Skip to main content
PLOS Computational Biology logoLink to PLOS Computational Biology
. 2021 Sep 17;17(9):e1009353. doi: 10.1371/journal.pcbi.1009353

STDP and the distribution of preferred phases in the whisker system

Nimrod Sherf 1,2,*, Maoz Shamir 1,2,3
Editor: Abigail Morrison4
PMCID: PMC8480728  PMID: 34534208

Abstract

Rats and mice use their whiskers to probe the environment. By rhythmically swiping their whiskers back and forth they can detect the existence of an object, locate it, and identify its texture. Localization can be accomplished by inferring the whisker’s position. Rhythmic neurons that track the phase of the whisking cycle encode information about the azimuthal location of the whisker. These neurons are characterized by preferred phases of firing that are narrowly distributed. Consequently, pooling the rhythmic signal from several upstream neurons is expected to result in a much narrower distribution of preferred phases in the downstream population, which however has not been observed empirically. Here, we show how spike timing dependent plasticity (STDP) can provide a solution to this conundrum. We investigated the effect of STDP on the utility of a neural population to transmit rhythmic information downstream using the framework of a modeling study. We found that under a wide range of parameters, STDP facilitated the transfer of rhythmic information despite the fact that all the synaptic weights remained dynamic. As a result, the preferred phase of the downstream neuron was not fixed, but rather drifted in time at a drift velocity that depended on the preferred phase, thus inducing a distribution of preferred phases. We further analyzed how the STDP rule governs the distribution of preferred phases in the downstream population. This link between the STDP rule and the distribution of preferred phases constitutes a natural test for our theory.

Author summary

The distribution of preferred phases of whisking neurons in the somatosensory system of rats and mice presents a conundrum: a simple pooling model predicts a distribution that is an order of magnitude narrower than what is observed empirically. Here, we suggest that this non-trivial distribution may result from activity-dependent plasticity in the form of spike timing dependent plasticity (STDP). We show that under STDP, the synaptic weights do not converge to a fixed value, but rather remain dynamic. As a result, the preferred phases of the whisking neurons vary in time, hence inducing a non-trivial distribution of preferred phases, which is governed by the STDP rule. Our results imply that the considerable synaptic volatility which has long been viewed as a difficulty that needs to be overcome, may actually be an underlying principle of the organization of the central nervous system.

Introduction

The whisker system is used by rats and mice to actively gather information about their proximal environment [14]. Information about whisker position, touch events, and texture is relayed downstream the somatosensory system via several tracks; in particular, the lemniscal pathway that relays information about both whisking and touch [511].

During whisking, the animal moves its vibrissae back and forth in a rhythmic manner, Fig 1a–1c. Neurons that track the azimuthal position of the whisker by firing in a preferential manner to the phase of the whisking cycle are termed whisking neurons. Whisking neurons in the ventral posteromedial nucleus (VPM) of the thalamus as well as inhibitory whisking neurons in layer 4 of the barrel cortex are characterized by a preferred phase at which they fire with the highest rate during the whisking cycle [1, 6, 1216], Fig 1d and 1e. The distribution of preferred phases is non-uniform and can be approximated by the circular normal (Von-Mises) distribution

Pr(ϕ)=eκcos(ϕψ)2πI0(κ) (1)

where ψ is the mean phase and I0(κ) is the modified Bessel function of order 0, Fig 2a. The parameter κ quantifies the width of the distribution; κ = 0 yields a uniform (flat) distribution, whereas in the limit of κ → ∞ the distribution converges to a delta function. Typical values for κ in the thalamus and for layer 4 inhibitory (L4I) whisking neurons are κVPMκL4I ≈ 1 where ψVPM ≈ 5π/6 [rad] and κL4I ≈ 0.5 [rad] [2, 13].

Fig 1. Representation of the whisking phase.

Fig 1

(a) Mice and rats can infer the azimuthal location of an object by touch. (b) The angular position of a whisker, β, (during whisking) is shown as a function of time. The angle is often modeled as β(t) = βmidpoint(t) + Δβ(t) cos(ϕ(t)), where βmidpoint(t) and Δβ(t) are the midpoint and the whisking amplitude, respectively. (c) The whisking phase ϕ as a function of time is ϕ(t) = (νt)mod2π, where ν is the angular frequency of the whisking. (d) & (e) Raster plot and normalized tuning curve of a neuron with a preferred phase near maximal retraction.

