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.

  • Letter
  • Published:

Voltage imaging and optogenetics reveal behaviour-dependent changes in hippocampal dynamics

Abstract

A technology that simultaneously records membrane potential from multiple neurons in behaving animals will have a transformative effect on neuroscience research1,2. Genetically encoded voltage indicators are a promising tool for these purposes; however, these have so far been limited to single-cell recordings with a marginal signal-to-noise ratio in vivo3,4,5. Here we developed improved near-infrared voltage indicators, high-speed microscopes and targeted gene expression schemes that enabled simultaneous in vivo recordings of supra- and subthreshold voltage dynamics in multiple neurons in the hippocampus of behaving mice. The reporters revealed subcellular details of back-propagating action potentials and correlations in subthreshold voltage between multiple cells. In combination with stimulation using optogenetics, the reporters revealed changes in neuronal excitability that were dependent on the behavioural state, reflecting the interplay of excitatory and inhibitory synaptic inputs. These tools open the possibility for detailed explorations of network dynamics in the context of behaviour.

a, Schematic of the paQuasAr3 construct. b, Photoactivation by blue light enhanced voltage signals excited by red light in cultured neurons that expressed paQuasAr3 (representative example of n = 4 cells). c, Model of the photocycle of paQuasAr3. d, Confocal images of sparsely expressed paQuasAr3 in brain slices. Scale bars, 50 μm. Representative images, experiments were repeated in n = 3 mice. e, Simultaneous fluorescence and patch-clamp recordings from a neuron expressing paQuasAr3 in acute brain slice. Top, magnification of boxed regions. Schematic shows brain slice, patch pipette and microscope objective. f, Simultaneous fluorescence and patch-clamp recordings of inhibitory post synaptic potentials in an L2–3 neuron induced by electrical stimulation of L5–6 in acute slice. g, Normalized change in fluorescence (ΔF/F) and SNR of optically recorded post-synaptic potentials (PSPs) as a function of the amplitude of the post-synaptic potentials. The voltage sensitivity was ΔF/F = 40 ± 1.7% per 100 mV. The SNR was 0.93 ± 0.07 per 1 mV in a 1-kHz bandwidth (n = 42 post-synaptic potentials from 5 cells, data are mean ± s.d.). Schematic shows brain slice, patch pipette, field stimulation electrodes and microscope objective. h, Optical measurements of paQuasAr3 fluorescence in the CA1 region of the hippocampus (top) and glomerular layer of the olfactory bulb (bottom) of anaesthetized mice (representative traces from n = 7 CA1 cells and n = 13 olfactory bulb cells, n = 3 mice). Schematics show microscope objective and the imaged brain region. i, STA fluorescence from 88 spikes in a CA1 oriens neuron. j, Frames from the STA video showing the delay in the back-propagating action potential in the dendrites relative to the soma. k, Sub-Nyquist fitting of the action potential delay and width shows electrical compartmentalization in the dendrites. Experiments in km were repeated in n = 2 cells from n = 2 mice.

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

Fig. 2: Optical recording of neuronal activity in hippocampus of walking mice.
Fig. 3: Behaviour-dependent intercellular correlations in the hippocampus.
Fig. 4: Simultaneous optogenetic stimulation and voltage imaging in hippocampal SST cells in walking mice.

Similar content being viewed by others

Data availability

The datasets generated during the current study are available from the corresponding author upon reasonable request.

Code availability

Custom data analysis codes are available from the corresponding author upon reasonable request.

References

  1. Petersen, C. C. H. Whole-cell recording of neuronal membrane potential during behavior. Neuron 95, 1266–1281 (2017).

    Article  CAS  Google Scholar 

  2. Lee, A. K. & Brecht, M. Elucidating neuronal mechanisms using intracellular recordings during behavior. Trends Neurosci. 41, 385–403 (2018).

    Article  CAS  Google Scholar 

  3. Lou, S. et al. Genetically targeted all-optical electrophysiology with a transgenic Cre-dependent Optopatch mouse. J. Neurosci. 36, 11059–11073 (2016).

    Article  CAS  Google Scholar 

  4. Gong, Y. et al. High-speed recording of neural spikes in awake mice and flies with a fluorescent voltage sensor. Science 350, 1361–1366 (2015).

