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:

Gamma-range synchronization of fast-spiking interneurons can enhance detection of tactile stimuli

Subjects

Abstract

We tested the sensory impact of repeated synchronization of fast-spiking interneurons (FS), an activity pattern thought to underlie neocortical gamma oscillations. We optogenetically drove 'FS-gamma' while mice detected naturalistic vibrissal stimuli and found enhanced detection of less salient stimuli and impaired detection of more salient ones. Prior studies have predicted that the benefit of FS-gamma is generated when sensory neocortical excitation arrives in a specific temporal window 20–25 ms after FS synchronization. To systematically test this prediction, we aligned periodic tactile and optogenetic stimulation. We found that the detection of less salient stimuli was improved only when peripheral drive led to the arrival of excitation 20–25 ms after synchronization and that other temporal alignments either had no effects or impaired detection. These results provide causal evidence that FS-gamma can enhance processing of less salient stimuli, those that benefit from the allocation of attention.

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: Task structure.
Figure 2: Gamma is expressed in distinct bouts of activity in the LFP, associated with enhanced phase locking of spiking activity.
Figure 3: Optogenetic stimulation to emulate FS-gamma.
Figure 4: Generation and characterization of naturalistic vibrissal stimuli.
Figure 5: Endogenous and entrained FS-gamma predict enhanced detection of less salient naturalistic stimuli.
Figure 6: Entrained FS-gamma enhances detection of less salient periodic stimuli at a specific optimal temporal offset between FS synchronization and sensory drive.
Figure 7: Interaction between optogenetically entrained gamma and the response to vibrissal deflections.

Similar content being viewed by others

References

  1. Fries, P., Reynolds, J.H., Rorie, A.E. & Desimone, R. Modulation of oscillatory neuronal synchronization by selective visual attention. Science 291, 1560–1563 (2001).

    CAS  PubMed  Google Scholar 

  2. Fries, P., Womelsdorf, T., Oostenveld, R. & Desimone, R. The effects of visual stimulation and selective visual attention on rhythmic neuronal synchronization in macaque area V4. J. Neurosci. 28, 4823–4835 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Bosman, C.A. et al. Attentional stimulus selection through selective synchronization between monkey visual areas. Neuron 75, 875–888 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Ray, S., Ni, A.M. & Maunsell, J.H. Strength of gamma rhythm depends on normalization. PLoS Biol. 11, e1001477 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Meador, K.J., Ray, P.G., Echauz, J.R., Loring, D.W. & Vachtsevanos, G.J. Gamma coherence and conscious perception. Neurology 59, 847–854 (2002).

    Article  CAS  PubMed  Google Scholar 

  6. Womelsdorf, T., Fries, P., Mitra, P.P. & Desimone, R. Gamma-band synchronization in visual cortex predicts speed of change detection. Nature 439, 733–736 (2006).

    Article  CAS  PubMed  Google Scholar 

  7. Hoogenboom, N., Schoffelen, J.M., Oostenveld, R. & Fries, P. Visually induced gamma-band activity predicts speed of change detection in humans. Neuroimage 51, 1162–1167 (2010).

    Article  PubMed  Google Scholar 

  8. Wang, X.J. & Buzsáki, G. Gamma oscillation by synaptic inhibition in a hippocampal interneuronal network model. J. Neurosci. 16, 6402–6413 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Börgers, C. & Kopell, N. Synchronization in networks of excitatory and inhibitory neurons with sparse, random connectivity. Neural Comput. 15, 509–538 (2003).

    Article  PubMed  Google Scholar 

  10. Tiesinga, P.H., Fellous, J.M., Salinas, E., José, J.V. & Sejnowski, T.J. Inhibitory synchrony as a mechanism for attentional gain modulation. J. Physiol. Paris 98, 296–314 (2004).

