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:

Engineering an allosteric transcription factor to respond to new ligands

Abstract

Genetic regulatory proteins inducible by small molecules are useful synthetic biology tools as sensors and switches. Bacterial allosteric transcription factors (aTFs) are a major class of regulatory proteins, but few aTFs have been redesigned to respond to new effectors beyond natural aTF-inducer pairs. Altering inducer specificity in these proteins is difficult because substitutions that affect inducer binding may also disrupt allostery. We engineered an aTF, the Escherichia coli lac repressor, LacI, to respond to one of four new inducer molecules: fucose, gentiobiose, lactitol and sucralose. Using computational protein design, single-residue saturation mutagenesis or random mutagenesis, along with multiplex assembly, we identified new variants comparable in specificity and induction to wild-type LacI with its inducer, isopropyl β-D-1-thiogalactopyranoside (IPTG). The ability to create designer aTFs will enable applications including dynamic control of cell metabolism, cell biology and synthetic gene circuits.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: General workflow for designing new ligand binding in an aTF.
Figure 2: Characterization of Rosetta-designed variants responding to new inducers.
Figure 3: Characterization of gentiobiose-responsive variants from the protein-wide single-amino-acid substitution library.
Figure 4: Ligand cross-reactivity of LacI variants.
Figure 5: Activity maturation of LacI.
Figure 6: Crystal structure and GFP induction with ligand of sucralose-binding LacI design variant (D149T,S193D,V150A,I156L).

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Weickert, M.J. & Adhya, S. A family of bacterial regulators homologous to Gal and Lac repressors. J. Biol. Chem. 267, 15869–15874 (1992).

    CAS  PubMed  Google Scholar 

  2. Schell, M.A. Molecular biology of the LysR family of transcriptional regulators. Annu. Rev. Microbiol. 47, 597–626 (1993).

    Article  CAS  PubMed  Google Scholar 

  3. Gallegos, M.T., Schleif, R., Bairoch, A., Hofmann, K. & Ramos, J.L. Arac/XylS family of transcriptional regulators. Microbiol. Mol. Biol. Rev. 61, 393–410 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Ramos, J.L. et al. The TetR family of transcriptional repressors. Microbiol. Mol. Biol. Rev. 69, 326–356 (2005).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  5. Lutz, R. & Bujard, H. Independent and tight regulation of transcriptional units in Escherichia coli via the LacR/O, the TetR/O and AraC/I1-I2 regulatory elements. Nucleic Acids Res. 25, 1203–1210 (1997).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  6. Dietrich, J.A., Shis, D.L., Alikhani, A. & Keasling, J.D. Transcription factor-based screens and synthetic selections for microbial small-molecule biosynthesis. ACS Synth. Biol. 2, 47–58 (2013).

    Article  CAS  PubMed  Google Scholar 

  7. Raman, S., Rogers, J.K., Taylor, N.D. & Church, G.M. Evolution-guided optimization of biosynthetic pathways. Proc. Natl. Acad. Sci. USA 111, 17803–17808 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Lu, T.K., Khalil, A.S. & Collins, J.J. Next-generation synthetic gene networks. Nat. Biotechnol. 27, 1139–1150 (2009).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  9. Dietrich, J.A., McKee, A.E. & Keasling, J.D. High-throughput metabolic engineering: advances in small-molecule screening and selection. Annu. Rev. Biochem. 79, 563–590 (2010).

    Article  CAS  PubMed  Google Scholar 

  10. Tang, S.-Y. & Cirino, P.C. Design and application of a mevalonate-responsive regulatory protein. Angew. Chem. Int. Edn Engl. 50, 1084–1086 (2011).

    Article  CAS  Google Scholar 

  11. Süel, G.M., Lockless, S.W., Wall, M.A. & Ranganathan, R. Evolutionarily conserved networks of residues mediate allosteric communication in proteins. Nat. Struct. Biol. 10, 59–69 (2003).

    Article  PubMed  Google Scholar 

  12. Markiewicz, P., Kleina, L.G., Cruz, C., Ehret, S. & Miller, J.H. Genetic studies of the lac repressor. XIV. Analysis of 4000 altered Escherichia coli lac repressors reveals essential and non-essential residues, as well as “spacers” which do not require a specific sequence. J. Mol. Biol. 240, 421–433 (1994).