Fig 2. Distribution of preferred phases.

Fig 2

(a) The von-Mises distribution, Eq (1), with ψ = 0 is shown for different values of κ as depicted by color. (b) The distribution of preferred phases in the uniform pooling model. The distribution was estimated from 10000 repetitions of drawing N = 100 neurons from the upstream population. The preferred phases in the upstream population were drawn in an i.i.d. manner following Eq (1), with κUpStream = 1 and ψUpStream = 0. The dashed red line depicts the mean phase. (c) Distribution width in the naive uniform/random pooling model. The width of the distribution of preferred phases, κL4I, in the downstream layer (L4I) is shown as a function of the number of pooled VPM neurons, N, for the uniform and random pooling in squares and circles, respectively. The width of the distribution, κL4I, was estimated from 10000 repetitions of drawing N preferred phases of the upstream population with κ = 1 (blue) and κ = 2 (red).

Throughout this work we ignore the possible contribution of recurrent connections to the shaping of phase selectivity in layer 4 (but see Discussion). Assuming the rhythmic input to L4I neurons originates solely from the VPM, the distribution of preferred phases of L4I neurons is determined by the distribution in the thalamus and the profile of thalamo-cortical synaptic weights. However, synaptic weights have been reported to be highly volatile [1720]. If synaptic plasticity is activity independent, then one would expect the synaptic weight profile to become unstructured. In the naive pooling model, which lacks activity-dependent plasticity, we consider two types of unstructured connectivity; namely, uniform pooling and random pooling.

Fig 2b shows the distribution of preferred phases in a naive model that pools the responses of N = 100 VPM neurons. The source of variability in the preferred phases of the downstream neuron is the random choice of N VPM neurons, each characterized by a preferred phase that is distributed in an i.i.d. manner following Eq (1). As a result, one would expect that the number, N, of pooled thalamic neurons would govern the width of the distribution. Fig 2c shows the expected distribution width, κL4I, as a function of N for uniform (squares) and random (circles) pooling. As can be seen from the figure, even a random pooling of N = 10 results in a distribution of preferred phases that is considerably narrower than empirically observed. This result is particularly surprising since the number of thalamic neurons synapsing onto a single L4 neuron was estimated to be on the order of 100, see e.g. [16, 21]. Moreover, in the naive pooling model, the mean preferred phase of the downstream population is given by the mean phase of the upstream population. The delay in the response of downstream neurons to their input should also be added. It is estimated to be around 1-10ms; hence, for whisking at 10Hz translates into ψVPMψL4I = 0.06-0.6 [rad]. This result also runs counter empirical findings reporting ψVPMψL4I ≈ 2.2 [rad]. Thus, the synaptic weight profile in a model that lacks activity-dependent plasticity is expected to converge to a random pooling (due to constant synaptic remodelling), which is inconsistent with the observed distribution of preferred phases.

Recently, the effects of rhythmic activity on the spike timing dependent plasticity (STDP) dynamics of feed-forward synaptic connections have been examined [22, 23]. It was shown that in this case the synaptic weights remain dynamic. As a result, the phase of the downstream neuron is arbitrary and drifts in time; thus, effectively, inducing a distribution of preferred phases in the downstream population. However, if the phases of the downstream population are arbitrary and drift in time, how can information about the whisking phase be transmitted?

Here we investigated the hypothesis that the distribution of phases in the downstream layer is governed by the interplay of the distribution in the upstream layer and the STDP rule. The remainder of this article is organized as follows. First, we define the network architecture and the STDP learning rule. We then derive a mean-field approximation for the STDP dynamics in the limit of a slow learning rate for a threshold-linear Poisson downstream neuron model. Next, we show that STDP dynamics can generate non-trivial distributions of preferred phases in the downstream population and analyze how the parameters characterizing the STDP govern this distribution. Finally, we summarize the results, discuss the limitations of this study and suggest essential empirical predictions deriving from our theory.

