Skip to main content
Log in

Dynamic branching in a neural network model for probabilistic prediction of sequences

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

Abstract

An important function of the brain is to predict which stimulus is likely to occur based on the perceived cues. The present research studied the branching behavior of a computational network model of populations of excitatory and inhibitory neurons, both analytically and through simulations. Results show how synaptic efficacy, retroactive inhibition and short-term synaptic depression determine the dynamics of selection between different branches predicting sequences of stimuli of different probabilities. Further results show that changes in the probability of the different predictions depend on variations of neuronal gain. Such variations allow the network to optimize the probability of its predictions to changing probabilities of the sequences without changing synaptic efficacy.

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.

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

Similar content being viewed by others

Availability of data and material

Not applicable.

Code availability

Available at https://github.com/elifkoksal/dynamicBranching

Notes

  1. This is the simplest way to form graphs of overlapping patterns. Preliminary results seem to indicate that the model can catch latching behavior in more general situations (overlaps with variable size). This is forthcoming work.

References

  • Aguilar, C., Chossat, P., Krupa, M., et al. (2017). Latching dynamics in neural networks with synaptic depression. PLoS ONE, 12(8), e0183710.

    Article  PubMed  PubMed Central  Google Scholar 

  • Albrengues, C., Lavigne, F., Aguilar, C., et al. (2019). Linguistic processes do not beat visuo-motor constraints, but they modulate where the eyes move regardless of word boundaries: Evidence against top-down word-based eye-movement control during reading. PLoS ONE, 14(7), e0219666.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Amari, S. (1972). Characteristics of random nets of analog neuron-like elements. IEEE Transactions on Systems, Man, and Cybernetics, 5, 643–57.

    Google Scholar 

  • Amit, D. J., & Brunel, N. (1997). Model of global spontaneous activity and local structured activity during delay periods in the cerebral cortex. Cerebral Cortex, 7(3), 237–52.

    Article  CAS  PubMed  Google Scholar 

  • Amit, D. J., Brunel, N., & Tsodyks, M. V. (1994). Correlations of cortical hebbian reverberations: Theory versus experiment. Journal of Neuroscience, 14, 6435–45.

    Article  CAS  PubMed  Google Scholar 

  • Amit, D. J., Bernacchia, A., & Yakovlev, V. (2003). Multiple-object working memory-a model for behavioral performance. Cerebral Cortex, 13(5), 435–43.

    Article  CAS  PubMed  Google Scholar 

  • Bak, P., & Chialvo, D. R. (2001). Adaptive learning by extrernal dynamics and negative feedback. Physical Review E, 63(3), 031912.

    Article  CAS  Google Scholar 

  • Bak, P., & Paczuski, M. (1995). Complexity, contingency and criticality. Proceedings of the National Academy of Sciences, 92(15), 6669–96.

    Article  Google Scholar 

  • Bastos, A. M., Lundqvist, M., Waite, A. S., et al. (2020). Layer and rhythm specificity for predictive routing. Proceedings of the National Academy of Sciences, 117(49), 31459–69.

    Article  CAS  Google Scholar 

  • Bell, A. H., Summerfield, C., Morin, E. L., et al. (2016). Encoding of stimulus probability in macaque inferior temporal cortex. Current Biology, 26(17), 2280–90.

    Article  CAS  PubMed  Google Scholar 

  • Bienenstock, E., & Lehmann, D. (1998). Regulated criticality in the brain? Advances in Complex Systems, 1(4), 361–84.

    Article  Google Scholar 

  • Bliss, T. V., & Collingridge, G. L. (1993). A synaptic model of memory: long-term potentiation in the hippocampus. Nature, 361, 31–39.

    Article  CAS  PubMed  Google Scholar 

  • Bliss, T. V., & Lomo, T. (1973). Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. Journal of Physiology, 232, 331–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Brunel, N. (1996). Hebbian learning of context in recurrent neural networks. Neural Computation, 15(8), 1677–710.

    Article  Google Scholar 

  • Brunel, N., & Lavigne, F. (2009). Semantic priming in a cortical network model. Journal of Cognitive Neuroscience, 21(2300-19).

  • Brunel, N., & Wang, X. J. (2001). Effects of neuromodulation in a cortical network model of object working memory dominated by recurrent inhibition. Journal of Computational Neuroscience, 11(1), 63–85.

    Article  CAS  PubMed  Google Scholar 

  • Bunge, S. A., Kahn, I., Wallis, J. D., et al. (2003). Neural circuits subserving the retrieval and maintenance of abstract rules. Journal of Neuophysiology, 90(5), 3419–28.

    Article  Google Scholar 

  • Busemeyer, J. R., & Townsend, J. T. (1993). Decision field theory: a dynamic-cognitive approach to decision making in an uncertain environment. Psychological Review, 100(3), 432.

    Article  CAS  PubMed  Google Scholar 

  • Chen, B., & Miller, P. (2020). Attractor-state itinerancy in neural circuits with synaptic depression. Journal of Mathematical Neuroscience, 10(1), 1–19.

    Article  Google Scholar 

  • Chialvo, D. R. (2010). Emergent complex neural dynamics. Nature Physics, 6(10), 744–50.

    Article  CAS  Google Scholar 

  • Cocchi, L., Gollo, L. L., Zalesky, A., et al. (2017). Criticality in the brain: A synthesis of neurobiology, models and cognition. Progress in Neurobiology, 158, 132–52.

