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:

Synthetic promoter designs enabled by a comprehensive analysis of plant core promoters

Abstract

Targeted engineering of plant gene expression holds great promise for ensuring food security and for producing biopharmaceuticals in plants. However, this engineering requires thorough knowledge of cis-regulatory elements to precisely control either endogenous or introduced genes. To generate this knowledge, we used a massively parallel reporter assay to measure the activity of nearly complete sets of promoters from Arabidopsis, maize and sorghum. We demonstrate that core promoter elements—notably the TATA box—as well as promoter GC content and promoter-proximal transcription factor binding sites influence promoter strength. By performing the experiments in two assay systems, leaves of the dicot tobacco and protoplasts of the monocot maize, we detect species-specific differences in the contributions of GC content and transcription factors to promoter strength. Using these observations, we built computational models to predict promoter strength in both assay systems, allowing us to design highly active promoters comparable in activity to the viral 35S minimal promoter. Our results establish a promising experimental approach to optimize native promoter elements and generate synthetic ones with desirable features.

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: STARR-seq measures core promoter strength in tobacco leaves and maize protoplasts.
Fig. 2: Plant core promoters span a wide range of activity.
Fig. 3: GC content affects promoter strength in tobacco leaves.
Fig. 4: The TATA box is a key determinant of promoter strength.
Fig. 5: Enhancer responsiveness of promoters depends on the TATA box and GC content.
Fig. 6: Promoter strength can be modulated by light.
Fig. 7: Design and validation of synthetic promoters.
Fig. 8: Computational models can predict promoter strength and enable in silico evolution of plant promoters.

Similar content being viewed by others

Data availability

All sequencing results are deposited in the NCBI Sequence Read Archive under the BioProject accession PRJNA714258.

Code availability

The code used in this study is available on Github (https://github.com/tobjores/Synthetic-Promoter-Designs-Enabled-by-a-Comprehensive-Analysis-of-Plant-Core-Promoters).

References

  1. Liu, W. & Stewart, C. N. Plant synthetic biology. Trends Plant Sci. 20, 309–317 (2015).

    Article  CAS  PubMed  Google Scholar 

  2. Lomonossoff, G. P. & D’Aoust, M.-A. Plant-produced biopharmaceuticals: a case of technical developments driving clinical deployment. Science 353, 1237–1240 (2016).

    Article  CAS  PubMed  Google Scholar 

  3. Smale, S. T. & Kadonaga, J. T. The RNA polymerase II core promoter. Annu. Rev. Biochem. 72, 449–479 (2003).

    Article  CAS  PubMed  Google Scholar 

  4. Andersson, R. & Sandelin, A. Determinants of enhancer and promoter activities of regulatory elements. Nat. Rev. Genet. 21, 71–87 (2020).

    Article  CAS  PubMed  Google Scholar 

  5. Ricci, W. A. et al. Widespread long-range cis-regulatory elements in the maize genome. Nat. Plants 5, 1237–1249 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Banerji, J., Rusconi, S. & Schaffner, W. Expression of a β-globin gene is enhanced by remote SV40 DNA sequences. Cell 27, 299–308 (1981).

    Article  CAS  PubMed  Google Scholar 

  7. Banerji, J., Olson, L. & Schaffner, W. A lymphocyte-specific cellular enhancer is located downstream of the joining region in immunoglobulin heavy chain genes. Cell 33, 729–740 (1983).

    Article  CAS  PubMed  Google Scholar 

  8. Grosschedl, R. & Birnstiel, M. L. Identification of regulatory sequences in the prelude sequences of an H2A histone gene by the study of specific deletion mutants in vivo. Proc. Natl Acad. Sci. USA 77, 1432–1436 (1980).

    Article  CAS  PubMed  Google Scholar 

  9. Wasylyk, B. et al. Specific in vitro transcription of conalbumin gene is drastically decreased by single-point mutation in T-A-T-A box homology sequence. Proc. Natl Acad. Sci. USA 77, 7024–7028 (1980).