Results

The upstream thalamic population

We model a system of N thalamic excitatory whisking neurons, synapsing in a feed-forward manner onto a single inhibitory L4 barrel cortical neuron. Unless stated otherwise, following [16], in our numerical simulations we used N = 150. The spiking activity of the thalamic neurons is modelled by independent inhomogeneous Poisson processes with an instantaneous firing rate that follows the whisking cycle:

ρk(t)=D(1+γcos[νtϕk]), (2)

where ρk(t) = ∑i δ(ttk,i), k ∈ {1, …N}, is the spike train of the kth thalamic neuron, with {tk,i}i=1 denoting its spike times. The parameter D is the mean firing rate during whisking (averaged over one cycle), γ is the modulation depth, ν is the angular frequency of the whisking, and ϕk is the preferred phase of firing of the kth thalamic neuron. We further assume that the preferred phases in the thalamic population, {ϕk}k=1N, are distributed i.i.d. according to the von-Mises distribution, Eq (1).

The downstream layer 4 inhibitory neuron model

To facilitate the analysis, we model the response of the downstream layer 4 inhibitory (L4I) neuron to its thalamic inputs by the linear Poisson model, which has been used frequently in the past [22, 2431]. Given the thalamic responses, the firing of the L4I neuron follows inhomogeneous Poisson process statistics with an instantaneous firing rate

rL4I(t)=1Nk=1Nwkρk(td), (3)

where d > 0 represents the characteristic delay, and wk is the synaptic weight of the kth thalamic neuron.

Due to the rhythmic nature of the thalamic inputs, Eq (2), and the linearity of the L4I neuron, Eq (3), the L4I neuron will exhibit rhythmic activity:

rL4I(t)=DL4I(1+γL4Icos[νtψL4I]), (4)

with a mean, DL4I, a modulation depth, γL4I, and a preferred phase ψL4I, that depend on global order parameters that characterize the thalamocortical synaptic weights population. For large N these order parameters are given by:

w¯(t)=02πPr(ϕ)w(ϕ,t)dϕ (5)

and

w˜(t)eiψ=02πPr(ϕ)w(ϕ,t)eiϕdϕ. (6)

where w is the mean synaptic weight and w˜eiψ is its first Fourier component. The phase ψ is determined by the condition that w˜ is real and non-negative. Consequently, the L4I neurons in our model respond to whisking with a mean DL4I=Dw, a modulation depth of γL4I=γw˜/w, and a preferred phase ψL4I = ψ + νd.

The STDP rule

We model the modification of the synaptic weight, Δw, following either a pre- or post-synaptic spike as the sum of two processes: potentiation (+) and depression (-) [3133], as

Δw=λ[f+(w)K+(Δt)f(w)K(Δt)]. (7)

The parameter λ denotes the learning rate. We further assume a separation of variables and write each term (potentiation and depression) as the product of the function of the synaptic weight, f±(w), and the temporal kernel of the STDP rule, K±t), where Δt = tposttpre is the time difference between pre- and post-synaptic spike times. Following Gütig et al. [33] the weight dependence functions, f±(w), were chosen to be:

f+(w)=(1w)μ (8a)
f(w)=αwμ, (8b)

where α > 0 is the relative strength of depression and μ ∈ [0, 1] controls the non-linearity of the learning rule.

The temporal kernels of the STDP rule are normalized: i.e., ∫K±t)dΔt = 1. Here, for simplicity, we assume that all pairs of pre and post spike times contribute additively to the learning process via Eq (7).

Empirical studies portray a wide variety of temporal kernels [3442]. Specifically, in our work, we used two families of STDP rules: 1. A temporally asymmetric kernel [35, 3739]; 2. A temporally symmetric kernel [38, 4042]. Both of these rules have been observed in the the barrel system of mice, at least for some developmental period [4345]. For the temporally asymmetric kernel we used the exponential model,

K±(Δt)=eΔt/τ±τ±Θ(±Δt), (9)

where Θ(x) is the Heaviside function, and τ± are the characteristic timescales of the potentiation (+) or depression (−). We take τ > τ+ as typically reported.

