Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

A general design strategy for protein-responsive riboswitches in mammalian cells

Abstract

RNAs are ideal for the design of gene switches that can monitor and program cellular behavior because of their high modularity and predictable structure-function relationship. We have assembled an expression platform with an embedded modular ribozyme scaffold that correlates self-cleavage activity of designer ribozymes with transgene translation in bacteria and mammalian cells. A design approach devised to screen ribozyme libraries in bacteria and validate variants with functional tertiary stem-loop structures in mammalian cells resulted in a designer ribozyme with a protein-binding nutR-boxB stem II and a selected matching stem I. In a mammalian expression context, this designer ribozyme exhibited dose-dependent translation control by the N-peptide, had rapid induction kinetics and could be combined with classic small molecule–responsive transcription control modalities to construct complex, programmable genetic 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: Expression platform with embedded modular ribozyme scaffold.
Figure 2: Validation of self-cleavage performance of HHRs with nutR-matching structural elements.
Figure 3: Bacterial cell–based strategy to select nutR-boxB–containing HHR variants with loop II sequences restoring the tertiary interloop structures.
Figure 4: Gene-switch performance of the designer ribozyme Env140-nutR-H6 in mammalian cells.
Figure 5: A three-input AND logic gate combining translation and transcriptional gene expression control.

Similar content being viewed by others

Accession codes

Primary accessions

NCBI Reference Sequence

References

  1. Tigges, M., Marquez-Lago, T.T., Stelling, J. & Fussenegger, M. A tunable synthetic mammalian oscillator. Nature 457, 309–312 (2009).

    Article  CAS  Google Scholar 

  2. Nissim, L. & Bar-Ziv, R.H. A tunable dual-promoter integrator for targeting of cancer cells. Mol. Syst. Biol. 6, 444 (2010).

    Article  Google Scholar 

  3. Xie, Z., Wroblewska, L., Prochazka, L., Weiss, R. & Benenson, Y. Multi-input RNAi-based logic circuit for identification of specific cancer cells. Science 333, 1307–1311 (2011).

    Article  CAS  Google Scholar 

  4. Ausländer, S., Ausländer, D., Müller, M., Wieland, M. & Fussenegger, M. Programmable single-cell mammalian biocomputers. Nature 487, 123–127 (2012).

    Article  Google Scholar 

  5. Bonnet, J., Yin, P., Ortiz, M.E., Subsoontorn, P. & Endy, D. Amplifying genetic logic gates. Science 340, 599–603 (2013).

    Article  CAS  Google Scholar 

  6. Siuti, P., Yazbek, J. & Lu, T.K. Synthetic circuits integrating logic and memory in living cells. Nat. Biotechnol. 31, 448–452 (2013).

    Article  CAS  Google Scholar 

  7. Ausländer, S. & Fussenegger, M. From gene switches to mammalian designer cells: present and future prospects. Trends Biotechnol. 31, 155–168 (2013).

    Article  Google Scholar 

  8. Bashor, C.J., Horwitz, A.A., Peisajovich, S.G. & Lim, W.A. Rewiring cells: synthetic biology as a tool to interrogate the organizational principles of living systems. Annu. Rev. Biophys. 39, 515–537 (2010).

    Article  CAS  Google Scholar 

  9. Benenson, Y. Biomolecular computing systems: principles, progress and potential. Nat. Rev. Genet. 13, 455–468 (2012).

    Article  CAS  Google Scholar 

  10. Khalil, A.S. & Collins, J.J. Synthetic biology: applications come of age. Nat. Rev. Genet. 11, 367–379 (2010).

    Article  CAS  Google Scholar 

  11. Wittmann, A. & Suess, B. Engineered riboswitches: expanding researchers' toolbox with synthetic RNA regulators. FEBS Lett. 586, 2076–2083 (2012).

    Article  CAS  Google Scholar 

  12. Breaker, R.R. Riboswitches and the RNA world. Cold Spring Harb. Perspect. Biol. 4, a003566 (2012).

    Article  Google Scholar 

  13. Carey, J., Lowary, P.T. & Uhlenbeck, O.C. Interaction of R17 coat protein with synthetic variants of its ribonucleic acid binding site. Biochemistry 22, 4723–4730 (1983).

