Richardson, J. P. Preventing the synthesis of unused transcripts by Rho factor. Cell 64, 1047–1049 (1991).

    Article  CAS  PubMed  Google Scholar 

  3. Landick, R., Carey, J. & Yanofsky, C. Translation activates the paused transcription complex and restores transcription of the trp operon leader region. Proc. Natl Acad. Sci. USA 82, 4663–4667 (1985).

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  4. Proshkin, S., Rahmouni, A. R., Mironov, A. & Nudler, E. Cooperation between translating ribosomes and RNA polymerase in transcription elongation. Science 328, 504–508 (2010).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  5. Burmann, B. M. B. et al. A NusE:NusG complex links transcription and translation. Science 328, 501–504 (2010).

    Article  CAS  PubMed  ADS  Google Scholar 

  6. Kohler, R., Mooney, R. A., Mills, D. J., Landick, R. & Cramer, P. Architecture of a transcribing-translating expressome. Science 356, 194–197 (2017).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  7. Fan, H. et al. Transcription-translation coupling: direct interactions of RNA polymerase with ribosomes and ribosomal subunits. Nucleic Acids Res. 45, 11043–11055 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Zhu, M., Mori, M., Hwa, T. & Dai, X. Disruption of transcription-translation coordination in Escherichia coli leads to premature transcriptional termination. Nat. Microbiol. 4, 2347–2356 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Webster, M. W. et al. Structural basis of transcription-translation coupling and collision in bacteria. Science https://doi.org/10.1126/science.abb5036 (2020).

  10. Wang, C. et al. Structural basis of transcription-translation coupling. Science https://doi.org/10.1126/science.abb5317 (2020).

  11. O’Reilly, F. J. et al. In-cell architecture of an actively transcribing-translating expressome. Science 369, 554–557 (2020).

    Article  PubMed  ADS  CAS  PubMed Central  Google Scholar 

  12. Roland, K. L., Liu, C. G. & Turnbough, C. L., Jr. Role of the ribosome in suppressing transcriptional termination at the pyrBI attenuator of Escherichia coli K-12. Proc. Natl Acad. Sci. USA 85, 7149–7153 (1988).

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  13. Yanofsky, C. Attenuation in the control of expression of bacterial operons. Nature 289, 751–758 (1981).

    Article  CAS  PubMed  ADS  Google Scholar 

  14. Landick, R., Turnbough, C. & Yanofsky, C. in Escherichia coli and Salmonella typhimurium: Cellular and Molecular Biology (eds. Neidhardt, F. et al.) 1276–1301 (American Society for Microbiology, 1996).

  15. Kervestin, S. & Jacobson, A. NMD: a multifaceted response to premature translational termination. Nat. Rev. Mol. Cell Biol. 13, 700–712 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Peters, J. M. et al. Rho and NusG suppress pervasive antisense transcription in Escherichia coli. Genes Dev. 26, 2621–2633 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Henkin, T. M. Control of transcription termination in prokaryotes. Annu. Rev. Genet. 30, 35–57 (1996).

    Article  CAS  PubMed  Google Scholar 

  18. Winkler, W. C. & Breaker, R. R. Regulation of bacterial gene expression by riboswitches. Annu. Rev. Microbiol. 59, 487–517 (2005).

    Article  CAS  PubMed  Google Scholar 

  19. Babitzke, P. & Yanofsky, C. Reconstitution of Bacillus subtilis trp attenuation in vitro with TRAP, the trp RNA-binding attenuation protein. Proc. Natl Acad. Sci. USA 90, 133–137 (1993).

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  20. Ingham, C. J., Dennis, J. & Furneaux, P. A. Autogenous regulation of transcription termination factor Rho and the requirement for Nus factors in Bacillus subtilis. Mol. Microbiol. 31, 651–663 (1999).

    Article  CAS  PubMed  Google Scholar 

  21. Shimotsu, H. & Henner, D. J. Construction of a single-copy integration vector and its use in analysis of regulation of the trp operon of Bacillus subtilis. Gene 43, 85–94 (1986).

