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:

Neural substrate of dynamic Bayesian inference in the cerebral cortex

Abstract

Dynamic Bayesian inference allows a system to infer the environmental state under conditions of limited sensory observation. Using a goal-reaching task, we found that posterior parietal cortex (PPC) and adjacent posteromedial cortex (PM) implemented the two fundamental features of dynamic Bayesian inference: prediction of hidden states using an internal state transition model and updating the prediction with new sensory evidence. We optically imaged the activity of neurons in mouse PPC and PM layers 2, 3 and 5 in an acoustic virtual-reality system. As mice approached a reward site, anticipatory licking increased even when sound cues were intermittently presented; this was disturbed by PPC silencing. Probabilistic population decoding revealed that neurons in PPC and PM represented goal distances during sound omission (prediction), particularly in PPC layers 3 and 5, and prediction improved with the observation of cue sounds (updating). Our results illustrate how cerebral cortex realizes mental simulation using an action-dependent dynamic model.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Auditory virtual navigation task.
Figure 2: Two-photon (2P) imaging of neuronal activity in PPC and PM during the task.
Figure 3: Action-dependent distance representation.
Figure 4: Probabilistic decoding.
Figure 5: Goal-distance neurons estimate distance with dynamic Bayesian inference.

Similar content being viewed by others

References

  1. Doya, K., Ishii, S., Pouget, A. & Rao, R.P. Bayesian Brain: Probabilistic Approaches to Neural Coding (MIT Press, 2007).

  2. Denk, W., Strickler, J.H. & Webb, W.W. Two-photon laser scanning fluorescence microscopy. Science 248, 73–76 (1990).

    Article  CAS  PubMed  Google Scholar 

  3. Larkum, M. A cellular mechanism for cortical associations: an organizing principle for the cerebral cortex. Trends Neurosci. 36, 141–151 (2013).

    Article  CAS  PubMed  Google Scholar 

  4. Noudoost, B., Chang, M.H., Steinmetz, N.A. & Moore, T. Top-down control of visual attention. Curr. Opin. Neurobiol. 20, 183–190 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Keller, G.B., Bonhoeffer, T. & Hübener, M. Sensorimotor mismatch signals in primary visual cortex of the behaving mouse. Neuron 74, 809–815 (2012).

    Article  CAS  PubMed  Google Scholar 

  6. Beck, J.M., Latham, P.E. & Pouget, A. Marginalization in neural circuits with divisive normalization. J. Neurosci. 31, 15310–15319 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Boerlin, M. & Denève, S. Spike-based population coding and working memory. PLoS Comput. Biol. 7, e1001080 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Shipp, S. Structure and function of the cerebral cortex. Curr. Biol. 17, R443–R449 (2007).

    Article  CAS  PubMed  Google Scholar 

  9. Murayama, M. & Larkum, M.E. Enhanced dendritic activity in awake rats. Proc. Natl. Acad. Sci. USA 106, 20482–20486 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Saleem, A.B., Ayaz, A., Jeffery, K.J., Harris, K.D. & Carandini, M. Integration of visual motion and locomotion in mouse visual cortex. Nat. Neurosci. 16, 1864–1869 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Franklin, K.B.J. & Paxinos, G. The Mouse Brain in Stereotaxic Coordinates (Academic Press, 2008).

  12. Wang, Q. & Burkhalter, A. Area map of mouse visual cortex. J. Comp. Neurol. 502, 339–357 (2007).

    Article  PubMed  Google Scholar 

  13. Wang, Q., Sporns, O. & Burkhalter, A. Network analysis of corticocortical connections reveals ventral and dorsal processing streams in mouse visual cortex. J. Neurosci. 32, 4386–4399 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Whitlock, J.R., Sutherland, R.J., Witter, M.P., Moser, M.B. & Moser, E.I. Navigating from hippocampus to parietal cortex. Proc. Natl. Acad. Sci. USA 105, 14755–14762 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Whitlock, J.R., Pfuhl, G., Dagslott, N., Moser, M.B. & Moser, E.I. Functional split between parietal and entorhinal cortices in the rat. Neuron 73, 789–802 (2012).