    Article  CAS  Google Scholar 

  14. Klein, D.J., Schmeing, T.M., Moore, P.B. & Steitz, T.A. The kink-turn: a new RNA secondary structure motif. EMBO J. 20, 4214–4221 (2001).

    Article  CAS  Google Scholar 

  15. Lazinski, D., Grzadzielska, E. & Das, A. Sequence-specific recognition of RNA hairpins by bacteriophage antiterminators requires a conserved arginine-rich motif. Cell 59, 207–218 (1989).

    Article  CAS  Google Scholar 

  16. Martinez, J., Patkaniowska, A., Urlaub, H., Lührmann, R. & Tuschl, T. Single-stranded antisense siRNAs guide target RNA cleavage in RNAi. Cell 110, 563–574 (2002).

    Article  CAS  Google Scholar 

  17. Patel, A.A. & Steitz, J.A. Splicing double: insights from the second spliceosome. Nat. Rev. Mol. Cell Biol. 4, 960–970 (2003).

    Article  CAS  Google Scholar 

  18. Goldfless, S.J., Belmont, B.J., de Paz, A.M., Liu, J.F. & Niles, J.C. Direct and specific chemical control of eukaryotic translation with a synthetic RNA-protein interaction. Nucleic Acids Res. 40, e64 (2012).

    Article  CAS  Google Scholar 

  19. Saito, H. et al. Synthetic translational regulation by an L7Ae-kink-turn RNP switch. Nat. Chem. Biol. 6, 71–78 (2010).

    Article  CAS  Google Scholar 

  20. Stripecke, R., Oliveira, C.C., McCarthy, J.E. & Hentze, M.W. Proteins binding to 5′ untranslated region sites: a general mechanism for translational regulation of mRNAs in human and yeast cells. Mol. Cell. Biol. 14, 5898–5909 (1994).

    Article  CAS  Google Scholar 

  21. Culler, S.J., Hoff, K.G. & Smolke, C.D. Reprogramming cellular behavior with RNA controllers responsive to endogenous proteins. Science 330, 1251–1255 (2010).

    Article  CAS  Google Scholar 

  22. Kashida, S., Inoue, T. & Saito, H. Three-dimensionally designed protein-responsive RNA devices for cell signaling regulation. Nucleic Acids Res. 40, 9369–9378 (2012).

    Article  CAS  Google Scholar 

  23. Yen, L. et al. Exogenous control of mammalian gene expression through modulation of RNA self-cleavage. Nature 431, 471–476 (2004).

    Article  CAS  Google Scholar 

  24. Hammann, C., Luptak, A., Perreault, J. & de la Peña, M. The ubiquitous hammerhead ribozyme. RNA 18, 871–885 (2012).

    Article  CAS  Google Scholar 

  25. Perreault, J. et al. Identification of hammerhead ribozymes in all domains of life reveals novel structural variations. PLoS Comput. Biol. 7, e1002031 (2011).

    Article  CAS  Google Scholar 

  26. Ausländer, S., Ketzer, P. & Hartig, J.S. A ligand-dependent hammerhead ribozyme switch for controlling mammalian gene expression. Mol. Biosyst. 6, 807–814 (2010).

    Article  Google Scholar 

  27. Soukup, G.A. & Breaker, R.R. Engineering precision RNA molecular switches. Proc. Natl. Acad. Sci. USA 96, 3584–3589 (1999).

    Article  CAS  Google Scholar 

  28. Wieland, M. & Hartig, J.S. Improved aptazyme design and in vivo screening enable riboswitching in bacteria. Angew. Chem. Int. Edn. Engl. 47, 2604–2607 (2008).

    Article  CAS  Google Scholar 

  29. Win, M.N. & Smolke, C.D. A modular and extensible RNA-based gene-regulatory platform for engineering cellular function. Proc. Natl. Acad. Sci. USA 104, 14283–14288 (2007).

    Article  CAS  Google Scholar 

  30. Wittmann, A. & Suess, B. Selection of tetracycline inducible self-cleaving ribozymes as synthetic devices for gene regulation in yeast. Mol. Biosyst. 7, 2419–2427 (2011).

    Article  CAS  Google Scholar 

  31. Hartig, J.S. et al. Protein-dependent ribozymes report molecular interactions in real time. Nat. Biotechnol. 20, 717–722 (2002).

