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.

ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation

Abstract

Channelrhodopsins (ChRs) are used to optogenetically depolarize neurons. We engineered a variant of ChR, denoted red-activatable ChR (ReaChR), that is optimally excited with orange to red light (λ 590–630 nm) and offers improved membrane trafficking, higher photocurrents and faster kinetics compared to existing red-shifted ChRs. Red light is less scattered by tissue and is absorbed less by blood than the blue to green wavelengths that are required by other ChR variants. We used ReaChR expressed in the vibrissa motor cortex to drive spiking and vibrissa motion in awake mice when excited with red light through intact skull. Precise vibrissa movements were evoked by expressing ReaChR in the facial motor nucleus in the brainstem and illumination with red light through the external auditory canal. Thus, ReaChR enables transcranial optical activation of neurons in deep brain structures without the need to surgically thin the skull, form a transcranial window or implant optical fibers.

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

Figure 1: Basic properties of ReaChR compared to those of C-VChR1 (VChR1 with the ChIEF N terminus) and C1V1E122T.
Figure 2: Kinetics of C-VChR1, C1V1E122T and ReaChR.
Figure 3: Characterization and comparison of different ChR variants in primary cultured hippocampal neurons.
Figure 4: The responses of neurons expressing the different ChR variants to a 10 Hz pulsed-light stimulation train.
Figure 5: In vivo expression and utilization of ReaChR for cortical control.
Figure 6: In vivo expression and utilization of ReaChR for brainstem activation.

Accession codes

Primary accessions

NCBI Reference Sequence

References

  1. Arrenberg, A.B., Stainier, D.Y., Baier, H. & Huisken, J. Optogenetic control of cardiac function. Science 330, 971–974 (2010).

    Article  CAS  Google Scholar 

  2. 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  Google Scholar 

  3. Bruegmann, T. et al. Optogenetic control of heart muscle in vitro and in vivo. Nat. Methods 7, 897–900 (2010).

    Article  CAS  Google Scholar 

  4. Nagel, G. et al. Channelrhodopsin-1: a light-gated proton channel in green algae. Science 296, 2395–2398 (2002).

    Article  CAS  Google Scholar 

  5. Nagel, G. et al. Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc. Natl. Acad. Sci. USA 100, 13940–13945 (2003).

    Article  CAS  Google Scholar 

  6. Adamantidis, A.R., Zhang, F., Aravanis, A.M., Deisseroth, K. & de Lecea, L. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature 450, 420–424 (2007).

    Article  CAS  Google Scholar 

  7. Lin, J.Y., Lin, M.Z., Steinbach, P. & Tsien, R.Y. Characterization of engineered channelrhodopsin variants with improved properties and kinetics. Biophys. J. 96, 1803–1814 (2009).

    Article  CAS  Google Scholar 

  8. Wen, L. et al. Opto-current-clamp actuation of cortical neurons using a strategically designed channelrhodopsin. PLoS ONE 5, e12893 (2010).

    Article  Google Scholar 

  9. Govorunova, E.G., Spudich, E.N., Lane, C.E., Sineshchekov, O.A. & Spudich, J.L. New channelrhodopsin with a red-shifted spectrum and rapid kinetics from Mesostigma viride. MBio. 2, e00115–e001111 (2011).

    Article  Google Scholar 

  10. Kleinlogel, S. et al. Ultra light-sensitive and fast neuronal activation with the Ca2+-permeable channelrhodopsin CatCh. Nat. Neurosci. 14, 513–518 (2011).

    Article  CAS  Google Scholar 

  11. Yizhar, O. et al. Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature 477, 171–178 (2011).

    Article  CAS  Google Scholar 

  12. Tromberg, B.J. et al. Non-invasive in vivo characterization of breast tumors using photon migration spectroscopy. Neoplasia 2, 26–40 (2000).

    Article  CAS  Google Scholar 

  13. Aravanis, A.M. et al. An optical neural interface: in vivo control of rodent motor cortex with integrated fiberoptic and optogenetic technology. J. Neural Eng. 4, S143–S156 (2007).

    Article  Google Scholar 

  14. Zhang, F. et al. Red-shifted optogenetic excitation: a tool for fast neural control derived from Volvox carteri. Nat. Neurosci. 11, 631–633 (2008).

    Article  Google Scholar 

  15. Lin, J.Y. A user's guide to channelrhodopsin variants: features, limitations and future developments. Exp. Physiol. 96, 19–25 (2011).