    Article  ADS  CAS  Google Scholar 

  5. Yang, H. H. et al. Subcellular imaging of voltage and calcium signals reveals neural processing in vivo. Cell 166, 245–257 (2016).

    Article  CAS  Google Scholar 

  6. Hochbaum, D. R. et al. All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins. Nat. Methods 11, 825–833 (2014).

    Article  CAS  Google Scholar 

  7. Madisen, L. et al. Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance. Neuron 85, 942–958 (2015).

    Article  CAS  Google Scholar 

  8. Millerd, J. E. et al. Improved sensitivity in blue-membrane bacteriorhodopsin films. Opt. Lett. 24, 1355–1357 (1999).

    Article  ADS  CAS  Google Scholar 

  9. Wagner, N. L., Greco, J. A., Ranaghan, M. J. & Birge, R. R. Directed evolution of bacteriorhodopsin for applications in bioelectronics. J. R. Soc. Interface 10, 20130197 (2013).

    Article  Google Scholar 

  10. Häusser, M., Spruston, N. & Stuart, G. J. Diversity and dynamics of dendritic signaling. Science 290, 739–744 (2000).

    Article  ADS  Google Scholar 

  11. Wu, C., Ivanova, E., Zhang, Y. & Pan, Z. H. rAAV-mediated subcellular targeting of optogenetic tools in retinal ganglion cells in vivo. PLoS ONE 8, e66332 (2013).

    Article  ADS  CAS  Google Scholar 

  12. Baker, C. A., Elyada, Y. M., Parra, A. & Bolton, M. M. Cellular resolution circuit mapping with temporal-focused excitation of soma-targeted channelrhodopsin. eLife 5, e14193 (2016).

    Article  Google Scholar 

  13. Buchanan, E. K. et al. Penalized matrix decomposition for denoising, compression, and improved demixing of functional imaging data. Preprint at https://www.biorxiv.org/content/10.1101/334706v4 (2018).

  14. Klausberger, T. et al. Brain-state- and cell-type-specific firing of hippocampal interneurons in vivo. Nature 421, 844–848 (2003).

    Article  ADS  CAS  Google Scholar 

  15. Harvey, C. D., Collman, F., Dombeck, D. A. & Tank, D. W. Intracellular dynamics of hippocampal place cells during virtual navigation. Nature 461, 941–946 (2009).

    Article  ADS  CAS  Google Scholar 

  16. Bittner, K. C. et al. Conjunctive input processing drives feature selectivity in hippocampal CA1 neurons. Nat. Neurosci. 18, 1133–1142 (2015).

    Article  CAS  Google Scholar 

  17. Epsztein, J., Brecht, M. & Lee, A. K. Intracellular determinants of hippocampal CA1 place and silent cell activity in a novel environment. Neuron 70, 109–120 (2011).

    Article  CAS  Google Scholar 

  18. Varga, C., Golshani, P. & Soltesz, I. Frequency-invariant temporal ordering of interneuronal discharges during hippocampal oscillations in awake mice. Proc. Natl Acad. Sci. USA 109, E2726–E2734 (2012).

    Article  ADS  CAS  Google Scholar 

  19. Buzsáki, G. Theta oscillations in the hippocampus. Neuron 33, 325–340 (2002).

    Article  Google Scholar 

  20. Poulet, J. F. & Petersen, C. C. Internal brain state regulates membrane potential synchrony in barrel cortex of behaving mice. Nature 454, 881–885 (2008).

    Article  ADS  CAS  Google Scholar 

  21. Gentet, L. J., Avermann, M., Matyas, F., Staiger, J. F. & Petersen, C. C. Membrane potential dynamics of GABAergic neurons in the barrel cortex of behaving mice. Neuron 65, 422–435 (2010).

    Article  CAS  Google Scholar 

  22. Lampl, I., Reichova, I. & Ferster, D. Synchronous membrane potential fluctuations in neurons of the cat visual cortex. Neuron 22, 361–374 (1999).

    Article  CAS  Google Scholar 

  23. Okun, M. & Lampl, I. Instantaneous correlation of excitation and inhibition during ongoing and sensory-evoked activities. Nat. Neurosci. 11, 535–537 (2008).

    Article  CAS  Google Scholar 

  24. Briggman, K. L., Abarbanel, H. D. & Kristan, W. B. Jr. From crawling to cognition: analyzing the dynamical interactions among populations of neurons. Curr. Opin. Neurobiol. 16, 135–144 (2006).

