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:

Breakdown of spatial coding and interneuron synchronization in epileptic mice

Abstract

Temporal lobe epilepsy causes severe cognitive deficits, but the circuit mechanisms remain unknown. Interneuron death and reorganization during epileptogenesis may disrupt the synchrony of hippocampal inhibition. To test this, we simultaneously recorded from the CA1 and dentate gyrus in pilocarpine-treated epileptic mice with silicon probes during head-fixed virtual navigation. We found desynchronized interneuron firing between the CA1 and dentate gyrus in epileptic mice. Since hippocampal interneurons control information processing, we tested whether CA1 spatial coding was altered in this desynchronized circuit, using a novel wire-free miniscope. We found that CA1 place cells in epileptic mice were unstable and completely remapped across a week. This spatial instability emerged around 6 weeks after status epilepticus, well after the onset of chronic seizures and interneuron death. Finally, CA1 network modeling showed that desynchronized inputs can impair the precision and stability of CA1 place cells. Together, these results demonstrate that temporally precise intrahippocampal communication is critical for spatial processing.

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. 1: Desynchronization of hippocampal inhibition in epileptic mice.
Fig. 2: Disrupted spatial coding in epileptic mice.
Fig. 3: Disrupted stability of place cells across sessions.
Fig. 4: Spatial instability emerges 6 weeks after pilocarpine treatment.
Fig. 5: Desynchronization disrupts place cell coding in a CA1 network model.

Similar content being viewed by others

Data availability

The experimental data that support the findings of this study are available from Peyman Golshani (pgolshani@mednet.ucla.edu) or Tristan Shuman (tristan.shuman@mssm.edu) upon reasonable request.

Code availability

The software and codes related to the CA1 network model and its analysis are available from the Poirazi lab (poirazi@imbb.forth.gr) on reasonable request. The model is available on ModelDB, accession number 256311. Data analysis scripts are available on reasonable request from Peyman Golshani (pgolshani@mednet.ucla.edu) or Tristan Shuman (tristan.shuman@mssm.edu).

References

  1. Bell, B., Lin, J. J., Seidenberg, M. & Hermann, B. The neurobiology of cognitive disorders in temporal lobe epilepsy. Nat. Rev. Neurol. 7, 154–164 (2011).

    Article  PubMed  Google Scholar 

  2. Kleen, J. K., Scott, R. C., Lenck-Santini, P. P. & Holmes, G. L. in Jasper’s Basic Mechanisms of the Epilepsies (eds Noebels, J. L., Avoli, M., Rogawski, M. A., Olsen, R. W. & Delgado-Escueda, A. V.) 915–929 (Oxford Univ. Press, 2012).

  3. Henshall, D. C. & Meldrum, B. S. in Jasper’s Basic Mechanisms of the Epilepsies Noebels, J. L., Avoli, M., Rogawski, M. A., Olsen, R. W. & Delgado-Escueda, A. V.) 362–376 (Oxford Univ. Press, 2012).

  4. Sloviter, R. S. Decreased hippocampal inhibition and a selective loss of interneurons in experimental epilepsy. Science 235, 73–76 (1987).

    Article  CAS  PubMed  Google Scholar 

  5. Houser, C. R. et al. Altered patterns of dynorphin immunoreactivity suggest mossy fiber reorganization in human hippocampal epilepsy. J. Neurosci. 10, 267–282 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Zhang, W. et al. Surviving hilar somatostatin interneurons enlarge, sprout axons, and form new synapses with granule cells in a mouse model of temporal lobe epilepsy. J. Neurosci. 29, 14247–14256 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Peng, Z. et al. A reorganized GABAergic circuit in a model of epilepsy: evidence from optogenetic labeling and stimulation of somatostatin interneurons. J. Neurosci. 33, 14392–14405 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kobayashi, M. & Buckmaster, P. S. Reduced inhibition of dentate granule cells in a model of temporal lobe epilepsy. J. Neurosci. 23, 2440–2452 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Dinocourt, C., Petanjek, Z., Freund, T. F., Ben-Ari, Y. & Esclapez, M. Loss of interneurons innervating pyramidal cell dendrites and axon initial segments in the CA1 region of the hippocampus following pilocarpine-induced seizures. J. Comp. Neurol. 459, 407–425 (2003).

    Article  PubMed  Google Scholar 

  10. Cossart, R. et al. Dendritic but not somatic GABAergic inhibition is decreased in experimental epilepsy. Nat. Neurosci.4, 52–62 (2001).

    Article  CAS  PubMed  Google Scholar 

  11. Hirsch, J. C. et al. Deficit of quantal release of GABA in experimental models of temporal lobe epilepsy. Nat. Neurosci. 2, 499–500 (1999).

    Article  CAS  PubMed  Google Scholar 

  12. Goldberg, E. M. & Coulter, D. A. Mechanisms of epileptogenesis: a convergence on neural circuit dysfunction. Nat. Rev. Neurosci. 14, 337–349 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Christenson Wick, Z., Leintz, C. H., Xamonthiene, C., Huang, B. H. & Krook-Magnuson, E. Axonal sprouting in commissurally projecting parvalbumin-expressing interneurons. J. Neurosci. 95, 2336–2344 (2017).

    CAS  Google Scholar 

  14. 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  CAS  PubMed  PubMed Central  Google Scholar 

  15. Royer, S. et al. Control of timing, rate and bursts of hippocampal place cells by dendritic and somatic inhibition. Nat. Neurosci. 15, 769–775 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  17. van Dijk, M. T. & Fenton, A. A. On how the dentate gyrus contributes to memory discrimination. Neuron 98, 832–845 (2018). e835.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Chauviere, L. et al. Early deficits in spatial memory and theta rhythm in experimental temporal lobe epilepsy. J. Neurosci. 29, 5402–5410 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Inostroza, M., Brotons-Mas, J. R., Laurent, F., Cid, E. & de la Prida, L. M. Specific impairment of “what-where-when” episodic-like memory in experimental models of temporal lobe epilepsy. J. Neurosci. 33, 17749–17762 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Shuman, T., Amendolara, B. & Golshani, P. Theta rhythmopathy as a cause of cognitive disability in TLE. Epilepsy Curr. 17, 107–111 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Lopez-Pigozzi, D. et al. Altered oscillatory dynamics of CA1 parvalbumin basket cells during theta-gamma rhythmopathies of temporal lobe epilepsy. eNeuro 3, (2016).

  22. Liu, X. et al. Seizure-induced changes in place cell physiology: relationship to spatial memory. J. Neurosci. 23, 11505–11515 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Lenck-Santini, P. P. & Holmes, G. L. Altered phase precession and compression of temporal sequences by place cells in epileptic rats. J. Neurosci. 28, 5053–5062 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Lenck-Santini, P. P. & Scott, R. C. Mechanisms responsible for cognitive impairment in epilepsy. Cold Spring Harb. Perspect. Med. 5, pii: a022772 (2015).

    Article  CAS  Google Scholar 

  25. Zhou, J. L., Shatskikh, T. N., Liu, X. & Holmes, G. L. Impaired single cell firing and long-term potentiation parallels memory impairment following recurrent seizures. Eur. J. Neurosci. 25, 3667–3677 (2007).

    Article  PubMed  Google Scholar 

  26. Cai, D. J. et al. A shared neural ensemble links distinct contextual memories encoded close in time. Nature 534, 115–118 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Ghosh, K. K. et al. Miniaturized integration of a fluorescence microscope. Nat. Methods 8, 871–878 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Liberti, W. A., Perkins, L. N., Leman, D. P. & Gardner, T. J. An open source, wireless capable miniature microscope system. J. Neural Eng. 14, 045001 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Barbera, G. et al. Spatially Compact Neural Clusters in the Dorsal Striatum Encode Locomotion Relevant Information. Neuron 92, 202–213 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Curia, G., Longo, D., Biagini, G., Jones, R. S. & Avoli, M. The pilocarpine model of temporal lobe epilepsy. J. Neurosci. Methods 172, 143–157 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. von Stein, A. & Sarnthein, J. Different frequencies for different scales of cortical integration: from local gamma to long range alpha/theta synchronization. Int. J. Psychophysiol. 38, 301–313 (2000).

    Article  Google Scholar 

  32. Turi, G. F. et al. Vasoactive intestinal polypeptide-expressing interneurons in the hippocampus support goal-oriented spatial learning. Neuron 101, 1150–1165 (2019). e1158.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Cichon, J. & Gan, W. B. Branch-specific dendritic Ca2+ spikes cause persistent synaptic plasticity. Nature 520, 180–185 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kelley, M. S., Jacobs, M. P., Lowenstein, D. H. & Stewards, N. E. B. The NINDS epilepsy research benchmarks. Epilepsia 50, 579–582 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Magnus, C. J. et al. Ultrapotent chemogenetics for research and potential clinical applications. Science 364, eaav5282 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Grienberger, C., Milstein, A. D., Bittner, K. C., Romani, S. & Magee, J. C. Inhibitory suppression of heterogeneously tuned excitation enhances spatial coding in CA1 place cells. Nat. Neurosci. 20, 417–426 (2017).

    Article  CAS  PubMed  Google Scholar 

  37. Grasse, D. W., Karunakaran, S. & Moxon, K. A. Neuronal synchrony and the transition to spontaneous seizures. Exp. Neurol. 248, 72–84 (2013).

    Article  PubMed  Google Scholar 

  38. Miri, M. L., Vinck, M., Pant, R. & Cardin, J. A. Altered hippocampal interneuron activity precedes ictal onset. eLife 7, e40750 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Mably, A. J., Gereke, B. J., Jones, D. T. & Colgin, L. L. Impairments in spatial representations and rhythmic coordination of place cells in the 3xTg mouse model of Alzheimer’s disease. Hippocampus 27, 378–392 (2017).

    Article  CAS  PubMed  Google Scholar 

  40. Cacucci, F., Yi, M., Wills, T. J., Chapman, P. & O’Keefe, J. Place cell firing correlates with memory deficits and amyloid plaque burden in Tg2576 Alzheimer mouse model. Proc. Natl Acad. Sci. USA 105, 7863–7868 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zaremba, J. D. et al. Impaired hippocampal place cell dynamics in a mouse model of the 22q11.2 deletion. Nat. Neurosci. 20, 1612–1623 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Rosenzweig, E. S. & Barnes, C. A. Impact of aging on hippocampal function: plasticity, network dynamics, and cognition. Prog. Neurobiol. 69, 143–179 (2003).

    Article  CAS  PubMed  Google Scholar 

  43. Raveau, M. et al. Alterations of in vivo CA1 network activity in Dp(16)1Yey Down syndrome model mice. eLife 7, e31543 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Buckmaster, P. S. & Haney, M. M. Factors affecting outcomes of pilocarpine treatment in a mouse model of temporal lobe epilepsy. Epilepsy Res. 102, 153–159 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Morris, R. Developments of a water-maze procedure for studying spatial learning in the rat. J. Neurosci. Methods 11, 47–60 (1984).

    Article  CAS  PubMed  Google Scholar 

  46. Sigurdsson, T., Stark, K. L., Karayiorgou, M., Gogos, J. A. & Gordon, J. A. Impaired hippocampal-prefrontal synchrony in a genetic mouse model of schizophrenia. Nature 464, 763–767 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Guo, Z. V. et al. Procedures for behavioral experiments in head-fixed mice. PLoS One 9, e88678 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. 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  CAS  PubMed  PubMed Central  Google Scholar 

  49. 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  PubMed  PubMed Central  Google Scholar 

  50. Bokil, H., Andrews, P., Kulkarni, J. E., Mehta, S. & Mitra, P. P. Chronux: a platform for analyzing neural signals. J. Neurosci. Methods 192, 146–151 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Lubenov, E. V. & Siapas, A. G. Hippocampal theta oscillations are travelling waves. Nature 459, 534–539 (2009).

    Article  CAS  PubMed  Google Scholar 

  52. Schomburg, E. W. et al. Theta phase segregation of input-specific gamma patterns in entorhinal-hippocampal networks. Neuron 84, 470–485 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Karlsson, K. A. & Blumberg, M. S. Temperature-induced reciprocal activation of hippocampal field activity. J. Neurophysiol. 91, 583–588 (2004).

    Article  PubMed  Google Scholar 

  54. Senzai, Y. & Buzsaki, G. Physiological properties and behavioral correlates of hippocampal granule cells and mossy cells. Neuron 93, 691–704 (2017). e695.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Willers, B. Multimodal sensory contributions to hippocampal spatiotemporal selectivity. Doctoral Dissertation, University of California Los Angeles. (2013).

  56. Csicsvari, J., Hirase, H., Czurko, A., Mamiya, A. & Buzsaki, G. Oscillatory coupling of hippocampal pyramidal cells and interneurons in the behaving Rat. J. Neurosci. 19, 274–287 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Berens, P. CircStat: a MATLAB toolbox for circular statistics. J. Stat. Softw. https://doi.org/10.18637/jss.v031.i10 (2009).

  58. Pennington, Z. T., et al. ezTrack: an open-source video analysis pipeline for the investigation of animal behavior. Preprint at bioRxiv https://www.biorxiv.org/content/10.1101/592592v1 (2019).

  59. Zhou, P. et al. Efficient and accurate extraction of in vivo calcium signals from microendoscopic video data. eLife 7, e28728 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Vogelstein, J. T. et al. Fast nonnegative deconvolution for spike train inference from population calcium imaging. J. Neurophysiol. 104, 3691–3704 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Ravassard, P. et al. Multisensory control of hippocampal spatiotemporal selectivity. Science 340, 1342–1346 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Mau, W. et al. The Same Hippocampal CA1 Population Simultaneously Codes Temporal Information over Multiple Timescales. Curr. Biol. 28, 1499–1508.e4 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Bezaire, M. J., Raikov, I., Burk, K., Vyas, D. & Soltesz, I. Interneuronal mechanisms of hippocampal theta oscillations in a full-scale model of the rodent CA1 circuit. eLife 5, e18566 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Cutsuridis, V., Cobb, S. & Graham, B. P. Encoding and retrieval in a model of the hippocampal CA1 microcircuit. Hippocampus 20, 423–446 (2010).

    CAS  PubMed  Google Scholar 

  65. Cutsuridis, V. & Poirazi, P. A computational study on how theta modulated inhibition can account for the long temporal windows in the entorhinal-hippocampal loop. Neurobiol. Learn. Mem. 120, 69–83 (2015).

    Article  PubMed  Google Scholar 

  66. Konstantoudaki, X., Papoutsi, A., Chalkiadaki, K., Poirazi, P. & Sidiropoulou, K. Modulatory effects of inhibition on persistent activity in a cortical microcircuit model. Front. Neural Circuits 8, 7 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  67. Danielson, N. B. et al. Sublayer-specific coding dynamics during spatial navigation and learning in hippocampal area CA1. Neuron 91, 652–665 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Solstad, T., Moser, E. I. & Einevoll, G. T. From grid cells to place cells: a mathematical model. Hippocampus 16, 1026–1031 (2006).

    Article  PubMed  Google Scholar 

  69. Leutgeb, S., Leutgeb, J. K., Treves, A., Moser, M. B. & Moser, E. I. Distinct ensemble codes in hippocampal areas CA3 and CA1. Science 305, 1295–1298 (2004).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  71. 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  PubMed  PubMed Central  Google Scholar 

  72. Smith, B. N. & Dudek, F. E. Short- and long-term changes in CA1 network excitability after kainate treatment in rats. J. Neurophysiol. 85, 1–9 (2001).

    Article  CAS  PubMed  Google Scholar 

  73. Hines, M. L. & Carnevale, N. T. The NEURON simulation environment. Neural Comput. 9, 1179–1209 (1997).

    Article  CAS  PubMed  Google Scholar 

  74. Ziv, Y. et al. Long-term dynamics of CA1 hippocampal place codes. Nat. Neurosci. 16, 264–266 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Ewell, L. A., Fischer, K. B., Leibold, C., Leutgeb, S. & Leutgeb, J. K. The impact of pathological high-frequency oscillations on hippocampal network activity in rats with chronic epilepsy. eLife 8, e42148 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank K. Maguire, J. Lou, A. Fariborzi, J. Daneshrad, S. Ghiaee, R. Manavi, C. Araradian, M. Song, B. Wei, C. Zhou, A. Meyer, H. Chen, J. Davis, N. Abduljawad, J. Hodson, I. Bachmutsky, L. Zilbermintz, H. Karbasforoushan, J. Friedman, T. Kotze, D. McCoy, K. Casale and E. Goldblatt (all of UCLA); and N. Berryman, G. Condori, M. Abdelmageed, C. Rosado, and B. Nunez (all of Mount Sinai) for excellent technical assistance and help with experiments. This work was supported by VA Merit Award 1 I01 BX001524–01A1, U01 NS094286, R01MH101198, R01 MH105427, U54 HD87101, R01NS099137, and NSF Neurotech Hub 1700408 to P.G.; a David Geffen School of Medicine Dean’s Fund for development of open-source miniaturized microscopes to B.S.K., A.J.S., and P.G.; a CURE Epilepsy Taking Flight Award, an American Epilepsy Society Junior Investigator Award, R03 NS111493, a Cellular Neurobiology Training Grant T32 NS710133, and an Epilepsy Foundation Postdoctoral Research Training Fellowship to T.S.; Neurobehavioral Genetics Training Grant T32 NS048004 and Neural Microcircuits Training Grant T32 NS058280 to D.A.; DP2 MH122399, a Klingenstein-Simons Fellowship, a McKnight Memory and Cognitive Disorder Award, a NARSAD Young Investigator Award, a Fay/Frank Seed Grant Program award, a One Mind Rising Star Research Award, National Research Service Award F32 MH97413, and Behavioral Neuroscience Training Grant T32 MH15795 to D.J.C.; DP1 MH104069 to B.S.K.; a McKnight Technological Innovations in Neuroscience Award to S.C.M.; and Dr. Miriam and Sheldon G. Adelson Medical Research Foundation funding to A.J.S. S.C., I.P. and P.P. were supported by the European Research Council Starting Grant dEMORY (GA 311435) and the Fondation Sante.

Author information

Authors and Affiliations

Authors

Contributions

T.S. and P.G. designed the calcium imaging and electrophysiology experiments. T.S., D.A., D.J.C., B.S.K., A.J.S., and P.G. developed the wire-free miniscope. D.A. designed and built the wire-free Miniscope. T.S., D.A., D.J.C., C.R.L. tested the wire-free miniscope. T.S., D.J.C., C.R.L., L.P.-H., L.M.V., Y.F., C.Y., I.M.-G., and M.L.-V. performed calcium imaging experiments. T.S., D.A., L.P.-H., and Z.T.P. analyzed calcium imaging experiments. T.S., M.J., C.C.K., M.S., and P.G. designed the virtual reality apparatus. T.S., S.E.F., K.C., M.J., C.C.K., and N.R. performed electrophysiology training and recording. T.S., K.I.B., S.C.M., P.G. built the in vivo electrophysiology system. T.S. and J.T. analyzed in vivo electrophysiology data. T.S., K.C., and C.C.K. performed spike sorting of single units. T.S., D.A., D.J.C., S.C., I.P., P.P., and P.G. designed modeling experiments. T.S., S.C., I.P., and P.P. performed and analyzed modeling experiments. T.S., D.J.C., and L.C. performed slice electrophysiology. T.S., L.P.-H., L.M.V., and Y.F. performed immunohistochemistry. T.S., D.A., D.J.C., S.C., P.P., and P.G. wrote the manuscript and all authors edited the manuscript.

Corresponding authors

Correspondence to Tristan Shuman, Panayiota Poirazi or Peyman Golshani.

Ethics declarations

Competing interests

The 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

Extended Data Fig. 1 Spatial memory deficits in epileptic mice.

a. Morris water maze task. Mice (at least 6 weeks after pilocarpine or control treatment) were trained to find a hidden platform for 6 days at random start locations. The next day, mice were given a probe trial to assess learning. On the probe trial, control mice spent more time in the target quadrant than epileptic mice, while epileptic mice spent more time in the opposite quadrant (n = 6 Control, n = 5 Epileptic, 2-way ANOVA FGroupXQuadrant(3,36) = 14.6, P < 0.001, Training Quadrant Post-hoc: P < 0.001, Opposite Quadrant Post-hoc: P = 0.02). b. Mice were trained to find a visible platform with a flag extending above the water. There were no differences between the control and epileptic mice on this task (n = 5 Control, n = 8 Epileptic, 2-way ANOVA FGroup(1,11) = 0.424, P = 0.53). c. Mice were trained on a cued, delayed alternation T-maze task for 5 days. On each trial, animals were cued to one direction (Cue Phase), returned to the start position for a 15 s delay (Delay Phase), and then released for the Test Phase. If the animal went to the opposite side it received a water reward. Epileptic mice performed worse than control mice (n = 5 Control, n = 6 Epileptic, 2-way ANOVA FGroup(1,45) = 44.98, P < 0.001). df. Activity during virtual reality recordings. No differences were found in trials per minute (n = 5 per group, Unpaired t-test, P = 0.16), percent time running (n = 5 animals per group, Unpaired t-test, P = 0.14), or running speed (n = 5 animals per group, Mann-Whitney test, P = 0.94). Error bars represent 1 S.E.M. *P < 0.05. g-i. Activity during calcium imaging on the linear track. No differences were found in the number of trials per minute (n = 5 animals per group, Mann-Whitney test, P = 0.15) or percent time running (n = 5 animals per group, Unpaired t-test, P = 0.64). Epileptic mice ran faster than control mice on the linear track (n = 5 animals per group, Unpaired t-test, P = 0.03). Error bars represent 1 S.E.M. *P < 0.05, ***P < 0.001.

Extended Data Fig. 2 Local field potential power across frequencies.

af. LFP power in beta, slow gamma, mid-gamma, fast gamma, ripple and fast ripple frequencies throughout CA1 and DG. No differences were found in beta (FGroupXRegion(6,48) = 1.23, P = 0.31, FGroup(1,8) = 0.47, P = 0.51), fast gamma (FGroupXRegion(6,48) = 1.88, P = 0.10, FGroup(1,8) = 2.03, P = 0.19), ripple (FGroupXRegion(6,48) = 0.93, P = 0.48, FGroup(1,8) = 0.67, P = 0.44), or fast ripple (FGroupXRegion(6,48) = 1.38, P = 0.24, FGroup(1,8) = 0.24, P = 0.64) power. Epileptic mice had reduced slow gamma and mid-gamma power in the hilus of DG (Slow Gamma: FGroupXRegion(6,48) = 6.76, P < 0.001, FGroup(1,8) = 7.31, P = 0.03, Hilus: P < 0.001; Mid-gamma: FGroupXRegion(6,48) = 4.49, P = 0.001, FGroup(1,8) = 3.48, P = 0.09, Hilus: P = 0.01). n = 5 animals per group for all graphs. Error bars represent 1 S.E.M. **P < 0.01, ***P < 0.001.

Extended Data Fig. 3 Theta phase locking changes are not caused by decreased power or specific to reference location.

a. Example clustering of single units from a single channel set in one animal using principal components, peak amplitude, and trough amplitude. b. Units were characterized as putative interneurons based on complex spike index (CSI), mean autocorrelogram (Mean AC), and mean firing rate. N = 5 Control, N = 5 Epileptic animals. c. Example phase locking in a CA1 interneuron. Top, mean waveforms from each recorded channel within the pyramidal layer. Middle, spike raster of 300 theta cycles during running, with the proportion of spikes occurring during each phase of theta. Bottom, Rose plot of firing phase relative to theta. This cell has a modulation index (r-value) of 0.41 and preferred firing phase (mu) of 189°. dg. Theta cycles were subsampled to match power between control and epileptic mice. This did not change the results as CA1 interneurons in epileptic mice were less phase locked than in control mice (d; n = 71 Control cells, n = 34 Epileptic cells, Unpaired t-test, P < 0.001), while their preferred phase was not different (e; Kuiper circular test, P > 0.05). There were no differences in the modulation of DG interneurons (f; n = 34 Control cells, n = 22 Epileptic cells, Unpaired t-test, P = 0.80) but there were differences in the preferred phase of DG interneurons (g; Kuiper circular test, P < 0.001). h, i. Phase locking of DG interneurons to theta in the lacunosum moleculare. No difference was found in modulation (h; n = 34 Control cells, n = 41 Epileptic cells, Unpaired t-test, P = 0.25) but the distributions of preferred phases differed between the groups (i; Kuiper circular test, P < 0.001). j-k. Phase locking of DG interneurons to the theta in hilus of DG. No difference was found in modulation (j; n = 34 Control cells, n = 41 Epileptic cells, Unpaired t-test, P = 0.64) but the distributions of preferred phases differed between the groups (k; Kuiper circular test, P < 0.001). N = 5 animals per group for all panels. Error bars represent 1 S.E.M. ***P < 0.001.

Extended Data Fig. 4 Altered gamma synchronization in epileptic mice.

a. Gamma power throughout CA1 and DG. Epileptic mice had reduced gamma power in the hilus of DG (2-way RM ANOVA, FGroupXRegion(6,48) = 5.26, P < 0.001, Hilus: P < 0.001). b. Gamma coherence between each channel pair along the recording probe in control (left) and epileptic (middle) mice. Right, p-value matrix for each pair of recording sites (unpaired t-tests, uncorrected for multiple comparisons). c. Phase locking to CA1 gamma for each interneuron in CA1 of control and epileptic mice. Each dot represents one interneuron and the data is double plotted for visualization. Closed circles were significantly phase locked (Raleigh test, P < 0.05) and open circles were not. d. Phase locking to CA1 gamma for each interneuron in DG of control and epileptic mice. e. Mean phase modulation (r) of CA1 interneurons to CA1 gamma. Interneurons in epileptic mice had increased phase modulation (n = 69 Control cells, n = 57 Epileptic cells, Mann-Whitney test, P < 0.001). f. Mean phase modulation (r) of DG interneurons to CA1 gamma. Interneurons in epileptic mice had no change in phase modulation (n = 34 Control cells, n = 41 Epileptic cells, Mann-Whitney test, P = 0.42). g. Rose plot of preferred firing phases of significantly phase locked CA1 interneurons. There were no differences between interneurons in control and epileptic mice (Kuiper circular test, P > 0.05). h. Rose plot of preferred firing phases of significantly phase locked DG interneurons. The distribution of preferred phases in control and epileptic mice were different (Kuiper circular test, P < 0.001). i. Aggregate inhibition of all interneurons in CA1 and DG relative to CA1 gamma phase (2-way RM ANOVA, CA1: FGroupXPhase(17,136) = 1.22, P = 0.25; DG: FGroupXPhase(17,136) = 2.196, P = 0.007). j. Pearson’s correlation between CA1 and DG aggregate inhibition relative to gamma. No difference was found between epileptic and control animals (Unpaired t-test, P = 0.19). All data in this figure came from 5 animals per group. Error bars represent 1 S.E.M. *P < 0.05, ***P < 0.001, ns: not significant.

Extended Data Fig. 5 Wire-free Miniscope design and behavior.

a. Schematic of wire-free Miniscope printed circuit board (PCB). b. Picture of top and bottom of wire-free sensor board. This camera sensor board attaches to the Miniscope body, which contains all optic components. c. The wire-free Miniscope can be used to probe naturalistic behaviors including social interaction. We tested social behavior with the wire-free Miniscope compared to no Miniscope or a wired version. d. During a social interaction test mice with either no Miniscope or the wire-free Miniscope chose a social cup over an empty cup, while mice with a wired Miniscope showed no significant preference (n = 3 per group, Unpaired t-tests, No Scope: P = 0.03, Wire-Free: P < 0.001, Wired: P = 0.10). Error bars represent 1 S.E.M. *P < 0.05, ***P < 0.001.

Extended Data Fig. 6 Stability of spatial representations.

a. Population Vector Correlation (PVC) of all cells in Control (left) and Epileptic (right) mice across all sessions recorded. b. Mean PVC as a function of offset distance (from the diagonal) across animals. Epileptic mice had higher PVC (that is, less distinct firing patterns) across distances of 30–150 cm (2-way RM ANOVA, FGroup(1,8) = 10.39, P = 0.01, post hoc P < 0.05 for each bin from 30–150 cm). c. Population Vector Correlation (PVC) of place cells in Control (left) and Epileptic (right) mice across all sessions recorded. d. Mean PVC as a function of offset distance (from the diagonal) across animals. Epileptic mice had higher PVC (that is, less distinct firing patterns) across distances of 32–150 cm (2-way RM ANOVA, FGroup(1,8) = 10.19, P = 0.01, post hoc P < 0.05 for each bin from 32–150 cm). N = 5 animals per group for all panels. Shading represents 1 S.E.M. *P < 0.05. e, f. Example cross registration in a Control (e) and Epileptic (f) mouse. Top, Aligned mean frame from each session (~ 550 um x 550 um). Bottom, overlaid cells from each session with white X indicating matched cells. g. Spatial correlation and centroid distance were calculated for all cell pairs. Dotted lines indicate thresholds used as matching criteria. All matched cells had spatial correlation ≥ 0.6 and centroid distance ≤ 4 pixels (~ 7 um). h. We assessed within-session stability of spatial representations in two ways. We first examined stability of the first half of trials against the second half of trials. We then examined stability of odd trials versus even trials. In both cases, we used the Fisher z-transformed correlation of the spatial firing rates between trials. We report stability as the average of the two different stability measures. i. We found a slight but significant difference between the two stability measures as the stability of odd/even trials was higher than in the first/second half of trials (2-way RM ANOVA FMeasurement(2,16) = 13.59, P < 0.001; post hoc tests: ***P < 0.001). j. The two stability measures were highly correlated in both control (Pearson’s correlation: All Cells: r = 0.83; Place Cells: r = 0.75) and epileptic (Pearson’s correlation: All Cells: r = 0.73; Place Cells: r = 0.72) neurons. N = 5 animals per group for all panels.

Extended Data Fig. 7 Example processing of one training session neuron for Bayesian decoding.

a. For each frame, temporal neural activity is calculated and classified into rightward trials, leftward trials, or non-trial times. b. Temporal neural activity is binarized and non-trial times are removed. c. The per trial binarized neural activity rate is calculated for rightward and leftward trials. d. The spatial probability function is constructed for each cell. A Gaussian distribution is first generated for each spatial bin using the mean and standard deviation of the binarized neural activity rate. The overall distribution is normalized across binarized neural activity rate rows and this data is entered into the Bayesian decoder for each cell. See Supplementary Video 3 for example decoding across all cells.

Extended Data Fig. 8 CA1 network model reveals that high levels of interneuron disruption can reduce information content, but not stability.

a. We tested how increasing amounts of desynchronization altered information content, stability, and place cell percent in our CA1 network model. We found significant reductions in all metrics with desynchronization of 15–35 ms (1-way ANOVA, Information Content: F(7,72) = 77.84, P < 0.001; Stability: F(7,72) = 70.02, P < 0.001; Place cell percent: F(7,72) = 99.44, P < 0.001; post hoc tests: *P < 0.05, **P < 0.01, ***P < 0.001). b. We next tested how reducing somatostatin-expressing neurons altered spatial coding in our model. We found that reducing SOM by 50–100% reduced information content (1-way ANOVA, F(4,45) = 38.01, P < 0.001, post hoc tests: ***P < 0.001) but not stability (1-way ANOVA, F(4,45) = 0.564, P = 0.69) or place cell percent (1-way ANOVA, F(4,45) = 2.369, P = 0.067). c. We next tested how reducing parvalbumin-expressing neurons altered spatial coding in our model. We found that reducing PV by 50–100% reduced information content (1-way ANOVA, F(4,45) = 134, P < 0.001, post hoc tests: **P < 0.01, ***P < 0.001) but also increased stability at 75–100% PV reduction (1-way ANOVA, F(4,45) = 18.26, P < 0.001, post hoc tests: ***P < 0.001). We also found that place cell percent was lower at 75–100% PV reduction (1-way ANOVA, F(4,45) = 41.78, P < 0.001, post hoc tests: ***P < 0.001).

Extended Data Fig. 9 Cell death and gliosis in epileptic mice.

a, b. Example immunohistochemistry staining for parvalbumin (PV), somatostatin (SOM), NeuN, and glial fibrillary acidic protein (GFAP) in dentate gyrus (a) and CA1 (b) of control and epileptic mice. For epileptic mice, tissue was collected at least 19 weeks after pilocarpine. c. PV staining was reduced in epileptic mice (2-way RM ANOVA, FGroup(1,14) = 7.086, P = 0.02, post hoc for CA1, CA3, DG: P > 0.05). N = 9 Control, N = 7 Epileptic. d. SOM staining was reduced in the DG of epileptic mice (2-way RM ANOVA, FGroup x Region(2,16) = 12.12, P < 0.001, post hoc for DG: P < 0.001; CA1, CA3: P > 0.05). N = 5 Control, N = 5 Epileptic. e. NeuN staining was reduced in CA3 of epileptic mice (2-way RM ANOVA, FGroup(1,16) = 5.581, P = 0.03, post hoc for CA3: P = 0.002; CA1, DG Hilus: P > 0.05; DG Blade: Unpaired t-test, t = 0.302, P = 0.77). N = 9 Control, N = 9 Epileptic. f. Because GFAP expression can be altered in the soma or processes we analyzed both the cell counts and percent of activated pixels within each image. We did not detect a difference in the number of GFAP + neurons (2-way RM ANOVA, FGroup(1,8) = 4.60, P = 0.06), however we did find increased percent of activated pixels in CA1, CA3, and DG (2-way RM ANOVA, FGroup(1,8) = 14.15, P = 0.005, post hoc for all regions: P < 0.05). N = 5 Control, N = 5 Epileptic. gi. In entorhinal cortex and area V1 we found no differences between groups in PV (2-way RM ANOVA, FGroup(1,7) = 0.371, P = 0.56, N = 5 Control, N = 4 Epileptic), SOM (2-way RM ANOVA, FGroup(1,8) = 0.815, P = 0.39, N = 5 Control, N = 5 Epileptic), or NeuN (2-way RM ANOVA, FGroup(1,8) = 0.288, P = 0.61, N = 5 Control, N = 4 Epileptic). j. We did find an increased number of GFAP positive neurons in EC and V1 (2-way RM ANOVA, FGroup(1,7) = 432.1, P < 0.001; post hoc: EC, P < 0.001; V1, P < 0.01, N = 5 Control, N = 4 Epileptic), and increased percent of activated pixels for GFAP in EC (2-way RM ANOVA, FGroup(1,7) = 31.1, P < 0.001; post hoc: EC, P < 0.001; V1, P = 0.31, N = 5 Control, N = 4 Epileptic). EC, entorhinal cortex; DG, dentate gyrus; Error bars denote 1 S.E.M. All sample sizes are by animal. *P < 0.05, **P < 0.01, ***P < 0.001.

Extended Data Fig. 10 Chemogenetic inhibition of PV + or SOM + interneurons does not alter spatial coding in CA1.

a. To selectively inhibit PV + and SOM + interneurons, we used PV-Cre and SOM-Cre mice and injected a cre-dependent virus expressing hM4Di (AAV5-Syn-DIO-hM4Di-mCherry) or control virus (AAV5-Syn-DIO-mCherry). We also injected virus to express GCaMP6f in all neurons (AAV1-Syn-GCaMP6f) to allow for calcium imaging. We then trained mice to run on a linear track and delivered CNO (5 mg/kg) or Vehicle (VEH) 45 min prior to imaging on the track. b. To confirm that CNO was effective in reducing firing in vitro we used whole-cell recordings of hippocampal mCherry + interneurons in acute brain slices and applied ACSF or ACSF with CNO during stimulation. Interneurons expressing hM4Di had reduced spiking to stimulation (middle, right) while control virus did not reduce spiking. c. We found no differences in information content between VEH and CNO in any of the groups examined (Paired t-tests – PV: t(4) = 1.769, P = 0.15; SOM: t(3) = 0.717, P = 0.53). d. We found no differences in stability between VEH and CNO in any of the groups examined (Paired t-tests – PV: t(4) = 2.047, P = 0.11; SOM: t(3) = 0.672, P = 0.55). e. We found no differences in place cell percent between VEH and CNO in any of the groups examined (Paired t-tests – PV: t(4) = 2.29, P = 0.08; SOM: t(3) = 0.169, P = 0.88). f. We found increased Activity with CNO compared to VEH in the PV-hM4Di mice, but not SOM mice (Paired t-tests – PV: t(4) = 2.897, P = 0.04; SOM: t(3) = 0.96, P = 0.41). N = 2 PV-Cre mCherry animals, N = 4 SOM-Cre hM4Di animals, N = 5 PV-Cre hM4Di animals. Error bars denote 1 S.E.M. *P < 0.05.

Supplementary information

Supplementary Information

Supplementary Table 1.

Reporting Summary

Supplementary Video 1

Mouse Running in Virtual Linear Track. Example mouse running through virtual linear track for water rewards during recording. Animals are head-fixed atop a Styrofoam ball and a virtual linear track is projected around them. As the animal runs on the ball the environment moves forward until it reach the end and the animal receives a water reward. Simultaneously, in vivo electrophysiology with silicon probes measures local field potentials and interneuron firing in the hippocampus.

Supplementary Video 2

Example Calcium Imaging on a Linear Track. Example Control and Epileptic mice running on the linear track during calcium imaging. The top row shows the behavior of the mouse. The middle row is the raw video synchronized with the behavior. Bottom row is the neuronal activity extracted using CNMF-E. Place cells are colored according to preferred position along the track (see top row) and non-place cells are gray.

Supplementary Video 3

Decoding Examples. Example decoding performed across two session separated by 30 minutes. Top row shows the animal’s position (white) with the decoded position in red (leftward) or green (rightward). Bottom row shows the extracted neurons used in the decoder and the distribution of decoder position error.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shuman, T., Aharoni, D., Cai, D.J. et al. Breakdown of spatial coding and interneuron synchronization in epileptic mice. Nat Neurosci 23, 229–238 (2020). https://doi.org/10.1038/s41593-019-0559-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-019-0559-0

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