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.

Genetically targeted magnetic control of the nervous system

Matters Arising to this article was published on 30 September 2019

Matters Arising to this article was published on 30 September 2019

Matters Arising to this article was published on 30 September 2019

This article has been updated

Abstract

Optogenetic and chemogenetic actuators are critical for deconstructing the neural correlates of behavior. However, these tools have several limitations, including invasive modes of stimulation or slow on/off kinetics. We have overcome these disadvantages by synthesizing a single-component, magnetically sensitive actuator, “Magneto,” comprising the cation channel TRPV4 fused to the paramagnetic protein ferritin. We validated noninvasive magnetic control over neuronal activity by demonstrating remote stimulation of cells using in vitro calcium imaging assays, electrophysiological recordings in brain slices, in vivo electrophysiological recordings in the brains of freely moving mice, and behavioral outputs in zebrafish and mice. As proof of concept, we used Magneto to delineate a causal role of striatal dopamine receptor 1 neurons in mediating reward behavior in mice. Together our results present Magneto as an actuator capable of remotely controlling circuits associated with complex animal behaviors.

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: Remote control of calcium signaling using Magneto2.0.
Figure 2: Electrophysiological characterization of Magneto2.0 in mouse brain slices.
Figure 3: Magnetic control over zebrafish tactile behavior in vivo.
Figure 4: Magnetogenetic control of the mammalian nervous system in vivo.
Figure 5: Activation of striatal D1R+ neurons is sufficient to control reward behavior.

Similar content being viewed by others

Change history

  • 28 March 2016

    In the version of this article initially published online, first, a relevant citation was omitted. To correct this, the sixth sentence of the introduction, which originally read "one study employed nonthermal magnetogenetic control of somatic tissues to regulate blood glucose11, but a genetically encoded, single-component magnetogenetic system has yet to be applied to the nervous system," has been rewritten to say "one study employed nonthermal magnetogenetic control of somatic tissues to regulate blood glucose11 and another utilized a naturally occurring iron-containing magnetoreceptor to trigger neuronal activity48, but a genetically encoded, single-component magnetogenetic system has yet to be applied to the nervous system of behaving vertebrates." Ref. 48 is provided as follows: Long, X., Ye, J., Zhao, D. & Zhang, S.J. Magnetogenetics: remote non-invasive magnetic activation of neuronal activity with a magnetoreceptor. Sci. Bull. (Beijing) 60, 2107–2119 (2015). Second, the author contribution statement incorrectly listed M.P.B as having performed experiments. This has been corrected to say that M.P.B. provided conceptual help during the development of the prototype channel. Finally, the reporter construct in the right panel of Figure 2e was mislabeled Camk2a::Cre-EGFP. This control construct should have been labeled Camk2a::EGFP. The errors have been corrected for the print, PDF and HTML versions of this article.

References

  1. Zemelman, B.V., Lee, G.A., Ng, M. & Miesenböck, G. Selective photostimulation of genetically chARGed neurons. Neuron 33, 15–22 (2002).

    Article  CAS  PubMed  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  PubMed  Google Scholar 

  3. Gradinaru, V., Mogri, M., Thompson, K.R., Henderson, J.M. & Deisseroth, K. Optical deconstruction of parkinsonian neural circuitry. Science 324, 354–359 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Sternson, S.M. & Roth, B.L. Chemogenetic tools to interrogate brain functions. Annu. Rev. Neurosci. 37, 387–407 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Alexander, G.M. et al. Remote control of neuronal activity in transgenic mice expressing evolved G protein-coupled receptors. Neuron 63, 27–39 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Güler, A.D. et al. Transient activation of specific neurons in mice by selective expression of the capsaicin receptor. Nat. Commun. 3, 746 (2012).

    Article  PubMed  CAS  Google Scholar 

  7. Bernstein, J.G., Garrity, P.A. & Boyden, E.S. Optogenetics and thermogenetics: technologies for controlling the activity of targeted cells within intact neural circuits. Curr. Opin. Neurobiol. 22, 61–71 (2012).

    Article  CAS  PubMed  Google Scholar 

  8. Hughes, S., McBain, S., Dobson, J. & El Haj, A.J. Selective activation of mechanosensitive ion channels using magnetic particles. J. R. Soc. Interface 5, 855–863 (2008).