    Article  CAS  Google Scholar 

  25. Cunningham, J. P. & Yu, B. M. Dimensionality reduction for large-scale neural recordings. Nat. Neurosci. 17, 1500–1509 (2014).

    Article  CAS  Google Scholar 

  26. Harris, K. D., Hirase, H., Leinekugel, X., Henze, D. A. & Buzsáki, G. Temporal interaction between single spikes and complex spike bursts in hippocampal pyramidal cells. Neuron 32, 141–149 (2001).

    Article  CAS  Google Scholar 

  27. Bezaire, M. J. & Soltesz, I. Quantitative assessment of CA1 local circuits: knowledge base for interneuron–pyramidal cell connectivity. Hippocampus 23, 751–785 (2013).

    Article  Google Scholar 

  28. Unal, G., Joshi, A., Viney, T. J., Kis, V. & Somogyi, P. Synaptic targets of medial septal projections in the hippocampus and extrahippocampal cortices of the mouse. J. Neurosci. 35, 15812–15826 (2015).

    Article  CAS  Google Scholar 

  29. Shaner, N. C. et al. Improving the photostability of bright monomeric orange and red fluorescent proteins. Nat. Methods 5, 545–551 (2008).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  31. Klapoetke, N. C. et al. Independent optical excitation of distinct neural populations. Nat. Methods 11, 338–346 (2014).

    Article  CAS  Google Scholar 

  32. Shepard, B. D., Natarajan, N., Protzko, R. J., Acres, O. W. & Pluznick, J. L. A cleavable N-terminal signal peptide promotes widespread olfactory receptor surface expression in HEK293T cells. PLoS ONE 8, e68758 (2013).

    Article  ADS  CAS  Google Scholar 

  33. Chow, B. Y. et al. High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature 463, 98–102 (2010).

    Article  ADS  CAS  Google Scholar 

  34. McIsaac, R. S. et al. Directed evolution of a far-red fluorescent rhodopsin. Proc. Natl Acad. Sci. USA 111, 13034–13039 (2014).

    Article  ADS  CAS  Google Scholar 

  35. Maclaurin, D., Venkatachalam, V., Lee, H. & Cohen, A. E. Mechanism of voltage-sensitive fluorescence in a microbial rhodopsin. Proc. Natl Acad. Sci. USA 110, 5939–5944 (2013).

    Article  ADS  CAS  Google Scholar 

  36. Daigle, T. L. et al. A suite of transgenic driver and reporter mouse lines with enhanced brain-cell-type targeting and functionality. Cell 174, 465–480 (2018).

    Article  CAS  Google Scholar 

  37. Molyneaux, B. J., Arlotta, P., Hirata, T., Hibi, M. & Macklis, J. D. Fezl is required for the birth and specification of corticospinal motor neurons. Neuron 47, 817–831 (2005).

    Article  CAS  Google Scholar 

  38. Krashes, M. J. et al. Rapid, reversible activation of AgRP neurons drives feeding behavior in mice. J. Clin. Invest. 121, 1424–1428 (2011).

    Article  CAS  Google Scholar 

  39. Adam, Y. & Mizrahi, A. Long-term imaging reveals dynamic changes in the neuronal composition of the glomerular layer. J. Neurosci. 31, 7967–7973 (2011).

    Article  CAS  Google Scholar 

  40. Dombeck, D. A., Harvey, C. D., Tian, L., Looger, L. L. & Tank, D. W. Functional imaging of hippocampal place cells at cellular resolution during virtual navigation. Nat. Neurosci. 13, 1433–1440 (2010).

    Article  CAS  Google Scholar 

  41. Goldey, G. J. et al. Removable cranial windows for long-term imaging in awake mice. Nat. Protoc. 9, 2515–2538 (2014).

    Article  CAS  Google Scholar 

  42. Kralj, J. M., Douglass, A. D., Hochbaum, D. R., Maclaurin, D. & Cohen, A. E. Optical recording of action potentials in mammalian neurons using a microbial rhodopsin. Nat. Methods 9, 90–95 (2012).

    Article  CAS  Google Scholar 

  43. Pnevmatikakis, E. A. & Giovannucci, A. NoRMCorre: an online algorithm for piecewise rigid motion correction of calcium imaging data. J. Neurosci. Methods 291, 83–94 (2017).