    Article  PubMed  Google Scholar 

  11. Börgers, C. & Kopell, N.J. Gamma oscillations and stimulus selection. Neural Comput. 20, 383–414 (2008).

    Article  PubMed  Google Scholar 

  12. Knoblich, U., Siegle, J.H., Pritchett, D.L. & Moore, C.I. What do we gain from gamma? Local dynamic gain modulation drives enhanced efficacy and efficiency of signal transmission. Front. Hum. Neurosci. 4, 185 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Shadlen, M.N. & Movshon, J.A. Synchrony unbound: a critical evaluation of the temporal binding hypothesis. Neuron 24, 67–77, 111–125 (1999).

    Article  CAS  PubMed  Google Scholar 

  14. Ray, S. & Maunsell, J.H. Differences in gamma frequencies across visual cortex restrict their possible use in computation. Neuron 67, 885–896 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Penttonen, M., Kamondi, A., Acsady, L. & Buzsaki, G. Gamma frequency oscillation in the hippocampus of the rat: intracellular analysis in vivo. Eur. J. Neurosci. 10, 718–728 (1998).

    Article  CAS  PubMed  Google Scholar 

  16. Whittington, M.A., Traub, R.D., Kopell, N., Ermentrout, B. & Buhl, E.H. Inhibition-based rhythms: experimental and mathematical observations on network dynamics. Int. J. Psychophysiol. 38, 315–336 (2000).

    Article  CAS  PubMed  Google Scholar 

  17. Hasenstaub, A. et al. Inhibitory postsynaptic potentials carry synchronized frequency information in active cortical networks. Neuron 47, 423–435 (2005).

    Article  CAS  PubMed  Google Scholar 

  18. Börgers, C., Epstein, S. & Kopell, N.J. Gamma oscillations mediate stimulus competition and attentional selection in a cortical network model. Proc. Natl. Acad. Sci. USA 105, 18023–18028 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Cardin, J.A. et al. Driving fast-spiking cells induces gamma rhythm and controls sensory responses. Nature 459, 663–667 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. 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 

  21. Buzsáki, G. & Wang, X.J. Mechanisms of gamma oscillations. Annu. Rev. Neurosci. 35, 203–225 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Moore, C.I., Carlen, M., Knoblich, U. & Cardin, J.A. Neocortical interneurons: from diversity, strength. Cell 142, 189–193 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Traub, R.D., Whittington, M.A., Colling, S.B., Buzsáki, G. & Jefferys, J.G. Analysis of gamma rhythms in the rat hippocampus in vitro and in vivo. J. Physiol. (Lond.) 493, 471–484 (1996).

    Article  CAS  Google Scholar 

  24. Cardin, J.A. et al. Targeted optogenetic stimulation and recording of neurons in vivo using cell-type-specific expression of Channelrhodopsin-2. Nat. Protoc. 5, 247–254 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Carlén, M. et al. A critical role for NMDA receptors in parvalbumin interneurons for gamma rhythm induction and behavior. Mol. Psychiatry 17, 537–548 (2012).

    Article  PubMed  CAS  Google Scholar 

  26. Cohen, J.D. & Castro-Alamancos, M.A. Detection of low salience whisker stimuli requires synergy of tectal and thalamic sensory relays. J. Neurosci. 30, 2245–2256 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Miyashita, T. & Feldman, D.E. Behavioral detection of passive whisker stimuli requires somatosensory cortex. Cereb. Cortex 23, 1655–1662 (2013).

    Article  PubMed  Google Scholar 

  28. Sachidhanandam, S., Sreenivasan, V., Kyriakatos, A., Kremer, Y. & Petersen, C.C. Membrane potential correlates of sensory perception in mouse barrel cortex. Nat. Neurosci. 16, 1671–1677 (2013).

    Article  CAS  PubMed  Google Scholar 

  29. Hamada, Y., Miyashita, E. & Tanaka, H. Gamma-band oscillations in the “barrel cortex” precede rat's exploratory whisking. Neuroscience 88, 667–671 (1999).

