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:

Transparent intracortical microprobe array for simultaneous spatiotemporal optical stimulation and multichannel electrical recording

Abstract

Optogenetics, the selective excitation or inhibition of neural circuits by light, has become a transformative approach for dissecting functional brain microcircuits, particularly in in vivo rodent models, owing to the expanding libraries of opsins and promoters. Yet there is a lack of versatile devices that can deliver spatiotemporally patterned light while performing simultaneous sensing to map the dynamics of perturbed neural populations at the network level. We have created optoelectronic actuator and sensor microarrays that can be used as monolithic intracortical implants, fabricated from an optically transparent, electrically highly conducting semiconductor ZnO crystal. The devices can perform simultaneous light delivery and electrical readout in precise spatial registry across the microprobe array. We applied the device technology in transgenic mice to study light-perturbed cortical microcircuit dynamics and their effects on behavior. The functionality of this device can be further expanded to optical imaging and patterned electrical microstimulation.

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

Access options

Buy this article

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

Figure 1: Transparent, electrically conductive ZnO MOA device for multichannel intracortical neural recording and optical stimulation.
Figure 2: Spatial distribution of light in brain tissue and dependence of neural activity on optical power.
Figure 3: Multichannel intracortical neural recordings by 16-channel ZnO MOA in an anesthetized transgenic mouse.
Figure 4: Optical proxy for natural stimulus in the low-power regime.
Figure 5: Stimulation in the high-power regime triggered body movements.

Similar content being viewed by others

References

  1. 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  PubMed  PubMed Central  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).

    CAS  PubMed  Google Scholar 

  3. Li, X. et al. Fast noninvasive activation and inhibition of neural and network activity by vertebrate rhodopsin and green algae channelrhodopsin. Proc. Natl. Acad. Sci. USA 102, 17816–17821 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. 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  PubMed  Google Scholar 

  5. Ishizuka, T., Kakuda, M., Araki, R. & Yawo, H. Kinetic evaluation of photosensitivity in genetically engineered neurons expressing green algae light-gated channels. Neurosci. Res. 54, 85–94 (2006).

    Article  CAS  PubMed  Google Scholar 

  6. Han, X. et al. Millisecond-timescale optical control of neural dynamics in the nonhuman primate brain. Neuron 62, 191–198 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Diester, I. et al. An optogenetic toolbox designed for primates. Nat. Neurosci. 14, 387–397 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Deisseroth, K. Optogenetics. Nat. Methods 8, 26–29 (2011).

    Article  CAS  PubMed  Google Scholar 

  9. Packer, A.M., Roska, B. & Häusser, M. Targeting neurons and photons for optogenetics. Nat. Neurosci. 16, 805–815 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. 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  PubMed  PubMed Central  Google Scholar 

  11. Royer, S. et al. Multi-array silicon probes with integrated optical fibers: light-assisted perturbation and recording of local neural circuits in the behaving animal. Eur. J. Neurosci. 31, 2279–2291 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Anikeeva, P. et al. Optetrode: a multichannel readout for optogenetic control in freely moving mice. Nat. Neurosci. 15, 163–170 (2012).

    Article  CAS  Google Scholar 

  13. Wang, J. et al. Integrated device for combined optical neuromodulation and electrical recording for chronic in vivo applications. J. Neural Eng. 9, 016001 (2012).

    Article  PubMed  Google Scholar 

  14. Hayashi, Y., Tagawa, Y., Yawata, S., Nakanishi, S. & Funabiki, K. Spatio-temporal control of neural activity in vivo using fluorescence microendoscopy. Eur. J. Neurosci. 36, 2722–2732 (2012).

    Article  PubMed  Google Scholar 

  15. Stark, E., Koos, T. & Buzsáki, G. Diode probes for spatiotemporal optical control of multiple neurons in freely moving animals. J. Neurophysiol. 108, 349–363 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Pisanello, F. et al. Multipoint-emitting optical fibers for spatially addressable in vivo optogenetics. Neuron 82, 1245–1254 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Canales, A. et al. Multifunctional fibers for simultaneous optical, electrical and chemical interrogation of neural circuits in vivo. Nat. Biotechnol. 33, 277–284 (2015).

