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:

Widespread divergent transcription from bacterial and archaeal promoters is a consequence of DNA-sequence symmetry

Abstract

Transcription initiates at promoters, DNA regions recognized by a DNA-dependent RNA polymerase. We previously identified horizontally acquired Escherichia coli promoters from which the direction of transcription was unclear. In the present study, we show that more than half of these promoters are bidirectional and drive divergent transcription. Using genome-scale approaches, we demonstrate that 19% of all transcription start sites detected in E. coli are associated with a bidirectional promoter. Bidirectional promoters are similarly common in diverse bacteria and archaea, and have inherent symmetry: specific bases required for transcription initiation are reciprocally co-located on opposite DNA strands. Bidirectional promoters enable co-regulation of divergent genes and are enriched in both intergenic and horizontally acquired regions. Divergent transcription is conserved among bacteria, archaea and eukaryotes, but the underlying mechanisms for bidirectionality are different.

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: TSS pairs within horizontally acquired genes.
Fig. 2: Widespread divergent transcription from bidirectional promoter sequences in E. coli.
Fig. 3: Reciprocal stimulation between divergent TSSs.
Fig. 4: Bidirectional promoter sequences are widespread in prokaryotes.
Fig. 5: Coordinated regulation of divergent transcription units from bidirectional promoter sequences.
Fig. 6: Promoter bidirectionality has a different basis in prokaryotes and eukaryotes.

Similar content being viewed by others

Data availability

The data that support these findings are available from the corresponding author upon request. The E. coli RNA-seq and B. subtilis cappable-seq data are available in ArrayExpress using accession nos. E-MTAB-9655 and E-MTAB-8582, respectively. Source data are provided with this paper.

References

  1. Mejía-Almonte, C. et al. Redefining fundamental concepts of transcription initiation in bacteria. Nat. Rev. Genet. https://doi.org/10.1038/s41576-020-0254-8 (2020).

  2. Browning, D. F. & Busby, S. J. W. The regulation of bacterial transcription initiation. Nat. Rev. Microbiol. 2, 57–65 (2004).

    Article  CAS  PubMed  Google Scholar 

  3. Haberle, V. & Stark, A. Eukaryotic core promoters and the functional basis of transcription initiation. Nat. Rev. Mol. Cell Biol. 19, 621–637 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Bae, B., Feklistov, A., Lass-Napiorkowska, A., Landick, R. & Darst, S. A. Structure of a bacterial RNA polymerase holoenzyme open promoter complex. eLife 4, e08504 (2015).

    Article  PubMed Central  Google Scholar 

  5. Feklistov, A. & Darst, S. A. Structural basis for promoter −10 element recognition by the bacterial RNA polymerase σ subunit. Cell 147, 1257–1269 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Kramm, K., Engel, C. & Grohmann, D. Transcription initiation factor TBP: old friend new questions. Biochem. Soc. Trans. 47, 411–423 (2019).

    Article  CAS  PubMed  Google Scholar 

  7. Butler, J. E. F. The RNA polymerase II core promoter: a key component in the regulation of gene expression. Genes Dev. 16, 2583–2592 (2002).

    Article  CAS  PubMed  Google Scholar 

  8. Core, L. J., Waterfall, J. J. & Lis, J. T. Nascent RNA sequencing reveals widespread pausing and divergent initiation at human promoters. Science 322, 1845–1848 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Seila, A. C. et al. Divergent transcription from active promoters. Science 322, 1849–1851 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Preker, P. et al. RNA exosome depletion reveals transcription upstream of active human promoters. Science 322, 1851–1854 (2008).

    Article  CAS  PubMed  Google Scholar 

  11. He, Y., Vogelstein, B., Velculescu, V. E., Papadopoulos, N. & Kinzler, K. W. The antisense transcriptomes of human cells. Science 322, 1855–1857 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Neil, H. et al. Widespread bidirectional promoters are the major source of cryptic transcripts in yeast. Nature 457, 1038–1042 (2009).