    Article  PubMed  Google Scholar 

  • Dasgupta, I., Schulz, E., Tenenbaum, J. B., et al. (2020). A theory of learning to infer. Psychological Review, 127(3), 412.

    Article  PubMed  Google Scholar 

  • Dehghani, N., Peyrache, A., Telenczuk, B., et al. (2016). Dynamic balance of excitation and inhibition in human and monkey neocortex. Scientific Reports, 6(1), 23176.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Delaney-Busch, N., Morgan, E., Lau, E. F., et al. (2017). Comprehenders rationally adapt semantic predictions to the statistics of the local environment: A Bayesian model of trial-by-trial N400 amplitudes. In 39th Annual Conference of the Cognitive Science Society. London, England.

  • DeLong, K. A., Urbach, T. P., & Kutas, M. (2005). Probabilistic word pre-activation during language comprehension inferred from electrical brain activity. Nature Neuroscience, 8(8), 1117–21.

    Article  CAS  PubMed  Google Scholar 

  • Ding, N., Melloni, L., Zhang, H., et al. (2015). ortical tracking of hierarchical linguistic structures in connected speech. Nature Neuroscience, 19(1), 158–64.

    Article  PubMed  PubMed Central  Google Scholar 

  • Erickson, C. A., & Desimone, R. (1999). Responses of macaque perirhinal neurons during and after visual stimulus association learning. Journal of Neuroscience, 19(10404-16).

  • FitzGerald, T., Dolan, R. J., & Friston, K. J. (2015). Dopamine, reward learning, and active inference. Frontiers in Computational Neuroscience, 9, 136.

    Article  PubMed  PubMed Central  Google Scholar 

  • Friston, K. J., Shiner, T., FitzGerald, T., et al. (2012). Dopamine, affordance and active inference. PLoS Computational Biology, 8(1), e1002327.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Fujimichi, R., Naya, Y., Koyano, K. W., et al. (2010). Unitized representation of paired objects in area 35 of the macaque perirhinal cortex. Europen Journal of Neuroscience, 32(4), 659–67.

    Article  Google Scholar 

  • Fuster, J. M., & Alexander, G. E. (1971). Neuron activity related to short-term memory. Science, 173(3997), 652–4.

    Article  CAS  PubMed  Google Scholar 

  • Gershman, S. J. (2019). How to never be wrong. Psychonomic Bulletin and Review, 26(1), 13–28.

    Article  PubMed  Google Scholar 

  • Gershman, S. J., & Uchida, N. (2019). Believing in dopamine. Nature Review. Neuroscience, 20(11), 703–14.

    CAS  PubMed  Google Scholar 

  • Gochin, P. M., Colombo, M., Dorfman, G. A., et al. (1994). Neural ensemble coding in inferior temporal cortex. Journal of Neuophysiology, 71, 2325–37.

    Article  CAS  Google Scholar 

  • Hahnloser, R. H., Kozhevnikov, A. A., & Fee, M. S. (2002). An ultra-sparse code underliesthe generation of neural sequences in a songbird. Nature, 419(6902), 65–70.

    Article  CAS  PubMed  Google Scholar 

  • Harnal, H., & Giraud, A. L. (2012). Cortical oscillations and sensory predictions. Trends in Cognitive Science, 16, 390–8.

    Article  Google Scholar 

  • Harvey, C. D., Coen, P., & Tank, D. W. (2012). Choice-specific sequences in parietal cortex during a virtual-navigation decision task. Nature, 484(7392), 62–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Hebb, D. (1949). The Organization of Behavior: A Neuropsychological Theory. New York, NY: Wiley and Sons.

    Google Scholar 

  • Hopfield, J. J. (1982). Neural networks and physical systems with emergent collective computational abilities. Proceedings of the National Academy of Sciences, 79(8), 2554–58.

    Article  CAS  Google Scholar 

  • Hung, C., Kreiman, G., Poggio, J., & DiCarlo, T. (2005). Fast read-out of object information in inferior temporal cortex. Science, 310, 863–6.

  • Hutchison, K., Heap, S., Neely, J., et al. (2014). Attentional control and asymmetric associative priming. Journal of Experimental Psychology: Learning, Memory, and Cognition, 40(3), 844–56.

    PubMed  Google Scholar 

  • Ison, M. J., Quian, Q. R., & Fried, I. (2015). Rapid encoding of new memories by individual neurons in the human brain. Neuron, 87(1), 220–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Kamiński, J., Sullivan, S., Chung, J. M., et al. (2017). Persistently active neurons in human medial frontal and medial temporal lobe support working memory. Nature neuroscience, 20(4), 590–601.

    Article  PubMed  PubMed Central  Google Scholar 

  • Kang, C. J., & Treves, A. (2019). The challenge of taming a latching network near criticality. The Functional Role of Critical Dynamics in Neural Systems (vol. 11, p. 81–94). Springer Series on Bio- and Neurosystems.

  • Kirkwood, A., & Bear, M. F. (1994). Homosynaptic long-term depression in the visual cortex. Neuroscience, 14, 3404–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Köksal Ersöz, E., Aguilar, C., Chossat, P., et al. (2020). Neuronal mechanisms for sequential activation of memory items: Dynamics and reliability. PLoS ONE, 15(4), e0231165.