    Article  CAS  PubMed  Google Scholar 

  18. Abaya, T.V.F., Blair, S., Tathireddy, P., Rieth, L. & Solzbacher, F. A 3D glass optrode array for optical neural stimulation. Biomed. Opt. Express 3, 3087–3104 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Zorzos, A.N., Boyden, E.S. & Fonstad, C.G. Multiwaveguide implantable probe for light delivery to sets of distributed brain targets. Opt. Lett. 35, 4133–4135 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Kim, T.-I. et al. Injectable, cellular-scale optoelectronics with applications for wireless optogenetics. Science 340, 211–216 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. McCall, J.G. et al. Fabrication and application of flexible, multimodal light-emitting devices for wireless optogenetics. Nat. Protoc. 8, 2413–2428 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Takahashi, K. Wide Bandgap Semiconductors: Fundamental Properties and Modern Photonic and Electronic Devices (Springer, 2007).

  23. Zhou, J., Xu, N. & Wang, Z.L. Dissolving behavior and stability of ZnO wires in biofluids: a study on biodegradability and biocompatibility of ZnO nanostructures. Adv. Mater. 18, 2432–2435 (2006).

    Article  CAS  Google Scholar 

  24. Li, Z. et al. Cellular level biocompatibility and biosafety of ZnO nanowires. J. Phys. Chem. C 112, 20114–20117 (2008).

    Article  CAS  Google Scholar 

  25. Hochberg, L.R. et al. Reach and grasp by people with tetraplegia using a neurally controlled robotic arm. Nature 485, 372–375 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Collinger, J.L. et al. High-performance neuroprosthetic control by an individual with tetraplegia. Lancet 381, 557–564 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Borton, D.A., Yin, M., Aceros, J. & Nurmikko, A. An implantable wireless neural interface for recording cortical circuit dynamics in moving primates. J. Neural Eng. 10, 026010 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Yin, M., Borton, D.A., Aceros, J., Patterson, W.R. & Nurmikko, A.V. A 100-channel hermetically sealed implantable device for chronic wireless neurosensing applications. IEEE Trans. Biomed. Circuits Syst. 7, 115–128 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Yin, M. et al. Wireless neurosensor for full-spectrum electrophysiology recordings during free behavior. Neuron 84, 1170–1182 (2014).

    Article  CAS  PubMed  Google Scholar 

  30. Campbell, P.K., Jones, K.E., Huber, R.J., Horch, K.W. & Normann, R.A. A silicon-based, three-dimensional neural interface: manufacturing processes for an intracortical electrode array. IEEE Trans. Biomed. Eng. 38, 758–768 (1991).

    Article  CAS  PubMed  Google Scholar 

  31. Bhandari, R., Negi, S., Rieth, L., Normann, R.A. & Solzbacher, F. A novel masking method for high aspect ratio penetrating microelectrode arrays. J. Micromech. Microeng. 19, 035004 (2009).

    Article  Google Scholar 

  32. Yoo, J.-M. et al. Excimer-laser deinsulation of Parylene-C coated Utah electrode array tips. Sens. Actuators B Chem. 166, 777–786 (2012).

    Article  Google Scholar 

  33. Gross, G.W., Wen, W.Y. & Lin, J.W. Transparent indium-tin oxide electrode patterns for extracellular, multisite recording in neuronal cultures. J. Neurosci. Methods 15, 243–252 (1985).

    Article  CAS  PubMed  Google Scholar 

  34. Ozden, I. et al. A coaxial optrode as multifunction write-read probe for optogenetic studies in non-human primates. J. Neurosci. Methods 219, 142–154 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Schoenenberger, P., Grunditz, Å., Rose, T. & Oertner, T.G. Optimizing the spatial resolution of channelrhodopsin-2 activation. Brain Cell Biol. 36, 119–127 (2008).

    Article  PubMed  Google Scholar 

  36. Grossman, N. et al. The spatial pattern of light determines the kinetics and modulates backpropagation of optogenetic action potentials. J. Comput. Neurosci. 34, 477–488 (2013).

    Article  PubMed  Google Scholar 

  37. Ayling, O.G., Harrison, T.C., Boyd, J.D., Goroshkov, A. & Murphy, T.H. Automated light-based mapping of motor cortex by photoactivation of channelrhodopsin-2 transgenic mice. Nat. Methods 6, 219–224 (2009).

    Article  CAS  PubMed  Google Scholar 

  38. Hira, R. et al. Transcranial optogenetic stimulation for functional mapping of the motor cortex. J. Neurosci. Methods 179, 258–263 (2009).