    Article  CAS  PubMed  Google Scholar 

  10. Smale, S. T. & Baltimore, D. The “initiator” as a transcription control element. Cell 57, 103–113 (1989).

    Article  CAS  PubMed  Google Scholar 

  11. Ince, T. A. & Scotto, K. W. A conserved downstream element defines a new class of RNA polymerase II promoters. J. Biol. Chem. 270, 30249–30252 (1995).

    Article  CAS  PubMed  Google Scholar 

  12. Burke, T. W. & Kadonaga, J. T. Drosophila TFIID binds to a conserved downstream basal promoter element that is present in many TATA box-deficient promoters. Genes Dev. 10, 711–724 (1996).

    Article  CAS  PubMed  Google Scholar 

  13. Lagrange, T., Kapanidis, A. N., Tang, H., Reinberg, D. & Ebright, R. H. New core promoter element in RNA polymerase II-dependent transcription: sequence-specific DNA binding by transcription factor IIB. Genes Dev. 12, 34–44 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Lewis, B. A., Kim, T.-K. & Orkin, S. H. A downstream element in the human β-globin promoter: evidence of extended sequence-specific transcription factor IID contacts. Proc. Natl Acad. Sci. USA 97, 7172–7177 (2000).

    Article  CAS  PubMed  Google Scholar 

  15. Lim, C. Y. et al. The MTE, a new core promoter element for transcription by RNA polymerase II. Genes Dev. 18, 1606–1617 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Deng, W. & Roberts, S. G. E. A core promoter element downstream of the TATA box that is recognized by TFIIB. Genes Dev. 19, 2418–2423 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Parry, T. J. et al. The TCT motif, a key component of an RNA polymerase II transcription system for the translational machinery. Genes Dev. 24, 2013–2018 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Molina, C. & Grotewold, E. Genome wide analysis of Arabidopsis core promoters. BMC Genom. 6, 25 (2005).

    Article  Google Scholar 

  19. Yamamoto, Y. Y. et al. Differentiation of core promoter architecture between plants and mammals revealed by LDSS analysis. Nucleic Acids Res. 35, 6219–6226 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bernard, V., Brunaud, V. & Lecharny, A. TC-motifs at the TATA box expected position in plant genes: a novel class of motifs involved in the transcription regulation. BMC Genom. 11, 166 (2010).

    Article  Google Scholar 

  21. Blake, M. C., Jambou, R. C., Swick, A. G., Kahn, J. W. & Azizkhan, J. C. Transcriptional initiation is controlled by upstream GC-box interactions in a TATAA-less promoter. Mol. Cell. Biol. 10, 6632–6641 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Patwardhan, R. P. et al. High-resolution analysis of DNA regulatory elements by synthetic saturation mutagenesis. Nat. Biotechnol. 27, 1173–1175 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Sharon, E. et al. Inferring gene regulatory logic from high-throughput measurements of thousands of systematically designed promoters. Nat. Biotechnol. 30, 521–530 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Lubliner, S. et al. Core promoter sequence in yeast is a major determinant of expression level. Genome Res. 25, 1008–1017 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Arnold, C. D. et al. Genome-wide assessment of sequence-intrinsic enhancer responsiveness at single-base-pair resolution. Nat. Biotechnol. 35, 136–144 (2017).

    Article  CAS  PubMed  Google Scholar 

  26. van Arensbergen, J. et al. Genome-wide mapping of autonomous promoter activity in human cells. Nat. Biotechnol. 35, 145–153 (2017).

    Article  PubMed  Google Scholar 

  27. Weingarten-Gabbay, S. et al. Systematic interrogation of human promoters. Genome Res. 29, 171–183 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. de Boer, C. G. et al. Deciphering eukaryotic gene-regulatory logic with 100 million random promoters. Nat. Biotechnol. 38, 56–65 (2020).