    Article  CAS  PubMed  Google Scholar 

  9. Huang, H., Delikanli, S., Zeng, H., Ferkey, D.M. & Pralle, A. Remote control of ion channels and neurons through magnetic-field heating of nanoparticles. Nat. Nanotechnol. 5, 602–606 (2010).

    Article  CAS  PubMed  Google Scholar 

  10. Stanley, S.A. et al. Radio-wave heating of iron oxide nanoparticles can regulate plasma glucose in mice. Science 336, 604–608 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Stanley, S.A., Sauer, J., Kane, R.S., Dordick, J.S. & Friedman, J.M. Remote regulation of glucose homeostasis in mice using genetically encoded nanoparticles. Nat. Med. 21, 92–98 (2015).

    Article  CAS  PubMed  Google Scholar 

  12. Chen, R., Romero, G., Christiansen, M.G., Mohr, A. & Anikeeva, P. Wireless magnetothermal deep brain stimulation. Science 347, 1477–1480 (2015).

    Article  CAS  PubMed  Google Scholar 

  13. Loukin, S., Zhou, X., Su, Z., Saimi, Y. & Kunsg, C. Wild-type and brachyolmia-causing mutant TRPV4 channels respond directly to stretch force. J. Biol. Chem. 285, 27176–27181 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Liedtke, W. et al. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell 103, 525–535 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Güler, A.D. et al. Heat-evoked activation of the ion channel, TRPV4. J. Neurosci. 22, 6408–6414 (2002).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Stanley, S. Biological nanoparticles and their influence on organisms. Curr. Opin. Biotechnol. 28, 69–74 (2014).

    Article  CAS  PubMed  Google Scholar 

  17. Iordanova, B., Robison, C.S. & Ahrens, E.T. Design and characterization of a chimeric ferritin with enhanced iron loading and transverse NMR relaxation rate. J. Biol. Inorg. Chem. 15, 957–965 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Lei, L. et al. A TRPV4 channel C-terminal folding recognition domain critical for trafficking and function. J. Biol. Chem. 288, 10427–10439 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Hofherr, A., Fakler, B. & Klöcker, N. Selective Golgi export of Kir2.1 controls the stoichiometry of functional Kir2.x channel heteromers. J. Cell Sci. 118, 1935–1943 (2005).

    Article  CAS  PubMed  Google Scholar 

  20. Gradinaru, V. et al. Molecular and cellular approaches for diversifying and extending optogenetics. Cell 141, 154–165 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Lytton, J., Westlin, M. & Hanley, M.R. Thapsigargin inhibits the sarcoplasmic or endoplasmic reticulum Ca-ATPase family of calcium pumps. J. Biol. Chem. 266, 17067–17071 (1991).

    CAS  PubMed  Google Scholar 

  22. Phan, M.N. et al. Functional characterization of TRPV4 as an osmotically sensitive ion channel in porcine articular chondrocytes. Arthritis Rheum. 60, 3028–3037 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Sohal, V.S., Zhang, F., Yizhar, O. & Deisseroth, K. Parvalbumin neurons and gamma rhythms enhance cortical circuit performance. Nature 459, 698–702 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Andermann, P., Ungos, J. & Raible, D.W. Neurogenin1 defines zebrafish cranial sensory ganglia precursors. Dev. Biol. 251, 45–58 (2002).

    Article  CAS  PubMed  Google Scholar 

  25. Douglass, A.D., Kraves, S., Deisseroth, K., Schier, A.F. & Engert, F. Escape behavior elicited by single, channelrhodopsin-2-evoked spikes in zebrafish somatosensory neurons. Curr. Biol. 18, 1133–1137 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Tian, L. et al. Imaging neural activity in worms, flies and mice with improved GCaMP calcium indicators. Nat. Methods 6, 875–881 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Wyart, C. et al. Optogenetic dissection of a behavioural module in the vertebrate spinal cord. Nature 461, 407–410 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Sagasti, A., Guido, M.R., Raible, D.W. & Schier, A.F. Repulsive interactions shape the morphologies and functional arrangement of zebrafish peripheral sensory arbors. Curr. Biol. 15, 804–814 (2005).