    Article  CAS  PubMed  Google Scholar 

  22. Yakhnin, H., Babiarz, J. E., Yakhnin, A. V. & Babitzke, P. Expression of the Bacillus subtilis trpEDCFBA operon is influenced by translational coupling and Rho termination factor. J. Bacteriol. 183, 5918–5926 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Schleif, R., Hess, W., Finkelstein, S. & Ellis, D. Induction kinetics of the l-arabinose operon of Escherichia coli. J. Bacteriol. 115, 9–14 (1973).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Vogel, U. & Jensen, K. F. The RNA chain elongation rate in Escherichia coli depends on the growth rate. J. Bacteriol. 176, 2807–2813 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Dai, X. et al. Reduction of translating ribosomes enables Escherichia coli to maintain elongation rates during slow growth. Nat. Microbiol. 2, 16231 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Artsimovitch, I., Svetlov, V., Anthony, L., Burgess, R. R. & Landick, R. RNA polymerases from Bacillus subtilis and Escherichia coli differ in recognition of regulatory signals in vitro. J. Bacteriol. 182, 6027–6035 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Li, R., Zhang, Q., Li, J. & Shi, H. Effects of cooperation between translating ribosome and RNA polymerase on termination efficiency of the Rho-independent terminator. Nucleic Acids Res. 44, 2554–2563 (2016).

    Article  PubMed  Google Scholar 

  28. Wright, J. J. & Hayward, R. S. Transcriptional termination at a fully rho-independent site in Escherichia coli is prevented by uninterrupted translation of the nascent RNA. EMBO J. 6, 1115–1119 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Reilman, E., Mars, R. A. T., van Dijl, J. M. & Denham, E. L. The multidrug ABC transporter BmrC/BmrD of Bacillus subtilis is regulated via a ribosome-mediated transcriptional attenuation mechanism. Nucleic Acids Res. 42, 11393–11407 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Dar, D. et al. Term-seq reveals abundant ribo-regulation of antibiotics resistance in bacteria. Science 352, aad9822 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Yakhnin, H. et al. NusG-dependent RNA polymerase pausing and tylosin-dependent ribosome stalling are required for tylosin resistance by inducing 23s rRNA methylation in Bacillus subtilis. MBio 10, 1–14 (2019).

    Article  Google Scholar 

  32. Goodson, J. R., Klupt, S., Zhang, C., Straight, P. & Winkler, W. C. LoaP is a broadly conserved antiterminator protein that regulates antibiotic gene clusters in Bacillus amyloliquefaciens. Nat. Microbiol. 2, 17003 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Garza-Sánchez, F., Schaub, R. E., Janssen, B. D. & Hayes, C. S. tmRNA regulates synthesis of the ArfA ribosome rescue factor. Mol. Microbiol. 80, 1204–1219 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Shimokawa-Chiba, N. et al. Release factor-dependent ribosome rescue by BrfA in the Gram-positive bacterium Bacillus subtilis. Nat. Commun. 10, 5397 (2019).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  35. Lalanne, J. B. et al. Evolutionary convergence of pathway-specific enzyme expression stoichiometry. Cell 173, 749–761.e38 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Gowrishankar, J. & Harinarayanan, R. Why is transcription coupled to translation in bacteria? Mol. Microbiol. 54, 598–603 (2004).

    Article  CAS  PubMed  Google Scholar 

  37. Nicolas, P. et al. Condition-dependent transcriptome reveals high-level regulatory architecture in Bacillus subtilis. Science 335, 1103–1106 (2012).

    Article  CAS  PubMed  ADS  Google Scholar 

  38. Nishida, M., Mine, Y., Matsubara, T., Goto, S. & Kuwahara, S. Bicyclomycin, a new antibiotic. 3. In vitro and in vivo antimicrobial activity. J. Antibiot. (Tokyo) 25, 582–593 (1972).

    Article  CAS  Google Scholar 

  39. D’Heygère, F., Rabhi, M. & Boudvillain, M. Phyletic distribution and conservation of the bacterial transcription termination factor Rho. Microbiology 159, 1423–1436 (2013).