    Article  CAS  Google Scholar 

  44. Mukamel, E. A., Nimmerjahn, A. & Schnitzer, M. J. Automated analysis of cellular signals from large-scale calcium imaging data. Neuron 63, 747–760 (2009).

    Article  CAS  Google Scholar 

  45. Ricciardi, L. M. & Sacerdote, L. The Ornstein–Uhlenbeck process as a model for neuronal activity. I. Mean and variance of the firing time. Biol. Cybern. 35, 1–9 (1979).

    Article  CAS  Google Scholar 

  46. Stujenske, J. M., Spellman, T. & Gordon, J. A. Modeling the spatiotemporal dynamics of light and heat propagation for in vivo optogenetics. Cell Rep. 12, 525–534 (2015).

    Article  CAS  Google Scholar 

  47. Podgorski, K. & Ranganathan, G. Brain heating induced by near-infrared lasers during multiphoton microscopy. J. Neurophysiol. 116, 1012–1023 (2016).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank M. S. Lee and V. Joshi for help with tissue culture, A. H. Gheorghe for help with spectroscopy experiments, N. Rollins and S. Brownsberger for technical help, L. Yapp for advice on hippocampal surgeries, A. Ruangkittisakul and K. Ballanyi for neuronal cultures, K. Podgorski for heating simulation code, G. Buzsaki and members of the Cohen laboratory for helpful discussions and D. Hochbaum, C. Straub, J. L. Saulnier, B. L. Sabatini, V. Kapoor and V. Murthy for help at early stages of the project. Y.A. was supported by fellowships from the Human Frontiers Science Program and the Edmund and Lili Safra Center for Brain Sciences. This work was supported by the Howard Hughes Medical Institute.

Reviewer information

Nature thanks Kenneth Harris, Michael Hausser, Samuel Andrew Hires and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

Y.A. performed the patch-clamp and imaging experiments in acute slices, cultured neurons and in vivo. J.J.K. performed patch-clamp measurements in HEK293T cells. Y.A. and Y.Z. performed protein engineering, with supervision by A.E.C. and R.E.C., respectively. Y.A. and D.B. performed spectroscopy experiments. Y.A. developed the imaging system with help from H.W., J.J.K., S.K. and V.P. S.L., L.M. and H.Z. developed and characterized the Ai155 Optopatch3 mice. M.A.M.-R. performed in utero electroporation, supervised by P.A. M.E.X. optimized and validated the penalized matrix decomposition–non-negative matrix factorization algorithm in collaboration with E.K.B., I.K., D.Z and L.P. S.C. and C.D.H. shared unpublished reagents for soma targeting of opsins. K.J.W. helped with molecular biology. B.G. performed the heating simulation. S.L.F. designed the CheRiff–HA construct. Y.A. and A.E.C. designed the project, analysed data and wrote the manuscript. A.E.C. supervised all aspects of the project.

Corresponding author

Correspondence to Adam E. Cohen.

Ethics declarations

Competing interests

A.E.C. is a founder of Q-State Biosciences. The other authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Development and characterization of QuasAr3.