    Article  CAS  PubMed  Google Scholar 

  29. Hersch, S.M. et al. Electron microscopic analysis of D1 and D2 dopamine receptor proteins in the dorsal striatum and their synaptic relationships with motor corticostriatal afferents. J. Neurosci. 15, 5222–5237 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Berke, J.D., Okatan, M., Skurski, J. & Eichenbaum, H.B. Oscillatory entrainment of striatal neurons in freely moving rats. Neuron 43, 883–896 (2004).

    Article  CAS  PubMed  Google Scholar 

  31. Wise, R.A. Dopamine, learning and motivation. Nat. Rev. Neurosci. 5, 483–494 (2004).

    Article  CAS  PubMed  Google Scholar 

  32. Tsai, H.-C. et al. Phasic firing in dopaminergic neurons is sufficient for behavioral conditioning. Science 324, 1080–1084 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Lobo, M.K. et al. Cell type-specific loss of BDNF signaling mimics optogenetic control of cocaine reward. Science 330, 385–390 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Zengin-Toktas, Y. et al. Motivational properties of D2 and D3 dopamine receptors agonists and cocaine, but not with D1 dopamine receptors agonist and L-dopa, in bilateral 6-OHDA-lesioned rat. Neuropharmacology 70, 74–82 (2013).

    Article  CAS  PubMed  Google Scholar 

  35. Gore, B.B. & Zweifel, L.S. Genetic reconstruction of dopamine D1 receptor signaling in the nucleus accumbens facilitates natural and drug reward responses. J. Neurosci. 33, 8640–8649 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Stuber, G.D., Britt, J.P. & Bonci, A. Optogenetic modulation of neural circuits that underlie reward seeking. Biol. Psychiatry 71, 1061–1067 (2012).

    Article  PubMed  Google Scholar 

  37. Jeong, J.W. et al. Wireless optofluidic systems for programmable in vivo pharmacology and optogenetics. Cell 162, 662–674 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. O'Neil, R.G. & Heller, S. The mechanosensitive nature of TRPV channels. Pflugers Arch. 451, 193–203 (2005).

    Article  CAS  PubMed  Google Scholar 

  39. Liedtke, W. & Kim, C. Functionality of the TRPV subfamily of TRP ion channels: add mechano-TRP and osmo-TRP to the lexicon! Cell. Mol. Life Sci. 62, 2985–3001 (2005).

    Article  CAS  PubMed  Google Scholar 

  40. Matthews, B.D. et al. Ultra-rapid activation of TRPV4 ion channels by mechanical forces applied to cell surface beta1 integrins. Integr. Biol. (Camb.) 2, 435–442 (2010).

    Article  CAS  Google Scholar 

  41. Kimmel, C.B., Ballard, W.W., Kimmel, S.R., Ullmann, B. & Schilling, T.F. Stages of embryonic development of the zebrafish. Dev. Dyn. 203, 253–310 (1995).

    Article  CAS  PubMed  Google Scholar 

  42. McFarland, T.J. et al. Evaluation of a novel short polyadenylation signal as an alternative to the SV40 polyadenylation signal. Plasmid 56, 62–67 (2006).

    Article  CAS  PubMed  Google Scholar 

  43. Wheeler, M.A. et al. TNF-α/TNFR1 signaling is required for the development and function of primary nociceptors. Neuron 82, 587–602 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Smith, C.J., Morris, A.D., Welsh, T.G. & Kucenas, S. Contact-mediated inhibition between oligodendrocyte progenitor cells and motor exit point glia establishes the spinal cord transition zone. PLoS Biol. 12, e1001961 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Hargus, N.J., Nigam, A., Bertram, E.H. III & Patel, M.K. Evidence for a role of Nav1.6 in facilitating increases in neuronal hyperexcitability during epileptogenesis. J. Neurophysiol. 110, 1144–1157 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Quintana, A. et al. Lack of GPR88 enhances medium spiny neuron activity and alters motor- and cue-dependent behaviors. Nat. Neurosci. 15, 1547–1555 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Chen, S., Chiu, C.N., McArthur, K.L., Fetcho, J.R. & Prober, D. TRP channel mediated neuronal activation and ablation in freely behaving zebrafish. Nat. Methods 13, 147–150 (2016).