    Article  CAS  PubMed  Google Scholar 

  13. Scruggs, B. S. et al. Bidirectional transcription arises from two distinct hubs of transcription factor binding and active chromatin. Mol. Cell 58, 1101–1112 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Rege, M. et al. Chromatin dynamics and the RNA exosome function in concert to regulate transcriptional homeostasis. Cell Rep. 13, 1610–1622 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Wu, X. & Sharp, P. A. XDivergent transcription: a driving force for new gene origination? Cell 155, 990–996 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Jin, Y., Eser, U., Struhl, K. & Churchman, L. S. The ground state and evolution of promoter region directionality. Cell 170, 889–898 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Dame, R. T., Rashid, F.-Z. M. & Grainger, D. C. Chromosome organization in bacteria: mechanistic insights into genome structure and function. Nat. Rev. Genet. 21, 227–242 (2019).

    Article  PubMed  CAS  Google Scholar 

  18. Browning, D. F. & Busby, S. J. W. Local and global regulation of transcription initiation in bacteria. Nat. Rev. Microbiol. 14, 638–650 (2016).

    Article  CAS  PubMed  Google Scholar 

  19. Singh, S. S. et al. Widespread suppression of intragenic transcription initiation by H-NS. Genes Dev. 28, 214–219 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Mitra, P., Ghosh, G., Hafeezunnisa, M. & Sen, R. Rho protein: roles and mechanisms. Annu. Rev. Microbiol. 71, 687–709 (2017).

    Article  CAS  PubMed  Google Scholar 

  21. Keseler, I. M. et al. The EcoCyc database: reflecting new knowledge about Escherichia coli K-12. Nucleic Acids Res. 45, D543–D550 (2017).

    Article  CAS  PubMed  Google Scholar 

  22. Ettwiller, L., Buswell, J., Yigit, E. & Schildkraut, I. A novel enrichment strategy reveals unprecedented number of novel transcription start sites at single base resolution in a model prokaryote and the gut microbiome. BMC Genom. 17, 199 (2016).

    Article  CAS  Google Scholar 

  23. Thomason, M. K. et al. Global transcriptional start site mapping using differential RNA sequencing reveals novel antisense RNAs in Escherichia coli. J. Bacteriol. 197, 18–28 (2015).

    Article  PubMed  CAS  Google Scholar 

  24. Singh, S. S., Typas, A., Hengge, R. & Grainger, D. C. Escherichia coli σ 70 senses sequence and conformation of the promoter spacer region. Nucleic Acids Res. 39, 5109–5118 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Warman, E. A., Singh, S. S., Gubieda, A. G. & Grainger, D. C. A non-canonical promoter element drives spurious transcription of horizontally acquired bacterial genes. Nucleic Acids Res. 48, 4891–4901 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Santos-Zavaleta, A. et al. A unified resource for transcriptional regulation in Escherichia coli K-12 incorporating high-throughput-generated binding data into RegulonDB version 10.0. BMC Biol. 16, 91 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Mendoza-Vargas, A. et al. Genome-wide identification of transcription start sites, promoters and transcription factor binding sites in E. coli. PLoS ONE 4, e7526 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Gill, E. E. et al. High-throughput detection of RNA processing in bacteria. BMC Genom. 19, 223 (2018).

    Article  CAS  Google Scholar 

  29. Papenfort, K., Förstner, K. U., Cong, J. P., Sharma, C. M. & Bassler, B. L. Differential RNA-seq of Vibrio cholerae identifies the VqmR small RNA as a regulator of biofilm formation. Proc. Natl Acad. Sci. USA 112, E766–E775 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Kröger, C. et al. The primary transcriptome, small RNAs and regulation of antimicrobial resistance in Acinetobacter baumannii ATCC 17978. Nucleic Acids Res. 46, 9684–9698 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Sharma, C. M. et al. The primary transcriptome of the major human pathogen Helicobacter pylori. Nature 464, 250–255 (2010).