a, Screening pipeline. Rationally designed constructs were cloned in the Optopatch configuration, expressed in primary neurons and tested for spike SNR using light-induced spikes. Constructs with improved SNR were then expressed in vivo using in utero electroporation and tested for spike SNR in acute slices. The process was repeated iteratively. b, Examples of SNR measurements in cultured neurons. Left, wide-field epifluorescence images of GFP (green) fused to CheRiff, an opsin with excellent membrane trafficking. Middle, fluorescence of QuasAr (red) mutants. Scale bar, 10 μm. Right, QuasAr fluorescence transients in response to optogenetically induced spikes (10 ms blue-light stimulation at 1 mW mm−2). Each construct was tested on at least five cultured neurons. c, Hierarchical screen for improved membrane trafficking of QuasAr variants (see Methods for details). Diagram, schematic of the FCK_DuEx1.0 construct and overview of the screening pipeline. E. coli colonies were transformed with libraries in FCK_DuEx1.0. The colonies with the brightest fluorescence were picked for lentivirus production and secondary screening in primary neuronal culture. Images, example images of the fluorescent protein (FP) channel of QuasAr2–FP fusions: (i) mOrange; (ii) mRuby2; (iii) mKate2; (iv) citrine. Scale bar, 10 μm. d, SNR of N-terminal modifications compared with QuasAr2. All constructs showed reduced SNR (see Methods for details). e, Replacing mOrange2 with Citrine as a fusion protein improved the trafficking only with two specific linkers. f, Adding additional TS sequences at the linker and C-terminal improved the spike SNR. g, The mutation K171R increased the expression level of QuasAr, quantified by normalizing QuasAr fluorescence by the fluorescence of the co-expressed CheRiff–GFP. dg, Data are mean ± s.e.m., one-tailed Student’s t-test. h, Top, diagram of the QuasAr2 and QuasAr3 constructs. Bottom, confocal images of brain slices expressing QuasAr2 and QuasAr3. Scale bar, 500 μm. Insets, single-cell bodies; scale bar, 10 μm. Representative images from n = 2 mice (QuasAr2) and n = 3 mice (QuasAr3). i, Confocal images of brain slices expressing Cre-dependent QuasAr3 with sparsity controlled by co-expression of hSyn-Cre. Scale bar, 50 μm. j, Simultaneous fluorescence and patch-clamp recordings from two neurons expressing QuasAr3 using AAV virus in acute brain slices. Left, image of QuasAr3 fluorescence in the soma. Scale bar, 10 μm. Middle, spiking during ramp current injection. Right, mean spike, overlay of fluorescence and voltage. Inset, boxed regions showing correspondence of optical and electrical recordings of subthreshold voltage overlaid. See Extended Data Fig. 3 for statistics.

Extended Data Fig. 2 Mapping neuronal excitability in Cre-on Optopatch3 transgenic mice (line Ai155).

a, Construct design for a Cre-dependent Optopatch3 transgenic mouse. b, Representative traces for all-optical electrophysiology recordings in acute brain slices from Optopatch3 transgenic mice crossed with different Cre driver lines. Scale bar, 10 μm. c, Confocal images showing citrine fluorescence from QuasAr3-citrine, in offspring of crosses between Optopatch3 mice and different Cre driver mice. Acute brain slices were prepared from mice aged 14 to 17 days and imaged in the cortex. Scale bar, 50 μm. d, Composite bright-field image of a coronal brain slice from an Rbp4-cre+/−;Optopatch3+/− transgenic mouse, with locations of optical recordings marked with white spots. e, Spike raster showing 94 cells recorded sequentially from a single Rbp4-cre+/−;Optopatch3+/− acute brain slice. f, Optogenetic stimulus intensity-dependent firing rates in acute slices with different Cre drivers. Left, slices that are homozygous for Optopatch3. Right, slices that are heterozygous for Optopatch3. g, Mean firing rate (F), during a 500-ms stimulus as a function of stimulus intensity (I) calculated from the data shown in f. Data are shown as mean ± s.e.m. In the mice with Optopatch3 expression driven by CamKII-Cre, the FI curve for the Optopatch3+/+ mice is compressed along the x axis relative to the Optopatch3+/− mice, which indicates a stronger optogenetic drive for a given optical stimulation strength in the mice homozygous for CheRiff. The decrease in firing rate at a strong stimulus in these mice is a signature of depolarization block. Data are from 128 cells from 5 slices from 2 SST-cre+/−;Optopatch3+/− mice; 25 cells from 1 slice from 1 CamKII-cre+/−;Optopatch3+/− mouse; 152 cells from 6 slices from 4 Rbp4-cre+/−;Optopatch3+/− mice; 89 cells from 2 slices from 2 mice for CamKII-cre+/−;Optopatch3+/− mice.

Extended Data Fig. 3 Photophysical properties, kinetics and electrical properties of paQuasAr3.

