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 small and highly sensitive red/far-red optogenetic switch for applications in mammals

Abstract

Optogenetic technologies have transformed our ability to precisely control biological processes in time and space. Yet, current eukaryotic optogenetic systems are limited by large or complex optogenetic modules, long illumination times, low tissue penetration or slow activation and deactivation kinetics. Here, we report a red/far-red light-mediated and miniaturized Δphytochrome A (ΔPhyA)-based photoswitch (REDMAP) system based on the plant photoreceptor PhyA, which rapidly binds the shuttle protein far-red elongated hypocotyl 1 (FHY1) under illumination with 660-nm light with dissociation occurring at 730 nm. We demonstrate multiple applications of REDMAP, including dynamic on/off control of the endogenous Ras/Erk mitogen-activated protein kinase (MAPK) cascade and control of epigenetic remodeling using a REDMAP-mediated CRISPR–nuclease-deactivated Cas9 (CRISPR–dCas9) (REDMAPcas) system in mice. We also demonstrate the utility of REDMAP tools for in vivo applications by activating the expression of transgenes delivered by adeno-associated viruses (AAVs) or incorporated into cells in microcapsules implanted into mice, rats and rabbits illuminated by light-emitting diodes (LEDs). Further, we controlled glucose homeostasis in type 1 diabetic (T1D) mice and rats using REDMAP to trigger insulin expression. REDMAP is a compact and sensitive tool for the precise spatiotemporal control of biological activities in animals with applications in basic biology and potentially therapy.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: Construction and optimization of the REDMAP system.
Fig. 2: Characterization of the REDMAP system performance.
Fig. 3: REDMAP-mediated dynamic control of Ras/Erk MAPK signaling.
Fig. 4: REDMAP-controlled CRISPR–dCas9 system (REDMAPcas) for epigenome engineering.
Fig. 5: REDMAP-mediated transgene expression in mice.
Fig. 6: REDMAP-mediated insulin production to control glucose homeostasis in T1D mice and rats.

Similar content being viewed by others

Data availability

All data associated with this study are present in the paper or the Supplementary Information. All genetic components related to this paper are available with a material transfer agreement and can be requested from H.Y. (hfye@bio.ecnu.edu.cn). Source data are provided with this paper.

References

  1. Yousefi, O. S. et al. Optogenetic control shows that kinetic proofreading regulates the activity of the T cell receptor. eLife 8, 42475 (2019).

    Article  Google Scholar 

  2. Ma, G. et al. Optogenetic engineering to probe the molecular choreography of STIM1-mediated cell signaling. Nat. Commun. 11, 1039 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Tkatch, T. et al. Optogenetic control of mitochondrial metabolism and Ca2+ signaling by mitochondria-targeted opsins. Proc. Natl Acad. Sci. USA 114, E5167–E5176 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Yu, D. et al. Optogenetic activation of intracellular antibodies for direct modulation of endogenous proteins. Nat. Methods 16, 1095–1100 (2019).

    Article  CAS  PubMed  Google Scholar 

  5. Ye, H. et al. Self-adjusting synthetic gene circuit for correcting insulin resistance. Nat. Biomed. Eng. 1, 0005 (2017).

    Article  CAS  PubMed  Google Scholar 

  6. Huang, Z. et al. Engineering light-controllable CAR T cells for cancer immunotherapy. Sci. Adv. 6, 9209 (2020).

    Article  Google Scholar 

  7. Wu, C.-Y., Roybal, K. T., Puchner, E. M., Onuffer, J. & Lim, W. A. Remote control of therapeutic T cells through a small molecule-gated chimeric receptor. Science 350, aab4077 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Anguela, X. M. & High, K. A. An edible switch for gene therapy. Nat. Biotechnol. 34, 824–825 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Roybal, K. T. & Lim, W. A. Synthetic immunology: hacking immune cells to expand their therapeutic capabilities. Annu. Rev. Immunol. 35, 229–253 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Chen, X., Chen, Y., Xin, H., Wan, T. & Ping, Y. Near-infrared optogenetic engineering of photothermal nanoCRISPR for programmable genome editing. Proc. Natl Acad. Sci. USA 117, 2395–2405 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Mansouri, M., Strittmatter, T. & Fussenegger, M. Light-controlled mammalian cells and their therapeutic applications in synthetic biology. Adv. Sci. 6, 1800952 (2018).