For the temporally symmetric learning rule we used a difference of Gaussians model,

K±(Δt)=1τ±2πe12(Δtτ±)2, (10)

where τ± are the temporal widths. In this case, the order of firing is not important, only the absolute time difference.

STDP dynamics in the limit of slow learning

In the limit of a slow learning rate, λ → 0, we obtain deterministic dynamics for the mean synaptic weights (see [32] for a detailed derivation)

w˙j(t)λ=Ij+(t)Ij(t) (11)

with

Ij±(t)=f±(wj(t))Γj,post(Δ)K±(Δ)dΔ, (12)

where Γj, L4I(Δ) is the cross-correlation function between the jth thalamic pre-synaptic neuron and the L4I post-synaptic neuron; see detailed derivation in Temporal correlations.

STDP dynamics of thalamocortical connectivity

We simulated the STDP dynamics of 150 upstream thalamic neurons synapsing onto a single L4I neuron in the barrel cortex, see Details of the numerical simulations & statistical analysis.

Fig 3a shows the temporal evolution of the synaptic weights (color-coded by their preferred phases). As can be seen from the figure, the synaptic weights do not relax to a fixed point; instead there is a continuous remodelling of the entire synaptic population. Examining the order parameters, Fig 3b and 3c, reveals that the STDP dynamics converge to a limit cycle.

Fig 3. Simulation of the STDP dynamics.

Fig 3

(a) Synaptic weight dynamics. Each trace depicts the time evolution of a single synaptic weight, differentiated by color according to its preferred phase, see legend. (b) & (c) Dynamics of the order parameters. The preferred phase of the downstream neuron, κL4I, (in (b)), and the mean, w, and the magnitude of the first Fourier component, w˜, (in red and black, respectively, in (c)) are shown as a function of time. (d) The distribution of preferred phases in the thalamic population that served as input to the downstream L4I neuron is presented as a polar histogram. The distribution followed Eq (1) with κVPM = 1 and ψVPM = 5π/6 rad. (e) The temporal distribution of preferred phases of the downstream L4I neuron is shown in a polar plot. Fitting the von-Mises distribution yielded κL4I = 1.1 and κL4I = 0.8 rad. In this simulation we used the following parameters: N = 150, ν=ν/(2π)=7hz, γVPM = 1, DVPM = 10hz, and d = 3ms. The temporally asymmetric STDP rule, Eq (9), was used with τ = 50ms, τ+ = 22ms, μ = 0.01, α = 1.1, and λ = 0.01.

The continuous remodelling of the synaptic weights causes the preferred phase of the downstream neuron, κL4I (see Eq (4)) to drift in time, Fig 3b. As can be seen from the figure, the drift velocity is not constant. Consequently, the downstream (L4I) neuron ‘spends’ more time in certain phases than others. Thus, the STDP dynamics induce a distribution over time for the preferred phases of the downstream neuron. One can estimate the distribution of the preferred phases of L4I neurons by tracking the phase of a single neuron over time. Alternatively, since our model is purely feed-forward, the preferred phases of different L4I neurons are independent; hence, this distribution can also be estimated by sampling the preferred phases of different L4I neurons at the same time. Fig 3d and 3e show the distribution of preferred phases of thalamic and L4I whisking neurons, respectively. Thus, STDP induces a distribution of preferred phases in the L4I population, which is linked to the temporal distribution of single L4I neurons.

Fig 4a shows the (non-normalized) distribution of preferred phases induced by STDP as a function of the number of pooled thalamic neurons, N. For large N, the STDP dynamics converge to the continuum limit of Eq (11), and the distribution converges to a limit that is independent of N. This contrasts with the naive pooling model lacking plasticity (cf. Fig 4b). In that model, the distribution of preferred phases in the cortical population results from a random process of pooling phases from the thalamic population. We shall refer to this type of variability as quenched disorder, since this randomness is frozen and does not fluctuate over time. This process is characterized by a distribution of preferred phases with vanishing width, κL4I → ∞, in the large N limit, Fig 4b. In addition to the different widths of the distribution, the mean of the preferred phases also differs greatly. Whereas in Fig 4a the mean preferred phase is determined by the STDP rule, without STDP, Fig 4b, the mean phases of L4I neurons is given by the mean preferred phase of the VPM population shifted by the delay, κL4I = ψVPM + .