a, Photoactivation by blue light. Red lines, fluorescence of HEK293T cells that express paQuasAr3 during voltage steps under constant red illumination (10 W mm−2) and variable blue illumination (average of n = 8 cells). Grey lines, same experiment in HEK293T cells that express QuasAr3 (average of n = 7 cells). b, PaQuasAr3 fluorescence compared to blue-light intensity at two membrane voltages (n = 8 cells, mean ± s.e.m.). Photoactivation showed saturation behaviour, with 50% maximum enhancement at I488 nm = 27 mW mm−2. c, Voltage-dependent near-infrared fluorescence of paQuasAr3 and QuasAr3 with and without blue light (150 mW mm−2). All fluorescence values are normalized to fluorescence with red-only illumination at Vm = −75 mV. d, Same data as c but each fluorescence trace was normalized to its value F0 at Vm = −75 mV. Blue illumination enhanced the absolute fluorescence and the absolute voltage sensitivity, but the fractional voltage sensitivity (ΔF/F0) was the same between QuasAr3 and paQuasAr3 and was not affected by blue illumination (n = 7 cells in each condition, P = 0.91, one-way analysis of variance). e, Kinetics of QuasAr3 and paQuasAr3 measured in HEK293T cells. Cells were subjected to a square wave from −60 mV to +40 mV at 5 Hz (see a). Response transients were fitted to a double exponential. QuasAr3, n = 5 cells; paQuasAr3, n = 9 cells. Data are mean ± s.e.m. Red intensity, 10 W mm−2; blue intensity, 150 mW mm−2. f, Response of paQuasAr3 and QuasAr3 to steps of blue illumination. The blue-light enhancement arose with a 50 ± 14-ms time constant and subsided with a 167 ± 26-ms time constant (mean ± s.d.). Blue light activated paQuasAr3 was approximately twofold brighter than QuasAr3 (mean of n = 10 cells). g, Action spectrum for photosensitization, measured in E. coli expressing paQuasAr3. Fluorescence was excited at λexc = 640 nm and emission was collected from λem = 660–740 nm. The activation wavelength was scanned from λact = 450–650 nm. Peak activation was at λact = 470 nm. h, Fluorescence excitation spectra with and without blue sensitization (40 mW mm−2). The sensitized state of paQuasAr3 had a fluorescence excitation spectrum similar to QuasAr3, with peak excitation at λexc = 580 nm. i, SNR of single spikes in acute slices for QuasAr3 (n = 10 cells) and paQuasAr3 (n = 10 cells) with either red light only or red and blue illumination (mean ± s.d. paQuasAr3 with and without blue light, paired Student’s t-test, paQuasAr3 red and blue illumination versus QuasAr3 red light only, Student’s t-test). j, Voltage-clamp recordings in CA1 pyramidal cell expressing paQuasAr3-s showed no photocurrents in response to illumination with red light (640 nm, 12 W mm−2), blue light (488 nm, 90 mW mm−2) or a combination of the two lights, both when the cell was held at −70 mV and at 0 mV (repeated in n = 2 cells). k, Electrical properties measured by patch clamp in acute slices. QuasAr3 (n = 11 cells, 6 slices, 2 mice) and paQuasAr3 (n = 9 cells, 6 slices, 3 mice) were expressed in the visual cortex and compared with non-expressing cortical cells (n = 9 cells, 7 slices, 7 mice). paQuasAr3-s (n = 7 cells, 7 slices, 5 mice) was expressed in the CA1 pyramidal cell layer, and compared with non-expressing cells in that layer (n = 8 cells, 6 slices, 5 mice). Data are mean ± s.d.

Extended Data Fig. 4 Patterned illumination improves brain imaging and minimizes brain heating.