    Article  PubMed  PubMed Central  Google Scholar 

  • Körding, K. P., & Wolpert, D. M. (2004). Bayesian integration in sensorimotor learning. Nature, 427(6971), 244–7.

    Article  PubMed  Google Scholar 

  • Kornblith, S., Quian Quiroga, R., Koch, C., et al. (2017). Persistent single-neuron activity during working memory in the human medial temporal lobe. Current Biology, 27(7), 1026–1032. https://doi.org/10.1016/j.cub.2017.02.013

    Article  CAS  PubMed  Google Scholar 

  • Kreiman, G., Hung, C. P., Kraskov, A., et al. (2006). Object selectivity of local field potentials and spikes in the macaque inferior temporal cortex. Neuron, 49, 433–45.

    Article  CAS  PubMed  Google Scholar 

  • Kutas, M., DeLong, K. A., & Smith, N. J. (2011). A look around at what lies ahead: Prediction and predictability in language processing. In Predictions in the Brain: Using Our Past to Generate a Future (pp. 190–207). Oxford University Press.

  • Lam, N., Borduqui, T., Hallak, J., et al. (2021). Effects of altered excitation-inhibition balance on decision making in a cortical circuit model. Journal of Neuroscience. https://doi.org/10.1523/JNEUROSCI.1371-20.2021

    Article  PubMed  Google Scholar 

  • Lau, E. F., Holcomb, P. J., & Kuperberg, G. R. (2013). Dissociating N400 effects of prediction from association in single-word contexts. Journal of Cognitive Neuroscience, 25(3), 484–502.

    Article  PubMed  Google Scholar 

  • Lavigne, F. (2004). Aim networks: autoincursive memory networks for anticipation toward learned goals. International Journal of Computing Anticipatory Systems, 14, 196–214.

    Google Scholar 

  • Lavigne, F., & Darmon, N. (2008). Dopaminergic neuromodulation of semantic priming in a cortical network model. Neuropsychologia, 46, 3074–87.

    Article  PubMed  Google Scholar 

  • Lavigne, F., & Denis, S. (2002). Neural network modeling of learning of contextual constraints on adaptive anticipations. International Journal of Computing Anticipatory Systems, 12, 253–68.

    Google Scholar 

  • Lavigne, F., Vitu, F., & d’Ydewalle, G. (2000). The influence of semantic context on initial eye landing sites in words. Acta Psychologica, 104(2), 191–214.

    Article  CAS  PubMed  Google Scholar 

  • Lavigne, F., Dumercy, L., & Darmon, N. (2011). Determinants of multiple semantic priming: A meta-analysis and spike frequency adaptive model of a cortical network. Journal of Cognitive Neuroscience, 23(6), 1447–74.

    Article  PubMed  Google Scholar 

  • Lavigne, F., Dumercy, L., Chanquoy, L., et al. (2012). Dynamics of the semantic priming shift: Behavioral experiments and cortical network model. Cogitive Neurodynamics, 6(6), 467–83.

    Article  Google Scholar 

  • Lavigne, F., Chanquoy, L., Dumercy, L., et al. (2013). Early dynamics of the semantic priming shift. Advances in Cognitive Psychology, 9(1), 1–14.

    Article  PubMed  PubMed Central  Google Scholar 

  • Lavigne, F., Avnaïm, F., & Dumercy, L. (2014). Inter-synaptic learning of combination rules in a cortical network model. Frontiers in Psychology, 5, 842.

    Article  PubMed  PubMed Central  Google Scholar 

  • Lazartigues, L., Mathy, F., & Lavigne, F. (2021). Statistical learning of unbalanced exclusive-or temporal sequences in humans. PLoS ONE, 16(2), e0246826.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Lerner, I., & Shriki, O. (2014). Internally and externally driven network transitions as a basis for automatic and strategic processes in semantic priming: theory and experimental validation. Frontiers in Psychology, 5(314), 00314.

    Google Scholar 

  • Lerner, I., Bentin, S., & Shriki, O. (2012). Spreading activation in an attractor network with latching dynamics: automatic semantic priming revisited. Cognitive Science, 36, 1339–82.

    Article  PubMed  PubMed Central  Google Scholar 

  • Levina, A., & Herrmann, M. (2006). Dynamical synapses give rise to a power-law distribution of neuronal avalanches. In Advances in Neural Information Processing Systems (pp 771–78. 18). MIT Press, Cambridge, MA, USA.

  • Luka, B. J., & Van Petten, C. (2014). Prospective and retrospective semantic processing: Prediction, time, and relationship strength in event-related potentials. Brain and Language, 135, 115–29.

    Article  PubMed  Google Scholar 

  • Magnasco, M. O., Piro, O., & Cecchi, G. A. (2009). Self-tuned critical anti-Hebbian networks. Physical Review Letters, 102(25), 258102.

    Article  PubMed  Google Scholar 

  • Messinger, A., Squire, L., Zola, S. M., et al (2001). Neuronal representations of stimulus associations develop in the temporal lobe during learning. Proceedings of the National Academy of Sciences, 98(12239-44).

  • Miller, E. K. (1999). The prefrontal cortex: complex neural properties for complex behavior. Neuron, 22, 15–17.

