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.

Fully integrated silicon probes for high-density recording of neural activity

Abstract

Sensory, motor and cognitive operations involve the coordinated action of large neuronal populations across multiple brain regions in both superficial and deep structures1,2. Existing extracellular probes record neural activity with excellent spatial and temporal (sub-millisecond) resolution, but from only a few dozen neurons per shank. Optical Ca2+ imaging3,4,5 offers more coverage but lacks the temporal resolution needed to distinguish individual spikes reliably and does not measure local field potentials. Until now, no technology compatible with use in unrestrained animals has combined high spatiotemporal resolution with large volume coverage. Here we design, fabricate and test a new silicon probe known as Neuropixels to meet this need. Each probe has 384 recording channels that can programmably address 960 complementary metal–oxide–semiconductor (CMOS) processing-compatible low-impedance TiN6 sites that tile a single 10-mm long, 70 × 20-μm cross-section shank. The 6 × 9-mm probe base is fabricated with the shank on a single chip. Voltage signals are filtered, amplified, multiplexed and digitized on the base, allowing the direct transmission of noise-free digital data from the probe. The combination of dense recording sites and high channel count yielded well-isolated spiking activity from hundreds of neurons per probe implanted in mice and rats. Using two probes, more than 700 well-isolated single neurons were recorded simultaneously from five brain structures in an awake mouse. The fully integrated functionality and small size of Neuropixels probes allowed large populations of neurons from several brain structures to be recorded in freely moving animals. This combination of high-performance electrode technology and scalable chip fabrication methods opens a path towards recording of brain-wide neural activity during behaviour.

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: The Neuropixels probe.
Figure 2: Recording from large neuronal populations with a single probe in an awake head-fixed mouse.
Figure 3: Recording from multiple brain structures in awake head-fixed mice.
Figure 4: Recordings from entorhinal and medial prefrontal cortices using chronic implants in unrestrained rats.

References

  1. Lewis, C. M., Bosman, C. A. & Fries, P. Recording of brain activity across spatial scales. Curr. Opin. Neurobiol. 32, 68–77 (2015)

    Article  CAS  Google Scholar 

  2. Bargmann, C. et al. BRAIN 2025: a Scientific Vision; Brain Research through Advancing Innovative Neurotechnologies (BRAIN) Working Group Report to the Advisory Committee to the Director, NIH. Available at: http://www.nih.gov/science/brain/2025/ (2014)

  3. Sofroniew, N. J., Flickinger, D., King, J. & Svoboda, K. A large field of view two-photon mesoscope with subcellular resolution for in vivo imaging. eLife 5, e14472 (2016)

    Article  Google Scholar 

  4. Hamel, E. J. O., Grewe, B. F., Parker, J. G. & Schnitzer, M. J. Cellular level brain imaging in behaving mammals: an engineering approach. Neuron 86, 140–159 (2015)

    Article  CAS  Google Scholar 

  5. Ahrens, M. B., Orger, M. B., Robson, D. N., Li, J. M. & Keller, P. J. Whole-brain functional imaging at cellular resolution using light-sheet microscopy. Nat. Methods 10, 413–420 (2013)

    Article  CAS  Google Scholar 

  6. Musa, S. Titanium nitride electrode. US patent 9,384,990 (2016)

  7. Buzsáki, G. Large-scale recording of neuronal ensembles. Nat. Neurosci. 7, 446–451 (2004)

    Article  Google Scholar 

  8. Stevenson, I. H. & Kording, K. P. How advances in neural recording affect data analysis. Nat. Neurosci. 14, 139–142 (2011)

    Article  CAS  Google Scholar 

  9. Meister, M., Wong, R. O. L., Baylor, D. A. & Shatz, C. J. Synchronous bursts of action potentials in ganglion cells of the developing mammalian retina. Science 252, 939–943 (1991)

    Article  ADS  CAS  Google Scholar 

  10. Wilson, M. A. & McNaughton, B. L. Dynamics of the hippocampal ensemble code for space. Science 261, 1055–1058 (1993)

    Article  ADS  CAS  Google Scholar 

  11. Seymour, J. P., Wu, F., Wise, K. D. & Yoon, E. State-of-the-art MEMS and microsystem tools for brain research. Microsyst. Nanoeng. 3, 16066 (2017)

    Article  Google Scholar 

  12. Harris, K. D., Henze, D. A., Csicsvari, J., Hirase, H. & Buzsáki, G. Accuracy of tetrode spike separation as determined by simultaneous intracellular and extracellular measurements. J. Neurophysiol. 84, 401–414 (2000)

