Skip to main content

Advertisement

Log in

Inference of synaptic connectivity and external variability in neural microcircuits

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

Abstract

A major goal in neuroscience is to estimate neural connectivity from large scale extracellular recordings of neural activity in vivo. This is challenging in part because any such activity is modulated by the unmeasured external synaptic input to the network, known as the common input problem. Many different measures of functional connectivity have been proposed in the literature, but their direct relationship to synaptic connectivity is often assumed or ignored. For in vivo data, measurements of this relationship would require a knowledge of ground truth connectivity, which is nearly always unavailable. Instead, many studies use in silico simulations as benchmarks for investigation, but such approaches necessarily rely upon a variety of simplifying assumptions about the simulated network and can depend on numerous simulation parameters. We combine neuronal network simulations, mathematical analysis, and calcium imaging data to address the question of when and how functional connectivity, synaptic connectivity, and latent external input variability can be untangled. We show numerically and analytically that, even though the precision matrix of recorded spiking activity does not uniquely determine synaptic connectivity, it is in practice often closely related to synaptic connectivity. This relation becomes more pronounced when the spatial structure of neuronal variability is jointly considered.

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

Similar content being viewed by others

References

  • Baker, C., Ebsch, C., Lampl, I., Rosenbaum, R. (2019). Correlated states in balanced neuronal networks. Physical Review E 99 5.

  • Barral, J., & D’Reyes, A. (2016). Synaptic scaling rule preserves excitatory-inhibitory balance and salient neuronal network dynamics. Nature Neuroscience, 19(12), 1690–1696.

    CAS  PubMed  Google Scholar 

  • Bishop, C.M. (2007). Pattern Recognition and Machine Learning.

  • Brinkman, B.A.W., Rieke, F., Shea-Brown, E., Buice, M.A. (2017). Predicting how and when hidden neurons skew measured synaptic interactions, 1–50.

  • Chambers, B., Levy, M., Dechery1, J.B., Maclean, J.N. (2017). Ensemble stacking mitigates biases in inference of synaptic connectivity. Network Neuroscience Ensemble stacking mitigates biases in inference of synaptic connectivity. JN.

  • Chiang, A.S., Lin, C.Y., Chuang, C.C., Chang, H.M., Hsieh, C.H., Yeh, C.W., Shih, C.T., Wu, J.J., Wang, G.T., Chen, Y.C., Wu, C.C., Chen, G.Y., Ching, Y.T., Lee, P.C., Lin, C.Y., Lin, H.H., Wu, C.C., Hsu, H.W., Huang, Y.A., Chen, J.Y., Chiang, H.J., Lu, C.F., Ni, R.F., Yeh, C.Y., Hwang, J.K. (2011). Three-dimensional reconstruction of brain-wide wiring networks in drosophila at single-cell resolution. Current Biology, 21(1), 1–11.

    CAS  PubMed  Google Scholar 

  • Cohen, M.R., & Kohn, A. (2011). Measuring and interpreting neuronal correlations. Nature Neuroscience, 14 (7), 811–819.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Cotton, R.J., Froudarakis, E., Storer, P., Saggau, P., Tolias, A. (2013). Three-dimensional mapping of microcircuit correlation structure. Frontiers in Neural Circuits.

  • Dayan, P., & Abbott, L.F. (2001). Theoretical neuroscience: computational and mathematical modeling of neural systems. Cambridge: MIT Press.

    Google Scholar 

  • Doiron, B., Litwin-Kumar, A., Rosenbaum, R., Ocker, G.K., Josic, K. (2016). The mechanics of state-dependent neural correlations. Nature Neuroscience, 19(3), 383–393.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Ebsch, C., & Rosenbaum, R. (2018). Imbalanced amplification: a mechanism of amplification and suppression from local imbalance of excitation and inhibition in cortical circuits. PLoS Computational Biology, 14(3), 1–28.

    Google Scholar 

  • Feldt, S., Bonifazi, P., Cossart, R. (2011). Dissecting functional connectivity of neuronal microcircuits: experimental and theoretical insights.

