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:

Impaired hippocampal place cell dynamics in a mouse model of the 22q11.2 deletion

Abstract

Hippocampal place cells represent the cellular substrate of episodic memory. Place cell ensembles reorganize to support learning but must also maintain stable representations to facilitate memory recall. Despite extensive research, the learning-related role of place cell dynamics in health and disease remains elusive. Using chronic two-photon Ca2+ imaging in hippocampal area CA1 of wild-type and Df(16)A+/− mice, an animal model of 22q11.2 deletion syndrome, one of the most common genetic risk factors for cognitive dysfunction and schizophrenia, we found that goal-oriented learning in wild-type mice was supported by stable spatial maps and robust remapping of place fields toward the goal location. Df(16)A+/− mice showed a significant learning deficit accompanied by reduced spatial map stability and the absence of goal-directed place cell reorganization. These results expand our understanding of the hippocampal ensemble dynamics supporting cognitive flexibility and demonstrate their importance in a model of 22q11.2-associated cognitive dysfunction.

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: Differences in learning performance between Df(16)A+/− and WT mice in GOL task.
Figure 2: Altered place cell properties in Df(16)A+/− mice.
Figure 3: Disrupted stability of place cell population in Df(16)A+/− compared to WT mice.
Figure 4: Lack of context change and task-dependent stability of spatial maps in Df(16)A+/− mice.
Figure 5: Place field enrichment of goal location.
Figure 6: Session-to-session place field shift dynamics.
Figure 7: Place field drift toward reward drives enrichment in the place field dynamics model.
Figure 8: Df(16)A+/− mice place fields do not drift toward goal and the model produces no enrichment.

Similar content being viewed by others

References

  1. Eichenbaum, H. A cortical-hippocampal system for declarative memory. Nat. Rev. Neurosci. 1, 41–50 (2000).

    Article  CAS  PubMed  Google Scholar 

  2. Leavitt, V.M. & Goldberg, T.E. Episodic memory in schizophrenia. Neuropsychol. Rev. 19, 312–323 (2009).

    Article  PubMed  Google Scholar 

  3. Ranganath, C., Minzenberg, M.J. & Ragland, J.D. The cognitive neuroscience of memory function and dysfunction in schizophrenia. Biol. Psychiatry 64, 18–25 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Karayiorgou, M., Simon, T.J. & Gogos, J.A. 22q11.2 microdeletions: linking DNA structural variation to brain dysfunction and schizophrenia. Nat. Rev. Neurosci. 11, 402–416 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Buzsáki, G. & Moser, E.I. Memory, navigation and theta rhythm in the hippocampal-entorhinal system. Nat. Neurosci. 16, 130–138 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. O'Keefe, J. & Nadel, L. The Hippocampus as a Cognitive Map (Clarendon, 1978).

  7. Squire, L.R. Memory and the hippocampus: a synthesis from findings with rats, monkeys, and humans. Psychol. Rev. 99, 195–231 (1992).

    Article  CAS  PubMed  Google Scholar 

  8. Tamminga, C.A., Stan, A.D. & Wagner, A.D. The hippocampal formation in schizophrenia. Am. J. Psychiatry 167, 1178–1193 (2010).

    Article  PubMed  Google Scholar 

  9. Harrison, P.J. & Eastwood, S.L. Neuropathological studies of synaptic connectivity in the hippocampal formation in schizophrenia. Hippocampus 11, 508–519 (2001).

    Article  CAS  PubMed  Google Scholar 

  10. Bast, T. The hippocampal learning-behavior translation and the functional significance of hippocampal dysfunction in schizophrenia. Curr. Opin. Neurobiol. 21, 492–501 (2011).

    Article  CAS  PubMed  Google Scholar 

  11. Zierhut, K.C. et al. Hippocampal CA1 deformity is related to symptom severity and antipsychotic dosage in schizophrenia. Brain 136, 804–814 (2013).

    Article  PubMed  Google Scholar 

  12. O'Keefe, J. & Dostrovsky, J. The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Res. 34, 171–175 (1971).

    Article  CAS  PubMed  Google Scholar 

  13. Hartley, T., Lever, C., Burgess, N. & O'Keefe, J. Space in the brain: how the hippocampal formation supports spatial cognition. Phil. Trans. R. Soc. Lond. B 369, 20120510 (2013).