    Article  CAS  PubMed  Google Scholar 

  • Miller, K. D., & Fumarola, F. (2012). Mathematical equivalence of two common forms of firing rate models of neural networks. Neural Computation, 24(1), 25–31.

    Article  PubMed  Google Scholar 

  • Minier, L., Fagot, J., & Rey, A. (2016). The temporal dynamics of regularity extraction in non-human primates. Cognitive Science, 40(4), 1019–30.

    Article  PubMed  Google Scholar 

  • Miyashita, Y. (1988). Neuronal correlate of visual associative long-term memory in the primate temporal cortex. Nature, 335.

  • Miyashita, Y., & Chang, H. S. (1988). Neuronal correlate of pictorial short-term memory in the primate temporal cortex. Nature, 331, 68–70.

    Article  CAS  PubMed  Google Scholar 

  • Mongillo, G., Amit, D. J., & Brunel, N. (2003). Retrospective and prospective persistent activity induced by hebbian learning in a recurrent cortical network. Europen Journal of Neuroscience, 18(7), 2011–24.

    Article  Google Scholar 

  • Mongillo, G., Rumpel, S., & Loewenstein, Y. (2018). Inhibitory connectivity defines the realm of excitatory plasticity. Nature Neuroscience, 21(10), 1463–70.

    Article  CAS  PubMed  Google Scholar 

  • Muhammad, R., Wallis, J. D., & Miller, E. K. (2006). A comparison of abstract rules in the prefrontal cortex, premotor cortex, inferior temporal cortex and striatum. Journal of Cognitive Neuroscience, 18(6), 974–89.

    Article  PubMed  Google Scholar 

  • Naya, Y., Yoshida, M., & Miyashita, Y. (2001). Backward spreading of memory-retrieval signal in the primate temporal cortex. Science, 291(5504), 661–64.

    Article  CAS  PubMed  Google Scholar 

  • Naya, Y., Yoshida, M., Takeda, M., et al. (2003). Delay-period activities in two subdivisions of monkey inferotemporal cortex during pair association memory task. The European Journal of Neuroscience, 18, 2915–8.

    Article  PubMed  Google Scholar 

  • Neely, J. H. (1991). Semantic priming effects in visual word recognition: A selective review of current findings and theories. In Basic Processes in Reading: Visual Word Recognition. Lawrence Erlbaum Associates, Inc. (pp. 264–336).

  • Pereira, U., & Brunel, N. (2020). Unsupervised learning of persistent and sequential activity. Frontiers in computational neuroscience, 13, 97.

    Article  PubMed  PubMed Central  Google Scholar 

  • Quian, Q. R. (2012). Concept cells: the building blocks of declarative memory functions. Nature Review Neuroscience, 13, 587–97.

    Article  Google Scholar 

  • Quian, Q. R. (2016). Neuronal codes for visual perception and memory. Neuropsychologia, 83, 227–41.

    Article  Google Scholar 

  • Quian, Q. R., & Kreiman, G. (2010). Measuring sparseness in the brain: comment on Bowers. Psychological Review, 117, 291–99.

    Article  Google Scholar 

  • Rainer, G., Rao, S. C., & Miller, E. K. (1999). Prospective coding for objects in primate prefrontal cortex. Journal of Neuroscience, 19, 5493–5505.

    Article  CAS  PubMed  Google Scholar 

  • Reddy, L., Poncet, M., Self, M. W., et al. (2015). Learning of anticipatory responses in single neurons of the human medial temporal lobe. Nature communications, 6(1), 1–8.

    Article  CAS  Google Scholar 

  • Rolls, E. T., & Tovee, M. J. (1995). Sparseness of the neuronal representation of stimuli in the primate temporal visual cortex. Journal of Neuophysiology, 73(2), 713–26.

    Article  CAS  Google Scholar 

  • Rolls, E. T., Loh, M., Deco, G., et al. (2008). Computational models of schizophrenia and dopamine modulation in the prefrontal cortex. Nature Review Neuroscience, 9, 696.

    Article  CAS  Google Scholar 

  • Sakai, K., & Miyashita, Y. (1991). Neural organization for the long-term memory of paired associates. Nature, 354, 152–55.

    Article  CAS  PubMed  Google Scholar 

  • Schaal, S., Mohajerian, P., & Ijspeert, A. (2007). Dynamics systems vs. optimal control-a unifying view. Progress in Brain Research, 165, 425–445.

    Article  PubMed  Google Scholar 

  • Tamura, H., & Tanaka, K. (2001). Visual response properties of cells in the ventral and dorsal parts of the macaque inferotemporal cortex. Cerebral Cortex, 11, 384–99.

    Article  CAS  PubMed  Google Scholar 

  • Tanaka, K. (1996). Inferotemporal cortex and object vision. Annual Review of Neuroscience, 19, 109–39.

    Article  CAS  PubMed  Google Scholar 

  • Tanaka, K. (2003). Columns for complex visual object features in the inferotemporal cortex: clustering of cells with similar but slightly different stimulus selectivities. Cereb Cortex, 13, 90–99.

    Article  PubMed  Google Scholar 

  • Thurley, K., Senn, W., & Luscher, H. R. (2008). Dopamine increases the gain of the input-output response of rat prefrontal pyramidal neurons. Journal of Neurophysiology, 99(6), 2985–97.