Fig 4. The effects of population size.

Fig 4

The distribution of the preferred phases of L4I neurons, κL4I, is shown as a function of N for different sources of variability. In each column (value of N), the non-normalized distribution: Pr(ψ)/max{Pr(ψ)}, is presented by color. (a) The temporal distribution of the preferred phase of a single L4I neuron as a result of STDP dynamics without quenched disorder or averaging (see Modeling pre-synaptic phase distributions) is presented. (b) Distribution due to quenched disorder without STDP. The distribution of L4I neurons in the randomly pooling model was estimated from 1000 trials of drawing the preferred phases of N VPM neurons in an i.i.d. manner from Eq (1). (c) The distribution due to both quenched disorder and STDP dynamics is shown. Unless stated otherwise, the parameters used in these graphs are as follows: ν=ν/(2π)=7Hz, κVPM = 1, γ = 0.9, D = 10hz, d = 3ms, ϕ0 = 5π/6. For the STDP we used the temporally exponential asymmetric kernel, Eq (9), with τ = 50ms, τ+ = 22ms, μ = 0.01, α = 1.1, and λ = 0.01.

For small N, N ⪅ 70, STDP induces a point measure distribution over time, Fig 4a. This is due to pinning in the noiseless STDP dynamics in the mean field limit, λ → 0. Stochastic dynamics, due to noisy neuronal responses, can overcome pinning. Fig 4c depicts the distribution of the preferred phases for different values of N, which results from both STDP dynamics and quenched averaging over different realizations of the preferred phases of the thalamic neurons. For small values of N, the distribution is dominated by the quenched statistics, in terms of its narrow width and mean that is dominated by the mean phase of the thalamic population. As N increases, the distribution widens and is centered around a preferred phase that is determined by the STDP. Thus, activity dependent plasticity helps to shape the distribution of preferred phases in the downstream population. Below, we examine how these different parameters affect the ways in which STDP shapes the distribution; hence, we will not average over the quenched disorder.

Parameters characterizing the upstream input

Fig 5a depicts the distribution of preferred phases as a function of the distribution width of their thalamic input, κVPM. For a uniform input distribution, κVPM = 0, the downstream distribution is also uniform, κL4I = 0, see [22]. As the distribution in the VPM becomes narrower, so does the distribution in the L4I population. If the distribution of the thalamic population is narrower than a certain critical value, STDP will converge to a fixed point and κL4I will diverge. Typically, we find that the width of the cortical distribution, κL4I, is similar to or larger than that of the upstream distribution, κVPM, Fig 5b. This sharpening is obtained via STDP by selectively amplifying certain phases while attenuating others. Consequently the rhythmic component, in terms of the modulation depth, γL4I, is also typically amplified relative to the uniform pooling model, Fig 5c.

Fig 5. The effect of the distribution width of the upstream population, κVPM.

Fig 5

(a) The (non-normalized) distribution of the preferred phases of L4I neurons, κL4I, is shown by color as a function of the width of the distribution of preferred phases in VPM, κVPM. (b) The distribution width in layer 4, κL4I, is shown as a function of κVPM (blue). The identity line is shown (dashed black) for comparison. (c) The modulation depth in the downstream population, γL4I, is shown as a function of κVPM (blue). For comparison, the modulation depth of the uniform pooling model is also presented (dashed red). The parameters used here are: ν=ν/(2π)=7Hz, κVPM = 1, γ = 1, D = 10hz, d = 3ms, ϕ0 = 5π/6. For the STDP we used the temporally exponential asymmetric kernel, Eq (9), with τ = 50ms, τ+ = 22ms, μ = 0.01, α = 1.1, and λ = 0.01.