    Article  PubMed  CAS  Google Scholar 

  40. Harwood, C. R. & Cutting, S. M. Molecular Biological Methods for Bacillus (John Wiley, 1990).

  41. Li, G.-W., Burkhardt, D., Gross, C. & Weissman, J. S. Quantifying absolute protein synthesis rates reveals principles underlying allocation of cellular resources. Cell 157, 624–635 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. DeLoughery, A., Lalanne, J.-B., Losick, R. & Li, G.-W. Maturation of polycistronic mRNAs by the endoribonuclease RNase Y and its associated Y-complex in Bacillus subtilis. Proc. Natl Acad. Sci. USA 115, E5585–E5594 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Zhu, M., Dai, X. & Wang, Y.-P. Real time determination of bacterial in vivo ribosome translation elongation speed based on LacZα complementation system. Nucleic Acids Res. 44, gkw698 (2016).

    Article  CAS  Google Scholar 

  44. Bonekamp, F., Clemmesen, K., Karlström, O. & Jensen, K. F. Mechanism of UTP-modulated attenuation at the pyrE gene of Escherichia coli: an example of operon polarity control through the coupling of translation to transcription. EMBO J. 3, 2857–2861 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Abe, H., Abo, T. & Aiba, H. Regulation of intrinsic terminator by translation in Escherichia coli: transcription termination at a distance downstream. Genes Cells 4, 87–97 (1999).

    Article  CAS  PubMed  Google Scholar 

  46. Unniraman, S., Prakash, R. & Nagaraja, V. Alternate paradigm for intrinsic transcription termination in eubacteria. J. Biol. Chem. 276, 41850–41855 (2001).

    Article  CAS  PubMed  Google Scholar 

  47. Kobayashi, K., Kuwana, R. & Takamatsu, H. kinA mRNA is missing a stop codon in the undomesticated Bacillus subtilis strain ATCC 6051. Microbiology 154, 54–63 (2008).

    Article  CAS  PubMed  Google Scholar 

  48. Guérout-Fleury, A. M., Shazand, K., Frandsen, N. & Stragier, P. Antibiotic-resistance cassettes for Bacillus subtilis. Gene 167, 335–336 (1995).

    Article  PubMed  Google Scholar 

  49. Rackham, O. & Chin, J. W. A network of orthogonal ribosome x mRNA pairs. Nat. Chem. Biol. 1, 159–166 (2005).

    Article  CAS  PubMed  Google Scholar 

  50. Chung, C. T., Niemela, S. L. & Miller, R. H. One-step preparation of competent Escherichia coli: transformation and storage of bacterial cells in the same solution. Proc. Natl Acad. Sci. USA 86, 2172–2175 (1989).

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  51. Peters, J. M., Vangeloff, A. D. & Landick, R. Bacterial transcription terminators: the RNA 3′-end chronicles. J. Mol. Biol. 412, 793–813 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bidnenko, V. et al. Termination factor Rho: from the control of pervasive transcription to cell fate determination in Bacillus subtilis. PLoS Genet. 13, e1006909 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Sáenz-Lahoya, S. et al. Noncontiguous operon is a genetic organization for coordinating bacterial gene expression. Proc. Natl Acad. Sci. USA 116, 1733–1738 (2019).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  54. Sesto, N., Wurtzel, O., Archambaud, C., Sorek, R. & Cossart, P. The excludon: a new concept in bacterial antisense RNA-mediated gene regulation. Nat. Rev. Microbiol. 11, 75–82 (2013).

    Article  CAS  PubMed  Google Scholar 

  55. Yan, B., Boitano, M., Clark, T. A. & Ettwiller, L. SMRT-Cappable-seq reveals complex operon variants in bacteria. Nat. Commun. 9, 3676 (2018).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  56. Albertini, A. M. & Galizzi, A. The sequence of the trp operon of Bacillus subtilis 168 (trpC2) revisited. Microbiology 145, 3319–3320 (1999).

    Article  CAS  PubMed  Google Scholar 

  57. Abe, K. et al. Developmentally-regulated excision of the SPβ prophage reconstitutes a gene required for spore envelope maturation in Bacillus subtilis. PLoS Genet. 10, e1004636 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Stragier, P., Kunkel, B., Kroos, L. & Losick, R. Chromosomal rearrangement generating a composite gene for a developmental transcription factor. Science 243, 507–512 (1989).

    Article  CAS  PubMed  ADS  Google Scholar 

  59. Ciampi, M. S. Rho-dependent terminators and transcription termination. Microbiology 152, 2515–2528 (2006).

    Article  CAS  PubMed  Google Scholar 

  60. O’Leary, N.A. et al. Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation. Nucleic Acids Res. 44, D733–D745 (2016).