    Article  Google Scholar 

  12. Kolar, K., Knobloch, C., Stork, H., Žnidarič, M. & Weber, W. OptoBase: a web platform for molecular optogenetics. ACS Synth. Biol. 7, 1825–1828 (2018).

    Article  CAS  PubMed  Google Scholar 

  13. Ye, H., Daoud-El Baba, M., Peng, R. W. & Fussenegger, M. A synthetic optogenetic transcription device enhances blood-glucose homeostasis in mice. Science 332, 1565–1568 (2011).

    Article  CAS  PubMed  Google Scholar 

  14. Strickland, D. et al. TULIPs: tunable, light-controlled interacting protein tags for cell biology. Nat. Methods 9, 379–384 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kennedy, M. J. et al. Rapid blue-light-mediated induction of protein interactions in living cells. Nat. Methods. 7, 973–975 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Yazawa, M., Sadaghiani, A. M., Hsueh, B. & Dolmetsch, R. E. Induction of protein–protein interactions in live cells using light. Nat. Biotechnol. 27, 941–945 (2009).

    Article  CAS  PubMed  Google Scholar 

  17. Chen, X., Wang, X., Du, Z., Ma, Z. & Yang, Y. Spatiotemporal control of gene expression in mammalian cells and in mice using the lighton system. Curr. Protoc. Chem. Biol. 5, 111–129 (2013).

    Article  CAS  PubMed  Google Scholar 

  18. Motta-Mena, L. B. et al. An optogenetic gene expression system with rapid activation and deactivation kinetics. Nat. Chem. Biol. 10, 196–202 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Müller, K. et al. A red/far-red light-responsive bi-stable toggle switch to control gene expression in mammalian cells. Nucleic Acids Res. 41, e77 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Kaberniuk, A. A., Shemetov, A. A. & Verkhusha, V. V. A bacterial phytochrome-based optogenetic system controllable with near-infrared light. Nat. Methods 13, 591–597 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Shao, J. et al. Smartphone-controlled optogenetically engineered cells enable semiautomatic glucose homeostasis in diabetic mice. Sci. Transl. Med. 9, eaal2298 (2017).

    Article  PubMed  Google Scholar 

  22. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G. & Deisseroth, K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8, 1263–1268 (2005).

    Article  CAS  PubMed  Google Scholar 

  23. Kawano, F., Suzuki, H., Furuya, A. & Sato, M. Engineered pairs of distinct photoswitches for optogenetic control of cellular proteins. Nat. Commun. 6, 6256 (2015).

    Article  CAS  PubMed  Google Scholar 

  24. Sharrock, R. A. & Quail, P. H. Novel phytochrome sequences in Arabidopsis thaliana: structure, evolution, and differential expression of a plant regulatory photoreceptor family. Genes Dev. 3, 1745–1757 (1989).

    Article  CAS  PubMed  Google Scholar 

  25. Jorissen, H. J., Quest, B., Lindner, I., Tandeau de Marsac, N. & Gärtner, W. Phytochromes with noncovalently bound chromophores: the ability of apophytochromes to direct tetrapyrrole photoisomerization. Photochem. Photobiol. 75, 554–559 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Kami, C. et al. Complementation of phytochrome chromophore-deficient Arabidopsis by expression of phycocyanobilin:ferredoxin oxidoreductase. Proc. Natl Acad. Sci. USA 101, 1099–1104 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Possart, A. & Hiltbrunner, A. An evolutionarily conserved signaling mechanism mediates far-red light responses in land plants. Plant Cell 25, 102–114 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Rausenberger, J. et al. Photoconversion and nuclear trafficking cycles determine phytochrome A’s response profile to far-red light. Cell 146, 813–825 (2011).

    Article  CAS  PubMed  Google Scholar 

  29. Hiltbrunner, A. et al. FHY1 and FHL act together to mediate nuclear accumulation of the phytochrome A photoreceptor. Plant Cell Physiol. 47, 1023–1034 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Shimizu-Sato, S., Huq, E., Tepperman, J. M. & Quail, P. H. A light-switchable gene promoter system. Nat. Biotechnol. 20, 1041–1044 (2002).