    Article  CAS  Google Scholar 

  13. Mora Lopez, C. et al. A neural probe with up to 966 electrodes and up to 384 configurable channels in 0.13 μm SOI CMOS. IEEE Trans. Biomed. Circuits Syst. 11, 510–522 (2017)

    Article  Google Scholar 

  14. Scholvin, J. et al. Close-packed silicon microelectrodes for scalable spatially oversampled neural recording. IEEE Trans. Biomed. Eng. 63, 120–130 (2016)

    Article  Google Scholar 

  15. Rios, G., Lubenov, E. V., Chi, D., Roukes, M. L. & Siapas, A. G. Nanofabricated neural probes for dense 3-D recordings of brain activity. Nano Lett. 16, 6857–6862 (2016)

    Article  ADS  CAS  Google Scholar 

  16. Shobe, J. L., Claar, L. D., Parhami, S., Bakhurin, K. I. & Masmanidis, S. C. Brain activity mapping at multiple scales with silicon microprobes containing 1,024 electrodes. J. Neurophysiol. 114, 2043–2052 (2015)

    Article  Google Scholar 

  17. Ludwig, K. A. et al. Poly(3,4-ethylenedioxythiophene) (PEDOT) polymer coatings facilitate smaller neural recording electrodes. J. Neural Eng. 8, 014001 (2011)

    Article  ADS  Google Scholar 

  18. Roth, M. M. et al. Thalamic nuclei convey diverse contextual information to layer 1 of visual cortex. Nat. Neurosci. 19, 299–307 (2016)

    Article  CAS  Google Scholar 

  19. Pachitariu, M., Steinmetz, N. A., Kadir, S. N., Carandini, M. & Harris, K. D. Fast and accurate spike sorting of high-channel count probes with KiloSort. Adv. Neural Inf. Process. Syst. 4448–4456 (2016)

  20. Rossant, C. et al. Spike sorting for large, dense electrode arrays. Nat. Neurosci. 19, 634–641 (2016)

    Article  CAS  Google Scholar 

  21. Yizhar, O., Fenno, L. E., Davidson, T. J., Mogri, M. & Deisseroth, K. Optogenetics in neural systems. Neuron 71, 9–34 (2011)

    Article  CAS  Google Scholar 

  22. Wang, Y., Romani, S., Lustig, B., Leonardo, A. & Pastalkova, E. Theta sequences are essential for internally generated hippocampal firing fields. Nat. Neurosci. 18, 282–288 (2015)

    Article  CAS  Google Scholar 

  23. Buzsáki, G., Horváth, Z., Urioste, R., Hetke, J. & Wise, K. High-frequency network oscillation in the hippocampus. Science 256, 1025–1027 (1992)

    Article  ADS  Google Scholar 

  24. Karumbaiah, L. et al. Relationship between intracortical electrode design and chronic recording function. Biomaterials 34, 8061–8074 (2013)

    Article  CAS  Google Scholar 

  25. Hafting, T., Fyhn, M., Molden, S., Moser, M.-B. & Moser, E. I. Microstructure of a spatial map in the entorhinal cortex. Nature 436, 801–806 (2005)

    Article  ADS  CAS  Google Scholar 

  26. Jun, J. J. et al. Real-time spike sorting platform for high-density extracellular probes with ground-truth validation and drift correction. Preprint at https://www.biorxiv.org/content/early/2017/01/19/101030 (2017)

  27. Harris, K. D. Neural signatures of cell assembly organization. Nat. Rev. Neurosci. 6, 399–407 (2005)

    Article  CAS  Google Scholar 

  28. Churchland, M. M., Yu, B. M., Sahani, M. & Shenoy, K. V. Techniques for extracting single-trial activity patterns from large-scale neural recordings. Curr. Opin. Neurobiol. 17, 609–618 (2007)

    Article  CAS  Google Scholar 

  29. Alivisatos, A. P. et al. The brain activity map project and the challenge of functional connectomics. Neuron 74, 970–974 (2012)

    Article  CAS  Google Scholar 

  30. Boccara, C. N. et al. Grid cells in pre- and parasubiculum. Nat. Neurosci. 13, 987–994 (2010)

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  32. Quiroga, R. Q., Nadasdy, Z. & Ben-Shaul, Y. Unsupervised spike detection and sorting with wavelets and superparamagnetic clustering. Neural Comput. 16, 1661–1687 (2004)