    Article  PubMed  Google Scholar 

  39. Reeves, G. & Harrison, H. Obtaining the specific contact resistance from transmission line model measurements. IEEE Electron Device Lett. 3, 111–113 (1982).

    Article  Google Scholar 

  40. Azimipour, M. et al. Extraction of optical properties and prediction of light distribution in rat brain tissue. J. Biomed. Opt. 19, 75001 (2014).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank W. Jiang and S. Kim for four-point probe measurement. This work was supported by the REPAIR Program (grant N66001-10-C-2010) from the US Defense Advanced Research Projects Agency, the US National Science Foundation (grants CBET-1402803 and CBET-1264816), and Emerging Frontiers in Research and Innovation from the US National Science Foundation (grant 0937848). The work was also partly supported by the Ministry of Science, ICT & Future Planning of Korea through the Center for Integrated Smart Sensors (grant CISS-2012366054204).

Author information

Authors and Affiliations

Authors

Contributions

A.V.N. and Y.-K.S. initiated the project. J.L. developed the ZnO MOA device and conducted benchtop characterizations. I.O. designed and led rodent experiments. I.O. and J.L. collected and analyzed neural recording data. I.O. performed Monte Carlo simulations. J.L., I.O. and A.V.N. wrote the manuscript. All authors contributed to the final version of the manuscript. A.V.N. contributed to and directed all aspects of this work.

Corresponding author

Correspondence to Arto V Nurmikko.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Optical transmittance of materials used in MOA.

(a) An electrically conducting 2 mm-thick bulk ZnO substrate shows excellent optical transparency across the visible spectrum. The transmittance curves of other constituent materials used in MOA fabrication are also shown. Note that the ITO and polyimide are thin films; the oscillations in the transmittance of the polyimide film originate from optical interference, as expected in this thickness range for highly uniform films.

Source data

Supplementary Figure 2 Step-by-step fabrication process of ZnO MOA structure starting from bulk crystal.

Rounded square symbols identify key processes which are magnified in the figures: (a) Square array of Ti/Au electrical pads defined on O-face of ZnO substrate. (b) Dicing of grooves between electrical pads for electrical isolation. Pads are intentionally aligned off-center to preserve full optical access. (c) A stamping method is used to cover only the top surface of ZnO with the photoresist (PR) layer. (d) Application of polymer adhesive for electrical isolation and mechanical support. Excessive polymer was removed by a rubber blade. (e) Oxygen plasma etching step to disconnect adhesive layer and its removal in acetone. (f) Defining the square pillar array by dicing saw on the opposite side of ZnO in precise spatial registry with steps (a)-(c). (g) Transformation of the square pillars to a tapered, sharp needle array by two-step wet etch process in FeCl3 and diluted H2SO4 solutions, respectively. (h) Conformal deposition of an electrical insulation layer of parylene-C. (i) Spin coating the separation layer with PR and (j) application of PDMS etch mask layer. (k) Plasma etch process is used to expose the ZnO tips, followed by (l) thin ITO layer deposition for recording site definition. (m) PDMS mask removal by resolving PR layer in acetone.

Supplementary Figure 3 Photographs and scanning electron microscope (SEM) images showing additional details of MOA fabrication and process quality.

(a) Large area view of freshly diced and etched ZnO “needle array”, tips of the individual elements are here still covered by the PDMS etch mask and photoresist. (b) SEM image of the tapered ZnO tip geometry following the tip de-insulation process by precisely controlled plasma and wet etch steps, respectively. (c) An SEM view of the ZnO MOA array, with a protective ITO sputtered thin film covering the active tip areas. (d) Close-up cross-sectional image of the materials layers in the tip region, made here visible by intentional breaking of the MOA. The SEM image shows how sputtered ITO layer provides good semi-conformal coating even across the very end of the sharp ZnO tip. (e) Photograph of ZnO bulk single crystal substrate and microfabricated 4×4 optoelectrode array for rodent experiments. (f) Photograph of the multichannel electrical wiring arrangement, taken through semitransparent ribbon cable. Yellow dots (indicated by green circle) show the approximate size of the electrical contact pads; red square shows the unobstructed light accessible optical aperture. Optical apertures of optoelectrodes are spatially separated from gold traces to minimize optically-induced electrical artifacts. (g) The multichannel electrical connection from a 4×4 MOA to a printed circuit board and connector by a custom-made polyimide ribbon cable. (h) MOA inserted in the mouse brain. A transparent sapphire plate was placed above the ribbon cable for temporary mechanical support.