    Article  CAS  PubMed  Google Scholar 

  30. Sirota, A. et al. Entrainment of neocortical neurons and gamma oscillations by the hippocampal theta rhythm. Neuron 60, 683–697 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jones, M.S. & Barth, D.S. Sensory-evoked high-frequency (gamma-band) oscillating potentials in somatosensory cortex of the unanesthetized rat. Brain Res. 768, 167–176 (1997).

    Article  CAS  PubMed  Google Scholar 

  32. Shaw, F.Z. & Chew, J.H. Dynamic changes of gamma activities of somatic cortical evoked potentials during wake-sleep states in rats. Brain Res. 983, 152–161 (2003).

    Article  CAS  PubMed  Google Scholar 

  33. Adesnik, H. & Scanziani, M. Lateral competition for cortical space by layer-specific horizontal circuits. Nature 464, 1155–1160 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Shao, Y.R. et al. Plasticity of recurrent l2/3 inhibition and gamma oscillations by whisker experience. Neuron 80, 210–222 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Baranauskas, G. et al. Origins of 1/f2 scaling in the power spectrum of intracortical local field potential. J. Neurophysiol. 107, 984–994 (2012).

    Article  PubMed  Google Scholar 

  36. Jia, X., Tanabe, S. & Kohn, A. Gamma and the coordination of spiking activity in early visual cortex. Neuron 77, 762–774 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Börgers, C., Epstein, S. & Kopell, N.J. Background gamma rhythmicity and attention in cortical local circuits: a computational study. Proc. Natl. Acad. Sci. USA 102, 7002–7007 (2005).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  38. Ritt, J.T., Andermann, M.L. & Moore, C.I. Embodied information processing: vibrissa mechanics and texture features shape micromotions in actively sensing rats. Neuron 57, 599–613 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Wolfe, J. et al. Texture coding in the rat whisker system: slip-stick versus differential resonance. PLoS Biol. 6, e215 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Pinto, D.J., Brumberg, J.C. & Simons, D.J. Circuit dynamics and coding strategies in rodent somatosensory cortex. J. Neurophysiol. 83, 1158–1166 (2000).

    Article  CAS  PubMed  Google Scholar 

  41. Gerdjikov, T.V., Bergner, C.G., Stüttgen, M.C., Waiblinger, C. & Schwarz, C. Discrimination of vibrotactile stimuli in the rat whisker system: behavior and neurometrics. Neuron 65, 530–540 (2010).

    Article  CAS  PubMed  Google Scholar 

  42. Jadhav, S.P., Wolfe, J. & Feldman, D.E. Sparse temporal coding of elementary tactile features during active whisker sensation. Nat. Neurosci. 12, 792–800 (2009).

    Article  CAS  PubMed  Google Scholar 

  43. Gross, J., Schnitzler, A., Timmermann, L. & Ploner, M. Gamma oscillations in human primary somatosensory cortex reflect pain perception. PLoS Biol. 5, e133 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Fellous, -M., Rudolph, M., Destexhe, A. & Sejnowski, J. Synaptic background noise controls the input/output characteristics of single cells in an in vitro model of in vivo activity. Neuroscience 122, 811–829 (2003).

    Article  CAS  PubMed  Google Scholar 

  45. Rudolph, M., Pospischil, M., Timofeev, I. & Destexhe, A. Inhibition determines membrane potential dynamics and controls action potential generation in awake and sleeping cat cortex. J. Neurosci. 27, 5280–5290 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. O'Connor, D.H. et al. Neural coding during active somatosensation revealed using illusory touch. Nat. Neurosci. 16, 958–965 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Histed, M.H. & Maunsell, J.H. Cortical neural populations can guide behavior by integrating inputs linearly, independent of synchrony. Proc. Natl. Acad. Sci. USA 111, E178–E187 (2014).

    Article  CAS  PubMed  Google Scholar 

  48. Doron, G., von Heimendahl, M., Schlattmann, P., Houweling, A. & Brecht, M. Spiking Irregularity and Frequency Modulate the Behavioral Report of Single-Neuron Stimulation. Neuron 81, 653–663 (2014).