    Article  PubMed  CAS  Google Scholar 

  61. Lorenz, R. et al. ViennaRNA Package 2.0. Algorithms Mol. Biol. 6, 26 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Gusarov, I. & Nudler, E. The mechanism of intrinsic transcription termination. Mol. Cell 3, 495–504 (1999).

    Article  CAS  PubMed  Google Scholar 

  63. Sipos, K., Szigeti, R., Dong, X. & Turnbough, C. L. Jr. Systematic mutagenesis of the thymidine tract of the pyrBI attenuator and its effects on intrinsic transcription termination in Escherichia coli. Mol. Microbiol. 66, 127–138 (2007).

    Article  CAS  PubMed  Google Scholar 

  64. Simão, F. A., Waterhouse, R. M., Ioannidis, P., Kriventseva, E. V. & Zdobnov, E. M. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31, 3210–3212 (2015).

    Article  PubMed  CAS  Google Scholar 

  65. Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  67. Revell, L. J. phytools: an R package for phylogenetic comparative biology (and other things). Methods Ecol. Evol. 3, 217–223 (2012).

    Article  Google Scholar 

  68. Parks, D. H. et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat. Biotechnol. 36, 996–1004 (2018).

    Article  CAS  PubMed  Google Scholar 

  69. The UniProt Consortium. UniProt: the universal protein knowledgebase. Nucleic Acids Res. 45 (D1), D158–D169 (2017).

    Article  CAS  Google Scholar 

  70. Lane, W. J. & Darst, S. A. Molecular evolution of multisubunit RNA polymerases: sequence analysis. J. Mol. Biol. 395, 671–685 (2010).

    Article  CAS  PubMed  Google Scholar 

  71. Yang, X. et al. The structure of bacterial RNA polymerase in complex with the essential transcription elongation factor NusA. EMBO Rep. 10, 997–1002 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank J. Taggart, M. Tien, and A. Grossman’s laboratory for providing plasmids; D. McCormick for help with strain generation; L. Herzel, J. Taggart, and M. Tien for help collecting cultures for measurements of induction kinetics; members of the G.-W.L. and A. Grossman laboratories for discussions; L. Herzel, D. J. Parker, A. Grossman, J. Peters and V. Siegel for comments on the manuscript; and members of the BioMicroCenter at MIT for help in performing DNA sequencing. This research was supported by NIH grant R35GM124732, the Pew Biomedical Scholars Program, a Sloan Research Fellowship, the Searle Scholars Program, the Smith Family Award for Excellence in Biomedical Research, an NSF graduate research fellowship (to G.E.J.), an NIH Pre-Doctoral Training Grant (T32 GM007287, to G.E.J. and M.L.P.), an NSERC graduate fellowship (to J.-B.L.), and an HHMI International Student Fellowship (to J.-B.L.).

Author information

Authors and Affiliations

Authors

Contributions

G.E.J., J.-B.L. and G.-W.L. designed experiments; G.E.J. performed induction kinetic experiments, performed Rho and polarity experiments, and analysed sequence features of Rho target RNAs; J.-B.L performed ORF extension experiments, identified intrinsic terminators from Rend-seq data, analysed nested antisense RNAs and expressed pseudogenes, and wrote the phylogenomic bioinformatic terminator identification pipeline; M.L.P. performed induction kinetic experiments in knockout backgrounds; and G.E.J., J.-B.L. and G.-W.L. wrote the manuscript.

Corresponding author

Correspondence to Gene-Wei Li.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Joseph Puglisi and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Transcription and translation kinetics in slow growth.

Induction time course of lacZ mRNA (top) and protein (bottom) as in Fig. 1b, d for WT B. subtilis grown in MOPS minimal media + 0.4% maltose (growth rate 0.65 h−1). Lines indicate linear fits after signals rise. Uncertainties are standard error of the mean (s.e.m.) among biological replicates (2).