    Article  CAS  PubMed  Google Scholar 

  13. Raman, S., Taylor, N., Genuth, N., Fields, S. & Church, G.M. Engineering allostery. Trends Genet. 30, 521–528 (2014).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  14. Collins, C.H., Arnold, F.H. & Leadbetter, J.R. Directed evolution of Vibrio fischeri LuxR for increased sensitivity to a broad spectrum of acyl-homoserine lactones. Mol. Microbiol. 55, 712–723 (2005).

    Article  CAS  PubMed  Google Scholar 

  15. Cebolla, A., Sousa, C. & de Lorenzo, V. Effector specificity mutants of the transcriptional activator NahR of naphthalene degrading Pseudomonas define protein sites involved in binding of aromatic inducers. J. Biol. Chem. 272, 3986–3992 (1997).

    Article  CAS  PubMed  Google Scholar 

  16. Wise, A.A. & Kuske, C.R. Generation of novel bacterial regulatory proteins that detect priority pollutant phenols. Appl. Environ. Microbiol. 66, 163–169 (2000).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  17. Galvão, T.C., Mencía, M. & de Lorenzo, V. Emergence of novel functions in transcriptional regulators by regression to stem protein types. Mol. Microbiol. 65, 907–919 (2007).

    Article  PubMed  Google Scholar 

  18. Scholz, O., Köstner, M., Reich, M., Gastiger, S. & Hillen, W. Teaching TetR to recognize a new inducer. J. Mol. Biol. 329, 217–227 (2003).

    Article  CAS  PubMed  Google Scholar 

  19. Tang, S.-Y., Fazelinia, H. & Cirino, P.C. AraC regulatory protein mutants with altered effector specificity. J. Am. Chem. Soc. 130, 5267–5271 (2008).

    Article  CAS  PubMed  Google Scholar 

  20. Jha, R.K., Chakraborti, S., Kern, T.L., Fox, D.T. & Strauss, C.E.M. Rosetta comparative modeling for library design: engineering alternative inducer specificity in a transcription factor. Proteins 10.1002/prot.24828 (13 May 2015).

  21. de Los Santos, E.L.C., Meyerowitz, J.T., Mayo, S.L. & Murray, R.M. Engineering transcriptional regulator effector specificity using computational design and in vitro rapid prototyping: developing a vanillin sensor. ACS Synth. Biol. 10.1021/acssynbio.5b00090 (19 August 2015).

  22. AbuOun, M. et al. Genome scale reconstruction of a Salmonella metabolic model: comparison of similarity and differences with a commensal Escherichia coli strain. J. Biol. Chem. 284, 29480–29488 (2009).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  23. Jiang, L. et al. De novo computational design of retro-aldol enzymes. Science 319, 1387–1391 (2008).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  24. Röthlisberger, D. et al. Kemp elimination catalysts by computational enzyme design. Nature 453, 190–195 (2008).

    Article  PubMed  Google Scholar 

  25. Tinberg, C.E. et al. Computational design of ligand-binding proteins with high affinity and selectivity. Nature 501, 212–216 (2013).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  26. Kosuri, S. et al. Scalable gene synthesis by selective amplification of DNA pools from high-fidelity microchips. Nat. Biotechnol. 28, 1295–1299 (2010).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  27. Swint-Kruse, L., Elam, C.R., Lin, J.W., Wycuff, D.R. & Shive Matthews, K. Plasticity of quaternary structure: twenty-two ways to form a LacI dimer. Protein Sci. 10, 262–276 (2001).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  28. Swint-Kruse, L., Zhan, H., Fairbanks, B.M., Maheshwari, A. & Matthews, K.S. Perturbation from a distance: mutations that alter LacI function through long-range effects. Biochemistry 42, 14004–14016 (2003).

    Article  CAS  PubMed  Google Scholar 

  29. Xu, J. & Matthews, K.S. Flexibility in the inducer binding region is crucial for allostery in the Escherichia coli lactose repressor. Biochemistry 48, 4988–4998 (2009).

    Article  CAS  PubMed  Google Scholar 

  30. DeVito, J.A. Recombineering with tolC as a selectable/counter-selectable marker: remodeling the rRNA operons of Escherichia coli. Nucleic Acids Res. 36, e4 (2008).