    CAS  PubMed  Google Scholar 

  • Friedrich, J., Zhou, P., Paninski, L. (2017). Fast online deconvolution of calcium imaging data. PLoS Computational Biology.

  • Garaschuk, O., Milos, R.I., Konnerth, A. (2006). Targeted bulk-loading of fluorescent indicators for two-photon brain imaging in vivo. Nature Protocols.

  • Gardiner, C. (2009). Stochastic methods - a handbook for the natural and social sciences.

  • Gerhard, F., Kispersky, T., Gutierrez, G.J., Marder, E., Kramer, M., Eden, U. (2013). Successful reconstruction of a physiological circuit with known connectivity from spiking activity alone. PLos Computational Biology, 9(7), e1003138.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Jiang, X., Shen, S., Cadwell, C.R., Berens, P., Sinz, F., Ecker, A., Patel, S., Tolias, A. (2016). Principles of connectivity among morphologically defined cell types in adult neocortex. Science, 350(6264), 1–21.

    Google Scholar 

  • Kadirvelu, B., Hayashi, Y., Nasuto, S.J. (2017). Inferring structural connectivity using Ising couplings in models of neuronal networks. Scientific Reports, 7(1), 1–12.

    CAS  Google Scholar 

  • Kalatsky, V.A., & Stryker, M.P. (2003). New paradigm for optical imaging: temporally encoded maps of intrinsic signal. Neuron.

  • Kohn, A. (2005). Stimulus dependence of neuronal correlation in primary visual cortex of the macaque. Journal of Neuroscience.

  • Krumin, M., Reutsky, I., Shoham, S. (2010). Correlation-based analysis and generation of multiple spike trains using hawkes models with an exogenous input. Frontiers in Computational Neuroscience 4.

  • Ladenbauer, J., McKenzie, S., English, D.F., Hagens, O., Ostojic, S. (2019). Inferring and validating mechanistic models of neural microcircuits based on spike-train data. Nat Communications, 10, 4933.

    Google Scholar 

  • Levy, R.B., & Reyes, A. (2012). . Mouse Primary Auditory Cortex, 32(16), 5609–5619.

    CAS  Google Scholar 

  • Lin, T.W., Das, A., Krishnan, G.P., Bazhenov, M., Sejnowski, T.J. (2017). Differential covariance: a new class of methods to estimate sparse connectivity from neural recordings. Neural Computation, 29(10), 2581–2632.

    PubMed  PubMed Central  Google Scholar 

  • Lütcke, H., Gerhard, F., Zenke, F., Gerstner, W., Helmchen, F. (2013). Inference of neuronal network spike dynamics and topology from calcium imaging data. Frontiers in Neural Circuits, 7(December), 1–20.

    Google Scholar 

  • Magrans de Abril, I., Yoshimoto, J., Doya, K. (2018). Connectivity inference from neural recording data: challenges, mathematical bases and research directions. Neural Networks, 102, 120–137.

    PubMed  Google Scholar 

  • Maswadeh, W.M., & Snyder, P.S. (2012). Multivariable and multigroup receiver operating characteristics curve analyses for qualitative and quantitative analysis. Edgewood Chemical Biological Center ECBC-TR-92(US Army Research, Development and Engineering Command).

  • Mishchencko, Y., Vogelstein, J., Paninski, L. (2007). a Bayesian Approach for Inferring Neuronal. Statistics.

  • Nykamp, D.Q. (2007). A mathematical framework for inferring connectivity in probabilistic neuronal networks. Mathematical Biosciences, 205(2), 204–251.

    PubMed  Google Scholar 

  • Paninski, L. (2004). Maximum likelihood estimation of cascade point-process neural encoding models. Network: Computation in Neural Systems.

    Google Scholar 

  • Pernice, V., & Rotter, S. (2013). Reconstruction of sparse connectivity in neural networks from spike train covariances. Journal of Statistical Mechanics: Theory and Experiment 2013(3).