Source data

Extended Data Fig. 2 Validation of β-gal assay.

a, Measurement of linear range of microplate reader. Fluorescence relative to input of dilutions of an induced culture of GLB503 (full-length lacZ) at steady-state (Methods). b, Effect of different stop solutions on stopping translation. Induction time courses of pycA-lacZα protein collected into a stop solution containing chloramphenicol and erythromycin (grey, all plots, from Fig. 1d) or with either flash freezing in liquid nitrogen (top), 15 μl toluene added to the stop solution (middle), or 50 μl 12.5 mg/ml lincomycin added to the stop solution (bottom), shown in red in each plot (as described in Methods). Lines indicate linear fits after signals rise and τTL is indicated. c, Induction time course of truncated pycA-lacZα mRNA (top) and protein (bottom) as in Fig. 1b, d. Lines indicate linear fits after signals rise. Uncertainties are standard error of the mean (s.e.m.) among biological replicates (2).

Source data

Extended Data Fig. 3 Contribution of non-essential RNAP subunits and transcription factors to fast transcription.

Induction time course of pycA-lacZα mRNA in various mutant backgrounds as in Fig. 1b, d. Time course of the same construct in WT from Fig. 1d also shown for reference. Lines indicate linear fits after signals rise. Uncertainties are standard error of the mean (s.e.m) among biological replicates (1 for ∆ykzG and 2 for all others). Time of appearance of full-length mRNA in mutants is not substantially different than that measured in WT (Supplementary Discussion SN4).

Source data

Extended Data Fig. 4 Phylogenetic distribution of domain architecture for NusG, NusA and RpoB. A.

Multiple sequence alignments (Methods) for NusA (602 columns), NusG (325 columns), and the β subunit of the RNAP RpoB (1732 columns) for species shown in Fig. 4. The alignments are visualized in a binary fashion to highlight presence/absence of certain domains: white indicates presence of an amino acid in the alignment, and black indicates presence of a gap. The alignments were trimmed by removing columns with >95% gaps. Species with no homologues, partial or pseudogene homologues, or multiple homologues are shown as grey lines. Phylogenetic tree and fraction of terminators with stop-to-stem distances within 12 nt from Fig. 4 are reproduced in linearized form. The position of domains from the E. coli protein are identified by bars above the alignments. For RpoB, conserved bacterial regions identified by70 (βb1 to βb16) are shown. The NusA C-terminal domain11,71 (orange box) is missing in a large fraction of Firmicutes (partly present in Mollicutes, which include Mycoplasma and Spiroplasma; red brace), Campylobacterota, Thermotogota, Fusobacteria, and Actinobacteria. NusG has a largely conserved domain architecture, with Actinobacteria showing N-terminal extension. As previously noted in detail70, the β subunit of the RNAP has multiple insertion domains in diverse bacteria. Insertion domain βSI2, recently implicated10 (green box) in transcription-translation coupling is lineage-specific and absent in many clades of Gram-positive bacteria, as noted in10. Dashed box in tree highlights clade containing Mycoplasma. b. Close-up view of our analysis of the clade containing Mycoplasma (indicated by black dots). Sub-tree includes species with n ≤ 20 identified terminators (marked in light red). Grayscale representation of stop-to-stem distributions and fraction of terminators with d ≤ 12 nt are the same as Fig. 4. M. pneumoniae is highlighted in cyan, and has no identified terminator (0/14) with d ≤ 12 nt. c. Cumulative distribution of stop-to-stem distance for bioinformatically identified terminators in M. pneumoniae.

Source data

Extended Data Fig. 5 Details of ORF extension constructs and transcription terminator readthrough vs. stop-to-stem distances.