    Article  PubMed  Google Scholar 

  • Tremblay, R., Lee, S., & Rudy, B. (2016). GABAergic interneurons in the neocortex: From cellular properties to circuits. Neuron, 91, 260–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Tsao, D. Y., Freiwald, W. A., Tootell, R. B., et al. (2006). A cortical region consisting entirely of face-selective cells. Science, 311(5761), 670–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Tsodyks, M. V. (1990). Hierarchical associative memory in neural networks with low activity level. Modern Physics Letters B, 4, 259–65.

    Article  Google Scholar 

  • Tsodyks, M. V., & Markram, H. (1997). The neural code between neocortical pyramidal neurons depends on neurotransmitter release probability. Proceedings of the National Academy of Sciences, 94, 719–23.

    Article  CAS  Google Scholar 

  • Van Petten, C. (2014). Examining the N400 semantic context effect item-by-item: Relationship to corpus-based measures of word co-occurrence. International Journal of Psychophysiology, 94, 407–19.

    Article  PubMed  Google Scholar 

  • Vander Weele, C. M., Siciliano, C. A., Matthews, G. A., et al. (2018). Dopamine enhances signal-to-noise ratio in cortical-brainstem encoding of aversive stimuli. Nature, 563, 397–401.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Varela, J. A., Sen, K., Gibson, J., et al. (1997). A quantitative description of short-term plasticity at excitatory synapses in layer 2/3 of rat primary visual cortex. Journal of Neuroscience, 17(20), 7926–40.

    Article  CAS  PubMed  Google Scholar 

  • Volman, V., Behrens, M. M., & Sejnowski, T. J. (2011). Downregulation of parvalbumin at cortical GABA synapses reduces network gamma oscillatory activity. Journal of Neuroscience, 31(49), 18137–48.

    Article  CAS  PubMed  Google Scholar 

  • Wallis, J. D., & Miller, E. K. (2003). From rule to response: neuronal processes in the premotor and prefrontal cortex. Journal of Neuophysiology, 90(3), 1790–806.

    Article  Google Scholar 

  • Wallis, J. D., Anderson, K. C., & Miller, E. K. (2001). Single neurons in prefrontal cortex encode abstract rules. Nature, 411(6840), 953–6.

    Article  CAS  PubMed  Google Scholar 

  • Wang, X. (2002). Probabilistic decision making by slow reverberation in cortical circuits. Neuron, 36, 955–68.

    Article  CAS  PubMed  Google Scholar 

  • Weinberger, N. M. (1998). Physiological memory in primary auditory cortex: characteristics and mechanisms. Neurobiology of Learning and Memory, 70(1–2), 226–51.

    Article  CAS  PubMed  Google Scholar 

  • Willems, R. M., Frank, S. L., Nijhof, A. D., et al. (2015). Prediction during natural language comprehension. Cerebral Cortex, 26(6), 2506–16.

    Article  PubMed  Google Scholar 

  • Wilson, H. R., & Cowan, J. D. (2012). Excitatory and inhibitory interactions in localized populations of model neurons. Biophysics Journal, 12(1), 1–24.

    Article  CAS  Google Scholar 

  • Wirth, S., Yanike, M., Frank, L. M., et al. (2003). Single neurons in the monkey hippocampus and learning of new associations. Science, 300(5625), 1578–81.

    Article  CAS  PubMed  Google Scholar 

  • Yakovlev, V., Fusi, S., E., B., et al. (1998). nter-trial neuronal activity in inferior temporal cortex: A putative vehicle to generate long-term visual associations. Nature Neuroscience, 1(4), 310–17.

  • Yizhar, O., Fenno, L., Prigge, M., et al. (2011). Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature, 477(7363), 171–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Yoshida, M., Naya, Y., & Miyashita, Y. (2003). Anatomical organization of forward fiber projections from area TE to perirhinal neurons representing visual long-term memory in monkeys. Proceedings of the National Academy of Sciences, 100(4257–62).

  • Young, M., & Yamane, S. (1992). Sparse population coding of faces in the inferotemporal cortex. Science, 256(5061), 1327–31.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors gratefully thank the reviewers for their invaluable comments to improve the paper.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the study conception, design, writing. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Elif Köksal Ersöz.

Ethics declarations

Ethics approval

Non applicable.

Consent to participate

Not applicable.

Consent for publication

Not applicable

Conflicts of interest

The authors declare that they have no conflict of interest.

Additional information

Action Editor: Nicolas Brunel

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: The latching dynamics

Appendix: The latching dynamics

1.1 A.1 Sketch of the analysis in Köksal Ersöz et al. (2020)

A sequence of \(N-1\) learned patterns \(\xi ^i=(\xi ^i_1\dots ,\xi ^i_{N-1})\) was considered in a network of N units, where \(\xi ^k_i=\xi ^k_{i+1}=1\) and the other coordinates are 0. These patterns sit on vertices of the hypercube \([0,1]^N\) and the trajectories of the activity Eq. (5) lie within the hypercube. The analysis will be simplified by the presence of the factor \(x_i(1-x_i)\) in (5), which forces the trajectories starting on one face of the hypercube to lie within that face, whatever the dimension of the face. In particular, the vertices are the steady-states of (5). The evolution of the system is governed by the activity Eq. (5) and the STD is driven by (1). The corresponding connectivity matrix is (7), according to the learning rule (3). In the following we sketch without proofs the analysis in (Köksal Ersöz et al., 2020).