    Article  CAS  Google Scholar 

  32. Vaish, N.K. et al. Monitoring post-translational modification of proteins with allosteric ribozymes. Nat. Biotechnol. 20, 810–815 (2002).

    Article  CAS  Google Scholar 

  33. Berschneider, B., Wieland, M., Rubini, M. & Hartig, J.S. Small-molecule-dependent regulation of transfer RNA in bacteria. Angew. Chem. Int. Edn. Engl. 48, 7564–7567 (2009).

    Article  CAS  Google Scholar 

  34. Wieland, M., Berschneider, B., Erlacher, M.D. & Hartig, J.S. Aptazyme-mediated regulation of 16S ribosomal RNA. Chem. Biol. 17, 236–242 (2010).

    Article  CAS  Google Scholar 

  35. Kumar, D., An, C.I. & Yokobayashi, Y. Conditional RNA interference mediated by allosteric ribozyme. J. Am. Chem. Soc. 131, 13906–13907 (2009).

    Article  CAS  Google Scholar 

  36. Chen, Y.Y., Jensen, M.C. & Smolke, C.D. Genetic control of mammalian T-cell proliferation with synthetic RNA regulatory systems. Proc. Natl. Acad. Sci. USA 107, 8531–8536 (2010).

    Article  CAS  Google Scholar 

  37. Ketzer, P. et al. Artificial riboswitches for gene expression and replication control of DNA and RNA viruses. Proc. Natl. Acad. Sci. USA 111, E554–E562 (2014).

    Article  CAS  Google Scholar 

  38. Chen, X. & Ellington, A.D. Design principles for ligand-sensing, conformation-switching ribozymes. PLoS Comput. Biol. 5, e1000620 (2009).

    Article  Google Scholar 

  39. Wieland, M., Ausländer, D. & Fussenegger, M. Engineering of ribozyme-based riboswitches for mammalian cells. Methods 56, 351–357 (2012).

    Article  CAS  Google Scholar 

  40. Khvorova, A., Lescoute, A., Westhof, E. & Jayasena, S.D. Sequence elements outside the hammerhead ribozyme catalytic core enable intracellular activity. Nat. Struct. Biol. 10, 708–712 (2003).

    Article  CAS  Google Scholar 

  41. Martick, M. & Scott, W.G. Tertiary contacts distant from the active site prime a ribozyme for catalysis. Cell 126, 309–320 (2006).

    Article  CAS  Google Scholar 

  42. Ferbeyre, G., Smith, J.M. & Cedergren, R. Schistosome satellite DNA encodes active hammerhead ribozymes. Mol. Cell. Biol. 18, 3880–3888 (1998).

    Article  CAS  Google Scholar 

  43. Legault, P., Li, J., Mogridge, J., Kay, L.E. & Greenblatt, J. NMR structure of the bacteriophage λ N peptide/boxB RNA complex: recognition of a GNRA fold by an arginine-rich motif. Cell 93, 289–299 (1998).

    Article  CAS  Google Scholar 

  44. Cilley, C.D. & Williamson, J.R. Analysis of bacteriophage N protein and peptide binding to boxB RNA using polyacrylamide gel coelectrophoresis (PACE). RNA 3, 57–67 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. de la Peña, M., Dufour, D. & Gallego, J. Three-way RNA junctions with remote tertiary contacts: a recurrent and highly versatile fold. RNA 15, 1949–1964 (2009).

    Article  Google Scholar 

  46. Jimenez, R.M., Delwart, E. & Lupták, A. Structure-based search reveals hammerhead ribozymes in the human microbiome. J. Biol. Chem. 286, 7737–7743 (2011).

    Article  CAS  Google Scholar 

  47. Saksmerprome, V., Roychowdhury-Saha, M., Jayasena, S., Khvorova, A. & Burke, D.H. Artificial tertiary motifs stabilize trans-cleaving hammerhead ribozymes under conditions of submillimolar divalent ions and high temperatures. RNA 10, 1916–1924 (2004).

    Article  CAS  Google Scholar 

  48. Oubridge, C., Ito, N., Evans, P.R., Teo, C.H. & Nagai, K. Crystal structure at 1.92 Å resolution of the RNA-binding domain of the U1A spliceosomal protein complexed with an RNA hairpin. Nature 372, 432–438 (1994).