The effect of the whisking frequency is shown in Fig 6a. For moderate rhythms that are on a similar timescale to that of the STDP rule, STDP dynamics can generate a wide distribution of preferred phases. However, in the high frequency limit, ν → ∞, the synaptic weights converge to a uniform solution with w(ϕ) = (1 + α1/μ)−1, ∀ϕ (see [22]). In this limit, due to the uniform pooling, there is no selective amplification of phases and the whisking signal is transmitted downstream due to the selectivity of the thalamic population, κVPM > 0. Consequently, at high frequencies, the distribution of preferred phases will be extremely narrow, O(1/N) because of the quenched disorder, and the rhythmic signal will be attenuated, γL4I = γVPM × I1(κ)/I0(κ) (where Ij(κ) is the modified Bessel function of order j), Fig 6b. The rate of convergence to the high frequency limit is governed by the smoothness of the STDP rule. Discontinuity in the STPD rule, such as in our choice of a temporally asymmetric rule, will induce algebraic convergence in ν to the high frequency limit, whereas a smooth rule, such as our choice of a temporally symmetric rule, will manifest in exponential convergence; compare Fig 6 with Fig 7, see also [22]. In our choice of parameters, τ+ ≈ 20ms and τ ≈ 50ms, STDP dynamics induced a wide distribution for frequencies in the α, β and low γ bands in the case of the asymmetric learning rule Fig 6, and around 5–15hz in the case of the symmetric learning rule Fig 7.

Fig 6. Effects of whisking frequency.

Fig 6

(a) The (non-normalized) distribution of L4I neuron phases, κL4I, is depicted by color as a function, ν=ν/(2π). (b) The width of the distribution, κL4I, of L4I neurons is shown as a function of ν. The parameters used in these graphs were: κVPM = 1, ψVPM = 5π/6 γ = 0.9, D = 10hz, and d = 3ms. We used the temporally asymmetric exponential learning rule, Eq (9), with τ = 50ms, τ+ = 22ms, μ = 0.01, α = 1.1, and λ = 0.01.

Fig 7. Effects of whisking frequency—The temporally symmetric STDP rule.

Fig 7

(a) The (non-normalized) distribution of L4I neuron phases, κL4I, is depicted by color as a function, ν=ν/(2π). (b) The width of the distribution, κL4I, of L4I neurons, is shown as a function of ν. Unless stated otherwise, the parameters used here were: κVPM = 1, ψVPM = 5π/6, γ = 0.9, D = 10hz, and d = 10ms. We used the temporally asymmetric exponential learning rule, Eq (10), with: τ = 50ms, τ+ = 22ms, μ = 0.01, α = 1.1, and λ = 0.01.

The effects of synaptic weight dependence, μ

Previous studies have shown that increasing μ weakens the positive feedback of the STDP dynamics, which generates multi-stability, and stabilizes more homogeneous solutions [22, 23, 33, 46]. This transition is illustrated in Fig 8: at low values of μ, the STDP dynamics converge to a limit cycle in which both the synaptic weights and the phase of the the L4I neuron cover their entire dynamic range, Fig 8a and 8b. As μ is increased, the synaptic weights become restricted in the limit cycle and no longer span their entire dynamic range, Fig 8c and 8d. A further increase of μ also restricts the phase of the L4I neuron along the limit cycle, Fig 8e and 8f. Finally, when μ is sufficiently large, STDP dynamics converge to a fixed point, Fig 8g and 8h. This fixed point selectively amplifies certain phases, yielding a higher value of γ than in the uniform solution. These results are summarized in Fig 9 that shows the (non-normalized) distribution of preferred phases and the relative strength of the rhythmic component, γL4I, as a function of μ. Note that except for a small value range of μ, around which the distribution of preferred phases is bi-modal (see, e.g. Fig 8e and 8f and μ ≈ 0.07 in Fig 9), STDP yields higher γL4I values than in the uniform pooling model.

Fig 8. Transition to a fixed point solution.

Fig 8

(a), (c), (e) & (f) show the synaptic weight dynamics for different values of μ. The synaptic weights are depicted by different traces, colored according to the preferred phase of the pre-synaptic neuron, see legend. (b), (d), (f) & (h) show the preferred phase of the downstream L4I neuron, κL4I, as a function of time. The parameters used in these simulation were: N = 150, κ = 1, ν=ν/(2π)=7Hz, γ = 0.9, D = 10Hz, and d = 3ms. We used here the temporally asymmetric learning rule, Eq (9), with: τ = 50ms, τ+ = 22ms, and α = 1.1. For the non-linearity parameter we used: μ = 0.01 in (a) & (b), μ = 0.06. in (c) & (d), μ = 0.07 in (e) & (f). and μ = 0.1 in (g) & (h).