The first question is about the stability of learned patterns \(\xi ^i\). This is determined by the sign of the eigenvalues at \(\xi ^i\) of the Jacobian matrix of the system (5), which we write \(\sigma ^i_j\) (\(j=1,\dots ,N\)). These eigenvalues are easily computed because the edges emanating from \(\xi ^i\) on the hypercube are the eigendirections. It can be seen that all eigenvalues with \(j<i\) or \(j>i+1\) are negative if the condition \(2\lambda +I>S\) is satisfied (\(S=(1+\rho )^{-1}<1\) is the value towards which the synaptic strength \(s_i\) decays as time elapses). The two remaining eigenvalues are

$$\begin{aligned} \begin{aligned}\sigma ^i_i&=\mu +2\lambda +I-ms_i-s_{i+1}\\\sigma ^i_{i+1}&=\mu +2\lambda +I-s_i-m's_{i+1} \end{aligned} \end{aligned}$$
(11)

where \(m=1\) if \(i=1\), \(m'=1\) if \(i=N-1\), and \(m=m'=2\) for all other values of i.

Without STD the learned patterns \(\xi ^i\) with \(1<i<N-1\) are stable under the mild condition \(S<2\lambda +I<3-\mu\). When \(i=1\) or \(N-1\) the condition becomes \(S<2\lambda +I<2-\mu\), which is more restrictive. We assumed in Köksal Ersöz et al. (2020) that \(I=0\). The results are still valid with \(I>0\) small enough. In this study we have set \(I=0\).

If STD is on, \(s_i\) and \(s_{i+1}\) decay towards \(0<S<1\), so that eventually \(\sigma ^i_i\) or \(\sigma ^i_{i+1}\) may become positive at finite time. When \(i=1\), (11) shows that \(\sigma ^1_1>\sigma ^1_2\), therefore \(\xi ^i\) becomes first unstable and the direction of this instability is \(x_1\). When \(i>1\), \(m=m'\) in (11), so that \(\sigma ^i_i\) and \(\sigma ^i_{i+1}\) may become positive simultaneously. However, thanks to the previous switch \(\xi ^{i-1}\rightarrow \xi ^i\), \(s_i\) has been decaying towards S while \(s_{i+1}\) was still close to 1 (\(x_{i+1}=0\) at \(\xi ^{i-1}\)). Hence in any case \(\xi ^i\) is first destabilized along \(x_i\).

We claim that “typical” trajectories starting near \(\xi ^i\) will converge towards \(\xi ^{i+1}\) by first decreasing \(x_i\) towards 0, so that the trajectory converges for some time towards the intermediate state \(\hat{\xi }^i=(0,1,0,\dots ,0)\) (which represents the “overlap” between the learned patterns \(\xi ^i\) and \(\xi ^{i+1}\)), then increasing \(x_{i+2}\) towards 1, so that \(\xi ^{i+1}\) is reached. Noise is an essential ingredient of this behavior as we will see. The proof of the claim relies on the following properties of the system: (i) the synaptic variables \(s_j\) are slow compared to the activity variables \(x_j\) and can be seen, in the limit of “stationary variables”, as bifurcation parameters for the \(x_j\)’s. This is the idea of “dynamic bifurcation”. These bifurcations drive the dynamics of the activity variables. (ii) The analysis can be restricted to the flow-invariant 2 dimensional face \(F_i\) on the hypercube, with vertices \(\xi ^{i}\), \(\hat{\xi }^{i}\) and \(\xi ^{i+1}\). In \(F_i\), \(x_{i+1}=1\), \(x_j=0\) when \(j<i\) or \(j>i+2\) and the coordinates are \(x_i\), \(x_{i+2}\). In these coordinates \(\xi ^i=(1,0)\), \(\xi ^{i+1}=(0,1)\) and \(\hat{\xi }^i=(0,0)\).

Without synaptic plasticity, the learned patterns are stable and the dynamics in \(F_i\) looks like Fig. 9. When STD is on, synaptic strengths \(s_i\) of active neurons diminish slowly, which triggers successive transitions in \(F_i\). The proof of the claim is provided in Köksal Ersöz et al. (2020) and relies upon slow-fast dynamical systems theory. Here we only describe the possible scenarios.

Fig. 9
figure 9

A typical phase portrait of the flow in the square \(F_i\) in the absence of synaptic depression. Green spots correspond to the learned patterns. Other spots are unstable steady-states

The eigenvalue at \(\xi ^i\) along coordinate \(x_i\) is given in (11) and note that \(\hat{\xi }^i\) has a double eigenvalue in \(F_i\), given by

$$\begin{aligned} \hat{\sigma }^i_i=\hat{\sigma }^i_{i+2}=-\lambda +s_{i+1} \end{aligned}$$
(12)

Initially, \(\xi ^i\) is stable. Let \(t_{(1,0)}\) be the time at which \(ms_i(t_{(1,0)})+s_{i+1}(t_{(1,0)})=\mu +2\lambda\) so that \(\sigma ^i_i\) becomes positive, and let \(t_{0,0}\) be the time at which the initially positive eigenvalue \(\hat{\sigma }^i_i\) becomes negative. This happens when \(s_{i+1}(t_{(0,0)})=\lambda\). Depending on the order of these two critical times two scenarios can occur.