    Article  PubMed  Google Scholar 

  29. Kotopka, B. J. & Smolke, C. D. Model-driven generation of artificial yeast promoters. Nat. Commun. 11, 2113 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Kumari, S. & Ware, D. Genome-wide computational prediction and analysis of core promoter elements across plant monocots and dicots. PLoS ONE 8, e79011 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Morton, T. et al. Paired-end analysis of transcription start sites in Arabidopsis reveals plant-specific promoter signatures. Plant Cell 26, 2746–2760 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Zhu, Q., Dabi, T. & Lamb, C. TATA box and initiator functions in the accurate transcription of a plant minimal promoter in vitro. Plant Cell 7, 1681–1689 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Kiran, K. et al. The TATA box sequence in the basal promoter contributes to determining light-dependent gene expression in plants. Plant Physiol. 142, 364–376 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Srivastava, R. et al. Distinct role of core promoter architecture in regulation of light-mediated responses in plant genes. Mol. Plant 7, 626–641 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Jores, T. et al. Identification of plant enhancers and their constituent elements by STARR-seq in tobacco leaves. Plant Cell 32, 2120–2131 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Cai, Y.-M. et al. Rational design of minimal synthetic promoters for plants. Nucleic Acids Res. 48, 11845–11856 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Srivastava, A. K., Lu, Y., Zinta, G., Lang, Z. & Zhu, J.-K. UTR-dependent control of gene expression in plants. Trends Plant Sci. 23, 248–259 (2018).

    Article  CAS  PubMed  Google Scholar 

  38. Fang, R. X., Nagy, F., Sivasubramaniam, S. & Chua, N. H. Multiple cis regulatory elements for maximal expression of the cauliflower mosaic virus 35S promoter in transgenic plants. Plant Cell 1, 141–150 (1989).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Benfey, P. N., Ren, L. & Chua, N. H. Tissue-specific expression from CaMV 35S enhancer subdomains in early stages of plant development. EMBO J. 9, 1677–1684 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Bruce, W. B., Christensen, A. H., Klein, T., Fromm, M. & Quail, P. H. Photoregulation of a phytochrome gene promoter from oat transferred into rice by particle bombardment. Proc. Natl Acad. Sci. USA 86, 9692–9696 (1989).

    Article  CAS  PubMed  Google Scholar 

  41. Yahraus, T., Chandra, S., Legendre, L. & Low, P. S. Evidence for a mechanically induced oxidative burst. Plant Physiol. 109, 1259–1266 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Walley, J. W. et al. Integration of omic networks in a developmental atlas of maize. Science 353, 814–818 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Wang, B. et al. A comparative transcriptional landscape of maize and sorghum obtained by single-molecule sequencing. Genome Res. 28, 921–932 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Mergner, J. et al. Mass-spectrometry-based draft of the Arabidopsis proteome. Nature 579, 409–414 (2020).

    Article  CAS  PubMed  Google Scholar 

  45. Singh, R., Ming, R. & Yu, Q. Comparative analysis of GC content variations in plant genomes. Trop. Plant Biol. 9, 136–149 (2016).

    Article  CAS  Google Scholar 

  46. Rensink, W. A. et al. Comparative analyses of six solanaceous transcriptomes reveal a high degree of sequence conservation and species-specific transcripts. BMC Genom. 6, 124 (2005).

    Article  Google Scholar 

  47. Tsai, F. T. F. & Sigler, P. B. Structural basis of preinitiation complex assembly on human Pol II promoters. EMBO J. 19, 25–36 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Gehrig, J. et al. Automated high-throughput mapping of promoter–enhancer interactions in zebrafish embryos. Nat. Methods 6, 911–916 (2009).

    Article  CAS  PubMed  Google Scholar 

  49. Yanai, I. et al. Genome-wide midrange transcription profiles reveal expression level relationships in human tissue specification. Bioinformatics 21, 650–659 (2005).