    Article  CAS  PubMed  Google Scholar 

  16. Nitz, D.A. Tracking route progression in the posterior parietal cortex. Neuron 49, 747–756 (2006).

    Article  CAS  PubMed  Google Scholar 

  17. Nitz, D.A. Spaces within spaces: rat parietal cortex neurons register position across three reference frames. Nat. Neurosci. 15, 1365–1367 (2012).

    Article  CAS  PubMed  Google Scholar 

  18. Parron, C. & Save, E. Evidence for entorhinal and parietal cortices involvement in path integration in the rat. Exp. Brain Res. 159, 349–359 (2004).

    Article  PubMed  Google Scholar 

  19. McNaughton, B.L., Battaglia, F.P., Jensen, O., Moser, E.I. & Moser, M.B. Path integration and the neural basis of the 'cognitive map'. Nat. Rev. Neurosci. 7, 663–678 (2006).

    Article  CAS  PubMed  Google Scholar 

  20. Wilber, A.A., Clark, B.J., Forster, T.C., Tatsuno, M. & McNaughton, B.L. Interaction of egocentric and world-centered reference frames in the rat posterior parietal cortex. J. Neurosci. 34, 5431–5446 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Zingg, B. et al. Neural networks of the mouse neocortex. Cell 156, 1096–1111 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Nakamura, K. Auditory spatial discriminatory and mnemonic neurons in rat posterior parietal cortex. J. Neurophysiol. 82, 2503–2517 (1999).

    Article  CAS  PubMed  Google Scholar 

  23. Wilber, A.A. et al. Cortical connectivity maps reveal anatomically distinct areas in the parietal cortex of the rat. Front. Neural Circuits 8, 146 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Oh, S.W. et al. A mesoscale connectome of the mouse brain. Nature 508, 207–214 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. 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 

  27. Foldiak, P. The 'ideal homunculus': statistical inference from neural population responses. in Computation and Neural Systems (eds. Eeckman, F.H. & Bower, J.M.) 55–60 (Kluwer Academic Publishers, 1993).

  28. Sanger, T.D. Probability density estimation for the interpretation of neural population codes. J. Neurophysiol. 76, 2790–2793 (1996).

    Article  CAS  PubMed  Google Scholar 

  29. Ma, W.J., Beck, J.M., Latham, P.E. & Pouget, A. Bayesian inference with probabilistic population codes. Nat. Neurosci. 9, 1432–1438 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Ernst, M.O. & Banks, M.S. Humans integrate visual and haptic information in a statistically optimal fashion. Nature 415, 429–433 (2002).

    Article  CAS  PubMed  Google Scholar 

  31. Körding, K.P. & Wolpert, D.M. Bayesian integration in sensorimotor learning. Nature 427, 244–247 (2004).

    Article  PubMed  CAS  Google Scholar 

  32. Brunton, B.W., Botvinick, M.M. & Brody, C.D. Rats and humans can optimally accumulate evidence for decision-making. Science 340, 95–98 (2013).

    Article  CAS  PubMed  Google Scholar 

  33. Jones, J.L. et al. Orbitofrontal cortex supports behavior and learning using inferred but not cached values. Science 338, 953–956 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. George, D. & Hawkins, J. Towards a mathematical theory of cortical micro-circuits. PLoS Comput. Biol. 5, e1000532 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Shipp, S., Adams, R.A. & Friston, K.J. Reflections on agranular architecture: predictive coding in the motor cortex. Trends Neurosci. 36, 706–716 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Beck, J.M. et al. Probabilistic population codes for Bayesian decision making. Neuron 60, 1142–1152 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Rao, R.P.N. & Ballard, D.H. Predictive coding in the visual cortex: a functional interpretation of some extra-classical receptive-field effects. Nat. Neurosci. 2, 79–87 (1999).

    Article  CAS  PubMed  Google Scholar 

  38. Denève, S., Duhamel, J.R. & Pouget, A. Optimal sensorimotor integration in recurrent cortical networks: a neural implementation of Kalman filters. J. Neurosci. 27, 5744–5756 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Boccara, C.N. et al. Grid cells in pre- and parasubiculum. Nat. Neurosci. 13, 987–994 (2010).