    Article  Google Scholar 

  33. Okun, M., Lak, A., Carandini, M. & Harris, K. D. Long term recordings with immobile silicon probes in the mouse cortex. PLoS One 11, e0151180 (2016)

    Article  Google Scholar 

Download references

Acknowledgements

We thank the support of the charities that fully funded this work: Howard Hughes Medical Institute’s Janelia Research Campus, Allen Institute for Brain Science, Gatsby Charitable Foundation (grant GAT3353), and the Wellcome Trust (grant 100154). We thank S. Caddick for early and continued enthusiastic support of the project. We thank G. Buzsáki for advice and D. Rinberg for early discussions and advocacy. C.M., S.L.G. and C.A.A. would like to thank NSG portal personnel for offering core-hour access to the San Diego Supercomputer Center, troubleshooting and support. The Allen Institute for Brain Science wishes to thank the enduring support of our founders, Paul G. Allen and Jody Allen, without whom this work could not have been accomplished. J.C., C.A. and V.B. were funded by NERF. Experiments and software development in the laboratory of M.C. and K.D.H. were supported by the Wellcome Trust (grants 095668 and 095669). M.C. holds the GlaxoSmithKline/Fight for Sight Chair in Visual Neuroscience. N.A.S. was supported by postdoctoral fellowships from the Human Frontier Science Program and the Marie Curie Actions of the EU.

Author information

Authors and Affiliations

Authors

Contributions

A.K.L., T.D.H. and B.D. conceived and originated the project. T.D.H., C.K., J.O., M.C., K.D.H., M.H., A.K.L. and K.S. secured project funding. T.D.H., A.K.L., B.D., S.M., C.M.L., J.P., J.O., C.K., T.J.B., J.J.J., N.A.S., J.H.S., D.J.D., M.Bau, B.B., D.A.G., K.D.H., M.H., B.K., P.L., P.D.R. and K.S. determined specifications. A.A., C.M.L., S.M., J.P., W.S. and T.D.H. designed and produced devices and firmware. J.J.J., N.A.S., J.H.S., D.J.D., B.B., T.J.B., B.K., P.L., C.M.L., S.M., J.P., W.S. and T.D.H. tested devices in vitro. N.A.S., J.H.S., D.J.D., M.Bar, V.B., C.A. and J.C. performed acute recordings in mouse or rat. J.J.J. and P.D.R. performed chronic recordings in the rat mPFC. M.Bau performed chronic recordings in the rat entorhinal cortex. N.A.S. and M.O. performed chronic recordings in mouse. J.J.J., N.A.S., J.H.S., M.Bau, B.B., T.J.B., C.M.L., S.M., J.P. and W.S. developed instrumentation or other materials. J.J.J., N.A.S., J.H.S., D.J.D., M.Bau, B.B., A.K.L., M.Bar, T.J.B., V.B., M.C., J.C., D.A.G., K.D.H., M.H., P.L., J.O., M.O., P.D.R., K.S. and T.D.H. designed experiments. B.K., J.P., J.H.S., D.J.D., J.J.J., M.P., K.D.H. and C.R. wrote software for data acquisition or analysis. C.A.A., S.L.G. and C.M. performed simulations. J.J.J., N.A.S., J.H.S., D.J.D., M.Bau, B.B., C.A. and J.C. analysed data. T.D.H., A.K.L., M.C., N.A.S., D.J.D., C.K., J.J.J., J.H.S., M.Bau, V.B., J.C., K.D.H., C.M.L., B.D. and J.O. wrote and edited the manuscript with input from all authors. J.J.J., N.A.S., J.H.S., D.J.D., M.Bau, B.B. and J.C. prepared figures. T.D.H., B.D., M.C., J.O., T.J.B., K.D.H., D.J.D., V.B., C.K. and K.S. supervised the work. T.D.H and S.M. managed the project.

Corresponding author

Correspondence to Timothy D. Harris.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks M. Maharbiz and the other anonymous reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Figure 1 Comparison of switchable and active probe options.