a, Sequence for terminators T1 and T2 for three variants (T1+: pupG original terminator, T1-: disrupted pupG terminator, ORF extension: original pupG with upstream ORF extended inside the loop of the terminator). For T1 and T2, blue and grey shading, respectively, marks the position of the terminator hairpin stems, with free energy of folding ∆G indicated. Black stars indicate introduced mutations. Downward carets () indicate the position of the 3′ ends associated with intrinsic terminators as determined by Rend-seq. Red dashed line indicates the complementary region of the northern blot probe to the readthrough product. b, Terminator readthrough fraction (defined as the Rend-seq read density after terminator divided by read density upstream of terminator, see ref.35 for details) as a function of stop-to-stem distance for E. coli intrinsic terminators from Fig. 2 for which readthrough could be reliably estimated (n = 392). Terminators with stop-to-stem distance d ≤ 12 nt are highlighted in red. c, Cumulative distribution function of terminator readthrough for terminators far (black, d>12 nt) from and close (red, d ≤ 12 nt) to stop codons. Terminator close to genes have significantly more readthrough (less termination), P < 10−3 (q30d>12 and q30d ≤ 12 indicate the 30th percentile in the readthrough distribution for the two categories of terminators, with fold-change F30: = q30d ≤ 12/ q30d>12, P-value determined as the fraction of bootstrap random sub-samplings of the readthrough distributions with q30d>12 > q30d ≤ 12, Methods) d, Terminator readthrough as a function of ∆GU, the U-tract DNA/RNA hybrid free energy (measure of U-tract quality, with larger ∆GU corresponding to U-rich U-tract). Grey shading indicates cutoff (∆GU>-5 kcal/mol) to select good U-tract terminators. e, Same as c, but restricting to good U-tract terminators, still showing significantly less termination for terminators near ORF, P < 10−3 (same as above, Methods). f-i, same as be, but with terminators from B. subtilis. Terminators close to ORF do not show less readthrough than their gene-distal counterparts (P > 0.3, p-value determined with same strategy as above, Methods).

Source data

Extended Data Fig. 6 Examples of identified nested antisense RNAs.

B. subtilis shows a number (n = 35, see Methods for selection criteria) of mRNAs with long untranslated regions fully encompassing genes in the antisense directions, which we call nested antisense RNAs (also termed non-contiguous operons53 or excludons54). The majority (n = 29/35) of these have a fold-change in mRNA level less than twofold upon rho deletion (Fig. 3b). a, Schematic of a nested antisense RNA with corresponding Rend-seq signal, with orange peaks and blue peaks marking 5′ and 3′ boundaries of the transcript. b, Representative examples of nested antisense RNAs with mRNA level fold change upon rho deletion less than 2. Rend-seq data (peak shadows removed, see35 for details on data processing) is shown. Orange and blue signal correspond to summed 5′-mapped reads and 3′-mapped reads, respectively (rpm: reads per million). Top trace corresponds to wild type, and bottom trace to ∆rho. Horizontal size marker provides positional scale (200 bp) on each subpanel. Sense and antisense genes are shown in dark and light grey, respectively. Double line breaks (//) indicate truncated Rend-seq signal at peaks. Dashed lines mark regions for which fold-change in read density for ∆rho/WT was estimated. The fold-change for each instance is indicated on the graph. c, Same as b, with representative examples of nested antisense RNAs with increased expression upon rho deletion (Fig. 3b). Three nested antisense RNAs were found in E. coli with identical criteria. See Supplementary Data 3 for a list of nested antisense RNAs identified.

Source data

Extended Data Fig. 7 Expressed pseudogenes with interrupted translation in B. subtilis show no polarity.

Expressed pseudogenes endogenously present in the extant genome were used as additional independent experiments to assess the prevalence of Rho-mediated nonsense polarity in B. subtilis in situations of obligately uncoupled transcription and translation. Concomitant Rend-seq (mapping operon architecture) and ribosome profiling (measurement of translation) provides stringent data to determine translational status and transcript integrity of mRNAs. a, Schematic of analysis: for expressed pseudogenes (see Methods for selection criteria) with translation disruption, polarity was assessed by (1) comparing the mRNA read density at start and end of transcription unit, with large changes (start/end 1) indicative of polarity, and (2) fold change of pseudogene transcript upon rho deletion. Position of translation disrupting mutation is shown by ▲ and X. Dark and pale grey indicates region prior and after translation disruption mutation. b, Rend-seq and ribosome profiling data for the 8 identified expressed pseudogenes. Each subpanel corresponds to a pseudogene region. Top traces show Rend-seq data (orange and blue signal correspond to summed 5′-mapped reads and 3′-mapped reads, peak shadows removed, see ref.35 for details on data processing). Orange peaks and blue peaks mark 5′ and 3′ boundaries of transcripts. Double line breaks (//) indicate truncated Rend-seq signal at peaks. Bottom traces show ribosome profiling data. Translation efficiency (ribosome profiling rpkm/Rend-seq rpkm) percentiles for each pseudogene sub-region (before and after translation disruption) are shown. Horizontal size marker provides positional scale (200 bp) on each subpanel. Nearby intact genes are shown in light blue. rpm: reads per million. Regions used to assess start to end decrease in RNA levels are marked by dashed lines. mRNA levels fold-changes (start/end, and ∆rho/WT) are shown. The ydzW region showed a second translation disruption the secondary frame, shown as a pale ▲ and X. See Methods, Fig. 3b and Supplementary Data 3 for details.