    Article  CAS  PubMed  Google Scholar 

  50. Heerah, S., Katari, M., Penjor, R., Coruzzi, G. & Marshall-Colon, A. WRKY1 mediates transcriptional regulation of light and nitrogen signaling pathways. Plant Physiol. 181, 1371–1388 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Cuperus, J. T. et al. Deep learning of the regulatory grammar of yeast 5′ untranslated regions from 500,000 random sequences. Genome Res. 27, 2015–2024 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Arnold, C. D. et al. Genome-wide quantitative enhancer activity maps Identified by STARR-seq. Science 339, 1074–1077 (2013).

    Article  CAS  PubMed  Google Scholar 

  53. Klein, J. C. et al. A systematic evaluation of the design and context dependencies of massively parallel reporter assays. Nat. Methods 17, 1083–1091 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Hong, C. K. & Cohen, B. A. Genomic environments scale the activities of diverse core promoters. Preprint at bioRxiv https://doi.org/10.1101/2021.03.08.434469 (2021).

  55. Dorrity, M. W. et al. The regulatory landscape of Arabidopsis thaliana roots at single-cell resolution. Nat. Commun. https://doi.org/10.1038/s41467-021-23675-y (2021).

  56. Marand, A. P., Chen, Z., Gallavotti, A. & Schmitz, R. J. A cis-regulatory atlas in maize at single-cell resolution. Cell https://doi.org/10.1016/j.cell.2021.04.014 (2021).

  57. Zhang, T.-Q., Chen, Y., Liu, Y., Lin, W.-H. & Wang, J.-W. Single-cell transcriptome atlas and chromatin accessibility landscape reveal differentiation trajectories in the rice root. Nat. Commun. 12, 2053 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Cheng, C.-Y. et al. Araport11: a complete reannotation of the Arabidopsis thaliana reference genome. Plant J. 89, 789–804 (2017).

    Article  CAS  PubMed  Google Scholar 

  59. McCormick, R. F. et al. The Sorghum bicolor reference genome: improved assembly, gene annotations, a transcriptome atlas, and signatures of genome organization. Plant J. 93, 338–354 (2018).

    Article  CAS  PubMed  Google Scholar 

  60. Mejía-Guerra, M. K. et al. Core promoter plasticity between maize tissues and genotypes contrasts with predominance of sharp transcription initiation sites. Plant Cell 27, 3309–3320 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Jiao, Y. et al. Improved maize reference genome with single-molecule technologies. Nature 546, 524–527 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Engler, C., Kandzia, R. & Marillonnet, S. A one pot, one step, precision cloning method with high throughput capability. PLoS ONE 3, e3647 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Hellens, R. P., Edwards, E. A., Leyland, N. R., Bean, S. & Mullineaux, P. M. pGreen: a versatile and flexible binary Ti vector for Agrobacterium-mediated plant transformation. Plant Mol. Biol. 42, 819–832 (2000).

    Article  CAS  PubMed  Google Scholar 

  64. Sheen, J. Metabolic repression of transcription in higher plants. Plant Cell 2, 1027–1038 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Masella, A. P., Bartram, A. K., Truszkowski, J. M., Brown, D. G. & Neufeld, J. D. PANDAseq: paired-end assembler for illumina sequences. BMC Bioinform. 13, 31 (2012).

    Article  CAS  Google Scholar 

  66. Raudvere, U. et al. g:Profiler: a web server for functional enrichment analysis and conversions of gene lists (2019 update). Nucleic Acids Res. 47, W191–W198 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Madeira, F. et al. The EMBL-EBI search and sequence analysis tools APIs in 2019. Nucleic Acids Res. 47, W636–W641 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Shahmuradov, I. A., Gammerman, A. J., Hancock, J. M., Bramley, P. M. & Solovyev, V. V. PlantProm: a database of plant promoter sequences. Nucleic Acids Res. 31, 114–117 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Fornes, O. et al. JASPAR 2020: update of the open-access database of transcription factor binding profiles. Nucleic Acids Res. 48, D87–D92 (2020).