    Article  CAS  Google Scholar 

  49. Ausländer, D., Wieland, M., Ausländer, S., Tigges, M. & Fussenegger, M. Rational design of a small molecule-responsive intramer controlling transgene expression in mammalian cells. Nucleic Acids Res. 39, e155 (2011).

    Article  Google Scholar 

  50. Schlatter, S., Rimann, M., Kelm, J. & Fussenegger, M. SAMY, a novel mammalian reporter gene derived from Bacillus stearothermophilus α-amylase. Gene 282, 19–31 (2002).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank P. Silver (Harvard Medical School) for providing the U1A-encoding vector pPS2042; M. Müller, F. Sedlmayer, M. Wieland, A. Bosshart and M. Jeschek for generous advice; and S. Stebler for experimental assistance. We are grateful to the single cell unit team V. Jäggin, M. Dessing and T. Horn for assistance with flow cytometry and fluorescence microscopy. This work was supported by a European Research Council (ERC) advanced grant (no. 321381) and in part by the Cantons of Basel and the Swiss Confederation within the INTERREG IV A.20 tri-national research program.

Author information

Authors and Affiliations

Authors

Contributions

S.A. designed the project. S.A., P.S., C.R., D.A., J.S.H. and M.F. analyzed results and wrote the manuscript. S.A., P.S., C.R. and D.A. performed the experimental work.

Corresponding author

Correspondence to Martin Fussenegger.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Design of the HHR expression platform.

PmeI/SpeI restriction endonucleases are used for simple one-step swapping of the HHR scaffold among bacterial and mammalian expression vectors. Whereas the bacterial expression vector contains a PT7-driven eGFP expression unit (pET16b-eGFP) with the HHR scaffold embedded in the 5’ untranslated region, the mammalian expression vector harbors a PSV40-driven SEAP expression unit (pSEAP2-control) with the HHR scaffold inserted in the 3’ untranslated region.

Supplementary Figure 2 Ribozyme-dependent EGFP expression profiles in HEK-293 cells.

GFP expression profiles correlating with the self-cleavage performance of Env25ac/inac (pDA230/pDA231), Env140ac/inac (pDA228/pDA229) and Env143ac/inac (pDA232/pDA233) HHRs in HEK-293 cells (set-up shown in Fig. 1c). Error bars are mean±SD of a representative experiment performed in triplicates.

Supplementary Figure 3 Detailed experimental guidelines for the selection of designer ribozymes.

The HHR scaffold expression library containing degenerated nucleotides in the desired loop structure is constructed by site-directed PCR-based mutagenesis using the bacterial expression vector as template. The HHR expression library is transformed into bacteria and functional hybrid ribozymes correlating with high-level eGFP expression are enriched by FACS-mediated sorting. The library subpopulation containing HHR variants with highest self-cleavage performance in bacteria is swapped into the corresponding mammalian expression vector and transformed again into bacteria to isolate clonal HHR variants which are subsequently transfected into mammalian cells for validation. The best-in-class designer ribozymes with optimal functional tertiary inter-loop contacts were identified in mammalian cells by correlating low-level SEAP expression and characterized by sequence analysis.

Supplementary Figure 4 General strategy for the design of protein-responsive mammalian riboswitches.

(a) Ribozyme candidates with high structural homology to a desired natural ribonucleoprotein (RNP) complex are identified in a comprehensive database of natural HHR sequences. (b) Replacement of the natural ribozyme’s stem loop by the RNP-derived stem loop creates a hybrid ribozyme whose self-cleavage performance is scored in a bacterial and mammalian expression context. (c) A complementary stem loop structure matching the RNP-derived stem loop to form functional inter-loop contacts is selected from a library of the hybrid ribozyme containing degenerated loop sequences by FACS-mediated selection in bacteria and validation in mammalian cells using a cross-kingdom ribozyme expression platform. The selected designer ribozyme enables protein-adjustable translation control in mammalian cells.

Supplementary Figure 5 Ribozyme performance of selected Env140-nutR-5Nac HHR clones in HEK-293 cells.

SEAP expression profiles correlating with the self-cleavage performance of selected Env140/nutR/5Nac HHR clones in HEK-293 cells relative to active Env140ac (100%) and inactive Env140inac (0%). The blue bar represents the parental hybrid ribozyme Env140/nutRac, which served as a template for the degenerated loop I library.