Scenario 1

\(t_{(1,0)}<t_{(0,0)}\). A (dynamic) bifurcation of a stable equilibrium occurs on the edge \(x_{i+2}=0\) from the equilibrium (1, 0). This equilibrium travels towards (0, 0), which is reached at time \(t_{(0,0)}\). A trajectory starting at initial time from the vicinity of \(\xi ^i=(1,0)\) will therefore follow the edge \(x_{i+2}=0\) until it arrives in a neighborhood of \(\hat{\xi }^i\).

Scenario 2

\(t_{(1,0)}>t_{(0,0)}\), so that an equilibrium bifurcates first from (0, 0) along the edge \(x_{i+2}=0\). The point (1, 0) is still stable but its basin of attraction shrinks. Then sufficient noise may allow the trajectory starting in the vicinity of (1, 0) to escape its basin of attraction and converge towards \(\hat{\xi }^i\).

In both scenarios, in order for the trajectory to converge to \(\xi ^{i+1}\), one must have \(\sigma ^{i+1}_{i+2}<0\), which by (11) requires \(s_{i+1}+m's_{i+2}>\mu +2\lambda\). Since \(s_{i+2}\) is still close to 1 this condition is easily satisfied, as was seen above, unless \(m'=1\). In this case \(i=N-1\), which implies that the last pattern of the chain can hardly be attained by the system.

Additional constraints shared by the two scenarios must be satisfied, notably that \(\hat{\xi }^i\) be stable in the transverse directions to \(F_i\). This gives the condition \(\mu <\lambda\) (see Fig. 10).

Fig. 10
figure 10

Snapshots of the transition \(\xi ^i\rightarrow \xi ^{i+1}\) in the square \(F_i\), in the scenarios 1 and 2. Bifurcated equilibria are blue. Red lines mark the boundary of the basin of attraction of \(\xi ^{i+1}\) and dashed circles mark the closest distance for a possible stochastic jump over it. Scenario 1: a \(\xi ^i\) is still stable. b Trajectory starting near \(\xi ^i\) follows the saddle which has dynamically bifurcated on the edge \(x_{i+2=0}\), until it reaches \(\hat{\xi }^i\). Then arbitrary small noise can suffice to allow trajectory to jump towards \(\xi ^{i+1}\). c \(\hat{\xi }^i\) has become stable by dynamic bifurcation of a saddle point on the edge \(x_{i+2=0}\), but trajectory can jump over its basin of attraction by effect of sufficiently strong noise and then converge to \(\xi ^{i+1}\). Scenario 2: a same as scenario 1. b A saddle has bifurcated on the edge \(x_{i+2}\), so that a trajectory starting near \(\xi ^i\) can only converge to \(\xi ^{i+1}\) by the effect of strong enough noise. c Eventually the trajectory can converge towards \(\hat{\xi }^i\) and the same situation as in scenario 1 holds

1.2 A.2 Latching dynamics in the case of n-ways branching

We analyze how the presence of a n-ways branching node in the graph of learned patterns affects the occurrence of latching dynamics. For the sake of clarity we concentrate on the example associated with the connectivity matrix (8), where the node \(\xi ^2\) has multiplicity 3 (Example 1, Sect. 2.2). The generalization to higher multiplicity is straightforward.

Now the equations for the neural activity are (10) where the term \(\nu _i\) is 0 if \(i\ne 3\) and \(\nu _3=\lambda\). Along chains which do not encounter the branching “choice” \(\xi ^2\rightarrow \xi ^3\) or \(\xi ^2\rightarrow \xi ^5\), the latching dynamics analysis proceeds as sketched in A.1. Indeed the higher excitation at unit 3, which in the factor of \(x_i(1-x_i)\) in (10) reads \(3s_3x_3\), is compensated by the higher inhibition \(-3\lambda x_3\). We therefore concentrate on the piece of the graph defined by the multi node \(\xi ^2\) and its adjacent patterns \(\xi ^3\) and \(\xi ^5\).

We expect \(\xi ^2\) becomes first unstable along its coordinate \(x_2\), converging towards the intermediate state \(\hat{\xi }^2=(0,0,1,0,\dots ,0)\). But then either \(x_4\) or \(x_6\) should rise to 1 (transitions \(\xi ^2\rightarrow \xi ^3\) or \(\xi ^2\rightarrow \xi ^5\). Therefore we need to consider the cube \(F_3\) defined by setting \(x_3=1\) and \(x_j=0\) for \(j\ne 2,4,6\). In \(F_3\), \(\xi ^2=(1,0,0)\), \(\xi ^3=(0,1,0)\), \(\xi ^5=(0,0,1)\) and \(\hat{\xi }^2=(0,0,0)\). The equations in \(F_3\) are

$$\begin{aligned} \begin{aligned} \dot{x_2} =&\ x_2(1-x_2)(-\mu x_2 -\lambda (x_2+x_4+x_6+1)\\&+ 2s_2x_2+s_3) \\ \dot{x_4} =&\ x_4(1-x_4)(-\mu x_4 -\lambda (x_2+x_4+x_6+1) \\&+ s_3+2s_4x_4) \\ \dot{x_6} =&\ x_6(1-x_6)(-\mu x_6 -\lambda (x_2+x_4+x_6+1) \\&+ s_3+2s_6x_6) \end{aligned} \end{aligned}$$
(13)

Observe that restricting further the equations to the squares defined by \(x_4=0\) or \(x_6=0\) leads to systems completely analogous to those in \(F_i\) in A.1, so that the same analysis applies to both faces. Moreover the transverse eigenvalues to these faces at \(\xi ^3\) and \(\xi ^5\) are, in the notations of A.1, \(\sigma ^3_6=\sigma ^5_4=-2\lambda +s_3<0\). Therefore the dynamics will closely follow one of the faces and the transitions \(\xi ^2\rightarrow \xi ^3\) and \(\xi ^2\rightarrow \xi ^5\) will occur with equal probability. Figure 11 illustrates the global dynamics in \(F_3\) in a case corresponding to scenarios 1-(c) or 2-(c) in Fig. 10.