    Article  Google Scholar 

  16. Wang, H. et al. Molecular determinants differentiating photocurrent properties of two channelrhodopsins from Chlamydomonas. J. Biol. Chem. 284, 5685–5696 (2009).

    Article  CAS  Google Scholar 

  17. Gunaydin, L.A. et al. Ultrafast optogenetic control. Nat. Neurosci. 13, 387–392 (2010).

    Article  CAS  Google Scholar 

  18. Nagel, G. et al. Light activation of channelrhodopsin-2 in excitable cells of Caenorhabditis elegans triggers rapid behavioral responses. Curr. Biol. 15, 2279–2284 (2005).

    Article  CAS  Google Scholar 

  19. Griesbeck, O., Baird, G.S., Campbell, R.E., Zacharias, D.A. & Tsien, R.Y. Reducing the environmental sensitivity of yellow fluorescent protein. Mechanism and applications. J. Biol. Chem. 276, 29188–29194 (2001).

    Article  CAS  Google Scholar 

  20. Mattis, J. et al. Principles for applying optogenetic tools derived from direct comparative analysis of microbial opsins. Nat. Methods 9, 159–172 (2012).

    Article  CAS  Google Scholar 

  21. Hill, D.N., Curtis, J.C., Moore, J.D. & Kleinfeld, D. Primary motor cortex reports efferent control of vibrissa motion on multiple timescales. Neuron 72, 344–356 (2011).

    Article  CAS  Google Scholar 

  22. Berg, R.W. & Kleinfeld, D. Vibrissa movement elicited by rhythmic electrical microstimulation to motor cortex in the aroused rat mimics exploratory whisking. J. Neurophysiol. 90, 2950–2963 (2003).

    Article  Google Scholar 

  23. Brecht, M., Schneider, M., Sakmann, B. & Margrie, T.W. Whisker movements evoked by stimulation of single pyramidal cells in rat motor cortex. Nature 427, 704–710 (2004).

    Article  CAS  Google Scholar 

  24. Haiss, F. & Schwarz, C. Spatial segregation of different modes of movement control in the whisker representation of rat primary motor cortex. J. Neurosci. 25, 1579–1587 (2005).

    Article  CAS  Google Scholar 

  25. Herfst, L.J. & Brecht, M. Whisker movements evoked by stimulation of single motor neurons in the facial nucleus of the rat. J. Neurophysiol. 99, 2821–2832 (2008).

    Article  Google Scholar 

  26. Martin, M.R. & Lodge, D. Morphology of the facial nucleus of the rat. Brain Res. 123, 1–12 (1977).

    Article  CAS  Google Scholar 

  27. Dörfl, J. The musculature of the mystacial vibrissae of the white mouse. J. Anat. 135, 147–154 (1982).

    PubMed  PubMed Central  Google Scholar 

  28. Berndt, A., Yizhar, O., Gunaydin, L.A., Hegemann, P. & Deisseroth, K. Bi-stable neural state switches. Nat. Neurosci. 12, 229–234 (2009).

    Article  CAS  Google Scholar 

  29. Berndt, A. et al. High-efficiency channelrhodopsins for fast neuronal stimulation at low light levels. Proc. Natl. Acad. Sci. USA 108, 7595–7600 (2011).

    Article  CAS  Google Scholar 

  30. Shichida, Y., Matuoka, S., Hidaka, Y. & Yoshizawa, T. Absorption spectra of intermediates of bacteriorhodopsin measured by laser photolysis at room temperatures. Biochim. Biophys. Acta 723, 240–246 (1983).

    Article  CAS  Google Scholar 

  31. Kato, H.E. et al. Crystal structure of the channelrhodopsin light-gated cation channel. Nature 482, 369–374 (2012).

    Article  CAS  Google Scholar 

  32. Xu, H.T., Pan, F., Yang, G. & Gan, W.B. Choice of cranial window type for in vivo imaging affects dendritic spine turnover in the cortex. Nat. Neurosci. 10, 549–551 (2007).

    Article  CAS  Google Scholar 

  33. Hauss-Wegrzyniak, B., Lynch, M.A., Vraniak, P.D. & Wenk, G.L. Chronic brain inflammation results in cell loss in the entorhinal cortex and impaired LTP in perforant path-granule cell synapses. Exp. Neurol. 176, 336–341 (2002).

    Article  CAS  Google Scholar 

  34. Grutzendler, J., Kasthuri, N. & Gan, W.B. Long-term dendritic spine stability in the adult cortex. Nature 420, 812–816 (2002).