  • Pernice, V., Staude, B., Cardanobile, S., Rotter, S. (2011). How structure determines correlations in neuronal networks. PLoS Computational Biology 7(5).

    CAS  PubMed  PubMed Central  Google Scholar 

  • Pfeffer, C.K., Xue, M., He, M., Huang, Z.J., Scanziani, M. (2013). Inhibition of inhibition in visual cortex: the logic of connections between molecularly distinct interneurons. Nature Neuroscience. https://doi.org/10.1038/nn.3446.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Pillow, J.W., Shlens, J., Paninski, L., Sher, A., Litke, A.M., Chichilnisky, E.J., Simoncelli, E.P. (2008). Spatio-temporal correlations and visual signalling in a complete neuronal population. Nature.

  • Pnevmatikakis, E.A., Soudry, D., Gao, Y., Machado, T.A., Merel, J., Pfau, D., Reardon, T., Mu, Y., Lacefield, C., Yang, W., Ahrens, M., Bruno, R., Jessell, T.M., Peterka, D.S., Yuste, R. (2017). Simultaneous denoising, deconvolution, and demixing of calcium imaging data. HHS Public Access, 89(2), 285–299.

    Google Scholar 

  • Poli, D., Pastore, V.P., Martinoia, S., Massobrio, P. (2016). From functional to structural connectivity using partial correlation in neuronal assemblies. Journal of Neural Engineering, 13(2), 26, 023.

    Google Scholar 

  • Pyle, R., & Rosenbaum, R. (2016). Highly connected neurons spike less frequently in balanced networks. Physical Review E, 93(4), 1–6.

    Google Scholar 

  • Renart, A., Rocha, J.D., Bartho, P., Hollender, L., Reyes, A., Harris, K.D. (2010). The asynchronus state in cortical circuits. Science, 327(5965), 587–590.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Rosenbaum, R., Smith, M.A., Kohn, A., Rubin, J.E., Doiron, B. (2017). The spatial structure of correlated neuronal variability. Nature Neuroscience, 20(1), 107–114.

    CAS  PubMed  Google Scholar 

  • Singh, R., Ghosh, D., Adhikari, R. (2017). Fast Bayesian inference of the multivariate Ornstein-Uhlenbeck process 012136:1–9.

  • Smith, M.A., & Kohn, A. (2008). Spatial and temporal scales of neuronal correlation in primary visual cortex. Journal of Neuroscience.

  • Song, S., Sjöström, P.J., Reigl, M., Nelson, S., Chklovskii, D.B. (2005). Highly nonrandom features of synaptic connectivity in local cortical circuits. In PLoS biology. https://doi.org/10.1371/journal.pbio.0030068.

    PubMed  PubMed Central  Google Scholar 

  • Soudry, D., Keshri, S., Stinson, P., Oh, M.H., Iyengar, G., Paninski, L. (2013). A shotgun sampling solution for the common input problem in neural connectivity inference, arXiv.

  • Stevenson, I.H., Rebesco, J.M., Hatsopoulos, N.G., Haga, Z., Miller, L.E., Körding, K.P. (2009). Bayesian inference of functional connectivity and network structure from spikes. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 17(3), 203–213.

    PubMed  Google Scholar 

  • Trousdale, J., Hu, Y., Shea-Brown, E., Josić, K. (2012). Impact of network structure and cellular response on spike time correlations. PLoS Computational Biology 8(3).

    CAS  PubMed  PubMed Central  Google Scholar 

  • Van Vreeswijk, C., & Sompolinsky, H. (1996). Chaos in neuronal networks with balanced excitatory and inhibitory activity. Science, 274(5293), 1724–1726. https://doi.org/10.1126/science.274.5293.1724.

    PubMed  Google Scholar 

  • van Vreeswijk, C., & Sompolinsky, H. (1998). Chaotic balanced state in a model of cortical circuits. Neural Computation, 10(6), 1321–1371.