    Article  CAS  PubMed  Google Scholar 

  49. Lee, S.H. et al. Activation of specific interneurons improves V1 feature selectivity and visual perception. Nature 488, 379–383 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Engel, A.K. & Singer, W. Temporal binding and the neural correlates of sensory awareness. Trends Cogn. Sci. 5, 16–25 (2001).

    Article  PubMed  Google Scholar 

  51. Boyden, E.S. et al. Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8, 1263–1268 (2005).

    Article  CAS  PubMed  Google Scholar 

  52. Tsien, J.Z. et al. Subregion- and cell type–restricted gene knockout in mouse brain. Cell 87, 1317–1326 (1996).

    Article  CAS  PubMed  Google Scholar 

  53. Nguyen, D.P. et al. Micro-drive array for chronic in vivo recording: tetrode assembly. J. Vis. Exp. 26, 1098 (2009).

    Google Scholar 

  54. Voigts, J., Siegle, J.H., Pritchett, D.L. & Moore, C.I. The flexDrive: An ultra-light implant for optical control and highly parallel chronic recording of neuronal ensembles in freely moving mice. Front. Sys. Neurosci. 7, 8 (2013).

  55. Siegle, J.H. Combining optical stimulation with extracellular electrophysiology in behaving mice. in Neuronal Network Analysis (eds. Fellin, T. & Halassa, M.) 357–372 (Springer, 2012).

  56. Carvell, G.E. & Simons, D.J. Membrane potential changes in rat SmI cortical neurons evoked by controlled stimulation of mystacial vibrissae. Brain Res. 448, 186–191 (1988).

    Article  CAS  PubMed  Google Scholar 

  57. Armstrong-James, M., Fox, K. & Das-Gupta, A. Flow of excitation within rat barrel cortex on striking a single vibrissa. J. Neurophysiol. 68, 1345–1358 (1992).

    Article  CAS  PubMed  Google Scholar 

  58. Moore, C.I. & Nelson, S.B. Spatio-temporal subthreshold receptive fields in the vibrissa representation of rat primary somatosensory cortex. J. Neurophysiol. 80, 2882–2892 (1998).

    Article  CAS  PubMed  Google Scholar 

  59. Brecht, M., Roth, A. & Sakmann, B. Dynamic receptive fields of reconstructed pyramidal cells in layers 3 and 2 of rat somatosensory barrel cortex. J. Physiol. (Lond.) 553, 243–265 (2003).

    Article  CAS  Google Scholar 

  60. Tolias, A.S. et al. Recording chronically from the same neurons in awake, behaving primates. J. Neurophysiol. 98, 3780–3790 (2007).

    Article  PubMed  Google Scholar 

  61. Schmitzer-Torbert, N., Jackson, J., Henze, D., Harris, K. & Redish, A.D. Quantitative measures of cluster quality for use in extracellular recordings. Neuroscience 131, 1–11 (2005).

    Article  CAS  PubMed  Google Scholar 

  62. Davidson, T.J., Kloosterman, F. & Wilson, M.A. Hippocampal replay of extended experience. Neuron 63, 497–507 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Goldberg, J.M. & Brown, P.B. Response of binaural neurons of dog superior olivary complex to dichotic tonal stimuli: some physiological mechanisms of sound localization. J. Neurophysiol. 32, 613–636 (1969).

    Article  CAS  PubMed  Google Scholar 

  64. Kumbhani, R.D., Nolt, M.J. & Palmer, L.A. Precision, reliability, and information-theoretic analysis of visual thalamocortical neurons. J. Neurophysiol. 98, 2647–2663 (2007).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank R. Clary, J. Feather, H. Farrow, R. Lichtin, S. Bechek, J. Klee, N. Padilla and C. Burley for help running experiments, J. Cardin, U. Knoblich, M. Halassa, J. Ritt, J. Voigts, C. Deister, B. Higashikuibo, D. Meletis and M. Carlén, and members of the Moore laboratory, M. Wilson, M. Andermann, R. Haslinger, N. Kopell, C. Börgers, R. Sekuler and D. Sheinberg for their comments on the manuscript. This study was supported by a grant from the US National Institutes of Health to C.I.M., a National Research Service Award Fellowship to D.L.P., and a National Defense and Science & Engineering Graduate Fellowship and a National Research Service Award Fellowship to J.H.S.