    CAS  Google Scholar 

  70. Tian, F., Yang, D.-C., Meng, Y.-Q., Jin, J. & Gao, G. PlantRegMap: charting functional regulatory maps in plants. Nucleic Acids Res. 48, D1104–D1113 (2020).

    CAS  PubMed  Google Scholar 

  71. Onimaru, K., Nishimura, O. & Kuraku, S. Predicting gene regulatory regions with a convolutional neural network for processing double-strand genome sequence information. PLoS ONE 15, e0235748 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank A. Gutierrez Diaz and E. Grotewold for providing maize TSS data, and A. Gallavotti for providing maize B73 seeds. This work was supported by the National Science Foundation (RESEARCH-PGR grant no. 1748843 to E.S.B., S.F. and C.Q.), the German Research Foundation (DFG; fellowship no. 441540116 to T.J.) and the National Institutes of Health (T32 training grant no. HG000035 to J.T. and R01-GM079712 to C.Q. and J.T.C.).

Author information

Authors and Affiliations

Authors

Contributions

All authors conceived and interpreted experiments and wrote the article. T.J. and J.T. performed experiments. T.J. analysed the data and prepared the figures. T.J. and T.W. did the in silico modelling.

Corresponding authors

Correspondence to Josh T. Cuperus, Stanley Fields or Christine Queitsch.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Plants thanks Philip Benfey, Shira Weingarten-Gabbay 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

Extended Data Fig. 1 Promoter strength and in vivo expression levels of corresponding genes are not correlated.

a, Correlation (Pearson’s r) between the promoter strength and expression levels of the corresponding genes in the indicated species. Each boxplot (centre line, median; box limits, upper and lower quartiles; whiskers, 1.5 × interquartile range; points, outliers) represents the correlation for all individual tissue samples in the RNA-seq dataset (see Methods). The number of samples in the RNA-seq dataset is indicated at the bottom of the plot. b,c, Examples of the correlation between gene expression (Arabidopsis adult cotyledon (b) or maize root cortex (c) samples) and promoter strength as determined in tobacco leaves (b) or maize protoplasts (c). These examples correspond to the highest correlations in (a).

Extended Data Fig. 2 Strength of maize promoters depends on the TATA box location in maize protoplasts.

a, Histogram showing the percentage of maize promoters with a TATA box at the indicated position (reproduced from Fig. 4). Three peaks in the distribution of TATA boxes are highlighted in grey. Peak 1 spans bases −72 to −65, peak 2 spans bases −59 to −50, and peak 3 spans bases −34 to −24. b, Violin plots, boxplots and significance levels (as defined in Fig. 2) of promoter strength for maize promoters without enhancer in the indicated assay system. Promoters without a TATA box (−) were compared to those with a TATA box outside (+/−) or within one of the three peaks highlighted in (a).

Extended Data Fig. 3 The BREu element is most active in maize protoplasts.

a-d, Violin plots of promoter strength in tobacco leaves (a,c) or maize protoplasts (b,d). Promoters with a strong or intermediate TATA box (motif score ≥ 0.7; see Methods) were grouped by GC content and split into promoters without (left half, darker colour) or with (right half, lighter colour) a BREu (a,b), or BREd (c,d) element. Violin plots, boxplots and significance levels are as defined in Fig. 2. Only one half is shown for violin plots. e,f, Logoplots for promoters with a BREu (e) or BREd (f) before (WT) and after (mut) introducing mutations that disrupt the elements. g, Logoplots for promoters without a BRE (WT) and with an inserted BREu (+ BREu) or BREd (+ BREd) element. h, Boxplots and significance levels (as defined in Fig. 4) for the relative strength of the promoter variants shown in (e-g). The corresponding WT promoter was set to 0 (horizontal black line).