    Article  PubMed  Google Scholar 

  31. Rogers, J.K. et al. Synthetic biosensors for precise gene control and real-time monitoring of metabolites. Nucleic Acids Res. 43, 7648–7660 (2015).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  32. Mirny, L.A. & Gelfand, M.S. Using orthologous and paralogous proteins to identify specificity-determining residues in bacterial transcription factors. J. Mol. Biol. 321, 7–20 (2002).

    Article  CAS  PubMed  Google Scholar 

  33. Pei, J., Cai, W., Kinch, L.N. & Grishin, N.V. Prediction of functional specificity determinants from protein sequences using log-likelihood ratios. Bioinformatics 22, 164–171 (2006).

    Article  CAS  PubMed  Google Scholar 

  34. Bell, C.E. & Lewis, M. A closer view of the conformation of the Lac repressor bound to operator. Nat. Struct. Biol. 7, 209–214 (2000).

    Article  CAS  PubMed  Google Scholar 

  35. Werstuck, G. & Green, M.R. Controlling gene expression in living cells through small molecule-RNA interactions. Science 282, 296–298 (1998).

    Article  CAS  PubMed  Google Scholar 

  36. Guntas, G., Mansell, T.J., Kim, J.R. & Ostermeier, M. Directed evolution of protein switches and their application to the creation of ligand-binding proteins. Proc. Natl. Acad. Sci. USA 102, 11224–11229 (2005).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  37. Licitra, E.J. & Liu, J.O. A three-hybrid system for detecting small ligand-protein receptor interactions. Proc. Natl. Acad. Sci. USA 93, 12817–12821 (1996).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  38. Maynard-Smith, L.A., Chen, L.-C., Banaszynski, L.A., Ooi, A.G.L. & Wandless, T.J. A directed approach for engineering conditional protein stability using biologically silent small molecules. J. Biol. Chem. 282, 24866–24872 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Qin, Y. et al. Screening and identification of a fungal β-glucosidase and the enzymatic synthesis of gentiooligosaccharide. Appl. Biochem. Biotechnol. 163, 1012–1019 (2011).

    Article  CAS  PubMed  Google Scholar 

  40. Gossen, M. & Bujard, H. Tight control of gene expression in mammalian cells by tetracycline-responsive promoters. Proc. Natl. Acad. Sci. USA 89, 5547–5551 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Wang, H.H. et al. Programming cells by multiplex genome engineering and accelerated evolution. Nature 460, 894–898 (2009).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  42. Datsenko, K.A. & Wanner, B.L. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc. Natl. Acad. Sci. USA 97, 6640–6645 (2000).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  43. Pédelacq, J.-D., Cabantous, S., Tran, T., Terwilliger, T.C. & Waldo, G.S. Engineering and characterization of a superfolder green fluorescent protein. Nat. Biotechnol. 24, 79–88 (2006).

    Article  PubMed  Google Scholar 

  44. Hawkins, P.C.D. et al. Conformer generation with OMEGA: algorithm and validation using high quality structures from the Protein Databank and the Cambridge Structural Database. J. Chem. Inf. Model. 50, 572–584 (2010).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  45. Hawkins, P.C.D. & Nicholls, A. Conformer generation with OMEGA: learning from the data set and the analysis of failures. J. Chem. Inf. Model. 52, 2919–2936 (2012).

    Article  CAS  PubMed  Google Scholar 

  46. Kabsch, W. XDS. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Strong, M. et al. Toward the structural genomics of complexes: crystal structure of a PE/PPE protein complex from Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. USA 103, 8060–8065 (2006).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  48. Emsley, P., Lohkamp, B., Scott, W.G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  50. Winn, M.D., Isupov, M.N. & Murshudov, G.N. Use of TLS parameters to model anisotropic displacements in macromolecular refinement. Acta Crystallogr. D Biol. Crystallogr. 57, 122–133 (2001).

    Article  CAS  PubMed  Google Scholar 

  51. Altschul, S.F. et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389–3402 (1997).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  52. Larkin, M.A. et al. Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–2948 (2007).

    CAS  PubMed  Google Scholar 

  53. Majumdar, A., Rudikoff, S. & Adhya, S. Purification and properties of Gal repressor:pL-galR fusion in pKC31 plasmid vector. J. Biol. Chem. 262, 2326–2331 (1987).

    CAS  PubMed  Google Scholar 

  54. Meinhardt, S. et al. Novel insights from hybrid LacI/GalR proteins: family-wide functional attributes and biologically significant variation in transcription repression. Nucleic Acids Res. 40, 11139–11154 (2012).