    Article  Google Scholar 

  14. Lever, C., Wills, T., Cacucci, F., Burgess, N. & O'Keefe, J. Long-term plasticity in hippocampal place-cell representation of environmental geometry. Nature 416, 90–94 (2002).

    Article  CAS  PubMed  Google Scholar 

  15. Kentros, C.G., Agnihotri, N.T., Streater, S., Hawkins, R.D. & Kandel, E.R. Increased attention to spatial context increases both place field stability and spatial memory. Neuron 42, 283–295 (2004).

    Article  CAS  PubMed  Google Scholar 

  16. Mankin, E.A. et al. Neuronal code for extended time in the hippocampus. Proc. Natl. Acad. Sci. USA 109, 19462–19467 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Thompson, L.T. & Best, P.J. Long-term stability of the place-field activity of single units recorded from the dorsal hippocampus of freely behaving rats. Brain Res. 509, 299–308 (1990).

    Article  CAS  PubMed  Google Scholar 

  18. 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 

  19. Breese, C.R., Hampson, R.E. & Deadwyler, S.A. Hippocampal place cells: stereotypy and plasticity. J. Neurosci. 9, 1097–1111 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Fyhn, M., Molden, S., Hollup, S., Moser, M.B. & Moser, E. Hippocampal neurons responding to first-time dislocation of a target object. Neuron 35, 555–566 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Gothard, K.M., Skaggs, W.E. & McNaughton, B.L. Dynamics of mismatch correction in the hippocampal ensemble code for space: interaction between path integration and environmental cues. J. Neurosci. 16, 8027–8040 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Hok, V. et al. Goal-related activity in hippocampal place cells. J. Neurosci. 27, 472–482 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Hollup, S.A., Kjelstrup, K.G., Hoff, J., Moser, M.B. & Moser, E.I. Impaired recognition of the goal location during spatial navigation in rats with hippocampal lesions. J. Neurosci. 21, 4505–4513 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Kobayashi, T., Nishijo, H., Fukuda, M., Bures, J. & Ono, T. Task-dependent representations in rat hippocampal place neurons. J. Neurophysiol. 78, 597–613 (1997).

    Article  CAS  PubMed  Google Scholar 

  25. Markus, E.J. et al. Interactions between location and task affect the spatial and directional firing of hippocampal neurons. J. Neurosci. 15, 7079–7094 1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Dupret, D., O'Neill, J., Pleydell-Bouverie, B. & Csicsvari, J. The reorganization and reactivation of hippocampal maps predict spatial memory performance. Nat. Neurosci. 13, 995–1002 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Etienne, A.S. & Jeffery, K.J. Path integration in mammals. Hippocampus 14, 180–192 (2004).

    Article  PubMed  Google Scholar 

  28. Buzsáki, G. Hippocampal sharp wave-ripple: a cognitive biomarker for episodic memory and planning. Hippocampus 25, 1073–1188 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Lee, I., Griffin, A.L., Zilli, E.A., Eichenbaum, H. & Hasselmo, M.E. Gradual translocation of spatial correlates of neuronal firing in the hippocampus toward prospective reward locations. Neuron 51, 639–650 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Mehta, M.R., Barnes, C.A. & McNaughton, B.L. Experience-dependent, asymmetric expansion of hippocampal place fields. Proc. Natl. Acad. Sci. USA 94, 8918–8921 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Monaco, J.D., Rao, G., Roth, E.D. & Knierim, J.J. Attentive scanning behavior drives one-trial potentiation of hippocampal place fields. Nat. Neurosci. 17, 725–731 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Altimus, C., Harrold, J., Jaaro-Peled, H., Sawa, A. & Foster, D.J. Disordered ripples are a common feature of genetically distinct mouse models relevant to schizophrenia. Mol Neuropsychiatry 1, 52–59 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Suh, J., Foster, D.J., Davoudi, H., Wilson, M.A. & Tonegawa, S. Impaired hippocampal ripple-associated replay in a mouse model of schizophrenia. Neuron 80, 484–493 (2013).

    Article  CAS  PubMed  Google Scholar 

  34. Colgin, L.L., Moser, E.I. & Moser, M.-B. Understanding memory through hippocampal remapping. Trends Neurosci. 31, 469–477 (2008).

    Article  CAS  PubMed  Google Scholar 

  35. Hardt, O., Nader, K. & Nadel, L. Decay happens: the role of active forgetting in memory. Trends Cogn. Sci. 17, 111–120 (2013).

    Article  PubMed  Google Scholar 

  36. Leutgeb, S. et al. Independent codes for spatial and episodic memory in hippocampal neuronal ensembles. Science 309, 619–623 (2005).

    Article  CAS  PubMed  Google Scholar 

  37. Cacucci, F., Wills, T.J., Lever, C., Giese, K.P. & O'Keefe, J. Experience-dependent increase in CA1 place cell spatial information, but not spatial reproducibility, is dependent on the autophosphorylation of the alpha-isoform of the calcium/calmodulin-dependent protein kinase II. J. Neurosci. 27, 7854–7859 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Karlsson, M.P. & Frank, L.M. Network dynamics underlying the formation of sparse, informative representations in the hippocampus. J. Neurosci. 28, 14271–14281 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  40. Lee, I. & Knierim, J.J. The relationship between the field-shifting phenomenon and representational coherence of place cells in CA1 and CA3 in a cue-altered environment. Learn. Mem. 14, 807–815 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Fenton, A.A. et al. Unmasking the CA1 ensemble place code by exposures to small and large environments: more place cells and multiple, irregularly arranged, and expanded place fields in the larger space. J. Neurosci. 28, 11250–11262 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Eichenbaum, H. Memory: organization and control. Annu. Rev. Psychol. 68, 19–45 (2017).

    Article  PubMed  Google Scholar 

  43. Mukai, J. et al. Molecular substrates of altered axonal growth and brain connectivity in a mouse model of schizophrenia. Neuron 86, 680–695 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Tamura, M., Mukai, J., Gordon, J.A. & Gogos, J.A. Developmental inhibition of Gsk3 Rescues behavioral and neurophysiological deficits in a mouse model of schizophrenia predisposition. Neuron 89, 1100–1109 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Crabtree, G.W. & Gogos, J.A. Synaptic plasticity, neural circuits, and the emerging role of altered short-term information processing in schizophrenia. Front. Synaptic Neurosci. 6, 28 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Debbané, M., Glaser, B. & Eliez, S. Encoding and retrieval processes in velo-cardio-facial syndrome (VCFS). Neuropsychology 22, 226–234 (2008).

    Article  PubMed  Google Scholar 

  47. McCabe, K., Rich, D., Loughland, C.M., Schall, U. & Campbell, L.E. Visual scanpath abnormalities in 22q11.2 deletion syndrome: is this a face specific deficit? Psychiatry Res. 189, 292–298 (2011).

    Article  CAS  PubMed  Google Scholar 

  48. Bassett, A.S. et al. The schizophrenia phenotype in 22q11 deletion syndrome. Am. J. Psychiatry 160, 1580–1586 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Kahn, R.S. & Keefe, R.S. Schizophrenia is a cognitive illness: time for a change in focus. JAMA Psychiatry 70, 1107–1112 (2013).

    Article  PubMed  Google Scholar 

  50. Vorstman, J.A. et al. Cognitive decline preceding the onset of psychosis in patients with 22q11.2 deletion syndrome. JAMA Psychiatry 72, 377–385 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Stark, K.L. et al. Altered brain microRNA biogenesis contributes to phenotypic deficits in a 22q11-deletion mouse model. Nat. Genet. 40, 751–760 (2008).

    Article  CAS  PubMed  Google Scholar 

  52. Lovett-Barron, M. et al. Dendritic inhibition in the hippocampus supports fear learning. Science 343, 857–863 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. 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 

  54. Mukai, J. et al. Palmitoylation-dependent neurodevelopmental deficits in a mouse model of 22q11 microdeletion. Nat. Neurosci. 11, 1302–1310 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Chen, T.-W. et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Kaifosh, P., Lovett-Barron, M., Turi, G.F., Reardon, T.R. & Losonczy, A. Septo-hippocampal GABAergic signaling across multiple modalities in awake mice. Nat. Neurosci. 16, 1182–1184 (2013).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Kaifosh, P., Zaremba, J.D., Danielson, N.B. & Losonczy, A. SIMA: Python software for analysis of dynamic fluorescence imaging data. Front. Neuroinform. 8, 80 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  60. Freund, T.F. & Buzsáki, G. Interneurons of the hippocampus. Hippocampus 6, 347–470 (1996).

    Article  CAS  PubMed  Google Scholar 

  61. Klausberger, T. & Somogyi, P. Neuronal diversity and temporal dynamics: the unity of hippocampal circuit operations. Science 321, 53–57 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Ambros-Ingerson, J. & Holmes, W.R. Analysis and comparison of morphological reconstructions of hippocampal field CA1 pyramidal cells. Hippocampus 15, 302–315 (2005).

    Article  PubMed  Google Scholar 

  63. Gulyás, A.I., Megías, M., Emri, Z. & Freund, T.F. Total number and ratio of excitatory and inhibitory synapses converging onto single interneurons of different types in the CA1 area of the rat hippocampus. J. Neurosci. 19, 10082–10097 (1999).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Papp, O.I., Karlócai, M.R., Tóth, I.E., Freund, T.F. & Hájos, N. Different input and output properties characterize parvalbumin-positive basket and axo-axonic cells in the hippocampal CA3 subfield. Hippocampus 23, 903–918 (2013).

    Article  CAS  PubMed  Google Scholar 

  65. 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 

  66. Lapray, D. et al. Behavior-dependent specialization of identified hippocampal interneurons. Nat. Neurosci. 15, 1265–1271 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. 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  PubMed  PubMed Central  Google Scholar 

  68. Jia, H., Rochefort, N.L., Chen, X. & Konnerth, A. In vivo two-photon imaging of sensory-evoked dendritic calcium signals in cortical neurons. Nat. Protoc. 6, 28–35 (2011).

    Article  CAS  PubMed  Google Scholar 

  69. Dombeck, D.A., Khabbaz, A.N., Collman, F., Adelman, T.L. & Tank, D.W. Imaging large-scale neural activity with cellular resolution in awake, mobile mice. Neuron 56, 43–57 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Danielson, N.B. et al. Distinct contribution of adult-born hippocampal granule cells to context encoding. Neuron 90, 101–112 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Rajasethupathy, P. et al. Projections from neocortex mediate top-down control of memory retrieval. Nature 526, 653–659 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Skaggs, W.E., Mcnaughton, B.L., Markus, E.J. & Gothard, K.M. An information-theoretic approach to deciphering the hippocampal code. In Advances in Neural Information Processing Systems 1030–1037 (Morgan Kaufmann, San Francisco,1993).

  73. Ahmed, O.J. & Mehta, M.R. The hippocampal rate code: anatomy, physiology and theory. Trends Neurosci. 32, 329–338 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Hastie, T.R. & Friedman, J. The Elements of Statistical Learning: Data Mining, Inference, and Prediction (Springer-Verlag, 2009).

  75. Wang, Y.W. & Wahba, G. Bootstrap condifedence itervals for smoothing splines and their comparison to bayesian confidence intervals. J. Stat. Comput. Simul. 51, 263–279 (1995).

    Article  Google Scholar 

Download references

Acknowledgements

We thank Z. He in the laboratory of I. Ionita-Laza for helpful discussions on the statistical analysis. J.D.Z. is supported by NIMH 1F31MH105169. N.B.D. is supported by NINDS F30NS090819. P.W.K. is supported by the Howard Hughes Medical Institute. F.T.S is supported by the Canadian Institutes of Health Research. J.A.G. is supported by NIMH MH097879 and NIMH MH096274. A.L. is supported by NIMH 1R01MH100631, NINDS 1U01NS090583, NINDS 1R01NS094668, the Searle Scholars Program, the Human Frontier Science Program, and the McKnight Memory and Cognitive Disorders Award.

Author information

Authors and Affiliations

Authors

Contributions

A.L. and J.A.G. conceived the project. J.D.Z. and A.D. designed and performed experiments with the help of F.T.S. J.D.Z. and A.D. analyzed data with the help of N.B.D., A.D.G., P.W.K., J.C.B. and Z.L. in analysis and technical infrastructure. J.D.Z., A.D., J.A.G., F.T.S. and A.L. wrote the manuscript.

Corresponding authors

Correspondence to Joseph A Gogos or Attila Losonczy.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Task performance by mouse.

Behavioral performance per mouse across conditions of the task for (a) WT and (b) Df(16)A± mice. Each line is an individual mouse showing the overall task performance each day.

Supplementary Figure 2 Comparison of behavior during initial learning in GOL.

a. Mean velocity (excluding stationary time) during Condition I for WT and Df(16)A± mice. (WT: 7.123 ± 1.002, n=6 mice; Df(16)A±: 6.778 ± 0.312, n=5 mice; independent samples T-test, t=0.302, p=0.769). b. Velocity during Condition I separated by session within each day (two-way ANOVA for session and genotype, all n.s.). c,d. Lap rate (as in a,b) (WT: 0.947 ± 0.147, n=6 mice; Df(16)A±: 1.164 ± 0.128, n=5 mice; independent samples T-test, t=-1.087, p=0.305; two-way ANOVA for session and genotype, all n.s.). e,f. Lick rate (as in a,b) (WT: 1.100 ± 0.101, n=6 mice; Df(16)A±: 1.362 ± 0.312, n=5 mice; independent samples T-test, t=-1.102, p=0.299; two-way ANOVA for session and genotype, all n.s.). Both WT and Df(16)A± mice show similar levels of activity during the initial learning period, as well as across sessions within each day.

Supplementary Figure 3 Anticipatory licking.

a. Task performance across all conditions of the task (Two-way mixed-effects RM ANOVA, genotype*condition interaction, p=0.060). b. Task performance and centroid shift correlation (centroid shift vs. fraction of licks in reward zone, Pearson’s correlation coefficient, WT: -0.282, p=0.015; Df(16)A±: -0.343, p=0.008). c. Task performance and goal zone place cell enrichment during Condition III (fraction of place cells near reward vs. fraction of licks in reward zone, Pearson’s correlation coefficient, WT: 0.418, p=0.008; Df(16)A±: -0.119, p=0.503).

Supplementary Figure 4 Comparison of lifetime place coding and place cell metrics between Df(16)A+/− and WT mice.

a. Lifetime place coding percentage, the fraction of ROIs that were ever identified as a place cell by the nth session imaged (lifetime place coding, Cox Regression, B=0.222, p=0.244). b. Fraction of all sessions imaged that an ROI was identified as a place cell (fraction of sessions a place cell; WT: 0.254 ± 0.004, n=3162 cells; Df(16)A±: 0.214 ± 0.004, n=3322 cells; Mann-Whitney U, U=4.55x106, p<0.0001), averaged within mice (inset; independent sample T-test, t=1.517, p=0.164). c. Transient sensitivity, defined as the fraction of laps in which a transient occurred in the place field (WT: 0.5786 ± 0.00218, n=12524 place cell*sessions; Df(16)A±: 0.5445 ± 0.0027, n=7664 place cell*sessions; Mann-Whitney U, U=4.4x107, p<0.0001), averaged within mice (inset, independent samples T-test, t=0.0142, p=0.989). d. Transient specificity, defined as the fraction of transients that occurred in the place field (WT: 0.795 ± 0.0161, n=12571 place cell*sessions; Df(16)A±: 0.872 ± 0.0018, n=7683 place cell*sessions; Mann-Whitney U, U=3.59x107, p<0.0001), averaged within mice (inset; Welch’s T-test, t=2.427, p=0.0544). e. Place field width (WT: 32.09 ± 0.125, n=14833 place fields; Df(16)A±: 29.25 ± 0.136, n=8529 place fields; Mann-Whitney U, U=5.5x107 p<0.0001), averaged within mice (inset, Welch’s T-test, t=1.990, p=0.0911). f. Single-cell sparsity (WT: 0.2325 ± 0.001, n=12571 ROI*sessions; Df(16)A+/: 0.186 ± 0.001, n=8683 ROI*sessions; Mann-Whitney U, U=3.35x107, p<0.0001), averaged within mice (inset, Welch’s T-test, t=2.064, p=0.0852). ***p<0.001.

Supplementary Figure 5 Place field correlation.

Place field correlation showed an overall similar effect as centroid shift (see Fig. 3d-f). a. Compared to WT mice, Df(16)A± mice show a significant overall decrease in place field correlation (WT: 0.232 ± 0.010, n=74 sessions; Df(16)A±: 0.130 ± 0.010, n=59 sessions; shuffle: 0.0238 ± 0.005, n=133; WT vs. shuffle: Welch’s T-test; t=18.87; p<0.0001; Df(16)A± vs. shuffle: Welch’s T-test, t=9.89, p<0.0001; WT vs. Df(16)A±, independent samples T-test, t=7.143, p<0.0001; inset aggregated by mouse: WT vs. Df(16)A±: independent samples T-test, t=2.584, p=0.0295). b. Place fields were more stable from session-to-session than day-to-day and the Df(16)A± mice were less stable across elapsed time (two-way ANOVA for time elapsed and genotype, main effect of genotype: F(1,152)=5.710, p=0.0181; main effect of elapsed time: F(1,152)=55.329, p<0.0001; elapsed time*genotype interaction: F(1,152)=8.074, p=0.00511; S-S, WT vs. Df(16)A±f: t=0.507, p=0.613; D-D, WT vs. Df(16)A±: t=4.455; p<0.0001). c. Task performance correlates with the session-mean place field correlation for WT mice (Spearman’s correlation coefficient=0.335, p=0.004) and trends similarly for Df(16)A± mice (Pearson’s correlation coefficient=0.224, p=0.088).

Supplementary Figure 6 Task performance, place cell recurrence, and spatial tuning stability by Condition.

Task performance and population stability by genotype follows similar trajectories across conditions; that is to say, performance and stability are similar in Condition I, slightly impaired in the Df(16)A± mice during Condition II and most different during Condition III (three-way ANOVA, genotype*metric*condition interaction: F(4,549)=0.484, p=0.747; condition*genotype interaction: F(2,549)=11.982, p<0.0001; metric*genotype interaction: F(2,549)=0.771, p=0.463; metric*condition interaction: F(4,549)=1.503, p=0.200; Condition I, all metrics, WT vs. Df(16)A±: independent samples T-test: t=-1.194, p=0.234; Condition II, all metrics, WT vs. Df(16)A±: independent samples T-test: t=2.67, p=0.0081; Condition III, all metrics, WT vs. Df(16)A±: Welch’s T-test: t=5.586, p<0.0001). Same a. Fraction of licks in the reward zone by Condition. Same data in Fig. 3g with values transformed to z-scores (across all conditions and both genotypes). b. Recurrence probability by Condition. Same data in Fig. 3h with values transformed to z-scores (across all conditions and both genotypes). c. Mean centroid shift by Condition. Same data in Fig. 3i with values transformed to -1 * z-scores (across all conditions and both genotypes; multiplying by -1 makes positive values represent relatively increased stability). Bonferroni-corrected post hoc tests comparing genotype per condition; *p<0.05, **p<0.01, ***p<0.0001.

Supplementary Figure 7 Impaired task performance during CA1 inactivation.

During initial learning of a reward location, local inactivation of CA1 lead to significantly reduced task performance (Days 1-3, muscimol to saline vs. saline to muscimol: Mann-Whitney U, U=126.5, p<0.0001). In addition, mice which received saline infusion during the first three initial learning days performed significantly worse on the fourth day when they were infused with muscimol (saline to muscimol, Days 1-3: 0.221 ± 0.053, n=36 sessions; Day 4: 0.084 ± 0.034, n=12 sessions; Mann-Whitney U, U=111, p=0.0235) and now performed at a similar level to mice which were initially infused with muscimol (Day 4, saline to muscimol vs. muscimol to saline: independent samples T-test, t=0.633, p=0.535).

Supplementary Figure 8 No place cells on a cue-free belt; spatial tuning near the fabric-transition is not more stable.

a. The 6 most spatially-tuned cells (lowest circular variance) on a burlap belt, plotted as in Fig. 2c. b. Place cell fraction on a ‘cue-rich’ and ‘cue-free’ belt during RF (cue-rich: 0.275 ± 0.017, n=56 sessions; cue-free: 0.057 ± 0.018, n=3 sessions; independent samples T-test, t=3.006, p=0.004). c. Transient circular variance on a ‘cue-rich’ and ‘cue-free belt during RF (cue-rich: 0.427 ± 0.003, n=7828 cell*sessions; cue-free: 0.746 ± 0.008, n=375 cell*sessions; Mann-Whitney U, U=4.99x105, p<0.0001). d. Centroid shift of cells from the last day of Condition I to the first session of Condition II separated by tuning preference relative to fabric transitions—the only features that remain constant between the two contexts. WT tuning is generally more stable (WT vs. Df(16)A±, independent samples T-test: t=-4.96, p<0.0001; see Fig. 4a), but neither genotype shows increased stability near the fabric transitions (two-way ANOVA, main effect of binned distance: F(2,24)=0.024, p=0.977). **p<0.01, ***p<0.001

Supplementary Figure 9 Sharp wave-ripples (SWR) are altered in Df(16)A+/– mice.

a. Mean SWR wavelet power for WT (left) and Df(16)A± (right) mice. b. Difference (Df(16)A± - WT) of mean SWR wavelet power in a. c. Rate of SWRs during stationary bouts (mean ± SD; WT: 0.106 ± 0.103, n=45 stationary intervals; Df(16)A±: 0.178 ± 0.120, n=61 stationary intervals; Wilcoxon rank-sum test, h=3777.5, p=0.00096). d. Mean wavelet power (mean ± SD; WT: 3.622 ± 1.133, n=145 sharp wave-ripples; Df(16)A±: 4.060 ± 0.838, n=357 sharp wave-ripples; Wilcoxon rank-sum test, h=98423, p<0.0001). e. Frequency with maximum power (mean ± SD; WT: 152.190 ± 17.174, n=145 sharp wave-ripples; Df(16)A±: 163.391 ± 29.485, n=357 sharp wave-ripples; Wilcoxon rank-sum test, h=94798, p=0.00066). f-h. Same as (c-e) for several SWR detection thresholds (significance by Wilcoxon rank-sum test as marked). *p<0.05, **p<0.001, ***p<0.0001. SWR-related place cell reactivation (for example, as in Wilson and McNaughton, Science, 1994; Lee and Wilson, Neuron, 2002; Foster and Wilson, Nature 2006) was not directly assessed.

Supplementary Figure 10 Possible enrichment mechanisms.

Comparison of three theoretical methods by which place cells could enrich a goal location. a. Place fields could be generally stable, but place cells near the reward are more likely to reoccur as place cells from session-to-session. (top) Recurrence probability as a function of distance from reward. (middle) Mean place field centroid shift as a function of distance from reward. (bottom) Mean place field shift variance as a function of distance from reward. b. Place cells could reoccur at equal probability along the belt, but place fields shift towards the reward location such that fields before the reward shift forward and fields after the reward shift backwards. Plots as in a. Place fields shifting towards the reward also leads to enrichment in our model. c. Place fields might not shift uniformly towards the reward, but if fields are generally stable, the ones near the reward could shift less than ones farther away. Plots as in a. d-f. Using the parameters from a-c, our enrichment model suggests all three hypothetical models could lead to enrichments: (d) increased place cell recurrence at reward position, (e) place fields shifting towards the reward location, or (f) place fields near the reward shifting less than ones away from the reward.

Supplementary Figure 11 Latent spatial tuning revealed across multiple sessions.

a. Scatter plots of original position versus position shift after 2 elapsed sessions for WT and Df(16)A± mice. Cells that were not a place cell in the middle session are marked in magenta and cyan for WT and Df(16)A± data, respectively. Even cells that were not a place cell in the intervening session still cluster around 0, suggesting that they retain some latent place preference that is either not expressed or not detectable. Vertical dashed line denotes reward location. Horizontal dashed line marks fields that do not shift at all. Green diagonal dashed line marks fields that remap directly to the reward location. b. Mean place field shift variance across all positions for cells paired by 1 session elapsed, 2 sessions elapsed, 2 sessions elapsed for cells that were not a place cell in the intervening session, and 2 iterations of the model. Two-session elapsed place cells are equally stable whether or not the cell was a place cell in the middle session, which is also the same as two iterations of the model.

Supplementary Figure 12 Modeled enrichment for all three Conditions.

Our model of place cell enrichment only produces goal location enrichment with parameters fit from WT mice during Condition III. a,b. Enrichment by iteration with parameters fit from each of the three Conditions of the task for (a) WT and (b) Df(16)A± mice (as in Fig. 7b, 8e). c,d. Final distribution of place fields after 8 iterations for each set of parameters fit from (c) WT and (d) Df(16)A± mice (as in Fig. 7c, 8f).

Supplementary Figure 13 Model enrichment when swapping individual parameters between WT and Df(16)A+/− values.

Mean enrichment by iteration (left, 90% confidence interval shading determined from 100 simulations) and histogram of distributions of place fields after the final iteration (right, data pooled across all simulations, vertical dashed line denotes reward location) for each parameter individually swapped between WT and flat models. a. No swap. b. Swap place cell recurrence probability (Precur). c. Swap session-to-session place field shift variance. d. Swap session-to-session place field shift offset. Swapping shift offset has the largest effect, as the flat model with only the WT shift offset parameters leads to enrichment, while the WT model with the flat shift offset parameters does not.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zaremba, J., Diamantopoulou, A., Danielson, N. et al. Impaired hippocampal place cell dynamics in a mouse model of the 22q11.2 deletion. Nat Neurosci 20, 1612–1623 (2017). https://doi.org/10.1038/nn.4634

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4634

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