Supplementary Figure 6 Schematic representation of N-peptide–triggered Env140-nutR-H6–mediated SEAP expression control in mammalian cells.

In the absence of N-peptide self-cleavage of the designer ribozyme Env140/nutR/H6 embedded in the 3’ untranslated region of a SEAP-encoding mRNA (pPST67[H6]) removes the poly(A) tail which leads to nuclease (red pacman) –mediated destruction and shut down of SEAP expression. However, binding of N-peptide (pSA776) to stem loop II of Env140/nutR/H6 prevents formation of the tertiary inter-loop structures which inactivates the designer ribozyme and SEAP is translated from an intact mRNA exported to the cytosol and efficiently secreted by the host cell.

Supplementary Figure 7 Fluorescence micrographs of mCherry-transfected and N-peptide–mCherry–transfected mammalian cells.

Fluorescence micrographs of different cell lines containing the PSV40-driven expression mCherry expression vector (pFS29) or a PSV40-driven expression unit encoding the N-peptide-mCherry fusion protein (pSA776). Micrographs show the bright field (BF) channel as well as the 561nm-laser channel (561nm). Micrographs were taken 48 hours after transfection.

Supplementary Figure 8 Impact of N-peptide on the performance of different active and isogenic inactive HHR constructs.

(a-b) Impact of N-peptide on the performance of different active (a) and isogenic inactive (b) HHR constructs and the corresponding SEAP expression levels. Black bars: mCherry (pFS29); blue bars: N-peptide-mCherry (pSA776). SEAP expression analysis was performed 22 hours after transfection. Ac., active; inac., inactive. Error bars are mean±SEM (n=3). Please see Supplementary Table 1 for plasmid design and Supplementary Table 2 for ribozyme sequence.

Supplementary Figure 9 SEAP expression kinetics of active Env140ac and inactive Env140-nutR-H6inac ribozymes in CHO-K1 cells.

(a) SEAP expression kinetics of CHO-K1 cells cotransfected with active Env140ac HHR-embedded SEAP expression vector (pSA741) and either the mCherry- (pFS29; black line) or the N-peptide-mCherry-encoding expression vector (pSA776; blue line). (b) SEAP expression kinetics of CHO-K1 cells cotransfected with inactive Env140/nutR/H6inac HHR-embedded SEAP expression vector (pPST69) and either mCherry- (pFS29; black line) or the N-peptide-mCherry encoding expression vector (pSA776; blue line). Ac., active; inac., inactive. Error bars are mean±SEM (n=3).

Supplementary Figure 10 In vitro cleavage kinetics of the Env140ac ribozyme.

In-vitro cleavage kinetics of the Env140ac ribozyme in the presence (blue squares) and absence (black circles) of 4μM N-peptide. The cleavage kinetics of the inactive Env140inac ribozyme is shown in red diamonds.

Supplementary Figure 11 Impact of N-peptidemut/nat on Env140ac– and Env140-nutR-H6ac–controlled EGFP expression in bacterial cells.

(a) Amino acid identity of native and mutated N-peptide (N-peptidenat/mut). (b) Flow cytometry-based histograms showing eGFP-mediated fluorescence correlating with the self-cleavage performance of Env140ac (pPST49) or Env140/nutR/H6ac (pPST106) in the presence of native (pPST102) or mutated (pPST171) N-peptide in E. coli. Red peak indicates co-transformation with native N-peptide, blue peak with mutated N-peptide.

Supplementary Figure 12 Engineering a U1Ap-responsive designer ribozyme.