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  55. Magoč, T. & Salzberg, S.L. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27, 2957–2963 (2011).

    Article  PubMed Central  PubMed  Google Scholar 

  56. Kent, W.J. BLAT--the BLAST-like alignment tool. Genome Res. 12, 656–664 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Bolstad, B.M., Irizarry, R.A., Astrand, M. & Speed, T.P. A comparison of normalization methods for high density oligonucleotide array data based on variance and bias. Bioinformatics 19, 185–193 (2003).

    Article  CAS  PubMed  Google Scholar 

  58. Hadley, W. ggplot2: Elegant Graphics for Data Analysis (Springer, 2009).

  59. Suckow, J. et al. Genetic studies of the Lac repressor. XV: 4000 single amino acid substitutions and analysis of the resulting phenotypes on the basis of the protein structure. J. Mol. Biol. 261, 509–523 (1996).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank B. Turczyk and D. Weigand for synthesizing the single-amino-acid substitution library on the Custom Array synthesizer, and G. Cuneo and V. Toxavidis for assistance with flow cytometry and FACS. We thank Rosetta@home participants for providing the computing resources necessary for this work. This work was supported by the US Department of Energy (DOE) (DE-FG02-02ER63445 to G.M.C.), a Wyss Technology Development Fellowship (to S.R.) and the US National Institute of General Medical Sciences (grant 1P41 GM103533 to S.F.). The sucralose-responsive LacI mutant was purified and crystallized with assistance from the UCLA-DOE Protein Expression Technology Center, the UCLA-DOE X-ray Crystallography Core Facility (both supported by DOE grant DE-FC02-02ER63421) and the UCLA Crystallization Core Facility; in particular we thank M. Collazo for help with protein crystallization. X-ray data collection was facilitated by M. Capel, K. Rajashankar, N. Sukumar, F. Murphy and I. Kourinov of the Northeastern Collaborative Access Team beamline 24-ID-C at the Advanced Photon Source of Argonne National Laboratory, which is supported by US National Institutes of Health grants P41 RR015301 and P41 GM103403. Use of the Advanced Photon Source is supported by the DOE under contract DE-AC02-06CH11357.

Author information

Authors and Affiliations

Authors

Contributions

N.D.T., F.J.I., G.M.C. and S.R. conceived the study. N.D.T., S.F., G.M.C. and S.R. designed experiments. N.D.T., A.S.G. and S.R. performed experiments and carried out bioinformatic studies. R.M. and D.B. generated computational protein design candidates. S.C., D.C., M.A.A. and S.K. solved the crystal structure of a sucralose-binding variant. S.K. helped with Agilent OLS chip library design. J.K.R. helped optimize screening protocols. N.D.T., A.S.G., S.F., G.M.C. and S.R. analyzed the data. N.D.T., A.S.G., S.F., G.M.C. and S.R. wrote the paper.

Corresponding author

Correspondence to Srivatsan Raman.

Ethics declarations

Competing interests

S.R., N.D.T. and G.M.C. have filed a patent application (PCT/US15/16868) covering biosensor design methods.

Integrated supplementary information

Supplementary Figure 1 Chemical structure of ligands.

Allolactose and IPTG (native and synthetic inducers of LacI, respectively), and the four new inducers – fucose, lactitol, sucralose and gentiobiose.

Supplementary Figure 2 Coverage of all single-amino-acid substitutions found in the pre-selection single-site saturation mutagenesis library at gentiobiose-responsive positions.

Residues are ordered by protein region as in Supplementary Fig. 8 for comparison. (a) For each position of wild-type LacI, 19 substitutions were synthesized for the single amino-acid substitution library. By next-generation sequencing we measured how many of the 19 possible substitutions were found either before or after negative selection for positions showing response to gentiobiose. Mutants missing from the input library are likely due to synthesis or cloning inefficiencies. (b) Most (>80%) of the positions involved in gentiobiose response were found to have at least 18 of 19 single amino-acid substitutions prior to positive selection. For all 360 positions of LacI (not shown), we found 195 (~54%) positions contained all 19 substitutions, 238 (~66%) contained at least 18, and 306 (85%) contained at least 14 substitutions.

Source data

Supplementary Figure 3 Flow cytometric characterization of aTF library screening.