    Article  PubMed  CAS  Google Scholar 

  48. Long, X., Ye, J., Zhao, D. & Zhang, S.J. Magnetogenetics: remote non-invasive magnetic activation of neuronal activity with a magnetoreceptor. Sci. Bull. (Beijing) 60, 2107–2119 (2015).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank members of the Condron, Beenhakker, Kucenas, Patel, Deppmann and Güler labs for comments and suggestions. In particular, we are thankful for technical assistance provided by A. Morris, P. Neff, A. Rainwater and S. Young. We are thankful for comments on the manuscript from M. Caterina, I. Cheng, B. Condron, K. Gamage and I. Provencio. We acknowledge the UVa Keck Center for Cellular Imaging staff for use of the Leica confocal microscopy system, supported by US National Institutes of Health (NIH) National Center for Research Resources RR025616. This work was supported by NIH National Institute of General Medical Sciences (NIGMS) 5T32GM008328, NIH NIGMS 5T32GM008136, a Neuroscience Center of Excellence fellowship and a Wagner Fellowship (M.A.W.); NIH National Institute of Neurological Disorders and Stroke (NINDS) F32NS087791 (C.J.S.); NIH NINDS R01NS072212 (S.K.); NIH NINDS R01NS075157 (M.K.P.); NIH NINDS R01NS072388 (C.D.D.); and UVa startup funds (A.D.G.).

Author information

Authors and Affiliations

Authors

Contributions

M.A.W. and A.D.G. designed the study. M.A.W., C.J.S., M.O., B.S.B., A.M.P., R.M.G., R.P.G. and A.J.S. performed the experiments. M.A.W., C.J.S., M.O., B.S.B., A.M.P., R.M.G., M.K.P., C.D.D. and A.D.G. analyzed the data. M.P.B. provided conceptual help during the development of the prototype channel. S.K., M.K.P., C.D.D. and A.D.G. supervised the research. M.A.W. and A.D.G. wrote the manuscript with input from coauthors.

Corresponding author

Correspondence to Ali D Güler.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Model of magnetic activation via Magneto.

(a) The cation channel, TRPV4, is gated by stretch (among other diverse classes of stimuli), to depolarize cells. For simplicity, only two of the four homomeric subunits are shown. (b) Coupling ferritin to the TRPV4 C-terminus converts TRPV4 to a magnetic field detector. Gating properties were extrapolated from published descriptions of TRPV1 and TRPA1 gating mechanisms48-50.

48. Cao et al. (2013) Nature 504, 113-118.

49. Liao et al. (2013) Nature 504, 107-112.

50. Paulsen et al. (2015) Nature 520, 511-517.

Supplementary Figure 2 Measurement of electromagnet strength over distance.

Empirical determination of the strength of several electromagnets over distance powered by an identical current. Dashed line represents distance between HEK cells and electromagnet during calcium imaging assays. A 3 cm diameter magnet was used for all calcium imaging assays. Δx represents distance between magnet and cells used in calcium imaging.

Source data

Supplementary Figure 3 In vitro calcium imaging using Magneto1.0.

(a) Mammalian expression vector schematic of Magneto1.0. (b-g) Representative images of HEK293 cells used for in vitro magnetic stimulation Fluo-4 calcium imaging. (h) Quantification of relative calcium fluorescence in response to magnetic stimulation of mCherry+ cells. Replicates are shown as individual coverslips equaling n=6 (TRPV4/ferritin), n=8 (Magneto1.0), and n=6 (Magneto1.0+RR). Total cells analyzed for each condition are n=545 (TRPV4/ferritin), n=565 (Magneto1.0), and n=437 (Magneto1.0+RR). One-way ANOVA, Bonferroni post-test, (F2,17=7.509, p=0.0046). (i) Representative images of temporal association between calcium fluorescence and magnetic field pulses in an individual Magneto1.0-expressing cell (arrow). Field was pulsed for alternating 10 second periods of on/off. *p<0.05. Data are shown as mean±SEM.

Source data

Supplementary Figure 4 Optimization of Magneto1.0 by improving cellular trafficking.