a, Cumulative distribution of the single unit peak amplitude from recordings using non-switchable (option 1, blue, n = 5 recordings) or switchable (option 3, green, n = 5 recordings) probes. Each distribution shows an individual recording session; sessions were spike sorted, and the mean waveform on every channel for each unit was computed. The largest absolute peak in the mean waveform on any channel was taken as the amplitude. b, Distribution of mean firing rate across the recording sessions for each sorted unit from non-switchable (blue, n = 119 cells) and switchable (green, n = 294 cells) probes; all recordings with each variant are combined. Switches did not have an appreciable effect on the amplitude or firing rate of recorded units. c, Artefacts in awake head-fixed locomoting mice are similar across passive and active probe options and can be removed by common (median) average referencing. Filtered traces (bandpass 300–3,000 Hz) of 4 channels sampled at different depths (grey) and corrected common average referenced (CAR) traces (red). The CAR corrected traces are obtained by subtracting from each trace the median across all 120 channels (bottom). Figure panel depicts representative examples of artefacts found on both probe options. d, Extracellular spike waveform amplitude (signal), noise amplitude, and spike signal-to-noise ratio (SNR) measured in the cortex in vivo using passive (black) or actively amplified (red) recording sites. Voltage amplitudes are sampled from 250,000 spikes across the entire probe for each experiment (4 experiments with passive (SNR = 8.78 ± 0.52, 95% confidence interval) and 4 with active (SNR = 8.95 ± 0.54) recording sites). Signal-to-noise (right) distributions are similar; P = 0.78, two-sided Wilcoxon rank-sum. e, f, Signal-to-noise ratio computed using 200,000 spikes during locomotion (red) or stationary (black) epochs. Recordings with passive sites (3 recordings) are depicted in e and recordings with active sites (4 recordings) in f. SNR distributions are similar during stationary and locomotion epochs with both probe options. Overall, the performance of probes with on-site buffer amplification is similar to that of probes with passive sites and offers no clear advantage in vivo.

Extended Data Figure 2 Impact and chewing transient rejection.

Reduction of interfering transients with CAR post-processing in software. Recordings using external reference on one animal, 6 or 12 trials for each technology option. a, Example signal traces for a phase 3 unbuffered, switched probe in a chronically implanted rat tapped with a cable tie. Shown is the median across all 374 recording sites (red), the raw traces (grey, from 4 adjacent sites), and traces corrected with local CAR (blue; see Methods). b, Impact transient magnitudes for all 4 probe options chronically implanted with external referencing, and corrected with global or local CAR (n = 6 trials, one animal, median ± interquartile range). Probes with unity-gain buffer amplifiers at the sites (options 2 and 4) offer no significant advantage for CAR-processed data over probes without buffers (options 1 and 3). c, Transients as in a, induced by chewing food. d, Comparison of all 4 options, as in b, one animal. Again, there was no significant advantage to buffer amplifiers after CAR correction (two-sample t-test, P = 0.1266, α = 0.05, one-tailed). The median and interquartile ranges are computed from 12 chewing events measured at 270 sites (option 4) or 374 sites (options 1–3). n in b and d represents the number of independent samples (number of sites × number of trials). See Methods for details.

Extended Data Figure 3 Light sensitivity tests.

a, b, A comparison of the light sensitivity of active and passive Neuropixels probes, along with ‘conventional’ Si probes. a, Plot of transient amplitude versus illumination power density for a 5-ms directly applied 473-nm light pulse. Active sites, especially unswitched, were most sensitive. Switched and unswitched passive probes were much less sensitive. In each measurement the 5-ms light pulse was applied singly or in a pulse train at 20 Hz for 1 s. b, Example response to illumination with a 5-ms pulse at the indicated power density. In all cases, probes were illuminated with light from a 100-μm core multimode fibre with the probe immersed in phosphate buffered saline. Below intensities of 10–20 mW mm−2 directly onto the probe, there is no impact on spike detection (see Extended Data Fig. 8) or LFP phase or magnitude determination for passive switched Neuropixels probes. Experiments in a and b were performed once for each probe type. ce, Neuropixels recording sites were immersed in saline and directly illuminated with blue (465 nm) LED light from a 200-μm diameter, 0.66 numerical aperture (NA) fibre optic cable. Light was delivered with 6 different pulse shapes (ramp, raised cosine, square 1 ms, 10 ms, 100 ms and 1 s) at 3 peak power levels. c, Light power measured with an amplified photodetector for all 18 pulse types. d, e, Artefacts on a representative channel for low (cyan traces), middle (magenta traces) and high (yellow traces) light levels. For reference, the input range of the LFP band (gain of 250×) is overlaid across the LFP plots.