RFU denotes relative fluorescence units. (a) Flow cytometry histogram of representative LacI variant library (red) before and (blue) after colicin E1 negative selection. (b) Flow cytometry histogram of representative LacI variant library with no inducer molecule present (blue) or exposed to a new target inducer molecule (red).

Supplementary Figure 4 Fold induction of WT LacI with IPTG inducer.

Dose-response curve of WT LacI with IPTG, fold induction shown on the y-axis and IPTG concentration (mM) on the x-axis.

Supplementary Figure 5 Sequence and fold induction of the top-scoring full-length Rosetta design variants.

Induction response in relative fluorescence units (RFU) with and without ligand for WT LacI and top five scoring full-length Rosetta design variants for sucralose, lactitol and fucose. WT LacI was induced with IPTG, and the full-length Rosetta design variants were induced with their respective target ligands. The mutations in each Rosetta design variant are shown above the bar graph. All ligands were supplemented at 10 mM.

Source data

Supplementary Figure 6 Comparison of fucose, lactitol and sucralose response versus single-amino-acid substitutions found after negative selection.

(a) Fucose responsive induction values are shown pink. The induction values show the maximum weighted fold-change of response after positive selection. The black outlines indicate depletion of next-generation sequencing reads for single amino-acid substitutions after negative selection. The depletion value is the log2 fold-change of reads prior to negative selection divided by the read counts after negative selection. Higher depletion values indicate position and side-chain combinations that are lost after negative selection. Read counts were quantile normalized between pre- and post-selection separately for each amplicon (see Online Methods). (b) Lactitol responsive induction values versus depletion values. (c) Sucralose responsive induction values versus depletion values. Negative depletion values are not shown.

Source data

Supplementary Figure 7 Comparison of conservation of amino acids and mutations found for fucose response.

Mutations conferring fucose response in LacI are shown as red outlines. (a) A set of 41 LacI orthologs were aligned and the frequency of amino acid utilization is shown in blue. (b) Five experimentally validated sequences of GalR/S known to bind fucose were aligned with E. coli LacI and shown with respect to LacI positions. Mutations at positions 79 and 273 overlap with preferentially conserved amino acids in the GalR set, shown with arrows. The highest inducer at position 291 was conserved in neither LacI or fucose-responding GalR/S.

Source data

Supplementary Figure 8 Comparison of gentiobiose response versus single-amino-acid substitutions found after negative selection.

Induction values for gentiobiose-responding mutants are shown in pink. The induction values show the maximum weighted fold-change of response after positive selection. The color shades outside the amino acid substitution profile denotes the location of the residue in the binding pocket, dimerization interface, DNA-binding domain or as unclassified. The black outlines indicate depletion of next-generation sequencing reads for single amino-acid substitutions. The depletion value is the log2 fold-change of reads prior to negative selection divided by the read counts after negative selection. Read counts were quantile normalized between pre- and post-selection separately for each amplicon (see Online Methods).

Supplementary Figure 9 Cross-reactivity of additional LacI variants toward three other untargeted inducers and IPTG.

For additional variants displayed in Fig. 2, a dose-response was determined for non-target ligands and IPTG. Values displayed represent the highest fold induction at any ligand concentration. Inducers are colored as follows: gentiobiose, red; fucose, green; lactitol, blue; sucralose, magenta; and IPTG, black. Variants displayed were designed for binding to (a) gentiobiose, (b) fucose, (c) lactitol, and (d) sucralose. Error bars represent standard deviation of fold induction from three biological replicates.

Supplementary Figure 10 User guide for aTF redesign.

A detailed flowchart that guides the user through the choice of mutagenesis methods based on the choice of the target ligand for aTF redesign. We offer general guidelines on what we consider acceptable fold induction and specificity values by target and native ligands, presented as proportion of WT aTF induction, after the two-stage enrichment screen and following activity maturation. These guidelines could be adjusted on a case-by-case basis depending on the number and quality of ligand-responsive variants after the two-stage enrichment screen, and the nature of downstream application.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10, Supplementary Tables 1–6 and Supplementary Note (PDF 1962 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Taylor, N., Garruss, A., Moretti, R. et al. Engineering an allosteric transcription factor to respond to new ligands. Nat Methods 13, 177–183 (2016). https://doi.org/10.1038/nmeth.3696

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.3696

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