a, A single oriens interneuron was illuminated with red and blue light projected precisely onto the cell (1×) or with an oval mask of which the area was either twice (2×) or 10 times bigger (10×) than the cell. Scale bar, 20 μm. Illumination intensity was held constant across measurements. b, Spontaneous spiking activity of the oriens interneuron from a in an awake resting mouse. c, Mean spike SNR was similar with the 2× mask compared to the 1× mask (P = 0.2, two-sided paired Student’s t-test) and significantly reduced with the 10× mask (P = 0.001, two-sided paired Student’s t-test, n = 10 cells from n = 2 mice, data are mean ± s.e.m.). d, Cross-section of near-infrared fluorescence of oriens or PCL cells imaged in anaesthetized mice and illuminated with either wide-field red illumination, patterned red illumination or patterned red and patterned blue illumination. The lower SBR in the PCL is attributed to the greater density of expressing cells. All cells visible in the focal plane were targeted with illumination, leading to higher background in the PCL. In the oriens, blue illumination increased the signal but not background. In the PCL, blue illumination modestly increased background, a consequence of light scattering between neighbouring cells. SBR values were as follows. PCL: wide field, 0.27 ± 0.02; red only, 0.95 ± 0.063; red and blue, 1.31 ± 0.08; n = 64 cells. Oriens: wide field, 0.27 ± 0.05; red only, 2.9 ± 0.19; red and blue, 4.5 ± 0.27; n = 7 cells. Data are mean ± s.e.m. (shown as line and shading). These data were used to calculate the improvement in SBR in Fig. 2d. e, Left, spontaneous activity of a representative oriens neuron in an awake, resting mouse illuminated with the indicated laser intensity. Right, population average shows no change in the firing rate in the illumination range tested (corresponding to 12.5–34 mW into the tissue, n = 10 cells from 2 mice, data are mean ± s.d., P = 0.22, two-sided paired Student’s t-test, N.S, not significant). f, FOVs with single oriens interneurons were imaged while illuminating with increasing number of masks around the cell to simulate the multicellular imaging conditions used in the PCL. Left, typical illumination pattern with the indicated number of masks. Middle, representative traces from an oriens FOV imaged in an awake, resting mouse with the indicated number of masks. Right, mean spontaneous firing rate as a function of the total laser power. Firing rates were stable with illumination of up to 10 masks (80 mW). Projection of 15 masks (120 mW) caused a modest increase in the spontaneous firing rates (n = 10 cells from 2 mice, data are mean ± s.d., P = 0.05, two-sided paired Student’s t-test). Subsequent experiments were restricted to ten or fewer masks. g, Simulated spatial temperature profile in brain tissue with an imaging cannula and immersion water. Profiles are shown for no illumination, 100-mW 640-nm illumination and 100-mW 920-nm illumination, corresponding to a two-photon imaging experiment.

Extended Data Fig. 5 Photostability of paQuasAr3-s in vivo during ‘all-optical’ excitability measurements.

a, Fluorescence in the mouse hippocampus was recorded for 10-min in total. Oriens SST cells expressed paQuasAr3-s and CheRiff-s and were illuminated with red light at 12 W mm−2 and stimulated with blue light (up to 10 mW mm−2) using the stimulation protocol presented at the top, during quiet and walking periods. The trace shows the fluorescence from two user-defined regions of interest after subtraction of the background from a cell-free region. The baseline signal photobleached in both cells by around 50% during this interval. Top, total acquisition. Dashed lines denote separate videos. Bottom, magnified views of the indicated regions from the top graph. Similar recordings were performed in five FOVs. b, The same two cells were imaged three weeks later. While at the end of the initial 10-min recording the SNR was low due to photobleaching, 3 weeks later the signal had recovered. Images show wide-field epifluorescence images of the cells in the two imaging sessions. Scale bars, 20 μm. c, Spike SNR was on average stable in repeated recordings over three weeks (P = 0.34, two-sided paired Student’s t-test).

Extended Data Fig. 6 Validation of the pipeline for signal extraction from dense PCL videos using PMD–NMF.

a, Illustration of the segmentation pipeline: raw videos were first corrected for motion, followed by photobleaching correction, PMD denoising, manual removal of blood vessels and demixing using NMF. The pipeline produced waveforms that correspond to individual cell traces and to the background (see Methods for details). b, Testing the pipeline using simulated data composed of two cells partially overlapping in space, and with varying levels of correlation in their subthreshold voltages and with the background. Poisson-distributed shot noise was added to each pixel to mimic experimental noise. Left, image of the input video pixel-wise s.d. and the output cell footprints. Middle, input waveforms and the pipeline output waveforms. Top right, correlation matrix of input signals (C1, C2, cells 1 and 2; B, background). Bottom right, cross-correlation of output waveforms with input waveforms, data are mean ± s.d. for n = 5 simulations. c, Performance of the pipeline as a function of input parameters (n = 5 simulations per condition). Top, output-to-input correlation and output C1-to-C2 cross-correlation as a function of the pixel noise level. Noise is scaled to the spike amplitude, input C1-to-C2 cross-correlation is 0.5 and the input correlation with the background is also 0.5. Middle, output-to-input correlation and output C1-to-C2 cross-correlation as a function of the correlation between the input and the background, at two noise levels. Input C1-to-C2 correlation is 0.5. Bottom, output-to-input correlation and output C1-to-C2 cross-correlation as a function of the cross-correlation between input C1 and C2 at two noise levels. Input correlation with background is 0.5. (data are mean ± s.e.m.). d, Testing the pipeline with composite videos composed from real data. We imaged FOVs with single oriens neurons spontaneously spiking in awake, resting mice. Each cell was imaged in the focal plane, and then at 20-μm defocus. The two videos were first processed with the pipeline to extract the ground-truth input signals and then the videos were summed such that the focused and defocused cells were about 50% overlapping. We then ran the blended videos through the pipeline and compared outputs to the input traces using cross-correlation analysis. e, Mean cross-correlograms of five FOVs processed as described in d, showing that the segmentation pipeline accurately reproduced the correlational structure of the inputs even under these challenging conditions. Data are mean ± s.e.m. (line and shading). f, Validation of the image segmentation pipeline using patch-clamp recording as the ground-truth. Top left, FOV with dense expression of paQuasAr3-s in CA1 PCL in an acute brain slice. The FOV was imaged while the voltage in the blue outlined cell was recorded by manual patch clamp. Bottom left, two of the spatial footprints identified by the PMD–NMF pipeline. Middle, ground-truth voltage recording (black), flat average region of interest around the cell (blue) and two PMD–NMF units (magenta and red). The flat average region of interest trace showed fluctuations that were not present in the patch-clamp recording, presumably from an out-of-focus cell. These events were absent in the magenta PMD–NMF demixed trace. Right, magnification of the indicated trace inset. Arrow indicates the out-of-focus event. This experiment was performed once.