Extended Data Fig. 4 The Y patch is a plant-specific core promoter element.

a, Histogram showing the percentage of promoters with a TATA box at the indicated position. b,c, Violin plots of promoter strength in tobacco leaves (b) or maize protoplasts (c). Promoters were grouped by GC content and split into promoters without (left half, darker colour) or with (right half, lighter colour) a Y patch. Violin plots, boxplots and significance levels are as defined in Fig. 2. Only one half is shown for violin plots.

Extended Data Fig. 5 Core promoter elements at the TSS influence promoter strength.

a-d, Violin plots of promoter strength in tobacco leaves (a,c) or maize protoplasts (b,d). Promoters were grouped by GC content and split into promoters without (left half, darker colour) or with (right half, lighter colour) an Inr (a,b), or TCT (c,d) element at the TSS. Violin plots, boxplots and significance levels are as defined in Fig. 2. Only one half is shown for violin plots.

Extended Data Fig. 6 Transcription factor binding sites contribute to promoter strength in an assay system-dependent manner.

a-d, Violin plots of promoter strength for libraries without enhancer in tobacco leaves (a,c) or maize protoplasts (b,d). Promoters were grouped by GC content and split into promoters without (left half, darker colour) or with (right half, lighter colour) a binding site for TCP (a,b) or HSF (c,d) transcription factors. Violin plots, boxplots and significance levels are as defined in Fig. 2. Only one half is shown for violin plots.

Extended Data Fig. 7 Transcription factor binding sites are more active upstream of the TATA box.

a-c, Histograms showing the number of promoters with a TCP (a), HSF (b), or NAC (c) transcription factor binding site at the indicated position. d-i, Violin plots, boxplots and significance levels (as defined in Fig. 2) of promoter strength for libraries without enhancer in tobacco leaves (d-f) or maize protoplasts (g-i). Promoters were grouped by the position of their TCP (d,g), HSF (e,h), or NAC (f,i) transcription factor binding site relative to the TATA box: either upstream (up) or downstream (down).

Extended Data Fig. 8 Promoter-proximal transcription factor binding sites influence enhancer responsiveness.

a-f, Violin plots of enhancer responsiveness in tobacco leaves (a,c,e) or maize protoplasts (b,d,f). Promoters were grouped by GC content and split into promoters without (left half, darker colour) or with (right half, lighter colour) a TCP (a,b), WRKY (c,d), or B3 (e,f) transcription factor binding site. Violin plots, boxplots and significance levels are as defined in Fig. 2. Only one half is shown for violin plots.

Extended Data Fig. 9 Mutations in transcription factor binding sites alter light-dependency.

a-c, One or two T > G mutations were introduced in binding sites for TCP (a,b) or WRKY (c) transcription factors. The orientation of a binding site in the wild type promoter determined the bases that were mutated. d, Boxplots and significance levels (as defined in Fig. 4) for the relative light-dependency of promoters harbouring mutations in the indicated transcription factor binding site as shown in (a-c). The corresponding wild type promoter was set to 0 (horizontal black line).

Extended Data Fig. 10 The in silico evolution of promoters is most effective in early rounds.

a,b, 150 native and 160 synthetic promoters were subjected to 10 rounds of in silico evolution and the strength of the evolved promoters was predicted with the tobacco model (a) or the maize model (b). The black line represents the median promoter strength after each round. c,d, Correlation (Pearson’s R2 and Spearman’s ρ) between the predicted and experimentally determined strength of promoters after 0, 3, or 10 rounds of in silico evolution. Promoter strengths measured in tobacco leaves were compared to predictions from the tobacco model (c) and the data from maize protoplasts was compared to the predictions from the maize model (d). The models used for the in silico evolution are indicated on each plot.

Supplementary information

Supplementary Information

Supplementary Figs. 1–8.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–11.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jores, T., Tonnies, J., Wrightsman, T. et al. Synthetic promoter designs enabled by a comprehensive analysis of plant core promoters. Nat. Plants 7, 842–855 (2021). https://doi.org/10.1038/s41477-021-00932-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41477-021-00932-y

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