Source data

Extended Data Fig. 8 Most expressed pseudogenes with interrupted translation in E. coli show polarity.

Similar to Extended Data Fig. 7. Expressed pseudogenes endogenously present in the extant genome were used as additional independent experiments to assess the prevalence of Rho-mediated nonsense polarity in E. coli in situations of obligately uncoupled transcription and translation. Concomitant Rend-seq (mapping operon architecture) and ribosome profiling (monitoring translation) provides stringent data to determine translational status and transcript integrity on mRNAs. a, Schematic of analysis: for expressed pseudogenes (see Methods for selection criteria) with translation disruption, polarity was assessed by comparing the mRNA read density at start and end of transcription unit, with large changes (start/end1) indicative of polarity. b, Rend-seq and ribosome profiling data for the identified expressed pseudogene with evidence of polarity. Each subpanel corresponds to a pseudogene region. Top traces correspond to Rend-seq data (orange and blue signal correspond to summed 5′-mapped reads and 3′-mapped reads, peak shadows removed, see ref.35 for details on data processing). Orange peaks and blue peaks mark 5′ and 3′ boundaries of transcripts. Double line breaks (//) indicate truncated Rend-seq signal at peaks. Bottom traces show ribosome profiling data. Translation efficiency (ribosome profiling rpkm/Rend-seq rpkm) percentiles for each pseudogene sub-region (before and after translation disruption) are shown. Horizontal size marker provides positional scale (200 bp) on each subpanel. Light blue arrows correspond to nearby intact genes. rpm: reads per million. Regions used to assess start to end decrease in RNA levels are marked by dashed lines. mRNA levels fold-changes (start/end) are shown. The gapC region showed sequential translation disruptions secondary frames, shown as a pale ▲ and X. c, same as b, but for the two cases with no evidence of polarity. The translation disruptions mutation in ykiA and cybC are deletion of the beginning of ORFs. See Methods, Fig. 3b and Supplementary Data 3 for details.

Source data

Extended Data Fig. 9 Analysis of C-to-G ratio for putative Rho-terminated RNAs.

a, Cumulative distributions of maximum C-to-G ratio (“Max C:G”) of 100 nt sliding windows within non Rho-terminated coding sequences (CDSs, blue, n = 2625) and Rho-terminated CDSs (magenta, n = 10). Median of Max C:G is higher for Rho-terminated CDSs (magenta) than non Rho-terminated CDSs (blue) (P < 10−5, less than one in 105 random sub-samplings (n = 10) of non Rho-terminated distribution had higher median maximum C-to-G ratio). b, Cumulative distributions as in a for asRNAs that are not terminated by Rho (blue, n = 112) and asRNAs that are terminated by Rho (magenta, n = 91). Median of Max C:G is higher for Rho-terminated asRNAs than non Rho-terminated asRNAs (P < 10−3, less than one in 103 random sub-samplings (n = 10) of non Rho-terminated distribution had higher median maximum C to G ratio compared to sub-sampling (n = 10) of Rho-terminated distribution) (Methods).

Source data

Extended Data Fig. 10 Illustration of terminator identification pipeline and analysis of stem-to-stop distribution stratified by phyla.