Extended Data Figure 4 Long-term stability studies: effects of shank width, TiN sites and active and switched sites.

ad, To compare the chronic performance of 50-μm wide and 70-μm wide shanks as well as TiN and poly(3,4-ethylenedioxythiophene (PEDOT)-coated gold sites, four-shank probes with two shanks of each width were constructed. One version had PEDOT-coated gold sites, and a second had TiN sites. These were tested in separate animals. a, A diagram of the probe geometry for both probes, two of four shanks, a photograph of the distal end of a TiN site shank, and an electron microscopy image of one TiN site. The site is 12 × 12 μm. These probes were chronically implanted in the rat mPFC and recorded in unrestrained animals, without advancing the electrodes, in multiple sessions over a period of 6–8 weeks (further implantation, recording and analysis details are described in the Methods). b, The event rate versus implant age for one of two rats with TiN probes in the mPFC. We define an event as time-coincident spikes recorded on a contiguous group of sites for which the maximum amplitude on any site in the group exceeds the threshold. c, The event rate versus implant age for four rats with one probe each in mPFC: two animals with TiN site probes, two animals with gold-coated PEDOT site probes. Differences in the behavioural state probably contribute to the variability in event rate. d, The event SNR, defined in the Methods. The stability of this signal quality metric suggests that the drop in event rate for PEDOT-1 was due to biological factors unrelated to probe site integrity. eg, Probes (phase 2) of all four shank designs—passive, active, passive switched and active switched—were implanted chronically in the rat mPFC. Recordings were made at least once per week. Data analysis, the implant method and recording procedures are described in detail in the Methods. There is no apparent downward trend in neural activity by either metric (e, f), thus all four shank technologies can achieve exceptional chronic stability. g, mPFC recording, 200 ms of activity traces and examples of waveforms from two sorted units. Insets, dark lines show average waveform overlaid with 30 randomly selected single event waveforms of a neuron located at the top (green) and bottom (red) of the most distal group of 128 sites.

Extended Data Figure 5 Stability of chronically implanted phase 3 active and switched probe technologies in rats.

a, b, Phase 3 fully integrated probes: two of four designs were implanted in the mPFC of rats as for phase 2 (Extended Data Fig. 4). One probe with amplified sites and two switchable, unamplified probes (the preferred technology) were implanted. Amplified, switchable probes were implanted but were damaged electronically before adequate recording data could be accumulated. There was no indication of activity loss over the 8 weeks of recording (linear regression t-test, one-tailed, P > 0.1) except for the top half of one probe (blue). This good performance was expected based on phase 2 results, but the active and active switched probes are 70-μm wide in phase 3 compared to 50-μm wide for phase 2. See Methods for details of data analysis, the implant method and recording procedures. We define an event as time-coincident spikes recorded on a contiguous group of sites for which the maximum amplitude on any site in the group exceeds the threshold. Across a population of 14 probes chronically implanted in the rat mPFC (phases 2 and 3), we did not observe degradation of spiking activity over 8 weeks (linear regression t-test, single-tailed, P > 0.1, n = 30 sessions). The sole exception to this observation is shown in a and b above (blue). In this case, the top half of the distal most 3.8 mm array lost nearly all activity over the first 30 days, whereas the bottom ~1.9 mm remained stable for the duration of the 8 weeks monitored. A few probe implants became detached from the skull after many months (3 out of 16 implants, 223–482 days after implantation) or surgery wound irritation required the subject animal to be euthanized (but always after at least 20 weeks). c, mPFC recording, 200 ms of activity traces and examples of waveforms from two sorted units. Insets, dark lines show the average waveform overlaid with 30 randomly selected single event waveforms of a neuron located at the top (blue) and bottom (red) of the distal most group of 384 sites.

Extended Data Figure 6 Further examples of LFP data quality.

a, Two example periods of LFP data, over the entire depth of the probe. The data were filtered with a third-order Butterworth low-pass filter with 300 Hz cut-off frequency, but otherwise unprocessed. Only one channel from each depth along the probe is shown, so the total number of plotted traces is 192 rather than 384. Left example shows an epoch without ripples, and right example shows an epoch with a ripple (indicated by red arrow). The ripple is expanded in the inset. b, Quantification of LFP signal amplitude (r.m.s.) across depth. At the right, the anatomical locations of the sections of the probe are depicted, which were determined from histology, part of which is shown in c, along with functional markers such as spike rates and amplitudes. c, Histological section depicting the location of the probe (stained with DiI) as it passed through the hippocampus. d, Theta-filtered LFP. The same data from a (left) were filtered with a third-order bandpass Butterworth filter between 4 and 10 Hz, and are here depicted as a colour map to emphasize the phase inversion at the CA1 pyramidal layer (indicated by the red dashed line).