Supplementary Figure 4 Non-sequential ray tracing analysis of light propagation in the MOA structure.

(a) 3D computer model (left) and cross section of 1000 example traces from Gaussian source (w = 60 µm). Escaped rays are captured by detectors surrounding ZnO shank. Different colors of rays used to denote the segmented rays by reflection/refraction/transmission while tracing (b) Calculated intensity distribution of backreflected light measured at the back surface. Reflected rays are well confined in ZnO, away from metal contact and traces. (c) Analyzed irradiance distribution with various input beam sizes. Coupling efficiency is consistent with experimentally measured value; importantly any residual optical power leakage to brain tissue (through shank sidewalls) is negligible.

Source data

Supplementary Figure 5 Schematic diagram of laser scanning setup for spatio-temporal optical stimulation of neural activity.

(a) Time-modulated laser pulses were delivered through an optical fiber link onto a 2-axis MEMS mirror (diameter = 0.8 mm) and electronically directed to pre-programmed target positions at the optical input apertures of ZnO MOA elements. The optical alignment of the projection system was actively monitored and controlled by a CCD camera throughout the in vivo experiments. (b) Measured laser beam profile at the input aperture plane of an MOA element. Note that the beam diameter is somewhat smaller than the size of the optical aperture of the individual optoelectrodes to ensure high coupling efficiency and to avoid light scattering at the aperture edges.

Source data

Supplementary Figure 6 Demonstrating the capability for fluorescence detection by the transparent ZnO MOA.

(a) In a Thy1-ChR2 mouse, we collected fluorescence images at a variety of depths by a CCD camera while driving the MOA in to the motor cortex. Top row shows the location of those optoelectrode(s) which were targeted by the laser beam (473 nm, 225 µW). The same exposure time was used for all fluorescence images. When the MOA was outside the brain (second row from the top) only weak fluorescence was present, due to fluorescence emanating from the semitransparent polyimide cable. As the array was lowered into the brain (bottom three rows) fluorescence collected from the optoelectrodes intensified notably, now dominated by the yellow fluorescent protein (YFP) expression in the brain. (b) To confirm that these signals are not artifacts, we compared the results in a wild-type mouse (lacking YFP expression) where we only detected the weak fluorescence signals from the polyimide cable. (c) To quantify the practical influence of the fluorescence signals from the polyimide cable in performing fluorescence imaging in general, we performed fluorescence detection in fluorescein dye solutions across a range of concentrations. We inserted the optoelectrodes about 1 mm into the dye solutions and recorded the intensity of fluorescence signals. Even at dye concentrations of only about 100 nM we could clearly identify and separate the fluorescence signals originating from the dye solution from the background.

Source data

Supplementary Figure 7 Further examples of light modulated multichannel intracortical recording traces in Thy1-ChR2 mouse.

Raw sample traces (filtered between 0.3-10 kHz) of 16-channel ZnO MOA in anesthetized transgenic mouse, Blue shaded region indicates when optical stimuli are applied.

Source data

Supplementary Figure 8 LFP responses at different power regimes.

Peristimulus time plots of local field potential activity obtained in a transgenic mouse at varying levels of input optical power delivered onto optoelectrode 16, here as a light pulse of 1 s long in duration.

Source data

Supplementary Figure 9 Assessment of possible residual light-induced artifacts in electrophysiological recordings.

To quantify any light-induced artifacts, we performed simultaneous light delivery and electrophysiology using MOA in the motor cortex of a wild-type mouse lacking opsin expression. We delivered 1 s long laser pulses (n = 50) at the input aperture of the optoelectrodes and recorded multiunit spiking and LFP activity. The figure shows peristimulus time plots of average multiunit spiking rates in (a), and LFP traces in (b) when laser pulses were targeting only optoelectrode 16. Even at highest power levels laser stimulus did not trigger spiking activity in the recorded neurons in the wild type animal, and LFP band showed slight deflections only at optoelectrode 16. At levels beyond approximately 14 μW in optical power, the artifacts became detectable over the spontaneous activity in the LFP signals (only). At this power level, the artifacts had an average peak negative deflection of −0.043 ± 0.022 mV (p < 0.05) within 60 ms after the stimulus onset. In the high power regime (~1.4 mW), the average amplitude of negative deflections was higher (–0.17 ± 0.13 mV, p < 0.001).