(a) Secondary structure of the U1hpII RNA motif. (b) Secondary RNA structures of stem loops I & II of the Env14 hammerhead ribozyme embedded on the ribozyme scaffold. (c) Flow cytometry-based histograms showing eGFP-mediated fluorescence correlating with the self-cleavage performance of Env14ac/inac (pPST172/pPST173) HHRs in bacteria (set-up shown in Fig. 1b). Red peaks indicate active HHR, blue peaks show inactive HHR and the dashed peak represents the autofluorescence of bacteria. (d) SEAP expression profiles correlating with the self-cleavage performance of Env14ac/inac (pSA909/pSA917) in CHO-K1 cells (set-up shown in Fig. 1c). (e) SEAP-based performance analysis of Env14ac stem loop I variants in CHO-K1 cells. Nucleotide mutations compared to the wild-type stem loop I are shown in blue. The ribozyme activity was normalized to the cleavage activity of the active (100%) and inactive Env14 (0%). (f) Profiling of the U1Ap-triggered SEAP expression in CHO-K1 cells co-transfected with the SEAP expression vector containing a 3’-UTR-embedded Env14ac (pSA909), Env14/A4Gac (pSA912), Env14/CGac (pSA914) Env14/U1hpIIac (pSA928) or Env140/nutR/H6ac (pPST67[H6]) and either mCherry (pFS20; black bars) or U1Ap-mCherry (pSA906; orange bars). The results for Env14/U1hpIIac (pSA928) and Env140/nutR/H6ac (pPST67[H6]) are also shown in Figure 4j and are here included for comparison. eGFP/SEAP expression was analyzed 22 hours after inoculation/transfection. Ac., active; inac., inactive. Error bars are mean±SEM (n=3).

Supplementary Figure 13 Two-input logic gates combining ribozyme-based N-peptide–triggered translation control with transcription switches.

(a) Genetic switchboard of the AND Boolean logic gate circuit consisting of the doxycycline-inducible Ptight-driven N-peptide-mCherry expression unit (pSA781) transactivated by rtTA (reverse tetracycline-dependent transactivator; pTet-ON) and a vanillic acid-inducible PVANON8-driven SEAP expression unit with 3’-UTR-embedded HHR (Env140/nutR/H6ac; pSA783) whose transcription is transrepressed by VanA4 (vanillic acid-dependent transrepressor; pCK188) and translation is modulated by N-peptide. (b) Combining the two input signals according to the truth table programs transfected HEK-293 cells to produce the enzymatic output SEAP only in the presence of both input signals doxycyline (Dox) and vanillic acid (Van). (c) Genetic switchboard of the B N-IMPLY A Boolean logic gate circuit consisting of an erythromycin-repressible PETR2-driven N-peptide-mCherry expression unit (pSA763) transactivated by ET1 (erythromycin-dependent transactivator; pWW35) a vanillic acid-inducible PVANON8-driven SEAP expression unit with 3’-UTR-embedded HHR (Env140/nutR/H6ac; pSA783) whose transcription is transrepressed by VanA4 (vanillic acid-dependent transrepressor; pCK188) and translation is modulated by N-peptide. (d) Combining the two input according to the truth table programs transfected HEK-293 cells to produce the enzymatic output SEAP only in the presence of the vanillic acid (Van) input and not in the presence of erythromycin (Ery).

Supplementary Figure 14 Dose dependence of the small-molecule–responsive transcription-control systems used for the assembly of logic gates.

(a) HEK-293 cells cotransfected with the vanillic acid-dependent transrepressor (VanA4; pCK188) and the VanA4-specific PvanON8-driven SEAP expression vector with embedded Env140/nutR/H6ac (PVanON8-SEAP-pA; pSA783) and either N-peptide-mCherry (pSA776; circles) or mCherry (pFS29; squares) were programmed with different concentrations of vanillic acid and profiled for SEAP expression after 40h. (b) HEK-293 cells were cotransfected with the reverse tetracycline-dependent transactivator (rtTA; pTet-ON) and the rtTA-specific Ptight-driven N-peptide-mCherry expression vector (Ptight-N-peptide-mCherry-pA; pSA781) and programmed with different concentrations of the antibiotic doxycyline. Red fluorescence was quantified by flow cytometry 40 hours after transfection. (c) HEK-293 cells cotransfected with the erythromycin-dependent transactivator (ET1; pWW35) and the ET1-specific PETR2-driven driven N-peptide-mCherry expression vector (PETR2-N-peptide-mCherry-pA; pSA763) and programmed with different concentrations of the antibiotic erythromycin. Red fluorescence was quantified using flow cytometry 40 hours after transfection. Error bars are mean±SEM (n=3).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–14, Supplementary Tables 1–3 and Supplementary Note (PDF 5289 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ausländer, S., Stücheli, P., Rehm, C. et al. A general design strategy for protein-responsive riboswitches in mammalian cells. Nat Methods 11, 1154–1160 (2014). https://doi.org/10.1038/nmeth.3136

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.3136

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