Author information

Authors and Affiliations

Authors

Contributions

J.H.S., D.L.P. and C.I.M. designed the experiments. J.H.S. designed the implants and performed the viral injections. D.L.P. and J.H.S. designed the behavioral rig and oversaw training. J.H.S. and D.L.P. analyzed the data. J.H.S., D.L.P. and C.I.M. made the figures and wrote the manuscript.

Corresponding author

Correspondence to Christopher I Moore.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Quantifying gamma events.

a, An example of a gamma event in the local field potential (bottom), wavelet power spectrum and spectrogram (top left), and Fourier transform power spectrum and spectrogram (top right). Note that the highlighted peak in the spectrogram is associated with a clearly visible 40 Hz oscillation in the raw voltage trace. The 30–80 Hz range for event centroids is shown in white. Similar events were observed in all animals, with an average rate of occurrence of 0.89 Hz during periods in which the animals were not licking. b, Examples of spontaneously arising gamma oscillations during the pre-stimulus period (similar to Figure 2b in the main text). Mouse 1 is the same subject used in Figure 2b. Gamma events also occurred on “miss” trials, but at a lower frequency. Mouse 1: 1.430 events/trial on hits vs. 1.407 events/trial on misses; Mouse 2: 1.309 events/trial on hits vs. 1.272 events/trial on misses; Mouse 3: 1.245 events/trial on hits vs. 1.205 events/trial on misses.

Supplementary Figure 2 AAV and histology.

a, Schematic of the viral vector, which interacts with Cre in parvalbumin-positive cells to induce expression of ChR2-mCherry. b, Example histological section, showing viral expression primarily in the upper layers of barrel cortex. Expression was typically distributed over 1–3 barrel columns (300–500 μm medial–lateral spread, 500–1000 μm anterior–posterior spread). Blue = DAPI stain, red = mCherry fluorescence. Viral expression was confirmed in all experimental animals.

Supplementary Figure 3 Naturalistic stimuli.

a, The 17 motion sequences of an ex vivo B2 vibrissa contacting a rotating drum covered in sandpaper that were used for natural stimulus presentation in the detection task. b, Impact of optical stimulation in a control (Cre-negative littermate) animal on the natural scenes task. This mouse was injected with ChR2 and subjected to the same training protocol as all other mice. In contrast with the four Cre-positive animals, no significant modulation by the laser was observed (R2 = 0.073, P = 0.30).

Supplementary Figure 4 d′ as an indicator of detection performance.

a, Illustration of how d′ converts hit rate and false alarm rate into a single measure of perceptual acuity. b, Performance of one animal over the course of 4407 trials from 11 sessions. Top plot shows lick times on individual trials (black dots), bottom plot shows change in d′ for blocks of 50 trials. The animal transitioned between periods of constant licking, accurately licking to the stimulus, and no licking. Only blocks of trials in which d′ was above a threshold of 1.25 (green lines) were included in further analysis.

Supplementary Figure 5 Propagation of signals from the periphery to barrel cortex.

a, Schematic showing three synapses between the periphery and the cortex. PrV = principal trigeminal nucleus; VPM = ventral posterior medial nucleus; SI = primary somatosensory cortex. Evoked responses initiated in the periphery take, on average, 8–10 ms to propagate to barrel cortex. b, Raw voltage trace recorded during a single trial. A sharp transient in LFP traces from single trials is apparent 8 ms after the onset of stimulation, indicating the arrival of excitatory EPSPs from the thalamus. c, Average peri-stimulus spike histogram for N = 48 RS and FS cells, showing the 8 ms delay between deflecting the vibrissae (T = 0 ms) and the onset of spike activity in barrel cortex.