(a-e) HEK293 cells transfected with mCherry-fused variants of Magneto1.0 with combinations of various inwardly rectifying K+ channel 2.1 (Kir2.1) trafficking signals. (a) Magneto1.0-mCherry shows diffuse cellular localization, poor membrane expression, and poor transfection efficiency. (b) Addition of ER export signal from Kir2.1 to C-terminus of Magneto1.0-mCherry peptide partially improves Magneto expression. (c) Addition of Kir2.1 membrane trafficking signal (TS) significantly improves membrane expression of Magneto. (d) Dual addition of membrane trafficking and ER export signals improves expression relative to Magneto1.0 but not relative to a single membrane trafficking signal. (e) Tandem Kir2.1 membrane trafficking/ER export signals on Magneto1.0 C-terminus improves expression but not relative to c. n=2 coverslips and >100 cells analyzed per trafficking modification examined.

Supplementary Figure 5 Viability of Magneto2.0-transfected mammalian cells.

(a-d) Viability of Magneto2.0 transfected HEK293 cells several days post transfection (DPT). Images show bright field and mCherry fluorescence. Zoom increased in (c-d) to increase single cell resolution following significant cell division. Images are representative of n>100 cells examined.

Supplementary Figure 6 Calcium imaging controls using thapsigargin.

(a) Graph of Fluo-4 fluorescence using HEK293 cells transfected with Magneto2.0-p2A-mCherry and treated with thapsigargin over a period of 60 minutes. Arrow indicates addition of 1 μM thapsigargin to the imaging chamber after a 30 second baseline recording of calcium fluorescence. Dashed box indicates analysis window for “thapsigargin” experiments in panel b. n=114 cells analyzed from 3 independent replicates. (b) Time course showing the magnetic activation of Magneto2.0 expressing cells in the presence and absence of thapsigargin. All cells from one replicate shown per condition, n=102 cells (Magnet) and n=52 cells (Thapsigargin). In the “thapsigargin” condition, cells were pre-treated with 1μM thapsigargin and calcium imaging was initiated 30 minutes post-thapsigargin treatment during the window (dashed box) shown in panel a. (c) Quantification of maximal calcium fluorescence of HEK293 cells expressing Magneto2.0 and subjected to the above conditions using Fluo-4 calcium imaging 24 hours post-transfection. Values shown are the average maximal Fluo-4 fluorescence values per cell relative to baseline for each condition. Data points are shown as total cell averages among individual coverslips. n=114 (Thapsigargin) and n=396 (Magnet) cells analyzed from n=3 (Thapsigargin) and n=5 (Magnet) independent replicates. Welch’s two-tailed unpaired t-test, (t2.882=4.457, p=0.0395). “Magnet” data are duplicated from Figure 1. *p<0.05. Data shown as mean±SEM.

Source data

Supplementary Figure 7 Control analyses for electrophysiological characterization of Magneto2.0.

(a) Representative trace showing that injection of depolarizing current evokes spikes in doubly transduced EGFP+ Magneto2.0 expressing neurons. (b) No change in AP latency between conditions of current injection or magnetic field application in transduced neurons (measured from time immediately preceding depolarization). Unpaired two-tailed t-test, (t22=1.628, p=0.1178) (threshold), (t22=1.676, p=0.1079) (peak). (c-g) Membrane properties are unchanged under conditions of either current injection or magnetic stimulation in hippocampal neurons doubly transduced with CMV::DIO-Magneto2.0 and CaMKIIα::Cre-EGFP. Unpaired two-tailed t-test, (t22=0.1926, p=0.8498) in c, (t22=1.335, p=0.1954) in d, (t22=0.1290, p=0.8985) in e, (t22=1.052, p=0.3042) in f, (t22=0.4086, p=0.6868) in g. (h) Injection of depolarizing current evokes APs in Cre-negative DIO-Magneto2.0 transduced EGFP+ neurons. n=12 neurons analyzed for each condition shown in (b-g). ns: not significant. Data shown as mean±SEM.

Source data

Supplementary Figure 8 Controls for magnetic stimulation in brain slice electrophysiology.

(a) Paired traces depicting the onset of action potentials following current injection (black) and magnetic stimulation (red) for the same neuron co-transduced with AAVs carrying CaMKIIα::Cre-EGFP and CMV::DIO-Magneto2.0. Overlay shows a modest delay of action potential onset (50-100 ms) when neurons are stimulated with static magnetic fields. (b) Magnified traces of the resting state from three additional neurons co-transduced with the above viruses. Neurons are shown immediately prior to action potential initiation as static magnetic fields are brought more closely to the cells using a micromanipulator, a process requiring roughly 1 second. Traces do not show interference coming from 50 mT static magnetic fields in close proximity to the recording apparatus.