Extended Data Figure 7 Spike band signal quality across brain areas and multiple acute recordings.

a, For each channel in each brain area, the number of neurons with a mean waveform greater than 20 μV was counted (‘density’). b, For each neuron recorded in each brain area, the number of channels on which that neuron produces a mean waveform above 20 μV (‘spread’) was counted. In a separate set of experiments, the same probe was used for a series of 12 acute in vivo recordings in 10 animals across 4 months. c, The signal quality of each recording was computed by measuring the amplitude of detected spikes relative to the detection threshold within a 30 s selection of the recording. The detection threshold used was always 5 times the standard deviation (σ) of the mean signal. d, The mean (±s.e.m.; n = 200) ratio of the ninetieth percentile (blue), fiftieth percentile (red) and tenth percentile (green) to the 5σ detection threshold across all channels with an event, for each recording.

Extended Data Figure 8 Recording during optogenetic stimulation of excitatory and inhibitory cell populations.

a, b, A Neuropixels probe was inserted into the primary visual cortex of a Rorb-Cre;Ai32 mouse expressing channelrhodopsin-2 (ChR2) in Rorb-positive cells, which are primarily excitatory neurons of layer 4. Stimulation was carried out with a 465 nm LED coupled to a 200-μm, 0.66 NA fibre optic cable placed on the surface of the cortex approximately 1 mm from the recording site. Each trial (n = 20) consisted of a raised cosine light pulse 1 s in duration, with a peak intensity of 300 mW mm−2. a, Average multiunit firing rate across 250 channels, aligned to the start of the optogenetic stimulus. Location of the brain surface is marked with a white line. b, Average LFP band response to the optogenetic stimulus for 125 channels along the left edge of the probe (a subset of the sites in a). c, d, A Neuropixels probe was inserted into the primary visual cortex of a PV-Cre;Ai32 mouse expressing ChR2 in parvalbumin (PV)-positive cells, which are primarily fast-spiking inhibitory neurons. Stimulation was carried out with a 470 nm laser focused to a ~150 μm diameter spot on the surface of the cortex at the recording site. Each trial (n = 15) consisted of a 500 ms, 40 Hz raised cosine light pulse train, with a peak intensity of 25 mW mm−2. c, Average multiunit firing rate (regular-spiking waveforms only) across 150 channels, aligned to the start of the optogenetic stimulus. Location of the brain surface is marked with a white line. d, Average LFP band response to the optogenetic stimulus for 75 channels along the left edge of the probe (a subset of the sites in c). ei, Light artefact and detection of optogenetically driven spikes in a PV-Cre;Ai32 mouse. e, Raw data for a single trial with optical transients at onset and offset of light stimulation. f, The same trial and channel as e, after the light-evoked transient was reduced by subtraction of the median of the activity across all channels as in Extended Data Fig. 1. Spike times identified by the spike sorting algorithm are indicated with circles below the trace. g, Zoomed in view of the period of light stimulation. h, The average number of spikes, for this unit, during light stimulation. Each bin is 1.7 ms. i, Overlaid waveforms for three spikes during light stimulation (blue) and outside of light stimulation (black).

Extended Data Figure 9 Stability of a probe implanted chronically in a mouse over 21 weeks.

a, A chronically implanted mouse, 85 days after probe implantation. The subject was a male C57BL/6J mouse, 92 days old at the time of probe implant. The probe is protected by a custom plastic enclosure. Total implant weight was approximately 3.0 g, including the metal headplate (bar with side holes) that was used to fix the head during recordings. b, Firing rate across the probe, measured by counting spikes with amplitude >50 μV trough-to-peak at each depth as a function of days since implantation. Day 0 was recorded in the anaesthetized condition during implantation; all subsequent days were recorded during wakefulness. c, Total summed firing rate across all channels, showing a decrease in the few days after implantation followed by mostly stable high-yield recordings for a period of 153 days, after which time the experiment was ended by experimenter decision. d, Example waveforms from three putative frontal cortical neurons and one neuron in the lateral septal nucleus recorded on day 153 after implantation. Probe icon at the right depicts the location of the waveforms on the probe.

Extended Data Table 1 Measured operating parameters of the 4 probe types reported (phase 3)

Supplementary information

Supplementary Information

This file contains the probe technical description and design phases 1–3. (PDF 356 kb)

Life Sciences Reporting Summary (PDF 72 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jun, J., Steinmetz, N., Siegle, J. et al. Fully integrated silicon probes for high-density recording of neural activity. Nature 551, 232–236 (2017). https://doi.org/10.1038/nature24636

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature24636

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