    Article  CAS  PubMed  Google Scholar 

  31. Uda, Y. et al. Efficient synthesis of phycocyanobilin in mammalian cells for optogenetic control of cell signaling. Proc. Natl Acad. Sci. USA 114, 11962–11967 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Kyriakakis, P. et al. Biosynthesis of orthogonal molecules using ferredoxin and ferredoxin-NADP+ reductase systems enables genetically encoded PhyB optogenetics. ACS Synth. Biol. 7, 706–717 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Foight, G. W. et al. Multi-input chemical control of protein dimerization for programming graded cellular responses. Nat. Biotechnol. 37, 1209–1216 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Gureasko, J. et al. Membrane-dependent signal integration by the Ras activator Son of Sevenless. Nat. Struct. Mol. Biol 15, 452–461 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Shao, J. et al. Synthetic far-red light-mediated CRISPR–dCas9 device for inducing functional neuronal differentiation. Proc. Natl Acad. Sci. USA 115, E6722–E6730 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Polstein, L. R. & Gersbach, C. A. A light-inducible CRISPR–Cas9 system for control of endogenous gene activation. Nat. Chem. Biol. 11, 198–200 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nihongaki, Y. et al. CRISPR–Cas9-based photoactivatable transcription systems to induce neuronal differentiation. Nat. Methods 14, 963–966 (2017).

    Article  CAS  PubMed  Google Scholar 

  38. Zalatan, J. G. et al. Engineering complex synthetic transcriptional programs with CRISPR RNA scaffolds. Cell 160, 339–350 (2015).

    Article  CAS  PubMed  Google Scholar 

  39. Qi, L. S. et al. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152, 1173–1183 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Perrin, A., Rousseau, J. & Tremblay, J. P. Increased expression of laminin subunit α 1 chain by dCas9–VP160. Mol. Ther. Nucleic Acids 6, 68–79 (2017).

    Article  CAS  PubMed  Google Scholar 

  41. Wang, D., Tai, P. W. L. & Gao, G. Adeno-associated virus vector as a platform for gene therapy delivery. Nat. Rev. Drug Discov. 18, 358–378 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Yin, J. et al. A green tea-triggered genetic control system for treating diabetes in mice and monkeys. Sci. Transl Med. 11, eaav8826 (2019).

    Article  CAS  PubMed  Google Scholar 

  43. Valon, L. et al. Predictive spatiotemporal manipulation of signaling perturbations using optogenetics. Biophys. J. 109, 1785–1797 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Repina, N. A., Rosenbloom, A., Mukherjee, A., Schaffer, D. V. & Kane, R. S. At light speed: advances in optogenetic systems for regulating cell signaling and behavior. Annu. Rev. Chem. Biomol. Eng. 8, 13–39 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Zhao, E. M. et al. Optogenetic regulation of engineered cellular metabolism for microbial chemical production. Nature 555, 683–687 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Xu, Y. et al. Optogenetic control of chemokine receptor signal and T-cell migration. Proc. Natl Acad. Sci. USA 111, 6371–6376 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Saito, H., Kubota, M., Roberts, R. W., Chi, Q. & Matsunami, H. RTP family members induce functional expression of mammalian odorant receptors. Cell 119, 679–691 (2004).

    Article  CAS  PubMed  Google Scholar 

  48. Marosi, K. et al. Metabolic and molecular framework for the enhancement of endurance by intermittent food deprivation. FASEB J. 32, 3844–3858 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zolotukhin, S. et al. Recombinant adeno-associated virus purification using novel methods improves infectious titer and yield. Gene Ther. 6, 973–985 (1999).

    Article  CAS  PubMed  Google Scholar 

  50. Xue, S. et al. A synthetic-biology-inspired therapeutic strategy for targeting and treating hepatogenous diabetes. Mol. Ther. 25, 443–455 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was financially supported by grants from the National Key R&D Program of China, Synthetic Biology Research (no. 2019YFA0904500), the National Natural Science Foundation of China (NSFC; no. 31971346, no.31861143016), the Science and Technology Commission of Shanghai Municipality (no. 18JC1411000), the Thousand Youth Talents Plan of China and the Fundamental Research Funds for the Central Universities to H.Y. This work was also partially supported by NSFC (no. 31901023 to N.G.). W.W. was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy, EXC-2189, project ID 390939984. We also thank the ECNU Multifunctional Platform for Innovation (011) for supporting the mouse experiments and the Instruments Sharing Platform of the School of Life Sciences, ECNU.

Author information

Authors and Affiliations

Authors

Contributions

H.Y. conceived the project. H.Y., Y.Z. and D.K. designed the experiment. Y.Z., D.K., X. Wang, X. Wu and G.Y. performed the experimental work. H.Y., Y.Z., N.G. and W.W. analyzed the results and wrote the manuscript. All authors edited and approved the manuscript.

Corresponding author

Correspondence to Haifeng Ye.

Ethics declarations

Competing interests

H.Y., Y.Z. and D.K. are inventors of patent applications (Chinese patent application number 202010443080.3) submitted by ECNU that cover the new optogenetic REDMAP system. All other authors declare that they have no competing interests.

Additional information

Peer review information Nature Biotechnology thanks Zhen Gu and Guoping Feng 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 Characterization of the performance of the REDMAP system.

a, SEAP production of HEK293 cells transfected with REDMAP-encoding plasmids. HEK293 cells transfected with REDMAP-encoding plasmids were illuminated with red light (660 nm, 1 mW/cm2) for 1 min. The control cells were transfected with different expression vectors to yield different combinatorial configurations for the presence (+) or absence (-) of ΔPhyA-Gal4, FHY1-VP64 and SEAP reporter. ***P < 0.0001. b, Constitutive gene expression of human cells exposed to far-red, red and blue light, respectively. HEK293 cells were transfected with 100 ng pSEAP2-Control, followed by illumination (1 mW/cm2, 660 nm, 730 nm, or 465 nm) for 1 h. ***P < 0.0001. c, Viability of human cells exposed to far-red, red and blue light, respectively. HEK293 cells bearing the REDMAP system were illuminated with different wavelengths of light (1 mW/cm2, 660 nm, 730 nm, or 465 nm) for 1 h. ***P < 0.0001. d, REDMAP–mediated transgene expression in different mammalian cell lines. Different mammalian cell lines were co-transfected with REDMAP-encoding plasmids. Twenty-four hours after transfection, cells were illuminated with red light (1 mW/cm2) for 1 min, and SEAP expression in the culture supernatant was scored 24 h after illumination. Data are expressed as means ± SD; n = 3 independent experiments. P values were calculated by two-tailed unpaired t test. N.S., not significant. Detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 2 Assessment of the compatibility of phycocyanobilin (PCB) with HEK-293 cells and Mice.

a, Viability of HEK293 cells after exposure to various concentrations of PCB. HEK293 cells were seeded into each well of a 96-well plate and cultivated in medium containing the indicated PCB concentrations (0–100 µM) for 24 h before cell viability was evaluated with CCK-8 assays. b, Viability of HEK293 cells after transfection with various amounts of PCB-producer plasmid pDQ55 (0–200 ng). HEK293 cells were seeded into each well of a 96-well plate and transfected with the indicated amounts of PCB-producer plasmid pDQ55. Twenty-four hours later, cell viability was evaluated with CCK-8 assays. c–m, Compatibility of PCB with mice. Mice were intraperitoneally injected with PCB (5 mg/kg/day) or PBS (once daily) for three weeks. c, Body weight was monitored every three days for a total of three weeks. d, Hematoxylin and eosin (H&E) staining for the liver tissues in mice treated with 5 mg/kg/day via intraperitoneal injection. H&E staining, × 200 magnification. Scale bars, 100 μm. e–m, On day 21, mouse blood was collected for the blood biochemical analysis, liver function and hepatic function analysis. (e) Counts for total white blood cells (WBC), (f) red blood cells (RBC), (g) hemoglobin (HGB), (h) platelet (PLT). i, j, Kidney function analysis. (i) blood urea nitrogen (BUN), (j) creatinine (CRE). k–m Hepatic function analysis. (k) aspartate aminotransferase (AST), (l) alanine aminotransferase (ALT), (m) albumin/globulin ratio (A/G). Data in a, b, are expressed as means ± SD; n = 3 independent experiments. Data in c, e–m are expressed as means ± SEM (n = 6 mice). P values were calculated by two-tailed unpaired t test. N.S., not significant. Detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 3 Quantification of REDMAP activity with different amounts of the PCB-producer plasmid (pDQ55).

HEK293 cells were co-transfected with pYZ181, pDQ16, pDL6, x ng pDQ55 (PhCMV-PcyA-P2A-HO-P2A-Fd-P2A-FNR-pA) and (200-x) ng pcDNA3.1(+). Twenty-four hours after transfection, the cells were illuminated with red light (1 mW/cm2) for 1 min and SEAP expression in the culture supernatant was quantified 24 h after illumination. ***P < 0.0001 for all comparisons. Data are expressed as means ± SD; n = 3 independent experiments. P values were calculated by two-tailed unpaired t test. N.S., not significant. Detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 4 Characterization of the performance of the REDMAP system.

a–g, Light-induced SEAP expression from REDMAP-engineered cells preincubated with PCB. a, Experimental procedure for evaluating the binding properties of PCB. HEK293 cells were transfected with REDMAP-encoding plasmids, and PCB was added into the culture medium at 24 h after transfection. After incubation for 30 min, the culture medium was replaced with fresh medium without PCB and the transfected cells were divided into three groups, which were cultivated for different times (6, 12, or 24 h) before illuminating with red light. b–d, Light-induced SEAP production from REDMAP-engineered cells preincubated with PCB. In bd, ***P < 0.0001. e, Experimental procedure for evaluating the binding properties of PCB. HEK293 cells transfected with REDMAP-encoding plasmids were incubated with PCB for 30 min and then illuminated with red light. The illuminated cells were divided into two groups. One group of cells were cultivated with fresh medium supplemented with PCB, while the other group without PCB. f, g, Light-induced SEAP production from REDMAP-engineered cells incubated with or without PCB after 1 min of illumination. In f, g, ***P < 0.0001. h, Fluorescence micrograph profiling of EGFP expression kinetics mediated by the REDMAP system. Scale bar, 100 μm. Representative data from three independent experiments. i, Comparison of REDMAP with other red/far-red light-inducible transgene expression systems. j, Illumination time-dependent REDMAP-mediated OFF kinetics of transgene expression. HEK293 cells were illuminated with red light. One hour after red light illumination, cells were exposed to far-red light (1 mW/cm2) for different time periods (0 to 10 min). ***P < 0.0001 for all groups versus 660 nm. k, Illumination intensity-dependent REDMAP-mediated OFF kinetics of transgene expression. HEK293 cells were illuminated with red light. One hour after red light illumination, cells were exposed to far-red light of various light intensities (0 to 1 mW/cm2) for 1 min. ***P < 0.0001 for all groups versus 660 nm. Data are expressed as means ± SD; n = 3 independent experiments. P values were calculated by two-tailed unpaired t test. N.S., not significant. Detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 5 Control experiments for main Fig. 3b.

HEK293 cells transfected with REDMAPSOS-Ras (ΔPhyA-CAAX and FHY1-SOScat) were illuminated with red light (1 mW/cm2) for 1 min. Ten minutes after illumination, the cells were harvested and probed for phospho-Erk1/2, total Erk1/2 (tErk) and GAPDH. Negative control cells were only transfected with pcDNA3.1(+). Representative data from two independent experiments. Detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 6 A comparison of endogenous gene activation with the REDMAPcas system and the far-red light-activated CRISPR-dCas9 (FACE) system.

a, Experimental procedure for light-controlled transcriptional activation of endogenous gene in HEK293 cells. b, Light-controlled transcriptional activation of endogenous gene ASCL1 in HEK293 cells mediated by REDMAPcas and FACE. HEK293 cells were transfected with REDMAPcas [pSZ69 (PhCMV-dCas9-pA), pYZ181, pDQ16, pDQ100 (P5×UAS-MS2-p65-HSF1-pA, P5×UAS, 5×UAS-PhCMVmin)] or FACE [pSZ69, pGY102 (PFRL1b-FGTA4-pA; PFRL1b, (whiG)2-PhCMVmin3G; FGTA4, MS2-P65-HSF1), pWS189, pGY32] together with two gRNAs targeting the ASCL1 gene [pSZ83 (PU6-sgRNA1ASCL1-pA) and pSZ84 (PU6-sgRNA2ASCL1-pA)], and were illuminated with red light (1 mW/cm2) or far-red light (1 mW/cm2) for 1 min. ***P < 0.0001. Data are expressed as means ± SD; n = 3 independent experiments. P values were calculated by two-tailed unpaired t test. N.S., not significant. Detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 7 Optimization and characterization of the REDMAP system in mice.