    PubMed  Google Scholar 

  • Vinci, G., Smith, M.A., Kass, R.E. (2018). Adjusted regularization of cortical covariance. Journal of Computational Neuroscience. https://doi.org/10.1007/s10827-018-0692-x.

    PubMed  PubMed Central  Google Scholar 

  • Vogelstein, J.T., Packer, A.M., Machado, T.A., Sippy, T., Yuste, R., Paninski l, Babadi B. (2012). Fast nonnegative deconvolution for spike train inference from population calcium imaging fast nonnegative deconvolution for spike train inference from population calcium imaging. Journal of Neurophysiology.

  • Widloski, J., Marder, M.P., Fiete, I.R. (2018). Inferring circuit mechanisms from sparse neural recording and global perturbation in grid cells. eLife https://doi.org/10.7554/eLife.33503.

  • Yaglom, A. (1962). An introduction to the theory of stationary random functions.

  • Yatsenko, D., Froudarakis, E., Ecker, A., Rosenbaum, R., Josić, K, Tolias, A. (2016). Strong functional connectivity of parvalbumin-expressing cortical interneurons. Computational and Systems Neuroscience Meeting (COSYNE 2016).

  • Yatsenko, D., Josić, K., Ecker, A., Froudarakis, E., Cotton, R.J., Tolias, A. (2015). Improved estimation and interpretation of correlations in neural circuits. PLoS Computational Biology, 11(3), 1–28.

    Google Scholar 

  • Zaytsev, Y.V., Morrison, A., Deger, M. (2015). Reconstruction of recurrent synaptic connectivity of thousands of neurons from simulated spiking activity. Journal of Computational Neuroscience, 39(1), 77–103.

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank Ryan Pyle (Notre Dame) and Krešimir Josić (University of Houston) for feedback and many productive discussions on this topic.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Cody Baker.

Ethics declarations

Conflict of interests

The authors declare no conflict of interest.

Additional information

Action Editor: Uri Eden

Publisher’s note

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

This work was supported by National Foundation of Science grants DMS-1517828, DMS-1654268, and NeuroNex DBI-1707400.

Electronic supplementary material

Below is the link to the electronic supplementary material.

(PDF 772 KB)

Appendix

Appendix

1.1 Experimental methods

All procedures were carried out in accordance with the ethical guidelines of the National Institutes of Health and were approved by the Institutional Animal Care and Use Committee (IACUC) of Baylor College of Medicine.

The animals (n = 5 PV-Cre/Ai9 crosses on a C57Bl/6 background, labeled with the fluorescent marker tdTomato) with an age range of p40 to p60 were initially anesthetized with Isoflurane (3%) and then anesthesia was maintained by either Isoflurane (2%) or a mixture of Fentanyl (0.05 mg/kg), Midazolam (5 mg/kg), and Medetomidin (0.5 mg/kg) with anesthesia boosts consisting half of the initial dose every three hours. The body temperature of the animal was maintained at 37C throughout the surgery using a homeothermic blanket system (Harvard Instruments). In some experiments we applied eye oil ointment (polydimethylsiloxane) to prevent dehydration of the cornea. Surgery and dye injections of the Oregon Green 488 BAPTA-1 AM (OGB1, Invitrogen) calcium indicator were performed as previously described (Garaschuk et al. 2006).

We used stereotactic information to locate our recordings to the primary visual cortex of the mouse (V1). In some experiments we used intrinsic imaging to verify the location of V1 (Kalatsky and Stryker 2003). We recorded calcium traces using a custom built two-photon microscope equipped with a Chameleon Ti-Sapphire laser (Coherent) tuned at 800 nm and a 20x, 1.0 NA Olympus objective. Scanning was controlled by a custom built acousto-optic deflector system (AODs) (Cotton et al. 2013). The average power out of the objective was kept less than 120mW. Calcium activity was typically sampled at a mean of 260Hz (min/max: 78-450 Hz). We recorded data from depths of 100-540 μ m below the cortical surface.

The measured fluorescent traces were preprocessed in order to reduce common mode noise related to small cardiovascular movements (Cotton et al. 2013) and the firing rates were estimated using by nonnegative deconvolution (Vogelstein et al. 2012).