    Article  CAS  Google Scholar 

  35. Sohler, T.P., Lothrop, G.N. & Forbes, H.S. The poial circulation of normal, non-anesthetized animals. Part 1. Description of a method of observation. J. Pharmacol. Exp. Ther. 71, 325–330 (1941).

    CAS  Google Scholar 

  36. Drew, P.J. et al. Chronic optical access through a polished and reinforced thinned skull. Nat. Methods 7, 981–984 (2010).

    Article  CAS  Google Scholar 

  37. Gradinaru, V. et al. Targeting and readout strategies for fast optical neural control in vitro and in vivo. J. Neurosci. 27, 14231–14238 (2007).

    Article  CAS  Google Scholar 

  38. Osakada, F. et al. New rabies virus variants for monitoring and manipulating activity and gene expression in defined neural circuits. Neuron 71, 617–631 (2011).

    Article  CAS  Google Scholar 

  39. Armstrong, C.M. & Gilly, W.F. Access resistance and space clamp problems associated with whole-cell patch clamping. Methods Enzymol. 207, 100–122 (1992).

    Article  CAS  Google Scholar 

  40. Knutsen, P.M., Derdikman, D. & Ahissar, E. Tracking whisker and head movements in unrestrained behaving rodents. J. Neurophysiol. 93, 2294–2301 (2005).

    Article  Google Scholar 

Download references

Acknowledgements

J.Y.L. was funded by the Foundation of Research, Science and Technology New Zealand, and P.M.K. was supported by a long-term fellowship from the Human Frontier Science Program (HFSP). The project was supported by grants to R.Y.T. from the US National Institutes of Health (NS027177) and the Howard Hughes Medical Institute and grants to D.K. from the US National Institutes of Health (DA029706, OD006831 and NS058668). We thank P. Tsai for suggesting the name ReaChR. We thank K. Deisseroth (Stanford University) for the hChR2H134R-eYFP construct, K. Svoboda (Howard Hughes Medical Institute Janelia Farm Research Campus) for the CAG promoter vector, E. Boyden (Massachusetts Institute of Technology) for the lentiviral vector, L. Tian (Howard Hughes Medical Institute Janelia Farm Research Campus) for the AAV2 vector and D. Trono (École Polytechnique Fédérale De Lausane) for the psPAX2 and pMD2.G lentivirus packaging vectors. AAV2-ReaChR-Citrine and pLenti-ReaChR-Citrine constructs can be requested from http://tsienlab.ucsd.edu/Samples.htm.

Author information

Authors and Affiliations

Authors

Contributions

J.Y.L. designed and developed ReaChR and conducted and analyzed the experiments in HEK293 and neuron cultures. P.M.K. and A.M. conducted and analyzed the in vivo experiments. D.K. and R.Y.T. contributed to the design and analysis of the experiments. All authors contributed to the writing and discussions of the manuscripts.

Corresponding author

Correspondence to John Y Lin.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1-10 (PDF 3215 kb)

Supplementary Data Set

Supplementary Data Set 1 (XLS 172 kb)

Supplementary Video 1

Example of whisker protraction evoked by inter-aural stimulation of a mouse expressing ReaChR in FN motoneurons. Movements were captured with high-speed video (800 fps), during isoflurane anesthesia, and tracked offline. FN was stimulated by placing a 617 nm LED at the opening of the external auditory canal (100 mW light output; red vertical bars; see Methods). Upon photo-stimulation the whiskers promptly protracted and remained protracted for the duration of the light pulse (100 msec). The whiskers then retracted to their initial reference position. (MOV 1115 kb)

Supplementary Video 2

Example of whisker retraction evoked by inter-aural stimulation of a mouse expressing ReaChR in FN motoneurons. Movements were captured with high-speed video (800 fps), and whisker position and angle tracked offline. As the mouse recovered from isoflurane anesthesia, the whiskers slowly protracted and remained in a protracted position for approximately one minute. During this time window, ReaChR expressing neurons in FN were stimulated by inter-aural illumination with a 617 nm LED (100 mW light output; red vertical bars). Upon photo-activation, the whiskers promptly retracted and remained retracted for the duration of the light pulse (100 msec). As the light turned off, the whiskers then returned to their initial protracted position. (MOV 1556 kb)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Lin, J., Knutsen, P., Muller, A. et al. ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation. Nat Neurosci 16, 1499–1508 (2013). https://doi.org/10.1038/nn.3502

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3502

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