    Article  CAS  PubMed  Google Scholar 

  40. Roome, C.J. & Kuhn, B. Chronic cranial window with access port for repeated cellular manipulations, drug application and electrophysiology. Front. Cell. Neurosci. 8, 379 (2014).

    PubMed  PubMed Central  Google Scholar 

  41. Phillips, D.P. & Cynader, M.S. Some neural mechanisms in the cat's auditory cortex underlying sensitivity to combined tone and wide-spectrum noise stimuli. Hear. Res. 18, 87–102 (1985).

    Article  CAS  PubMed  Google Scholar 

  42. Allen, T.A. et al. Imaging the spread of reversible brain inactivations using fluorescent muscimol. J. Neurosci. Methods 171, 30–38 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Martin, J.H. Autoradiographic estimation of the extent of reversible inactivation produced by microinjection of lidocaine and muscimol in the rat. Neurosci. Lett. 127, 160–164 (1991).

    Article  CAS  PubMed  Google Scholar 

  44. Erlich, J.C., Brunton, B.W., Duan, C.A., Hanks, T.D. & Brody, C.D. Distinct effects of prefrontal and parietal cortex inactivations on an accumulation of evidence task in the rat. Elife http://dx.doi.org/10.7554/eLife.05457 (2015).

  45. Adesnik, H., Bruns, W., Taniguchi, H., Huang, Z.J. & Scanziani, M. A neural circuit for spatial summation in visual cortex. Nature 490, 226–231 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Bishop, C.M. Pattern Recognition and Machine Learning, Vol. 1 (Springer, 2006).

  47. Ito, M. & Doya, K. Distinct neural representation in the dorsolateral, dorsomedial, and ventral parts of the striatum during fixed- and free-choice tasks. J. Neurosci. 35, 3499–3514 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Pfeiffer, B.E. & Foster, D.J. Hippocampal place-cell sequences depict future paths to remembered goals. Nature 497, 74–79 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Averbeck, B.B., Latham, P.E. & Pouget, A. Neural correlations, population coding and computation. Nat. Rev. Neurosci. 7, 358–366 (2006).

    Article  CAS  PubMed  Google Scholar 

  50. Graf, A.B., Kohn, A., Jazayeri, M. & Movshon, J.A. Decoding the activity of neuronal populations in macaque primary visual cortex. Nat. Neurosci. 14, 239–245 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the GENIE Program and the Janelia Research Campus for distributing GCaMP6f. We thank S.D. Aird for editing the manuscript and K. Mori for technical assistance. This work was supported by a Grant-in-Aid for Scientific Research on Innovative Areas: Prediction and Decision Making (23120007) (K.D.), KAKENHI 26730124 (A.F.) and 15H01452 (A.F.), and internal funding from the Okinawa Institute of Science and Technology Graduate University (K.D. and B.K.). We are grateful for generous support from the Okinawa Institute of Science and Technology Graduate University to the Neural Computation and Optical Neuroimaging Units.

Author information

Authors and Affiliations

Authors

Contributions

A.F. designed the study, built the setup, collected and analyzed data, and wrote the paper. B.K. designed the study, built the setup and wrote the paper. K.D. designed the study and wrote the paper.

Corresponding author

Correspondence to Kenji Doya.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Action-dependent licking behavior

During no-sound zones in intermittent 1 and intermittent 2 conditions, linear regression analysis compared the increase of licking between slow- and fast-locomotion trials. Trials in each session were equally divided into slow and fast trials. The regression analysis was defined as follows: Lick(t) = β0 + β1t, where β0-1 were regression coefficients. Lick(t) was the licking frequency at time t. Slope of licking (β1) was significantly steeper in fast trials than slow trials, suggesting that mice use action-dependent distance estimation (left, mean ± s.e.m.; right, central mark in box: median, edge of box: 25th and 75th percentiles, whiskers: most extreme data points not considered outliers (beyond 1.5× the inter-quartile range); * P < 0.01 in Wilcoxon signed-rank test, intermittent 1, P = 1.27 × 10-4, z = 3.83, intermittent 2, P = 9.81 × 10-8, z = 5.33) (number of sessions: intermittent 1, n = 94; intermittent 2, n = 94).

Supplementary Figure 2 Muscimol injection in PPC

(a) Injection locations in 3 mice detected with DiI (top). Section is 1.7 mm posterior from bregma11. Spread of muscimol estimated with fluorophore-conjugated muscimol (muscimol-BODIPYR TMR-X conjugate) (bottom). Muscimol-BODIPY was injected in 5 mice (1 left, 2 right, 2 both hemispheres). Injected locations are superimposed on the right hemisphere. Standard muscimol spreads 2 or 3 times farther than muscimol-BODIPY42,43, suggesting that muscimol in our study spread mainly in PPC. (b) Licking behavior in one session after PBS or muscimol injection (first licks red, others black). After muscimol injection, the lick-initiation point became variable. Also, in intermittent 1 condition, mouse tended to receive rewards after overrunning the no-sound zone. (c) Correlation between the start-to-lick distance and start-to-goal distance. Correlation coefficients were significantly lower in muscimol-injected sessions than PBS sessions in both the continuous and intermittent conditions (Mann-Whitney U-test, continuous, P = 4.78 × 10-4, z = 3.49; intermittent, P = 1.96 × 10-4, z = 3.72). (d) Average distance of licking for reward acquisition. Silencing of PPC increased overrun (start licking over 8.3 cm past the goal) (* P < 0.01 in Mann-Whitney U-test, continuous, P = 4.26 × 10-5, z = 4.09; intermittent 1, P = 1.79 × 10-15, z = 7.96; intermittent 2, P = 0.0221, z = 2.29) (Number of trials in PBS and muscimol sessions: continuous, n = 1,407, 1,188; intermittent 1, n = 293, 240; intermittent 2, n = 259, 230).

Supplementary Figure 3 Imaging locations

Imaging locations in each mouse. At each location, neuronal activity was imaged in layers 2, 3 and 5. Some recorded regions covered secondary visual cortex mediomedial area (V2MM) or mediomedial cortex (MM) of the visual area. We analyzed them as a part of the posteromedial cortex (PM).

Supplementary Figure 4 Sound-responsive neurons

(a) Example traces of activity in a single, sound-responsive neuron in response to sound presentation (color indicates sound azimuth, 20 averages). Inset shows sound-azimuth coding in which minimum and maximum activity after sound presentations was normalized to 0 and 1. (b) Sound-responsive neurons. Proportion of sound-responsive neurons with maximum activity in each azimuth (left). PPC or PM neurons preferred 0 and 2π/3 or 0 degrees, respectively. Proportion of sound-responsive neurons per field of view (middle, box plot as in Supplementary Fig. 1, * P < 0.05 in Mann-Whitney U-test, layers 2, 3, 5, P = 0.942, 0.0402, 0.817, z = 0.0731, 2.05, 0.232) (number of sessions in PPC and PM, n = [17, 16], [15, 15], [14, 17]). Latencies of sound responsive neurons (right, * P < 0.05 in Mann-Whitney U-test, P = 0.876, 0.135, 0.0170, z = 0.156, 1.50, 2.39) (number of sound responsive neurons in PPC and PM, n = [356, 373], [108, 334], [61, 63]).

Supplementary Figure 5 Example traces of task-irrelevant neurons

Activity traces of task-irrelevant neurons from the Fig. 2b data set.

Supplementary Figure 6 Neuronal activity represents goal distance but not sound intensity

(a) Regression analysis was used to investigate whether the activity of goal-distance neurons correlated with goal distance (Distance), the distance held constant during no-sound zones (Step), or the sound intensity (Sound). The regression analysis was defined as follows: y(t) = β0 + β1φ(variable(t)), where β0-1 were regression coefficients. y(t) was the unfiltered neural activity trace . variable(t) was either Distance, Step or Sound at frame t. φ is the Gaussian basis function in equation (3). (b) Root mean squared error (RMSE) was compared between the regression analysis of Distance and Step or Distance and Sound. RMSE of Distance was significantly smaller than of other variables both in goal-distance neurons and after-reaching neurons, indicating goal distance representation (box plot as in Supplementary Fig. 1, * P < 0.05 in Wilcoxon signed-rank test, P = 0.0428 to 0, z = 2.03 to 27.0) (number of goal-distance neurons in layers 2, 3, 5: PPC, n = 1,116, 1,054, 483, PM, n = 601, 479, 129) (number of after-reaching neurons: PPC, n = 744, 854, 367, PM, n = 193, 204, 54).

Supplementary Figure 7 After-reaching neurons in individual mice

Every trace shows the average activity of after-reaching neurons in one mouse (PPC n = 6, PM n = 6). Activity traces in continuous and intermittent conditions are shown in different panels (presentation as in Fig. 3c).

Supplementary Figure 8 Regression analysis of after-reaching neurons

(a) Linear regression analysis investigated whether the activity of each after-reaching neuron had a positive slope before the mouse reached the goal (equation (5) and Fig. 3c). Slopes of activity between 25.1 and 8.4 cm, and between 8.4 and 0 cm are shown. Neurons with significantly positive or negative slopes are shown with dark colors (P < 0.05 in two-sided student t-test). Slopes were significantly positive with and without sound in PPC in all conditions, but in PM in intermittent 1 condition they were not significantly positive without sound. (* P < 0.01 in Wilcoxon signed-rank test; PPC with sound, P = 0, z = 13.6 to 25.0; without sound, P = 1.19 × 10-6 to 0, z = 4.86 to 14.1; PM with sound, P = 2.53 × 10-6 to 0, z = 4.71 to 11.7; without sound, P = 0.497 to 9.97 × 10-10, z = 0.679 to 6.11) (number of after-reaching neurons in layers 2, 3, 5: PPC, n = 744, 854, 367; PM, n = 193, 204, 54). (b) Comparison of slopes with and without sound. Colored and white boxes show slopes with and without sound, respectively. Although slopes were significantly positive even without sound in a, they were less steep than those with sound (box plot as in Supplementary Fig. 1, ** P < 0.01 in Wilcoxon signed-rank test, PPC layer 2, P = 0, 0, z = 14.0, 9.89, layer 3, P = 0, 0, z = 18.9, 15.3, layer 5, P = 0, 0, z = 12.3, 9.72; PM layer 2, P = 9.51 × 10-7, 4.24 × 10-9, z = 4.90, 5.87, layer 3, P = 0.00559, 2.47 × 10-7, z = 2.77, 5.16, layer 5, P = 0.00202, 0.118, z = 3.09, 1.56). (c) Comparison of slopes between PPC and PM. In layers 2 and 3, PPC had steeper slopes than PM without sound (*P < 0.05, ** P < 0.01 in Mann-Whitney U-test; 25.1 – 8.4 cm, layer 2, P = 0.142, 0.185, z = 1.47, 1.32, layer 3, P = 2.60 × 10-14, 0.0142, z = 7.62, 2.45, layer 5, P = 0.357, 0.758, z = 0.922, 0.308; 8.4 – 0 cm, layer 2, P = 0.00320, 0.122, z = 2.95, 1.55, layer 3, P = 2.15 × 10-7, 2.79 × 10-6, z = 5.19, 4.69, layer 5, P = 0.209, 0.350, z = 1.26, 0.935).

Supplementary Figure 9 Stability of distance representation among conditions

(a) Neural activity traces. Before-reaching neurons were categorized into 4 groups, depending on the distance preferences in the continuous condition (top 4 columns) (PPC layer 2, n = 82, 53, 66, 171; layer 3, n = 14, 19, 22, 145; layer 5, n = 26, 8, 16, 66; PM layer 2, n = 60, 38, 87, 223; layer 3, n = 23, 16, 76, 160; layer 5, n = 17, 4, 18, 36). Activity of after-reaching neurons is shown in the bottom column (PPC layers 2, 3, 5, n = 744, 854, 367; PM, n = 193, 204, 54). Averaged trace of every neuron was normalized before calculating the mean population activity. (b) Stability of distance representations. Plots show the difference in distance preference between continuous and intermittent conditions and compare PPC and PM (box-plot as in Supplementary Fig. 1, * P < 0.05, ** P < 0.01 in Mann-Whitney U-test; intermittent1, P = 0.808 to 6.82 × 10-14, w = 50, z = 0.249 to 7.49; intermittent2, P = 0.955 to 0.00222, z = 0.0561 to 3.06).

Supplementary Figure 10 Probabilistic decoding in layers 3 and 5

Data presentation as in Figs. 4b–e, but for layers 3 and 5 (* P < 0.01 in Mann-Whitney U-test) (layer 3, Root mean squared error, continuous, P = 0.853 to 0, z = 0.185 to 17.1, intermittent 1, P = 0.956 to 3.33 × 10-16, z = 0.0546 to 8.17, intermittent 2, P = 0.810 to 1.15 × 10-9, z = 0.240 to 6.09; MAP, continuous, P = 0.694 to 1.57 × 10-12, z = 0.394 to 7.07, intermittent 1, P = 0.786 to 0.0308, z = 0.272 to 2.16, intermittent 2, P = 0.901 to 7.32 × 10-4, z = 0.124 to 3.38; s.d. of posterior, P = 3.33 × 10-16 to 0, z = 8.15 to 24.2) (layer 5, Root mean squared error, continuous, P = 4.10 × 10-9 to 0, z = 5.88 to 19.1, intermittent 1, P = 0.197 to 0, z = 1.29 to 11.4, intermittent 2, P = 0.0239 to 0, z = 2.26 to 9.09; MAP, continuous, P = 0.838 to 0, z = 0.204 to 10.5, intermittent 1, P = 0.270 to 1.15 × 10-6, z = 1.10 to 4.86, intermittent 2, P = 0.977 to 2.18 × 10-7, z = 0.0285 to 5.18; s.d. of posterior, P = 0, z = 10.2 to 26.2) (number of trials in PPC and PM: layer 3 continuous, n = 1,581, 1,571; intermittent 1, n = 331, 348; intermittent 2, n = 334, 331; layer 5 continuous, n = 1,445, 1,438; intermittent 1, n = 347, 331; intermittent 2, n = 308, 331).

Supplementary Figure 11 Sound presentation decreases decoding errors

(a) Root mean squared error (RMSE) of decoding. Data is the same as in Fig. 4c and Supplementary Fig. 10, but RMSE was binned each 0.42 cm (median ± robust standard error). (b) Difference of root mean squared error (ΔRMSE). ΔRMSE was defined as the difference of median RMSE between intermittent and continuous conditions. Linear regression analysis tested how ΔRMSE changed with and without sound. The regression analysis was defined as follows: y(distance) = β0 - β1distance, where β0-1 were regression coefficients. y(distance) was ΔRMSE at a distance bin. In some distance zones, slopes of ΔRMSE (β1) significantly decreased with sound or increased without sound (median ± robust standard error, * P < 0.05 in Wilcoxon signed-rank test, P = 0.0353 to 6.10 × 10-4, w = 19 to 102) (number of sessions in layers 2, 3, 5: PPC, n = 17, 15, 14; PM, n = 16, 15, 14). (c) Comparison of ΔRMSE during sound and no-sound zones. Box plot as in Supplementary Fig. 1 (* P < 0.01 in Mann-Whitney U-test: PPC layers 2, 3, 5, P = 2.31 × 10-5, 0.00295, 1.56 × 10-6, z = 4.23, 2.97, 4.80; PM, P = 0.00135, 6.27 × 10-4, 0.188, z = 3.20, 3.42, 1.32) (number of zones in layers 2, 3, 5: PPC, n = 68 (that is, 17 sessions times 4), 60, 56; PM, n = 64, 60, 56).

Supplementary Figure 12 Probabilistic decoding without moving average of neural activity

Data presentation as in Fig. 5, but without moving average of neural activity. In Fig. 5, a 3-frame moving average (97.1 ms) of neural activity was used (Online Methods). Decoding performances without smoothing the activity were similar to the one with 3-frame moving averaged activity, suggesting that distance predictions during no-sound zones were not an artifact of filtering. (a, c) * P < 0.05 in Wilcoxon signed-rank test (median ± robust standard error; a, P = 0.916 to 0, z = 0.106 to 34.4; c, P = 0.989 to 0, z = 0.0141 to 19.1) (number of trials in continuous, intermittent 1 and intermittent 2 conditions: PPC layer 2, n = 1,785, 398, 367; layer 3, n = 1,585, 331, 334; layer 5, n = 1,445, 347, 308; PM layer 2, n = 1,721, 362, 317; layer 3, n = 1,571, 348, 331; layer 5, n = 1,438, 331, 331). (b, d) * P < 0.05, ** P < 0.01 in Mann-Whitney U-test (box plot as in Supplementary Fig. 1; (b) PPC layer 2, P = 0, 0, z = 12.3, 10.7, layer 3, P = 3.33 × 10-16, 0, z = 8.18, 9.15, layer 5, P = 0, 0, z = 11.2, 9.30; PM layer 2, P = 6.21 × 10-8, 8.40 × 10-14, z = 5.41, 7.46, layer 3, P = 1.89 × 10-15, 6.66 × 10-16, z = 7.95, 8.07, layer 5, P = 0.192, 0.00599, z = 1.31, 2.75; (d) PPC layer 2, P = 4.46 × 10-9, 1.99 × 10-5, z = 5.87, 4.27, layer 3, P = 4.14 × 10-6, 0.0183, z = 4.60, 2.36, layer 5, P = 4.95 × 10-12, 1.10 × 10-7, z = 6.91, 5.31; PM layer 2, P = 0.214, 0.142, z = 1.24, 1.47, layer 3, P = 0.0199, 0.00137, z = 2.33, 3.20, layer 5, P = 0.979, 0.434, z = 0.0264, 0.782).

Supplementary Figure 13 Action-dependent distance estimation

Linear regression analysis investigated whether the estimated distance of decoder (maximum a posterior: MAP) depended on the locomotion speed of mice during no-sound zones. The regression analysis was defined as follows: y(t) = β0 + β1t, where β0-1 were regression coefficients. y(t) was MAP at frame t. Trials in each session were equally divided into fast-locomotion and slow-locomotion trials. Slopes of MAPs were steeper in fast trials than slow trials, suggesting action-dependent distance prediction of mice (box plot as in Supplementary Fig. 1, * P < 0.05, ** P < 0.01 in Mann-Whitney U-test, PPC layer 2, P = 3.97 × 10-4, 8.60 × 10-4, z = 3.54, 3.33, layer 3, P = 0.00284, 0.00229, z = 2.98, 3.05, layer 5, P = 0.0869, 0.377, z = 1.71, 0.883; PM layer 2, P = 3.23 × 10-4, 0.616, z = 3.60, 0.502, layer 3, P = 0.00335, 0.821, z = 2.93, 0.226, layer 5, P = 0.0332, 0.0704, z = 2.13, 1.81) (number of slow and fast trials in layers 2, 3, 5 in intermittent 1 condition: PPC, n = [194, 204], [163, 168], [171, 176]; PM, n = [174, 188], [167, 181], [160, 171]) (slow and fast trials in intermittent 2 condition: PPC, n = [177, 190], [162, 172], [148, 160], PM, n = [153, 164], [161, 170], [162, 169]).

Supplementary Figure 14 Dynamic Bayesian inference in cortical microcircuits

Overall neural representations, prediction and updating of decoding were similar between PPC and PM. These regions have similar anatomical connections of motor and sensory inputs. Distance estimation was better in PPC than PM. This might be related to the stronger motor inputs to PPC than PM. Aud: auditory cortex; M2: secondary motor cortex; RSP: retrosplenial cortex.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–14 and Supplementary Table 1 (PDF 3043 kb)

Supplementary Methods Checklist (PDF 1110 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Funamizu, A., Kuhn, B. & Doya, K. Neural substrate of dynamic Bayesian inference in the cerebral cortex. Nat Neurosci 19, 1682–1689 (2016). https://doi.org/10.1038/nn.4390

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4390

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