The peak amplitude of the deflections in response to 1.36 mW light pulses (high power) was approximately 0.18 mV, some 20 fold smaller than the LFP deflections observed during high-power stimulation in the motor cortex of the transgenic mice. Thus the very small optically induced artifacts had no impact to the recorded activity observed in this work.

The noise levels with and without light in our recordings were also analyzed. To appreciate the individual contributions of the recording system and background neural activity to the overall noise level, we quantified the noise levels under three distinctly different conditions: (1) Thy1-ChR2 transgenic mouse (n = 3 animals); (2) wild type mouse (n = 1); and (3) a physiologically and anatomically dead Thy1-ChR2 transgenic mouse (n = 1, this data is collected about 1 hour after euthanizing the animal). The corresponding RMS noise levels without and with light were (1) 8.5 ± 2.9 µV and 10.1 ± 4.2 µV in transgenic mouse, (2) 7.0 ± 1.0 µV and 7.2 ± 1.5 µV in wild type mouse, and (3) 4.4 ± 0.3 µV and 4.3 ± 0.2 µV in the dead transgenic mouse, respectively.

Source data

Supplementary Figure 10 Medium power regime to explore local network connectivity.

(a) Peristimulus multiunit firing rate plots (red traces, 2 ms sliding window) and spike rasters (gray ticks) in the barrel cortex in response to 10 ms long light pulses at 5 Hz (n = 500 in 3 animals) delivered to optoelectrode 1. For each optoelectrode the response probability and latency is indicated if the neural activity at that optoelectrode showed statistically significant modulation in response to stimulation. Latencies were determined from PSTHs as the first time point after the stimulus onset where the PSTH amplitude reaches 5 standard deviations above the mean of PSTH before the stimulus onset. The response probabilities were determined as the probability of appearance of a spike within 20 ms after the stimulus (pstim), corrected against the spontaneous rate of firing pspon, i.e. (pstimpspon)/(1 – pspon). The significance of the responses (p < 0.01) was determined by boot-straping, i.e. generating 10,000 data sets through random selections from the data. The response probability at optoelectrode 1 was 1.0 with a response latency of τ = 3.8 ms. Some neighboring optoelectrodes also responded to stimulus but with lower response probabilities and longer latencies. Assuming a spike transmission delay of about 1-5 ms during synaptic transmission in the cortex1,2, we infer that sites at optoelectrode 2, 5 and 16 might locate one synapse away from the illuminated site (τ = 7.3, 7.8 and 8.5 ms, respectively). (b) Same plot as in (a), but light pulses were delivered at optoelectrode 16 for comparison. In this case, only optoelectrodes 8, 11, 12 and 15 showed a response. An inference from these two cases is that while neurons at optoelectrode 1 have a strong modulatory influence on neurons at optoelectrode 16, the reverse is not the case.

1. Thomson, A.M. & Lamy, C. Functional maps of neocortical local circuitry. Front. Neurosci. 1, 19 (2007).

2. Fujisawa, S., Amarasingham, A., Harrison, M.T. & Buzsáki, G. Behavior-dependent short-term assembly dynamics in the medial prefrontal cortex. Nat. Neurosc. 11, 823-833 (2008).

Source data

Supplementary Figure 11 Comparison of functional motor cortical map obtained by different modalities of stimulation.

We performed joint mapping studies by deploying electrical stimulation (10 monopolar pulses of 50-125 µA at 333 Hz, 100 µs pulse width) and optogenetic stimulation (1 mW, 20-50 ms pulse width), sequentially in the motor cortex. The recorded maps overlapped significantly where the motor responses were identical, at 73 ± 4% (out of n = 48) of the optoelectrode sites. 21 ± 4% of the stimulation sites showed partial overlap in motor responses, i.e. one type of stimulus generated muscle twitches of two body area whereas muscle twitches at one these body areas was obtained in response to the other type of stimulus.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–11 (PDF 5925 kb)

Spatiotemporal light delivery through ZnO MOA and localized light output in fluorescent dye solution. (MP4 2476 kb)

41592_2015_BFnmeth3620_MOESM482_ESM.mp4

Sequential control of forelimb movement of transgenic mouse: behavior generation by optical stimulus at approximately layer V. (MP4 6347 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, J., Ozden, I., Song, YK. et al. Transparent intracortical microprobe array for simultaneous spatiotemporal optical stimulation and multichannel electrical recording. Nat Methods 12, 1157–1162 (2015). https://doi.org/10.1038/nmeth.3620

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.3620

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing