Skip to main content
Log in

A neural mass model of place cell activity: theta phase precession, replay and imagination of never experienced paths

  • Published:
Journal of Computational Neuroscience Aims and scope Submit manuscript

Abstract

Recent results on hippocampal place cells show that the replay of behavioral sequences does not simply reflect previously experienced trajectories, but may also occur in the reverse direction, or may even include never experienced paths. In order to elucidate the possible mechanisms at the basis of this phenomenon, we have developed a model of sequence learning. The present model consists of two layers of place cell units. Long-range connections among units implement heteroassociation between the two layers, trained with a temporal Hebb rule. The network was trained assuming that a virtual rat moves within a virtual maze. This training leads to the formation of bidirectional synapses between the two layers, i.e. synapses connecting a neuron both with its previous and subsequent element in the path. Subsequently, two distinct conditions were simulated with the trained network. During an exploratory phase, characterized by a similar consideration to the external environment and to the internal representation, the model simulates the occurrence of theta precession in the forward path and the temporal compression. During an imagination phase, when there is no consideration to the external location, the model produces trains of gamma oscillations, without the presence of a theta rhythm, and simulates the occurrence of both direct and reverse replay, and the imagination of never experienced paths. The new paths are built by combining bunches of previous trajectories. The main mechanisms at the basis of this behavior are explained in detail, and lines for future improvements (e.g., to simulate preplay) are discussed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  • Almeida, L., Idiart, M., & Lisman, J.E. (2007) Memory retrivial time and memory capacity of the CA3 network: role of gamma frequency oscillations. Learning and Memory 14, 795–806.

  • Battaglia, F. P., Sutherland, G. R., Cowen, S. L., Mc Naughton, B. L., & Harris, K. D. (2005). Firing rate modulation: a simple statistical view of memory trace reactivation. Neural Networks, 18(9), 1280–1291. doi:10.1016/j.neunet.2005.08.011.

    Article  PubMed  Google Scholar 

  • Bendor, D., & Wilson, M. A. (2012). Biasing the content of hippocampal replay during sleep. Nature Neuroscience, 15(10), 1439–1444. doi:10.1038/nn.3203.

    Article  CAS  PubMed  Google Scholar 

  • Bi, G. Q., & Poo, M. M. (1998). Synaptic modifications in cultured Hippocampal neurons: dependence on spike timing, synaptic strength, and postsynaptic cell type. Journal of Neuroscience, 18(24), 10464–72.

    CAS  PubMed  Google Scholar 

  • Bi, G. Q., & Poo, M. M. (2001). Synaptic modification by correlated activity: Hebb’s postulate revisited. Annual Review of Neuroscience, 24, 139–66.

    Article  CAS  PubMed  Google Scholar 

  • Buhry, L., Azizi, A. H., & Cheng, S. (2011). Reactivation, replay, and preplay: how it might all fit together. Neural Plasticity, 2011, 203462. doi:10.1155/2011/203462.

    Article  PubMed Central  PubMed  Google Scholar 

  • Buzsáki, G. (1989). Two-stage model of memory trace formation: a role for “noisy” brain states. Neuroscience, 31(3), 551–570.

    Article  PubMed  Google Scholar 

  • Buzsáki, G. (2002). Theta oscillations in the hippocampus. Neuron, 33(3), 325–340.

    Article  PubMed  Google Scholar 

  • Cappaert, N. L. M., Lopes da Silva, F. H., & Wadman, W. J. (2009). Spatio-temporal dynamics of theta oscillations in hippocampal-entorhinal slices. Hippocampus, 19(11), 1065–1077. doi:10.1002/hipo.20570.

    Article  CAS  PubMed  Google Scholar 

  • Carr, M. F., Karlsson, M. P., & Frank, L. M. (2012). Transient slow gamma synchrony underlies hippocampal memory replay. Neuron, 75(4), 700–713. doi:10.1016/j.neuron.2012.06.014.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Colgin, L. L., Denninger, T., Fyhn, M., Hafting, T., Bonnevie, T., Jensen, O., Moser, M., & Moser, E. (2009). Frequency of gamma oscillations routes flow of information in the hippocampus. Nature, 462(7271), 353–7. doi:10.1038/nature08573.

    Article  CAS  PubMed  Google Scholar 

  • Cona, F., & Ursino, M. (2013). A multi-layer neural-mass model for learning sequences using theta/gamma oscillations. International Journal of Neural Systems, 23(3), 1–18.

    Article  Google Scholar 

  • Cona, F., Zavaglia, M., & Ursino, M. (2012). Binding and segmentation via a neural mass model trained with Hebbian and anti-Hebbian mechanisms. International Journal of Neural Systems, 22(02), 1–20. doi:10.1142/S0129065712500037.

    Article  Google Scholar 

  • Cona, F., Lacanna, M., & Ursino, M. (2014). A thalamo-cortical neural mass model for the simulation of brain rhythms during sleep. Journal of Computational Neuroscience, 37(1), 125–48. doi:10.1007/s10827-013-0493-1.

    Article  CAS  PubMed  Google Scholar 

  • Crick, F., & Mitchison, G. (1983). The function of dream sleep. Nature, 304(5922), 111–114.

    Article  CAS  PubMed  Google Scholar 

  • Csicsvari, J., O’Neill, J., Allen, K., & Senior, T. (2007). Place-selective firing contributes to the reverse-order reactivation of CA1 pyramidal cells during sharp waves in open-field exploration. The European Journal of Neuroscience, 26(3), 704–716. doi:10.1111/j.1460-9568.2007.05684.x.

    Article  PubMed Central  PubMed  Google Scholar 

  • Cutsuridis, V., Cobb, S., & Graham, B. P. (2010). Encoding and retrieval in a model of the hippocampal CA1 microcircuit. Hippocampus, 20(3), 423–446. doi:10.1002/hipo.20661.

    CAS  PubMed  Google Scholar 

  • David, O., Harrison, L., & Friston, K. J. (2005). Modelling event-related responses in the brain. NeuroImage, 25(3), 756–770. doi:10.1016/j.neuroimage.2004.12.030.

    Article  PubMed  Google Scholar 

  • Davidson, T. J., Kloosterman, F., & Wilson, M. A. (2009). Hippocampal replay of extended experience. Neuron, 63(4), 497–507. doi:10.1016/j.neuron.2009.07.027.

    Article  CAS  PubMed  Google Scholar 

  • Diba, K., & Buzsáki, G. (2007). Forward and reverse hippocampal place-cell sequences during ripples. Nature Neuroscience, 10(10), 1241–1242. doi:10.1038/nn1961.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Dockendorf, K., & Srinivasa, N. (2013). Learning and prospective recall of noisy spike pattern episodes. Frontiers in Computational Neuroscience, 7, 80. doi:10.3389/fncom.2013.00080.

    Article  PubMed Central  PubMed  Google Scholar 

  • Dragoi, G., & Buzsáki, G. (2006). Temporal encoding of place sequences by hippocampal cell assemblies. Neuron, 50(1), 145–157. doi:10.1016/j.neuron.2006.02.023.

    Article  CAS  PubMed  Google Scholar 

  • Dragoi, G., & Tonegawa, S. (2011). Preplay of future place cell sequences by hippocampal cellular assemblies. Nature, 469(7330), 397–401. doi:10.1038/nature09633.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Dragoi, G., & Tonegawa, S. (2013). Distinct preplay of multiple novel spatial experiences in the rat. Proceedings of the National Academy of Sciences of the United States of America, 110(22), 9100–9105. doi:10.1073/pnas.1306031110.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Ekstrom, A. D., Meltzer, J., McNaughton, B. L., & Barnes, C. A. (2001). NMDA receptor antagonism blocks experience-dependent expansion of hippocampal “place fields.”. Neuron, 31(4), 631–638.

    Article  CAS  PubMed  Google Scholar 

  • Felleman, D. J., & Van Essen, D. C. (1991). Distributed hierarchical processing in the primate cerebral cortex. Cerebral Cortex, 1(1), 1–47. doi:10.1093/cercor/1.1.1.

    Article  CAS  PubMed  Google Scholar 

  • Foster, D. J., & Wilson, M. A. (2006). Reverse replay of behavioural sequences in hippocampal place cells during the awake state. Nature, 440(7084), 680–683. doi:10.1038/nature04587.

    Article  CAS  PubMed  Google Scholar 

  • Foster, D. J., & Wilson, M. A. (2007). Hippocampal theta sequences. Hippocampus, 17(11), 1093–1099. doi:10.1002/hipo.20345.

    Article  PubMed  Google Scholar 

  • Gupta, A. S., van der Meer, M. A. A., Touretzky, D. S., & Redish, A. D. (2010). Hippocampal replay is not a simple function of experience. Neuron, 65(5), 695–705. doi:10.1016/j.neuron.2010.01.034.

    Article  CAS  PubMed  Google Scholar 

  • Harvey, C. D., Collman, F., Dombeck, D. A., & Tank, D. W. (2009). Intracellular dynamics of hippocampal place cells during virtual navigation. Nature, 461(7266), 941–946. doi:10.1038/nature08499.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Hasselmo, M. E., Bodelón, C., & Wyble, B. P. (2002). A proposed function for hippocampal theta rhythm: separate phases of encoding and retrieval enhance reversal of prior learning. Neural Computation, 14(4), 793–817. doi:10.1162/089976602317318965.

    Article  PubMed  Google Scholar 

  • Hopfield, J. J. (2010). Neurodynamics of mental exploration. Proceedings of the National Academy of Sciences of the United States of America, 107(4), 1648–1653. doi:10.1073/pnas.0913991107.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Huxter, J. R., Senior, T. J., Allen, K., & Csicsvari, J. (2008). Theta phase specific codes for two-dimensional position, trajectory and heading in the hippocampus. Nature Neuroscience, 11(5), 587–594.

    Article  CAS  PubMed  Google Scholar 

  • Jadhav, S. P., Kemere, C., German, P. W., & Frank, L. M. (2012). Awake hippocampal sharp-wave ripples support spatial memory. Science, 336(6087), 1454–1458. doi:10.1126/science.1217230.

    Article  CAS  PubMed  Google Scholar 

  • Jeewajee, A., Lever, C., Burton, S., O’Keefe, J., & Burgess, N. (2008). Environmental novelty is signaled by reduction of the Hippocampal theta frequency. Hippocampus, 18(4), 340–348.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Jeewajee, A., Barry, C., Douchamps, V., Manson, D., Lever, C., & Burgess, N. (2013). Theta phase precession and grid and place cell firing in open environments. Phil. Trans. Rol. Soc. B, 369(1635): doi: 10.1098/rstb.2012.0532.

  • Jensen, O., & Lisman, J. E. (1996a). Hippocampal CA3 region predicts memory sequences: accounting for the phase precession of place cells. Learning & Memory, 3(2–3), 279–287.

    Article  CAS  Google Scholar 

  • Jensen, O., & Lisman, J. E. (1996b). Theta/gamma networks with slow NMDA channels learn sequences and encode episodic memory: role of NMDA channels in recall. Learning & Memory, 3(2–3), 264–278.

    Article  CAS  Google Scholar 

  • Jensen, O., & Lisman, J. E. (2005). Hippocampal sequence-encoding driven by a cortical multi-item working memory buffer. Trends in Neurosciences, 28(2), 67–72. doi:10.1016/j.tins.2004.12.001.

    Article  CAS  PubMed  Google Scholar 

  • Karlsson, M. P., & Frank, L. M. (2009). Awake replay of remote experiences in the hippocampus. Nature Neuroscience, 12(7), 913–918. doi:10.1038/nn.2344.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Kloosterman, F., van Haeften, T., & Lopes da Silva, F. H. (2004). Two reentrant pathways in the hippocampal-entorhinal system. Hippocampus, 14(8), 1026–1039.

    Article  PubMed  Google Scholar 

  • Lee, A. K., & Wilson, M. A. (2002). Memory of sequential experience in the hippocampus during slow wave sleep. Neuron, 36(6), 1183–1194.

    Article  CAS  PubMed  Google Scholar 

  • Lengyel, M., Szatmáry, Z., & Erdi, P. (2003). Dynamically detuned oscillations account for the coupled rate and temporal code of place cell firing. Hippocampus, 13(6), 700–714. doi:10.1002/hipo.10116.

    Article  PubMed  Google Scholar 

  • Lisman, J. E. (1999). Relating hippocampal circuitry to function: recall of memory sequences by reciprocal dentate-CA3 interactions. Neuron, 22(2), 233–242.

    Article  CAS  PubMed  Google Scholar 

  • Lisman, J. E., & Buzsáki, G. (2008). A neural coding scheme formed by the combined function of gamma and theta oscillations. Schizophrenia Bulletin, 34(5), 974–980. doi:10.1093/schbul/sbn060.

    Article  PubMed Central  PubMed  Google Scholar 

  • Lisman, J. E., & Redish, A. D. (2009). Prediction, sequences and the hippocampus. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 364(1521), 1193–1201. doi:10.1098/rstb.2008.0316.

    Article  PubMed Central  PubMed  Google Scholar 

  • Lisman, J. E., Talamini, L. M., & Raffone, A. (2005). Recall of memory sequences by interaction of the dentate and CA3: a revised model of the phase precession. Neural Networks, 18(9), 1191–1201. doi:10.1016/j.neunet.2005.08.008.

    Article  PubMed  Google Scholar 

  • Louie, K., & Wilson, M. A. (2001). Temporally structured replay of awake hippocampal ensemble activity during rapid eye movement sleep. Neuron, 29(1), 145–156.

    Article  CAS  PubMed  Google Scholar 

  • Manns, J. R., Zilli, E. A., Ong, K. C., Hasselmo, M. E., & Eichenbaum, H. (2007). Hippocampal CA1 spiking during encoding and retrieval: relation to theta phase. Neurobiology of Learning and Memory, 87(1), 9–20. doi:10.1016/j.nlm.2006.05.007.

    Article  PubMed  Google Scholar 

  • Markram, H., Lübke, J., Frotscher, M., & Sakmann, B. (1997). Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science, 275(5297), 213–215. doi:10.1126/science.275.5297.213.

    Article  CAS  PubMed  Google Scholar 

  • Maurer, A. P., & McNaughton, B. L. (2007). Network and intrinsic cellular mechanisms underlying theta phase precession of hippocampal neurons. Trends in Neurosciences, 30(7), 325–333. doi:10.1016/j.tins.2007.05.002.

    Article  CAS  PubMed  Google Scholar 

  • Mehta, M. R., Barnes, C. A., & McNaughton, B. L. (1997). Experience-dependent, asymmetric expansion of hippocampal place fields. Proceedings of the National Academy of Sciences of the United States of America, 94(16), 8918–8921.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Mitchell, S. J., & Ranck, J. B., Jr. (1980). Generation of theta rhythm in medial entorhinal cortex of freely moving rats. Brain Research, 189(1), 49–66.

    Article  CAS  PubMed  Google Scholar 

  • Mizuseki, K., Sirota, A., Pastalkova, E., & Buzsáki, G. (2009). Theta oscillations provide temporal windows for local circuit computation in the entorhinal-Hippocampal loop. Neuron, 64(2), 267–280.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Montgomery, S. M., Betancur, M. I., & Buzsáki, G. (2009). Behavior-dependent coordination of multiple theta dipoles in the hippocampus. Journal of Neuroscience, 29(5), 1381–1394.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Morris, R. G., Garrud, P., Rawlins, J. N., & O’Keefe, J. (1982). Place navigation impaired in rats with hippocampal lesions. Nature, 297(5868), 681–683.

    Article  CAS  PubMed  Google Scholar 

  • Nádasdy, Z., Hirase, H., Czurkó, A., Csicsvari, J., & Buzsáki, G. (1999). Replay and time compression of recurring spike sequences in the hippocampus. The Journal of Neuroscience, 19(21), 9497–9507.

    PubMed  Google Scholar 

  • O’Keefe, J., & Dostrovsky, J. (1971). The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Research, 34(1), 171–175.

    Article  PubMed  Google Scholar 

  • O’Keefe, J., & Recce, M. L. (1993). Phase relationship between hippocampal place units and the EEG theta rhythm. Hippocampus, 3(3), 317–330. doi:10.1002/hipo.450030307.

    Article  PubMed  Google Scholar 

  • O’Neill, J., Senior, T., & Csicsvari, J. (2006). Place-selective firing of CA1 pyramidal cells during sharp wave/ripple network patterns in exploratory behavior. Neuron, 49(1), 143–155. doi:10.1016/j.neuron.2005.10.037.

    Article  PubMed  Google Scholar 

  • O'Keefe, J., & Burgess, N. (2005). Dual phase and rate coding in hippocampal place cells: theoretical significance and relationship to entorhinal grid cells. Hippocampus, 15(7), 853–66. doi:10.1002/hipo.20115.

    Article  PubMed Central  PubMed  Google Scholar 

  • Scharfman, H. E. (2007). The CA3 “backprojection” to the dentate gyrus. Progress in Brain Research, 163, 627–37. doi:10.1016/S0079-6123(07)63034-9.

    Article  PubMed Central  PubMed  Google Scholar 

  • Singer, A. C., & Frank, L. M. (2009). Rewarded outcomes enhance reactivation of experience in the hippocampus. Neuron, 64(6), 910–921. doi:10.1016/j.neuron.2009.11.016.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Skaggs, W. E., McNaughton, B. L., Wilson, M. A., & Barnes, C. A. (1996). Theta phase precession in hippocampal neuronal populations and the compression of temporal sequences. Hippocampus, 6(2), 149–172.

    Article  CAS  PubMed  Google Scholar 

  • Squire, L. R., & Alvarez, P. (1995). Retrograde amnesia and memory consolidation: a neurobiological perspective. Current Opinion in Neurobiology, 5(2), 169–177.

    Article  CAS  PubMed  Google Scholar 

  • Taube, J. S., Muller, R. U., & Ranck, J. B. J. (1990). Head-direction cells recorded from the postsubiculum in freely moving rats. I. Description and quantitative analysis. The Journal of Neuroscience, 10(2), 420–435.

    CAS  PubMed  Google Scholar 

  • Tsodyks, M. V., Skaggs, W. E., Sejnowski, T. J., & McNaughton, B. L. (1996). Population dynamics and theta rhythm phase precession of hippocampal place cell firing: a spiking neuron mode. Hippocampus, 6(3), 271–280.

    Article  CAS  PubMed  Google Scholar 

  • Ursino, M., Cona, F., & Zavaglia, M. (2010). The generation of rhythms within a cortical region: analysis of a neural mass model. NeuroImage, 52(3), 1080–1094.

    Article  PubMed  Google Scholar 

  • Wallenstein, G. V., & Hasselmo, M. E. (1997). GABAergic modulation of hippocampal population activity: sequence learning, place field development, and the phase precession effect. Journal of Neurophysiology, 78(1), 393–408.

    CAS  PubMed  Google Scholar 

  • Wilson, M. A., & McNaughton, B. L. (1993). Dynamics of the hippocampal ensemble code for space. Science, 261(5124), 1055–1058.

    Article  CAS  PubMed  Google Scholar 

Download references

Conflicts of interest

The authors declare that they have no conflict of interest.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Filippo Cona.

Additional information

Action Editor: Alessandro Treves

Appendices

Appendix A: Equations of a single unit

The model of a single cortical column consists of four neural populations (Fig. 1), which represent pyramidal neurons (subscript p), excitatory interneurons (subscript e), and inhibitory interneurons with slow and fast synaptic kinetics (GABAA, slow and GABAA, fast, subscripts s and f, respectively). All populations are described with a similar mathematical formalism. Briefly, each population computes an average post-synaptic membrane potential (v, in (A4), (A8), (A12) and (A16)) from other neural populations and from external inputs, and converts this membrane potential into an average density of spikes fired by the neurons (z, in (A3), (A7), (A11) and (A15)). In order to account for the presence of inhibition (when potential is below a given threshold) and saturation (when potential is high) this conversion is simulated with a static sigmoidal relationship.

To model a whole cortical column, the four populations are connected via excitatory and inhibitory synapses, with impulse response h e (t), h s (t) or h f (t), assuming that synapses from pyramidal neurons and excitatory interneurons have similar dynamics. The kinetics of each synapse are described with a second-order system, but with different parameter values. By denoting with z the input to a synapse, and with y the synapse output, its dynamics can be described through a second order differential equation (equivalent to two first-order differential equations, i.e., Eqs (A1)-(A2), (A5)-(A6), (A9)-(A10) and (A13)-(A14) for the four populations.

This corresponds to the following equations

Pyramidal neurons

$$ \frac{d{y}_p(t)}{dt}={x}_p(t) $$
(A1)
$$ \frac{d{x}_p(t)}{dt}=\frac{G_e}{\tau_e}{z}_p(t)-\frac{2}{\tau_e}{x}_p(t)-\frac{y_p(t)}{\tau_e^2} $$
(A2)
$$ {z}_p(t)=\frac{2{e}_o}{1+{e}^{r\left( so-{v}_p\right)}} $$
(A3)
$$ {v}_p(t)={C}_{pe}{y}_e(t)-{C}_{ps}{y}_s(t)-{C}_{pf}{y}_f(t)+E(t) $$
(A4)

Excitatory interneurons

$$ \frac{d{y}_e(t)}{dt}={x}_e(t) $$
(A5)
$$ \frac{d{x}_e(t)}{dt}=\frac{G_e}{\tau_e}\left({z}_e(t)+\frac{u_p(t)}{C_{pe}}\right)-\frac{2}{\tau_e}{x}_e(t)-\frac{y_e(t)}{\tau_e^2} $$
(A6)
$$ {z}_e(t)=\frac{2{e}_o}{1+{e}^{r\left({s}_o-{v}_e\right)}} $$
(A7)
$$ {v}_e(t)={C}_{ep}{y}_p(t) $$
(A8)

Slow inhibitory interneurons

$$ \frac{d{y}_s(t)}{dt}={x}_s(t) $$
(A9)
$$ \frac{d{x}_s(t)}{dt}=\frac{G_s}{\tau_s}{z}_s(t)-\frac{2}{\tau_s}{x}_s(t)-\frac{y_s(t)}{\tau_s^2} $$
(A10)
$$ {z}_s(t)=\frac{2{e}_o}{1+{e}^{r\left({s}_o-{v}_s\right)}} $$
(A11)
$$ {v}_s(t)={C}_{sp}{y}_p(t) $$
(A12)

Fast inhibitory interneurons

$$ \frac{d{y}_f(t)}{dt}={x}_f(t) $$
(A13)
$$ \frac{d{x}_f(t)}{dt}=\frac{G_f}{\tau_f}{z}_f(t)-\frac{2}{\tau_f}{x}_f(t)-\frac{y_f(t)}{\tau_f^2} $$
(A14)
$$ {z}_f(t)=\frac{2{e}_o}{1+{e}^{r\left({s}_o-{v}_f\right)}} $$
(A15)
$$ {v}_f(t)={C}_{fp}{y}_p(t)-{C}_{fs}{y}_s(t)-{C}_{ff}{y}_f(t)-{y}_l(t)+I(t) $$
(A16)
$$ \frac{d{y}_l(t)}{dt}={x}_l(t) $$
(A17)
$$ \frac{d{x}_l(t)}{dt}=\frac{G_e}{\tau_e}{u}_f(t)-\frac{2}{\tau_e}{x}_l(t)-\frac{y_l(t)}{\tau_e^2} $$
(A18)

In the previous equations, the term E(t) (Eq. (A4)) represents the overall input entering into the pyramidal neurons of that unit, from units in the other layer (inter-layer connections). The term I(t) (Eq. (A16)) represents the overall input entering into GABAA,fast interneurons of the unit from pyramidal neurons of other units in the same layer (intra-layer lateral connections). The terms u p (t) (Eq. (A6)) and u f (t) (Eq. (A18)) represent exogenous inputs. They include Gaussian noise and (when available) information on the present position of the rat (see Appendix C).

To improve model reliability, we assumed that the two inputs (u p (t) and u f (t)) affect the target neurons through the typical dynamics of excitatory synapses. This is justified by the idea that all long range connections in the brain (including exogenous inputs) come from pyramidal neurons. To this end, the quantity u p (t) is included in Eq. (A6), i.e., we exploit the same dynamics of an excitatory interneuron to affect the target pyramidal neurons. This truck avoids the introduction of two further state variables. Conversely, the same truck was not possible for what concerns the input quantity u f (t) affecting the fast inhibitory interneurons (in fact, these interneurons do not receive any specific excitation). Hence, we introduced two additional state variables (Eqs. (A17) and (A18)) to describe the synaptic kinetics of the u f (t) input.

For the meaning of other quantities, see Tables 1 and 2.

Table 1 Parameters of units
Table 2 Variables of units

In the subsequent paragraph, quantities belonging to the first and second layers will be denoted through the superscripts L1 and L2, respectively. The position within the layer will be denoted with a subscript (i or j).

Appendix B: The connections between units

Two types of long-range connections are used, both coming from pyramidal neurons of the pre-synaptic unit. Connections of type W affect the membrane potential v Lx p,i of the pyramidal neurons in the target unit by an additive term E Lx i (t). Connections of type A, instead, affect the membrane potential v Lx f,i of the GABAA,fast interneurons in the target unit by an additive term I Lx i (t) (see Eqs. (A4) and (A16) and also Cona et al. (2012) for more mathematical details). In the following, two superscripts and two subscripts are used for each connection. The superscripts represent the layers to which the postsynaptic and the presynaptic units respectively belong. The two subscripts represent the positions of the postsynaptic and of the presynaptic units within these layers.

The expressions of E Lx i (t) and I Lx i (t) for the two layers are:

$$ {E}_i^{L1}(t)={\sum}_j{W}_{ij}^{L1L2}\cdot {y}_{p,j}^{L2}(t) $$
(B1)
$$ {I}_i^{L1}(t)={\sum}_{j\ne i}{A}_{ij}^{L1L1}\cdot {z}_{p,j}^{L1}(t) $$
(B2)
$$ {E}_i^{L2}(t)={W}_{ii}^{L2L1}\cdot {y}_{p,i}^{L1}(t) $$
(B3)
$$ {I}_i^{L2}(t)={\sum}_{j\ne i}{A}_{ij}^{L2L2}\cdot {z}_{p,j}^{L2}(t) $$
(B4)

The previous equations can be explained as follows. Eq. (B1) signifies that pyramidal neurons in L1 receive excitatory inputs from pyramidal neurons in the L2. These connections are learnt during the maze exploration (see below) and implement a heteroassociative network that enables units in L2 to activate units in L1 encoding for near areas of the maze. According to Eq. (B2), GABAA,fast interneurons in L1 receive inputs through synapses with fast (AMPA) kinetics from the pyramidal neurons of other units in the same layer. Eq. (B3) means that pyramidal neurons in L2 receive an input from pyramidal neurons in L1 located at the same position. Eq. (B4) has the same meaning as Eq. (B2), but with reference to L2.

These expressions describe how the firing rates in source populations are mixed with different connection strengths to provide depolarizing or hyperpolarizing (excitatory or inhibitory) inputs to target populations. For what concerns the connections W, they are mediated by glutamatergic synapses, hence the postsynaptic potential y p (t) was used as driving quantity (Eqs. (B1) and (B3)). Conversely, in order to characterize the faster AMPA synapses that mediate connections of type A in Eqs. (B2) and (B4), we used directly the firing rate z p (t) of the presynaptic neuron, thus simulating a synapse with a negligible time constant (Cona and Ursino 2013; Cona et al. 2012).

Appendix C: The external input

As shown in Eqs (A6) and (A18) above, both pyramidal neurons and GABAA,fast interneurons in all the units in both layers receive an additional external input. In particular, all units receive a white Gaussian noise input with zero mean value, as in previous works (Cona and Ursino 2013; Cona et al. 2012). Moreover, pyramidal neurons in layer L1 also receive a sensory input, (say S i(t)) which indicates where the rat is located at a given time.

Hence, the quantities u p (t) and u f (t) have the following expressions:

$$ {u}_{p,i}^{L1}(t)={S}_i(t)+{n}_{p,i}^{L1}(t) $$
(C1)
$$ {u}_{p,i}^{L2}(t)={n}_{p,i}^{L2}(t) $$
(C2)
$$ {u}_{f,i}^{L1}(t)={n}_{f,i}^{L1}(t) $$
(C3)
$$ {u}_{f,i}^{L2}(t)={n}_{f,i}^{L2}(t) $$
(C4)

where the quantities n(t) represent independent Gaussian noises with zero mean and assigned variance (see Table 3 for the values used during the present simulations).

Table 3 Parameters involving maze exploration and inputs

The input S i (t) depends on the simulation phase (training, exploration or imagination) as described below.

Training phase - While the rat moves along the maze during the training phase, the layer L1 receives a sensory stimulus ST(x, y; t) that indicates where the rat is located at a given time, and has therefore the shape of a bi-dimensional impulse:

$$ ST\left(x,y;t\right)=\delta \left(x-{r}_x(t),y-{r}_y(t)\right) $$
(C4)

where r x (t) and r y (t) are the rat coordinates at time t, and δ(x, y) represents a bidimensional Dirac delta function.

Each unit in L1 responds to this stimulus according to a Gaussian shaped receptive field:

$$ R{F}_i\left(x,y\right)={G}_{RF}\cdot exp\left[-\frac{{\left(x-R{F}_{x,i}\right)}^2+{\left(y-R{F}_{y,i}\right)}^2}{2{\sigma}_{RF}^2}\right] $$
(C5)

where RF x,i and RF y,i are the central coordinates of the receptive field of unit i, while G RF and σ RF are the receptive field’s gain and standard deviation. RF x,i and RF y,i for the units in L1 have been set to fully cover the maze with a rectangular grid (Fig. 2).

The input S i (t) during training is then computed by calculating the convolution between the sensory input and the receptive field:

$$ {S}_i(t)=\int \int ST\left(x,y;t\right)\cdot R{F}_i\left(x,y\right) dxdy={G}_{RF}\cdot exp\left[-\frac{{\left({r}_x(t)-R{F}_{x,i}\right)}^2+{\left({r}_y(t)-R{F}_{y,i}\right)}^2}{2{\sigma}_{RF}^2}\right] $$
(C6)

which is sufficient to activate a few neurons that encode for areas around the point in the maze where the rat is at time t, while not affecting neurons that encode for distant areas.

Exploration phase - As described in the text, to mimic the exploration phase, when the rat is familiar with the environment, we assumed a decrease in the consideration toward the external stimulus, and an increased impact of the internal representation (i.e., on the activity of pyramidal neurons in L1). This was simulated by reducing the gain G RF to half the value used during the training phase, and adding an offset term, \( \frac{1}{2}{S}_{offset} \), to all units. It is worth noting that the presence of this offset is the same as to assume a reduced threshold for all neurons. Moreover, when exploring a known environment, the rat must choose a given direction of movement to imagine a future path. For instance, this may correspond to head direction. To this end, we included a further receptive field obtained as the directional derivative of RF i (x, y) along the direction in which the rat is moving \( \widehat{v}(t) \). Thus, the input in this case becomes

$$ {S}_i(t)=\frac{1}{2}\int \int ST\left(x,y;t\right)\cdot \left[R{F}_i\left(x,y\right)-{D_{\hat{v}}}_{(t)}R{F}_i\left(x,y\right)\right] dxdy+\frac{1}{2}{S}_{offset} $$
(C7)
$$ \hat{v}(t)=\frac{\left[\frac{\partial {r}_x(t)}{\partial t},\frac{\partial {r}_y(t)}{\partial t}\right]}{\sqrt{\left(\frac{\partial {r}_x{(t)}^2}{\partial t}+\frac{\partial {r}_y{(t)}^2}{\partial t}\right)}} $$
(C8)

where −\( {D}_{\widehat{v}(t)}R{F}_i\left(x,y\right) \) gives a negative contribution to the units that encode for the maze locations visited in the near past and thus warrants that the sequence is evoked in the same direction in which the rat is moving.

Imagination phase - During the imagination phase, we assumed that the rat completely neglects the external stimulus, but it pays stronger consideration to its internal representation of the maze, to spontaneously evoke neuronal activity. This has been simulated by doubling the offset term, compared with that used in the exploration phase. We have

$$ {S}_i(t)={S}_{offset} $$
(C9)

The increase of this offset term is the same as to assume a moderate leftward shift in the sigmoidal relationship of the pyramidal populations, i.e., all the neurons are more excitable.

The values of all of the parameters are found in Table 3.

Appendix D: The Hebbian learning

While the virtual rat moves through the virtual maze its units in L1 get activated by the inputs given in Eq. (C1). This activation is transferred to L2 through the W L2L1 ii connections, according to Eq. (B3). When units in the two layers are activated in close time synchronization, Hebbian learning between the units is possible. In this work Hebbian learning is implemented to train excitatory connections from L2 to L1 according to the following equations:

$$ {\tau}_{train}\frac{d{W}_{ij}^{L1L2}}{dt}={\gamma}_{ij}{\left[{z}_{p,i}^{L1}-\theta \right]}^{+}{\left[{m}_{p,j}^{L2}-\theta \right]}^{+} $$
(D1)

where γ ij is a learning rate (depending on the particular synapse value), z L1 p,i (t) is the spike density of the pyramidal population in the postsynaptic unit (in L1), m L2 p,j (t) is the average activity of pyramidal populations in the presynaptic unit (in L2) computed in a previous temporal window consisting of N s samples. Hence

$$ {m}_{p,j}^{L2}(t)=\frac{{\displaystyle {\sum}_{i=0}^{Ns-1}{z}_{p,j}^{L2}}\left(t-l\cdot Ts\right)}{N_s} $$
(D2)

It is worth noting that the right hand member of Eq. (D1) can be rewritten introducing the synchronization between the presynaptic and postsynaptic units, defined as follows

$$ sy{n}_{ij}^{L1L2}(t)=\frac{{\left[\frac{z_{p,i}^{L1}}{2{e}_o}-{\theta}_L\right]}^{+}{\left[\frac{m_{p,j}^{L2}}{2{e}_o}-{\theta}_L\right]}^{+}}{{\left(1-{\theta}_L\right)}^2} $$
(D3)

where 2e o is the maximal activity of a population (see Eqs. (A3), (A7), (A11) and (A15)). Eq. (D3) means that if the activity of the postsynaptic unit at time t and the average activity of the presynaptic unit in the near past are both above a given threshold, then the synchronization term gets a high value. The denominator is a normalizing factor to have values in the [0;1] range. Comparing Eqs. (D1) and (D3), we have

$$ {\tau}_{train}\frac{d{W}_{ij}^{L1L2}}{dt}=\left({W}_{\max }-{W}_{ij}^{L1L2}\right)sy{n}_{ij}^{L1L2}(t) $$
(D4)

with θ = 2e o θ L and \( {\gamma}_{ij}=\frac{W_{\max }-{W}_{ij}^{L1L2}}{{\left(2{e}_o-\theta \right)}^2} \).

In Eq. (D4) we assumed that the learning rate γ ij decreases progressively as the connection reaches its maximum value. Finally, at each step, once the values of W L1L2 ij for all the i,j pairs are updated, these values are normalized so that the sum of all connections entering each unit i has a fixed value W norm :

$$ {W}_{ij}^{L1L2}=\frac{W_{ij}^{L1L2}}{{\displaystyle {\sum}_j{W}_{ij}^{L1L2}}}{W}_{norm} $$
(D5)

Appendix E: Parameters assignment

The values of most parameters of the units (Table 1) are identical to those used in our previous works (Cona and Ursino 2013; Cona et al. 2012). In more detail, the only parameters of the units that changed in this work are the connection strengths from GABAA,slow interneurons to pyramidal neurons C ps , and from pyramidal neurons to GABAA,fast interneurons C fp . Both these changes were necessary to account for differences in the network architecture as commented in section Discussion.

The parameters concerning the maze, its exploration and the corresponding inputs (Table 3), such as the dimension of the receptive fields or the rat velocity, were mostly assigned following indications from real values measured in experimental settings (Dragoi and Buzsáki 2006). The value of the sensory input (G RF ) during training was assigned to have a significant activation of units in L1 as soon as the input enters the periphery of the receptive field. The value of neural activation during the imagination phase (S offset ) was chosen so that a few neurons in each simulation can be activated randomly by the internal noise, thus starting an imagined sequence; this value must not be excessively high, to allow inhibition from slow GABAergic synapses to induce a refractory period which avoids the presence of reflected activation. The values of the inputs used for the exploration of the familiar maze were arbitrarily assigned as G RF /2 and S offset /2 respectively, so that exploration of the familiar maze can be considered an average behavior between the exploration of the unknown maze (when the rat is just paying consideration to the external world) and the imagination of the known maze (when neuron activity completely depends on the internal dynamics). This simple choice provided satisfactory results that agree with what was expected from this modality.

The values of the long range connections (Table 4) were mostly inherited from our previous work (Cona and Ursino 2013). It should be noted that these connection strengths have been multiplied by a factor of 36, since in the previous work each unit received connections from neural assemblies consisting of 36 units, while in the present work each neural assembly consists of a single unit.

Table 4 Parameters of long range connections

Further slight adjustments were then required to compensate for the differences in the connectivity pattern of the present network, such as the absence of the inter-layer connections between units encoding for different features of the same object and the connection matrix W L1L2 ij that is now symmetric.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cona, F., Ursino, M. A neural mass model of place cell activity: theta phase precession, replay and imagination of never experienced paths. J Comput Neurosci 38, 105–127 (2015). https://doi.org/10.1007/s10827-014-0533-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-014-0533-5

Keywords

Navigation