    Article  CAS  PubMed  Google Scholar 

  32. Cortes, T. et al. Genome-wide mapping of transcriptional start sites defines an extensive leaderless transcriptome in Mycobacterium tuberculosis. Cell Rep. 5, 1121–1131 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Jeong, Y. et al. The dynamic transcriptional and translational landscape of the model antibiotic producer Streptomyces coelicolor A3(2). Nat. Commun. 7, 11605 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Fan, B. et al. dRNA-seq reveals genomewide TSSs and noncoding RNAs of plant beneficial rhizobacterium Bacillus amyloliquefaciens FZB42. PLoS ONE 10, e0142002 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Decker, K. B. & Hinton, D. M. Transcription regulation at the core: similarities among bacterial, archaeal, and eukaryotic RNA polymerases. Annu. Rev. Microbiol. 67, 113–139 (2013).

    Article  CAS  PubMed  Google Scholar 

  36. Grünberger, F. et al. Next generation DNA-seq and differential RNA-seq allow re-annotation of the Pyrococcus furiosus DSM 3638 genome and provide insights into archaeal antisense transcription. Front. Microbiol. 10, 1603 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Babski, J. et al. Genome-wide identification of transcriptional start sites in the haloarchaeon Haloferax volcanii based on differential RNA-Seq (dRNA-Seq). BMC Genom. 17, 629 (2016).

    Article  CAS  Google Scholar 

  38. Jäger, D., Förstner, K. U., Sharma, C. M., Santangelo, T. J. & Reeve, J. N. Primary transcriptome map of the hyperthermophilic archaeon Thermococcus kodakarensis. BMC Genom. 15, 684 (2014).

    Article  CAS  Google Scholar 

  39. Lamberte, L. E. et al. Horizontally acquired AT-rich genes in Escherichia coli cause toxicity by sequestering RNA polymerase. Nat. Microbiol. 2, 16249 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Chen, J. et al. Stepwise promoter melting by bacterial RNA polymerase. Mol. Cell 78, 275–288.e6 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Miller, J. Experiments in Molecular Genetics (Cold Spring Harbor Laboratory, 1972).

  42. Haycocks, J. R. J. & Grainger, D. C. Unusually situated binding sites for bacterial transcription factors can have hidden functionality. PLoS ONE 11, e0157016 (2016).

  43. Dugar, G. et al. High-resolution transcriptome maps reveal strain-specific regulatory features of multiple Campylobacter jejuni Isolates. PLoS Genet. 9, e1003495 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Singh, N. & Wade, J. T. Identification of regulatory RNA in bacterial genomes by genome-scale mapping of transcription start sites. Methods Mol. Biol. 1103, 1–10 (2014).

    Article  CAS  PubMed  Google Scholar 

  45. Crooks, G. E., Hon, G., Chandonia, J.-M. & Brenner, S. E. WebLogo: a sequence logo generator. Genome Res. 14, 1188–1190, https://doi.org/10.1101/gr.849004 (2004).

  46. Haycocks, J. R. J. J. et al. The quorum sensing transcription factor AphA directly regulates natural competence in Vibrio cholerae. PLoS Genet. 15, e1008362 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Kolb, A., Kotlarz, D., Kusano, S. & Ishihama, A. Selectivity of the Escherichia coli RNA polymerase Eσ38 for overlapping promoters and ability to support CRP activation. Nucleic Acids Res. 23, 819–826 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Savery, N. J. et al. Transcription activation at class II CRP-dependent promoters: identification of determinants in the C-terminal domain of the RNA polymerase α subunit. EMBO J. 17, 3439–3447 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Stead, M. B. et al. RNAsnapTM: a rapid, quantitative and inexpensive, method for isolating total RNA from bacteria. Nucleic Acids Res. 40, e156 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Afgan, E. et al. The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2018 update. Nucleic Acids Res. 46, W537–W544 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Wang, L., Wang, S. & Li, W. RSeQC: quality control of RNA-seq experiments. Bioinformatics 28, 2184–2185 (2012).