The terminator identification pipeline selects for strong hairpins immediately upstream of long U-tract found downstream of genes. Thresholds on hairpin folding free energy are determined on a species-by-species basis based on properties of randomly selected regions in respective genomes. The case of V. cholerae is illustrated in a-c. a, Results of folding 104 regions of 40 nt chosen at random positions in the genome. Left panel shows the 2D distribution as a heatmap (dark positions corresponding to more density) of hairpin geometrical parameters (number of base pairs in stem Nbp, length of loop). Geometric thresholds are highlighted with blue dashes (5 bp ≤ Nbp ≤ 15 bp, 3 nt ≤ Loop ≤ 8 nt) and retained region by blue shading. Right panel shows the 2D distribution as a heatmap (dark positions correspond to more density) of hairpin free energy of folding ∆Ghairpin and fraction of bases paired in stem f. Thresholds ∆G1 and ∆G2 on ∆Ghairpin are chosen such the total fraction of hairpin from random regions meeting geometrical (blue shading in left panel) and thermodynamic thresholds are 1% (orange, ∆Ghairpin ≤ ∆G1 and f ≥ 0.95) and 1.5% (red, ∆Ghairpin ≤ ∆G2 and f ≥ 0.9). b, Similar as for a, but for regions seeded by U-tracts (stretch of 5 or more consecutive T’s in the genome downstream of genes). Note the excess density of hairpins with strong energy of folding and large fraction of bases paired, corresponding to putative intrinsic terminators. c, Distribution of stop-to-stem distances for terminators passing thresholds shown in b. See Supplementary Data 2, Supplementary Data 3, and Methods for details of computational pipeline. d and e, Phylum stratified analysis on the stop-to-stem distribution. d, Each subpanel shows as a 2D greyscale the fraction of species within each phylum (shown in Fig. 4) for which more than fraction F (y-axis) of terminators have stop-to-stem distances less than or equal to D (x-axis). Black regions correspond to no species in the phylum, white all species. The contour line in the (D,F) space marks points where 50% of species in the phylum have fraction ≥ F of their terminators with stop-to-stem distance ≤ D. The yellow stars mark the thresholds used in Fig. 4 (D = 12 nt, F = 30%). For example, about 50% of species analysed in the Firmicutes have more than 30% of their terminators within 12 nt of upstream ORF (red contour line intersecting yellow star). e, The 50% species contour lines from d reported to the same panel, showing clear separation between phyla.

Source data

Supplementary information

Supplementary Information

Contains Supplementary Discussion (on previous approaches to bioinformatically identify intrinsic terminators, on the possibility of translation initiation on B. subtilis nascent mRNAs, on transcription elongation rates in mutant backgrounds, and on the phylogeny of domain architecture in NusA, NusG, and RpoB). Supplementary Figure 1: raw Northern blot data.

Reporting Summary

41586_2020_2638_MOESM3_ESM.pdf

Supplementary Data Supplementary Data 1: sequence for lacZ and cssSAS. Sequence of lacZ adapted from E. coli for expression in B. subtilis and sequence of cssSAS site.

41586_2020_2638_MOESM4_ESM.xlsx

Supplementary Data Supplementary Data 2: high-confidence list of intrinsic terminators. List of terminators with strong experimental evidence in B. subtilis and E. coli. Restricted subset from the list in Ref. 35, with additional quality criteria (Methods). See Fig. 2 and Extended Data Fig. 5.

41586_2020_2638_MOESM5_ESM.xlsx

Supplementary Data Supplementary Data 3: details for nested antisense RNAs and expressed pseudogenes. List of nested antisense RNAs and pseudogenes (nested antisense RNAs, final list of pseudogenes) and related properties. See Fig. 3 and Extended Data Fig. 6-8.

41586_2020_2638_MOESM6_ESM.xlsx

Supplementary Data Supplementary Data 4: species-by-species summary of identified intrinsic terminators. Summary of identified putative intrinsic terminators for the 1648 RefSeq species considered. See Fig. 4, Extended Data Fig. 10.

41586_2020_2638_MOESM7_ESM.txt

Supplementary Data Supplementary Data 5: list of all identified terminators. Complete list of all 301817 identified putative terminators for the 1648 RefSeq species. See Fig. 4 and Extended Data Fig. 10. We recommend parsing this file computationally.

41586_2020_2638_MOESM8_ESM.xlsx

Supplementary Data Supplementary Data 6: strain and plasmid details. Details of strains and plasmid used, with additional information on strain construction.

41586_2020_2638_MOESM9_ESM.xlsx

Supplementary Data Supplementary Data 7: list of oligonucleotides. List of oligonucleotides used in current work, with short description of use.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Johnson, G.E., Lalanne, JB., Peters, M.L. et al. Functionally uncoupled transcription–translation in Bacillus subtilis. Nature 585, 124–128 (2020). https://doi.org/10.1038/s41586-020-2638-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2638-5

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