Fig. 11
figure 11

A sketch of the dynamics in \(F_3\) corresponding to Fig. 10, scenario 1-(c) or scenario 2-(c)

1.3 A.3 Relation between the number of excitatory connections and self-inhibition

Consider that the system is in \(F_3\) and \((x_2, x_4, x_6) = (1,0,0)\), which corresponds to \(\xi ^2\). As \(s_i\) variables varies in slow time, the system (13) will undergo several bifurcations that leading to either \(\xi ^2 \rightarrow \hat{\xi }^2 \rightarrow \xi ^3\) or \(\xi ^2 \rightarrow \hat{\xi }^2 \rightarrow \xi ^5\) transitions, with equal probabilities. The conditions for the \(\xi ^2 \rightarrow \hat{\xi }^2 \rightarrow \xi ^3\) transition can be derived form the eigenvalue expressions:

  • Initially stable equilibrium point \((x_2, x_4, x_6) = (1,0,0)\) becomes unstable at \(t = t^3_{(1, 0, 0)}\):

    $$2 s_2(t^3_{(1, 0, 0)}) + s_3(t^3_{(1, 0, 0)}) = \mu + 2\lambda .$$
  • Initially unstable equilibrium point \((x_2, x_4, x_6) = (0,0,0)\) becomes stable at \(t = t^3_{(0, 0, 0)}\):

    $$s_3(t^3_{(0, 0, 0)}) = \lambda .$$

The next transition is \(\xi ^3 \rightarrow \hat{\xi }^3 \rightarrow \xi ^4\) in \(F_4\) defined by setting \(x_4 = 1\) and \(x_j = 0\) for \(j\ne 3,5\). In \(F_4\) the equations of the neural activity are reduced to

$$\begin{aligned} \begin{aligned} \dot{x_3} =&\ x_3(1-x_3)(-\mu x_3 -\lambda (x_3+x_5+1)\\&-\nu _3x_3+ 3s_3x_3+s_4)\\ \dot{x_5} =&\ x_5(1-x_5)(-\mu x_5 -\lambda (x_3+x_5+1)\\&-\nu _5x_5 +s_4+s_5x_5) \end{aligned} \end{aligned}$$
(14)

The conditions for the transition \(\xi ^3 \rightarrow \hat{\xi }^3 \rightarrow \xi ^4\) are

  • Initially stable equilibrium point \((x_3, x_5) = (1,0)\) becomes unstable at \(t = t^4_{(1, 0)}\):

    $$3 s_3(t^4_{(1, 0)}) + s_4(t^4_{(1, 0)}) = \mu + 2\lambda + \nu _3.$$
  • Initially unstable equilibrium point \((x_3, x_5) = (0,0)\) becomes stable at \(t = t^4_{(0, 0)}\):

    $$s_4(t^4_{(0, 0)}) = \lambda .$$

Assume that \(t^3_{(1, 0, 0)} < t^3_{(0, 0, 0)}\) and \(t^4_{(1, 0)} < t^4_{(0, 0)}\). To have equally distributed pattern duration along a sequence, the \(s_i\) values at the bifurcation moments should read \(s_2(t^3_{(1, 0, 0)}) = s_3(t^4_{(1, 0)})\) and \(s_3(t^3_{(1, 0, 0)}) = s_4(t^4_{(1, 0)})\). With this restriction, the following algebraic condition can be deduced:

$$\begin{aligned} \begin{aligned} 2 s_3(t^4_{(1, 0)}) + s_4(t^4_{(1, 0)})= & {} \mu + 2\lambda ,\\ 3 s_3(t^4_{(1, 0)}) + s_4(t^4_{(1, 0)})= & {} \mu + 2\lambda + \nu _3, \end{aligned} \end{aligned}$$

that is

$$\begin{aligned} s_3(t^4_{(1, 0)})= & {} \nu _3. \end{aligned}$$

More generally, the relation between the number excitatory connections \(d_i\) and the level of self-inhibition \(\nu _i\) is given by:

$$\begin{aligned} (d_i-2) s_i (t_{(1, 0)})= & {} \nu _i, \end{aligned}$$

with \(s_i (t_{(1, 0)})\) being the value of \(s_i\) at the moment where a learned pattern changes stability. This implies that to succeed in sequential activation between the learned patterns, over-excitation imposed by the learning should be balanced by self-inhibition.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Köksal Ersöz, E., Chossat, P., Krupa, M. et al. Dynamic branching in a neural network model for probabilistic prediction of sequences. J Comput Neurosci 50, 537–557 (2022). https://doi.org/10.1007/s10827-022-00830-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-022-00830-y

Keywords

Navigation