Source data

Supplementary Figure 9 Application of Magneto1.0 to zebrafish behavior in vivo.

(a) Schematic of trans cardiac myosin light chain 2 (cmcl2)::GFP element and its expression in 24 hpf zebrafish embryos for positive transgenic selection. n>100 fish examined. (b) Schematic of Magneto1.0 construct used: Tol2: Tol2 transposon sites; ß-Actin: promoter; IRES: internal ribosomal entry site; nls-EGFP: nuclear localized enhanced GFP. (c) Quantification of the number of coils in WT (uninjected) and ß-actin::Magneto1.0 expressing 24 hpf zebrafish embryos in response to magnetic stimulation. n=43 WT, n=25 ß-actin::Magneto1.0 fish. Statistics determined by Chi-squared analysis, (Chi23=36.51, p<0.0001). (d) Quantification of coiling rate in WT (uninjected) and ß-actin::Magneto1.0 expressing zebrafish. Replicates (number of individual fish) shown in columns. Statistics determined by one-way ANOVA, Bonferroni post-test, (F3,64=3.89, p=0.0129). ***p<0.001, *p<0.05. Data are shown as mean±SEM.

Source data

Supplementary Figure 10 Analysis of Magneto2.0 in live zebrafish.

(a) Maximal GCaMP3 calcium fluorescence change of mCherry+ (n=20 from 5 fish) and mCherry- (n=33 from 5 fish) neurons in response to magnetic field stimulation. Dashed line indicates average GCaMP3 fluorescence value for mCherry- neurons. (17/20 mCherry+ neurons exceed this fluorescence value). Unpaired two-tailed t-test, (t51=3.373, p=0.0014). (b) Schematic of behavioral paradigm for induction of zebrafish coiling behaviors using magnetic stimulation. (c) Schematic of Rohon-Beard neuron projections. (d) Magneto2.0 expression construct. Tol2: transposon site; ngn1: neurogenin-1 promoter; IRES: internal ribosomal entry site; nls: nuclear localization signal; EGFP: enhanced green fluorescent protein; polyA: polyadenylation signal. (e-f) In vivo imaging of Rohon-Beard neuron projections into the skin, n=10 fish examined per genotype. Inset: Magneto2.0+ (EGFP+/RFP+) and Magneto2.0- (EGFP–/RFP+) neurons. Data pooled from 2 injections per genotype. **p<0.01. Data shown as mean±SEM.

Source data

Supplementary Figure 11 Mouse behavioral controls.

(a) Quantification of the change in firing rate relative to baseline for low-frequency and high-frequency firing single units in the striatum in response to the D1R agonist SKF81297, n=7 (<5 Hz), n=8 (>5 Hz) units examined from one Drd1a::Cre mouse transduced with CMV::DIO-Magneto2.0, unpaired two-tailed t-test, (t13=2.192, p=0.0472). (b) Picture of magnetic open field behavioral chamber. (c) Quantification of change in linear velocity in open field for both groups (n=6 per genotype), unpaired two-tailed t-test, (t10=0.08856, p=0.9312). *p<0.05, ns: not significant. Data shown as mean±SEM.

Source data

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–11 and Supplementary Table 1 (PDF 2785 kb)

Supplementary Methods Checklist (PDF 421 kb)

In vivo manipulation of zebrafish behavior using Magneto2.0 (baseline)

Representative movies of ngn1::Magneto2.0-expressing zebrafish coiling behavior without magnetic field. Individual embryos are each roughly 0.5 mm in diameter. Movies shown at 8× speed. Length of original behavioral analysis is 2-3 min per video. (MP4 10104 kb)

In vivo manipulation of zebrafish behavior using Magneto2.0 (magnetic stimulation)

Representative movies of ngn1::Magneto2.0-expressing zebrafish coiling behavior with magnetic field. Individual embryos are each roughly 0.5 mm in diameter. Movies shown at 8× speed. Length of original behavioral analysis is 2-3 min per video. (MP4 11210 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wheeler, M., Smith, C., Ottolini, M. et al. Genetically targeted magnetic control of the nervous system. Nat Neurosci 19, 756–761 (2016). https://doi.org/10.1038/nn.4265

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4265

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