1.2 OU identity for normal \(\bar {\mathbf {K}}\)

For the OU system defined in Eq. (1), if KKT = KTK, G = gI, and QQT = σI we have

$$ \begin{array}{@{}rcl@{}} \mathcal{R}(0) &=& \frac{\sigma^{2} g^{2}}{{\tau^{2}_{r}}} \int\limits_{-\infty}^{0} {\exp}\left[ {\frac{1}{\tau_{r}} (\mathbf{I} - g \bar{\mathbf{K}}) s} \right]\\ &&\times {\exp} \left[ {\frac{1}{\tau_{r}} (\mathbf{I} - g \bar{\mathbf{K}})^{\mathrm{T}} s} \right] ds \\ &=& \frac{\sigma^{2} g^{2}}{{\tau^{2}_{r}}} \int\limits_{-\infty}^{0} {\exp} \left[ {\frac{1}{\tau_{r}} (2 \mathbf{I} - g \bar{\mathbf{K}} - g \bar{\mathbf{K}}^{\mathrm{T}}) s} \right] ds \\ &=& \sigma^{2} g^{2} \tau_{r} (2 \mathbf{I} - g \bar{\mathbf{K}} - g \bar{\mathbf{K}}^{\mathrm{T}})^{-1} \end{array} $$
$$ \Rightarrow \qquad \bar{\mathbf{K}} + \bar{\mathbf{K}}^{\mathrm{T}} = \frac{2}{g} \mathbf{I} - \sigma^{2} g \tau_{r} \mathcal{R}^{-1}(0) $$

1.3 Relation between AUROC and discriminability for normal distributions

It is well known that the area under an ROC curve may be parameterized and solved to yield the identity

$$ \text{AUROC} = \mathbb{P}(X_{1} > X_{0}) $$

for normally distributed scores in the positive and negative classes X1 and X0, respectively. If these scores are normally distributed then closure properties imply

$$ \mathbb{P}(X_{1} > X_{0}) = \mathbb{P}(X_{1} - X_{0} > 0) = \mathbb{P}(Y > 0) $$
$$ = 1 - \mathbb{P}(Y \leq 0) = 1 - F_{Y}(y) $$
$$ Y \sim \mathcal{N}(\mu_{1} - \mu_{0}, {\sigma^{2}_{1}} + {\sigma^{2}_{0}}) $$

and so

$$ \text{AUROC} = 1 - \frac{1}{2} \left[ 1 + \text{erf} \left( -\frac{\mu_{1} - \mu_{0}}{\sqrt{2} \sqrt{{\sigma^{2}_{1}} + {\sigma^{2}_{0}}}} \right) \right] $$

and so by appropriately defining the discriminability D and re-arranging we obtain

$$ \text{AUROC} = \frac{1}{2} \text{erfc} \left( -\frac{|D|}{\sqrt{2}} \right) $$

where the absolute value restricts the AUROC to \(\left [ \frac {1}{2},1 \right ]\) which corrects for anti-classifiers. The result is the bijective mapping from D to AUROC, leading to the natural invertibility between the two which is necessary for our arguments.

1.4 Central limit of precision with independent external input

Precision values from the case of independent external input can be expressed element-wise as

$$ \mathbf{P}^{a b}_{\alpha \beta} = \mathbf{W}^{a b}_{\alpha \beta} + \mathbf{W}^{b a}_{\beta \alpha} - {\sum}_{\gamma = 1}^{N} \mathbf{W}^{c a}_{\gamma \alpha} {\mathbf{W}}_{\gamma \beta}^{c b} $$

which in the thermodynamic limit of network size (\(N \rightarrow \infty \)) leads the summation term to converge to the limiting normal distribution, so long as all elements of W are independent with finite variance. So long as we enforce normal distributions on and zero-variance gains on the synaptic strengths, and since the normal distribution is closed under summation, elements of P will also be normally distributed. All these assumptions hold under the standard ER case as well as the CER case, whereby correlations are specially constructed so that pairs of the form \(\mathbf {W}^{c a}_{\gamma \alpha } {\mathbf {W}}_{\gamma \beta }^{c b}\) are independent across γ though not across pairs of α, β. The only breakdown of the CLT independence conditions occurs in the HDout case, as all such pairs are now highly correlated across γ, giving an apparent power law limiting distribution instead.