a, Schematic showing the time schedule and experimental procedure for REDMAP-mediated transgene expression in mice. b, REDMAP-mediated transgene expression in mice. Mice were hydrodynamically injected with pDQ59 (PHYA-GAL4::FHY1-VP64), pYZ360 (5×UAS-PhCMVmin) and intraperitoneally injected with PCB (3.65 mg/kg) before illumination. **P = 0.0014. c, Co-delivery of the PCB-producer plasmid in mice. Mice hydrodynamically injected with pDQ59, pYZ360, and a PCB-producer plasmid pDQ55 were illuminated with red light. ***P = 0.0006. d, An iterative REDMAP-mediated transgene expression in mice. Mice were hydrodynamically injected with pDQ59, pYZ450 (5×UAS-PTATA), and pDQ55, and illuminated with red light. **P = 0.0018. Data in bd are expressed as means ± SEM (n = 4 mice). P values were calculated by two-tailed unpaired t test. N.S., not significant. e, PCB concentration-dependent REDMAP-mediated luciferase expression kinetics in mice. Mice were hydrodynamically injected with pDQ59, pYZ450. Eight hours after injection, mice were intraperitoneally injected with various amounts of PCB (0–5 mg/kg) before illumination. ***P = 0.0001, ***P = 0.0007, ***P < 0.0001, ***P < 0.0001, left to right. Data are expressed as means ± SEM (n = 5 mice). P values were calculated by two-tailed unpaired t test. f, The activation and deactivation performance of the REDMAP system in mice. Mice were hydrodynamically injected with pDQ59, pYZ450, and pDQ55, and mice were then divided into three groups. The control group was kept in the dark. Eight hours after injection, the mice in group A and B were illuminated with red light to activate the REDMAP system. Mice in group B were immediately illuminated with far-red light to deactivate REDMAP system. ***P < 0.0001 for both comparisons. Data are expressed as means ± SEM (n = 6 mice). P values were calculated by two-tailed unpaired t test. Detailed images of the bioluminescence signal are provided in Supplementary Fig. 3 and detailed description of genetic components and transfection mixtures are provided in Supplementary Table 1 and 5.

Source data

Extended Data Fig. 8 Serum inflammatory cytokine and whole blood analysis of AAV REDMAP mice.

Mice were injected via tail vein with AAV encoding the REDMAP system. Control mice were injected with PBS. Twelve weeks after injection, mouse blood was collected for the quantification of (a) serum IL-6, (b) serum IFN-γ, and (c) serum TNF-α production by ELISA. d Counts for total white blood cells (WBC), blood lymphocytes, and blood monocytes, in mice. Data are expressed as means ± SEM (n = 6 mice). P values were calculated by two-tailed unpaired t test. N.S., not significant.

Source data

Extended Data Fig. 9 REDMAP-mediated SEAP production in animals.

a, HEK-293 cells transfected with REDMAP system were microencapsulated into coherent alginate-poly- (L-lysine)-alginate beads (400 μm; 200 cells per capsule) and subcutaneously implanted into mice, rats, and rabbits. The animals were illuminated from the top with red light (20 mW/cm2) and blood SEAP was quantified 48 h after the first illumination. bd, Light-induced SEAP expression in mice (b), rats (c), and rabbits (d). b, Mice implanted with 2.5×106 microencapsulated transgenetic HEK-293 cells and illuminated daily for 1 min. ***P < 0.0001. c, Rats implanted with 1.2×107 microencapsulated transgenetic HEK-293 cells and illuminated twice daily for 5 min. ***P < 0.0001. d, Rabbits implanted with 3×107 microencapsulated HEK-293 transgenic cells and illuminated twice daily for 5 min. Control animals were implanted with microencapsulated wild-type HEK293 cells. ***P < 0.0001. Data are presented as means ± SEM (n = 5 animals). P values were calculated by two-tailed unpaired t test. N.S., not significant.

Source data

Extended Data Fig. 10 Characterization of the REDMAP-mediated SEAP and insulin expression stable cell lines.

The HEKREDMAP-SEAP-P2A-mINS cell line, transgenic for red light-inducible SEAP expression, was constructed by co-transfecting HEK293 cells with pYZ474 (ITR-P5×UAS*-SEAP-P2A-insulin-pA::PmPGK-ZeoR-pA-ITR) and pYZ484 (ITR-PhCMV-ΔPhyA-Gal4-P2A-FHY1-VP64-pA::PhCMV-PcyA-P2A-HO-P2A-Fd-P2A-FNR-pA::PmPGK-Puro-pA-ITR), and selected with puromycin (1 μg/mL) and zeocin (100 μg/mL) for two weeks. Twenty randomly selected cell clones were profiled for their light-inducible SEAP production performance by cultivating them for 24 h after illumination with red light (1 mW/cm2) for 1 min.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–4 and Tables 1–6.

Reporting Summary

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, Y., Kong, D., Wang, X. et al. A small and highly sensitive red/far-red optogenetic switch for applications in mammals. Nat Biotechnol 40, 262–272 (2022). https://doi.org/10.1038/s41587-021-01036-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41587-021-01036-w

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research