Supplementary Figure 6 Comparing d′ and hit rate.

a, Average d′ as a function of the temporal offset between sensory stimulus and entrained gamma for “threshold” trials with periodic stimulus presentation (same as Figure 6a in the main text). b, Average d′ as a function of temporal offset between sensory stimulus and entrained gamma for “max” trials (same as Figure 6b in the main text). c, Average hit rate as a function of temporal offset between sensory stimulus and entrained gamma for “threshold” trials. d, Average hit rate as a function of temporal offset between sensory stimulus and entrained gamma for “max” trials. All panels are for the same N = 8 animals as in Figure 6. Error bars indicate s.e.m.

Supplementary Figure 7 Electrophysiological methods.

a, Diagram of the fiber-optic–electrode implant used to record single unit activity in barrel cortex. b, Example projection plots for waveforms recorded on the same electrode over the course of a month. Spikes from the same cell are color-coded across days. c, Assertions of cell identity, used to minimize chances of double-counting the same cells across days, were analyzed with a waveform similarity metric. Orange dots indicate distance between waveforms identified as coming from the same cell; blue dots indicate distance between waveforms identified as coming from different cells. Distance from the origin indicates degree of waveform similarity. d, Measures used to identify cells as FS (fast-spiking) or RS (regular spiking). Cells in the present sample were separated into two groups based on peak–trough ratio and peak–trough separation.

Supplementary Figure 8 Additional spike quantification.

a, Mean PSTH of multi-unit activity from 15 electrodes during the peri-stimulus period (black lines = baseline condition). The condition with a 12.5 ms temporal offset (Figure 6a) is visibly more entrained to the 40 Hz stimulus. b, Wavelet transform of the mean PSTH for each temporal offset (colors map to plots in a). c, Power at 40 Hz for each of 5 temporal offsets, taken from the plot in panel b. d, Mean rate and peak rate for 0–25 ms and 30–150 ms, relative to baseline (N = 35 well-isolated RS cells).

Supplementary Figure 9 FS-gamma enhances perception when peripheral sensory drive arrives in neocortex during a beneficial window 20–25 milliseconds after FS synchronization.

Prior computational and experimental studies have predicted that FS-gamma may create an optimal window 20-25 milliseconds after FS synchronization, benefitting sensory relay when inputs arrive during this period. The conceptual diagram shows the timing of potential spikes generated by high-velocity vibrissae micro-motions (filled triangles) relative to the laser pulses that define the FS-gamma cycle (blue rectangles). In the naturalistic stimulation condition (top), stochastic vibrissal deflections at 180–200 Hz generated many high velocity micro-motions within a 25 ms cycle. As shown in Figure 4, these events had no systematic relationship to the FS-gamma cycle, and afferent spikes generated by these stimuli should arrive during the beneficial window. In the periodic stimulation condition (bottom), only one temporal offset (green, 12.5 ms) was predicted to generate stimulus-driven spikes inside the beneficial window, given the predicted lag of 8–10 milliseconds between peripheral drive and arrival in neocortex.

Supplementary Figure 10 Proposed schema of gamma’s effects on evoked spiking in pyramidal cells.

In this example, high-velocity deflections arrive while the cortex is in one of two “states,” one in which low synchrony among FS leads to generalized inhibition of firing and desynchronized firing in pyramidal cells (left), and one in which high FS synchrony creates a localized gamma oscillation (right). In the former state, disorganized but high rates of FS activity create a persistent state of inhibitory tone, diminishing the probability of spiking. In contrast, FS synchronization creates a greater transient depth of inhibition, but recovery from this hyperpolarization creates a window in which a host of excitatory currents are more readily recruited. This relief from synchronized inhibition concentrates firing of local neurons to this period. Motion patterns (bottom) measured from an ex vivo vibrissa; FS spike times and pyramidal cell membrane potential are simulated.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Siegle, J., Pritchett, D. & Moore, C. Gamma-range synchronization of fast-spiking interneurons can enhance detection of tactile stimuli. Nat Neurosci 17, 1371–1379 (2014). https://doi.org/10.1038/nn.3797

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3797

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