    Article  CAS  PubMed  Google Scholar 

  53. Forrest, D., James, K., Yuzenkova, Y. & Zenkin, N. Single-peptide DNA-dependent RNA polymerase homologous to multi-subunit RNA polymerase. Nat. Commun. 8, 15774 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Podell, S., Gaasterland, T. & Allen, E. E. A database of phylogenetically atypical genes in archaeal and bacterial genomes, identified using the DarkHorse algorithm. BMC Bioinform. 9, 419 (2008).

    Article  CAS  Google Scholar 

  55. Kahramanoglou, C. et al. Direct and indirect effects of H-NS and Fis on global gene expression control in Escherichia coli. Nucleic Acids Res. 39, 2073–2091 (2011).

    Article  CAS  PubMed  Google Scholar 

  56. Narayanan, A. et al. Cryo-EM structure of Escherichia coli σ 70 RNA polymerase and promoter DNA complex revealed a role of σ non-conserved region during the open complex formation. J. Biol. Chem. 293, 7367–7375 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was funded by a Leverhulme Trust project grant (no. RPG-2018-198) and a Wellcome Trust Investigator award (no. 212193/Z/18/Z) to D.C.G.

Author information

Authors and Affiliations

Authors

Contributions

D.C.G. conceived and supervised the work. D.F. did the B. subtilis TSS mapping and in vitro analysis of different TSS spacings. T.G. and J.R.J.H. did the analysis of the V. cholerae VC1303VC1304 regulatory region. E.A.W. did all other experimental work. D.C.G., J.T.W. and D.F. did the computational analysis. All authors contributed to data analysis and interpretation. D.C.G. wrote the manuscript with input from all the authors.

Corresponding author

Correspondence to David C. Grainger.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Microbiology thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

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

Extended data

Extended Data Fig. 1 Not all σ70 binding sites align with an RNA 5′ end.

Binding patterns for H-NS (peach) and σ70 (purple or green) are derived from ChIP-seq assays19. The RNA 5′ ends associated with σ70 binding (purple or green) were identified by PPP-seq19. Genes are shown as red block arrows. Each ChIP-seq dataset is plotted on a scale of 0 to 360 reads on each strand. The TSS data show a read depth of between 0 and 100 on each strand. The AT-content fluctuates between 30 % and 75 %.

Extended Data Fig. 2 Directional transcription from canonical promoters.

a, β-galactosidase activity derived from canonical promoters listed by Ecocyc21 and not having a transcription start site in the reverse orientation detectable by any of three separate studies19,22,23. Data are presented as mean values (n = 3 independent experiments) +/- SD and individual data points are overlaid as dot plots. b, Direction of transcription from cloned DNA fragments. c, Average forward or reverse β-galactosidase activity of all DNA fragments.

Source data

Extended Data Fig. 3 Sequences of cryptic RNA polymerase binding sites associated with divergent transcription.

The figure shows promoter DNA sequences (black typeface) and part of the plasmid DNA backbone (grey typeface). The promoter −10 (red) and −35 (green) elements are highlighted on each DNA strand and transcription start sites are denoted by a bent arrow. Sites of mutations and deletions (Δ) are boxed. The sequences are in the ‘a’ orientation as indicated in Fig. 1. When in the ‘b’ orientation the DNA sequence encompassed by black typeface is the reverse complement. Oligonucleotide D49724, used in primer extension analysis, is indicated by a half arrow and binds to the corresponding sequence in grey bold typeface.