Extended Data Fig. 7 Simultaneous optical recording from seven spiking cells in the PCL of an anaesthetized mouse.

The mouse expressed paQuasAr3-s. a, Left, wide-field epifluorescence image of citrine fluorescence. Middle, same FOV with patterned blue illumination. Right, same FOV with patterned red illumination. Numbers indicate individual cells. b, Simultaneous fluorescence recordings from seven cells. Magnifications show synchronized complex spikes and action potentials that occur on top of subthreshold oscillations. This experiment was performed once.

Extended Data Fig. 8 Chronic recordings of oriens neurons in hippocampus.

a, Top left, wide-field epifluorescence and two-photon images showing two oriens interneurons. Scale bar, 50 μm. Middle, fluorescence recordings from the 2 cells during 3 consecutive repeats of a 65-s protocol. The protocol consisted of 10 s of rest, followed by 15-s epochs of walking at speeds of 5, 7.5 and 10 cm s−1, successively, followed by 10 s of rest. Inset (top right) shows the optical traces with clearly resolved spikes and subthreshold events. Bottom, average spike rate across three trials of the two cells (data are mean ± s.d.). b, Same two neurons and same protocol as in a, recorded seven days later. Trial 1 was only spontaneous activity without locomotion.

Extended Data Fig. 9 Examples of brain-state-dependent intercellular correlations and STAs for single pairs of cells.

ac, Left, magnified sections of recordings from trios of cells in the PCL (a, b) and oriens (c). Top, recordings during quiet. Bottom, recordings from the same cells during walking. Right, auto- and cross-correlations of the fluorescence traces, calculated from the complete 9-s recording in each brain region and behavioural state. The auto- and cross-correlations clearly show enhanced θ-rhythm in both brain regions during walking and differing cross-correlations between simultaneously recorded pairs of cells. d, e, Distribution of equal-time correlation coefficients between pairs of simultaneously recorded cells in the PCL (d) and oriens (e). f, Three examples of single-pair STA fluorescence during quiet (left) and walking (right) in the PCL (top) and oriens (bottom).

Extended Data Fig. 10 Optopatch measurements of hippocampal SST neurons.

a, Fluorescence waveforms of all 25 SST cells shown in Fig. 4. b, Cell-by-cell comparison of the change in θ-band (6.7 to 8.3 Hz) power between walking and quiet brain states. Left, without optogenetic stimulation. Middle, same cells with tonic optogenetic stimulation. Right, ratio of θ-band enhancement (walking versus quiet) with optogenetic stimulation versus without (cell-by-cell comparison for the data in Fig. 4h). c, Spike raster of 19 SST cells imaged at week 0 and week 3 and stimulated with the blue light protocol at the top (full dataset for the data in Fig. 4j–k).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adam, Y., Kim, J.J., Lou, S. et al. Voltage imaging and optogenetics reveal behaviour-dependent changes in hippocampal dynamics. Nature 569, 413–417 (2019). https://doi.org/10.1038/s41586-019-1166-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1166-7

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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