While the above requirements on J hold exactly only if it is normally distributed, as long as Ω is Erdos-Renyi this theory will still hold as a fair approximation since the \(\mathcal {O}(1)\) part of the precision distributions are determined by the summation term with the direct strengths offering only a \(\mathcal {O}(1 \slash \sqrt {N})\) deviation. Even if synaptic weights are specified from a one-sided distribution of finite variance, the induced asymmetries against the limiting Gaussian will dissipate in the large N limit. Any variance present in the gains will also lead to small errors with an exact Gaussian, but these effects will again decay for large N and so the discriminability theory outlined in the paper should still hold as a good approximation for the expected AUROC.

1.5 Simulation and figure parameters

For all simulations, we use an alternative parameterization of synaptic strength which makes modulation easier to control. We express the average synaptic weights as

$$ \left[\begin{array}{ll} j_{ee} & j_{ei} \\ j_{ie} & j_{ii} \end{array}\right] = k_{1} \left[\begin{array}{ll} \psi_{ee} & \psi_{ei} \\ \psi_{ie} & \psi_{ii} \end{array}\right] $$

where the synaptic proportionsψ are normalized by the inhibitory component as ψab = jab/jii and the mean synaptic magnitudek1 is then modulated while holding the proportions fixed. Similarly, the synaptic variance is parameterized as

$$ \left[\begin{array}{ll} v_{ee} & v_{ei} \\ v_{ie} & v_{ii} \end{array}\right] = k_{2} \left[\begin{array}{ll} \psi^{2}_{ee} & \psi^{2}_{ei} \\ \psi^{2}_{ie} & \psi^{2}_{ii} \end{array}\right] $$

where the magnitude of synaptic variancek2 is modulated. For all figures and simulations, we use a version of the synaptic proportions used in Pyle and Rosenbaum (2016) that have been perturbed in order to give non-zero real part to the eigenvalues. These proportions are

$$ \left[\begin{array}{ll} \psi_{ee} & \psi_{ei} \\ \psi_{ie} & \psi_{ii} \end{array}\right] = \left[\begin{array}{ll} 0.1 & 0.6 \\ 0.45 & 1 \end{array}\right] $$

as well as the fixed synaptic ratios of qe = 0.8, qi = 0.2 and the same fixed density of pab = p = 1.

In Fig. 1, we used σa = σ = 1, ga = g = 1, and ua = u = 0 for both rows in order to simplify the interpretation. For the top row we took k1 = 2.5, k2 = 6.25 × 10− 5 and for the bottom row, k1 = 12.5 and k2 = 1.25. The same exact networks used in Fig. 1 are used in Fig. 2, though they are partitioned according to the more informative mask.

In Fig. 3, all shared network parameters are identical to Fig. 1, except k1 = 2.5, k2 = 1.25. The model specific parameters are ρ = 0.2 for the CER model and μ = 5, ξ = 0.25 for the HD models. As mentioned, σ for the HD model is estimated numerically using the bisection method to find the fixed point.

In Fig. 4, the same recurrent networks were used across all cases. Shared recurrent parameters are consistent with the top row of Fig. 1. Shared external input parameters are \(p_{ax} = p_{x} = 0.1, j_{a x} = \psi _{a x} k_{x,1}, v_{a x} = \psi ^{2}_{a x} k_{x,2}, \psi _{e x} = 1.333, \psi _{i x} = 1\). Individually modified parameters are kx,1 = 0.2571, kx,2 = 0.6571,〈sx, sxαα = rx = 10,〈sx, sxαβ = c = 0.1 for αβ in both the full-rank correlated state and the correlated part of the first combination. The first combination also had Γind = I for the independent part. The second combination used kx,1 = 250, kx,2 = 0, rx = 10, c = 0, qx = 0.2, Γind = 10I. Full-rank effects in the correlated case are induced by taking qx = 7 to give Nx = 14, 000 total external neurons, making Γ invertible with high probability even without regularization from the independent external source.

In Fig. 5, shared parameters are consistent with the top row of Fig. 1 and used a spatial width of ςab = ς = 0.2. The additional spatial variability inherent to these networks accounts for the difference in AUROC using only precision as a marginal metric between Fig. 2 and Table 1.

The membrane potential dynamics for the AdEx model are

$$ C_{m} \frac{d \mathbf{V}}{d t} = -g_{L} (\mathbf{V} - E_{L}) + g_{L} {\Delta}_{T} e^{\frac{\mathbf{V} - \mathbf{V}_{T}}{{\Delta}_{T}}} + \mathbf{T} - \mathbf{w} $$
$$ \tau_{w} \frac{d \mathbf{w}}{d t} = a (\mathbf{V} - E_{L}) - \mathbf{w} $$

subject to the rule that if a voltage exceeds the threshold (\(\mathbf {V}^{a}_{\alpha }(t) \geq V_{th}\)) then it returns to reset (\(\mathbf {V}^{a}_{\alpha }(t) \rightarrow V_{re}\)), its adaptation current is incremented (\(\mathbf {w}^{a}_{\alpha } \rightarrow \mathbf {w} + b\)), and a spike is recorded. The input terms stemming from recurrent sources R and external sources X may be expressed as

$$ \mathbf{T} = C_{m} (\mathbf{X} + \mathbf{R}) + \mathbf{Q} \ \frac{d\mathcal{W}_{t}}{dt} $$
$$ \mathbf{R}^{a}_{\alpha} = {\sum}_{c = e,i} {\sum}_{\gamma \in c} \mathbf{K}^{a c}_{\alpha \gamma} {\sum}_{n} \eta_{c} (t - t^{c,\gamma}_{n}) $$
$$ \mathbf{X}^{a}_{\alpha} = {\sum}_{\gamma \in x} (\mathbf{K}_{x})^{a}_{\alpha \gamma} {\sum}_{n} \eta_{x} (t - t^{x,\gamma}_{n}) $$

where \(t_{n}^{c,\gamma }\) is the nth spike time of neuron γ in population c = {e, i, x} and synaptic kinetics are modeled by the filter \(\eta _{c}(t) = \frac {1}{\tau _{c}} e^{-\frac {t}{\tau _{c}}}{\Theta }(t)\) where Θ(t) is the Heaviside step function.

In Fig. 7, network parameters for the OU model are identical to the top row of Fig. 1 and have independent noise level σ = 0.1 and timescale τr = 1, discretized at the level of dt = 0.1. Shared network parameters for the EIF model are the same except for k1 = 250 and k2 = 0, as well as the fact that the statistics of the gains are no longer specifiable and are a consequence of the non-linearity of the system. AdEx-specific parameters are as follows: Cm = 1, gL = 0.0667, EL = − 72, Vth = − 50, Vre = − 75,ΔT = 1, VT = − 55, τw = 150, τe = 8, τi = 4, τx = 10, a = 0, and b = 0.1. External input followed: qx = 0.2, kx,1 = 250, kx,2 = 0, and identical rx, c, ψ, px as from Fig. 4 purple. The balanced network exhibited less than 20% relative error to both the balanced rate approximation and the mean-field covariance approximation from Baker et al. (2019).

All code pertaining to simulations and analysis may be found at https://github.com/cb239/Inference-of-Synaptic-Connectivity.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Baker, C., Froudarakis, E., Yatsenko, D. et al. Inference of synaptic connectivity and external variability in neural microcircuits. J Comput Neurosci 48, 123–147 (2020). https://doi.org/10.1007/s10827-020-00739-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-020-00739-4

Keywords

Navigation