Extended Data Fig. 4 Spacing optima between divergent transcription start sites.

a, The graph shows the number of divergent TSSs separated by different distances. The majority of bottom strand transcription start sites occur 18 bp upstream of top strand RNA initiation sites. However, peaks in the occurrence of divergent TSSs also occur elsewhere. These positions are denoted by green data points. The red data point indicates a sharp decrease in the occurrence of divergent TSSs. The symmetry score increases at spacing intervals where the promoter PWM identifies matching sequences overlapping on each DNA strand. b, DNA sequence motifs associated with each preferred spacing between divergent TSSs. Motifs were generated by aligning sequences according to the top strand TSS. The configuration of promoter −10 elements and TSSs, indicated by each motif, is shown below the respective sequence logo. Key positions within the −10 elements, and TSSs, are underlined. c, Products of in vitro transcription using the illustrated DNA templates. The RNAI transcript is derived from the replication origin of the plasmid DNA template. A representative example of two independent experiments is shown. d, Products of in vitro transcription from the intragenic yigG promoter. Note that products produced in vitro match those produced in vivo (Fig. 1d). A representative example of two independent experiments is shown.

Extended Data Fig. 5 Properties of transcription start sites in Bacillus subtilis.

We mapped transcription start sites globally in Bacillus subtilis using cappable-seq. The general properties of B. subtilis promoters were compared with those identified in E. coli. a, Distances between promoter −10 elements and transcription start sites (TSSs) in Escherichia coli and B. subtilis. b,c, DNA sequence motifs associated with unidirectional TSSs in E. coli and B. subtilis. d,e, Positioning of TSSs with respect to coding DNA sequences in E. coli and B. subtilis.

Extended Data Fig. 6 Properties of bidirectional promoters in different bacteria.

The figure shows the number of divergent transcription start sites separated by different distances (black line in graph). The data point in each graph, corresponding to the preferred configuration of divergent start sites, is green. The pale blue data indicate predicted promoter overlap (that is symmetry) derived from a position weight matrix (PWM) search of each DNA strand. The R2 values indicate the degree of correlation between computational prediction and experimental data shown. The DNA motifs adjacent to each graph were generated by aligning promoter −10 hexamer sequences. For V. cholerae, P. aeruginosa, A. baumannii, M. tuberculosis, and S. coelicolor we aligned −10 elements from start sites separated by 17, 18 and 19 bp. In the case of H. pylori, we aligned −10 elements from those start sites 15, 16 or 17 bp apart. Note that all of these distances typically involve the same configuration of −10 elements because the distance between the −10 hexamer and transcription start site is variable (see Supplementary information Fig. 4). For B. subtilis and B. amyloliquefaciens we aligned start sites separated by 10, 11 or 12 bp.

Source data

Extended Data Fig. 7 Bidirectional promoters in the archaea Thermococcus kodakarensis and Haloferax volcanii have a shared TATA box.

a,b, The figure shows the number of divergent transcription start sites separated by different distances (black line in graph). The data points in each graph, corresponding to the preferred configurations of divergent start sites, are green. The pale blue data indicate promoter sequence symmetry. The R2 values indicate the degree of correlation between computational prediction and experimental data shown. b, DNA sequences associated with divergent TSSs in Haloferax volcanii separated by 62, 53 or 34 bp were aligned according to the position of the TSS on the top DNA strand. The inf configuration of key elements is shown above each motif.

Source data

Extended Data Fig. 8 Bidirectional promoters are enriched in horizontally acquired genes.

a, Detection of directional and bidirectional promoters by PPP-seq19. The pie charts show the fractions of each promoter type detected in H-NS bound or H-NS free regions of the E. coli genome in the presence and absence of H-NS. b, Distribution of position weight matrix (PWM) scores for bidirectional promoters in different sections of the E. coli genome. Higher scores indicate a better match to the PWM describing bidirectional promoters. The bounds of the box represent the first and third quartiles and the centre line is the median. Whiskers extend to 1.5 times the interquartile range.

Extended Data Fig. 9 The ratio of directional to bidirectional promoters is similar in different bacteria.

We used multiple Escherichia coli TSS maps to identify bidirectional promoters (corresponding to divergent TSSs separated by between 25 and 7 bp). We noticed that the proportion of TSSs from bidirectional promoters was much smaller in datasets with fewer total TSSs. We reasoned that this was logical; the chance of detecting both transcripts from a given bidirectional promoter is much smaller for less complete TSS maps. For instance, 19 % of TSSs in the combined E. coli TSS map (28,107 TSSs in total) were derived from a bidirectional promoter. In contrast, this value was only 5 % for TSSs identified by PPP-seq19 (4,846 TSSs in total). The number of total and divergent TSSs, for each E. coli TSS map, is plotted in orange; the relationship is not linear. For comparison, we generated a probability model using a mock TSS map for E. coli. The artificial map consisted of 28,107 randomly selected E. coli genome co-ordinates as TSSs. Of these, 19 % of positions on the bottom strand were set to be between 7 and 25 bp upstream of a top strand co-ordinate (that is the mock data exactly emulated the combined TSS composition of the genuine experimental data for E. coli). We then randomly selected sub-populations of genome co-ordinates from the mock TSS map and determined how many pairs of top and bottom stand positions remained separated by between 7 and 25 bp. These data are plotted in pale blue. Consistent with our logic, the relationship was not linear and resembled the real experimental data in orange. We also determined the number of divergent TSSs pairs amongst those TSSs detected by both dRNA-seq and cappable-seq (5,593 TSSs in total). This data point also fell precisely on the trend line generated by the individual and combined data sets. Hence, excluding TSSs not identified by multiple methods does not alter the frequency at which divergent TSS pairs are detected. Finally, we plotted experimentally determined TSS maps for different bacteria (all TSS numbers were normalised for genome size). Crucially, these organisms have been subject to much less scrutiny than E. coli. Hence, the total number of TSSs identified for each bacterium is comparatively small. Even so, it is clear that all data points fall close to the orange and pale blue trend lines. Hence, the fraction of promoters that are bidirectional must be broadly similar in different bacterial species.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Supplementary Table 1. Location and sequence of bidirectional promoters in E. coli.

Supplementary Tables 2–4

Supplementary Table 2. Enrichment of TSSs in RegulonDB among TSSs in the combined RNA-seq dataset. Supplementary Table 3. Positions of TSSs across the B. subtilis genome (NC_000964.3). Supplementary Table 4. Strains, plasmids and oligonucleotides.

Peer Review File

Source data

Source Data Fig. 1

Numerical source data for Fig. 1a,f.

Source Data Fig. 1

Unprocessed gel image for Fig. 1d.

Source Data Fig. 3

Unprocessed gel image for Fig. 3c.

Source Data Fig. 5

Unprocessed gel images for Fig. 5b,c,f.

Source Data Fig. 5

Numerical source data for Fig. 5d.

Source Data Extended Data Fig. 2

Numerical source data for Extended Data Fig. 2a.

Source Data Extended Data Fig. 4

Unprocessed gel images for Extended Data Fig. 4c,d.

Source Data Extended Data Fig. 6

Numerical source data.

Source Data Extended Data Fig. 7

Numerical source data.

Source Data Extended Data Fig. 8

Numerical source data for Extended Data Fig. 8b,c.

Source Data Extended Data Fig. 9

Numerical source data.

Source Data Extended Data Fig. 4

Numerical source data for Extended Data Fig. 4a.

Source Data Extended Data Fig. 5

Numerical source data for Extended Data Fig. 5a.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Warman, E.A., Forrest, D., Guest, T. et al. Widespread divergent transcription from bacterial and archaeal promoters is a consequence of DNA-sequence symmetry. Nat Microbiol 6, 746–756 (2021). https://doi.org/10.1038/s41564-021-00898-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-021-00898-9

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: Microbiology