Fig 9. The effect of the non-linearity parameter, μ.

Fig 9

(a) The distribution of preferred phases of L4I neurons, κL4I, is shown by color as a function of μ. (b) The modulation depth, γL4I, is depicted as a function of μ. The parameters used here were: κVPM = 1, ψVPM = 5π/6, γ = 0.9, D = 10hz, and d = 3ms. The temporally asymmetric STDP rule, Eq (9), was used with: τ = 50ms, τ+ = 22ms, α = 1.1 and λ = 0.01.

The relative strength of depression, α

As one might expect, decreasing α beyond a certain value will result in a potentiation dominated STDP dynamics, saturating all synapses close to their maximal value, thus, approaching the uniform solution, which is characterized by a narrow preferred phase distribution centered around the mean preferred phase of the VPM neurons shifted by the delay, and low values of γL4I, Fig 10. Increasing α strengthens the competitive winner-take-all like nature of the STDP dynamics. Initially, this competition will generate a fixed point with non-uniform synaptic weights, thus increasing γL4I. A further increase of α results in a limit cycle solution to the STDP dynamics that widens the distribution of the preferred phases. Increasing α beyond a certain critical value will result in depression dominated STDP dynamics, thus driving the synaptic weights to zero, Fig 10.

Fig 10. The effect of the relative strength of depression, α.

Fig 10

(a) The mean input phase κL4I as a function of α. Probability of occurrences is depicted by color as shown in the right color bar. (b) The order parameters w and w˜ are shown as a function of α. Here, we used: κVPM = 1, ψ0 = 5π/6, γ = 1, D = 10hz, and d = 3ms. The temporally asymmetric STDP rule, Eq (9), was used with: τ = 50ms, τ+ = 22ms, μ = 0.01 and λ = 0.01.

Parameters characterizing the temporal structure of STDP

Fig 11 shows the effect of the temporal structure of the STDP on the distribution of preferred phases in the downstream population. As can be seen from the figure, varying the characteristic timescales of potentiation and depression, τ+ and τ, induces quantitative effects on the distribution of the preferred phases. However, qualitative effects result from changing the nature of the STDP rule. Comparing the temporally asymmetric rule, Fig 11a and 11b, to the temporally symmetric rule, Fig 11c and 11d, reveals a dramatic difference in the mean preferred phase of L4I neurons, dashed black lines.

Fig 11. The effect of the temporal structure of the STDP rule.

Fig 11

The distribution of preferred phases of L4I neurons, κL4I, is shown as a function of the characteristic timescales of the STDP: τ in (a) & (c) and τ + in (b) & (d), for the temporally asymmetric rule in (a) & (b) and the symmetric rule in (c) & (d). The dashed black lines depict the mean phase. Unless stated otherwise, the parameters used here were: κVPM = 1, γ = 1, D = 10hz, τ = 50ms, τ+ = 22ms, d = 10ms, ψVPM = 5π/6, μ = 0.01, α = 1.1 and λ = 0.01.

Discussion

We examined the possible contribution of STDP to the diversity of preferred phases of whisking neurons. Whisking neurons can be found along different stations throughout the somatosensory information processing pathway [1, 46, 9, 10, 13, 16, 47]. Here we focused on L4I whisking neurons that receive their whisking input mainly via excitatory inputs from the VPM, and suggested that the non-trivial distribution of preferred phases of L4I neurons results from a continuous process of STDP.

STDP has been reported in thalamocortical connections in the barrel system. However, STDP has only been observed during the early stages of development [4345]. Is it possible that the contribution of STDP to shaping thalamocortical connectivity is restricted to the early developmental stages? This is an empirical question that can only be answered experimentally. Nevertheless, several comments can be made.

First, Inglebert and colleagues recently showed that pairing single pre- and post-synaptic spikes, under normal physiological conditions, does not induce plastic changes [48]. On the other hand, activity dependent plasticity was observed when stronger activity was induced. Thus, whisking activity, which is a strong collective and synchronized activity, may induce STDP of thalamocortical connectivity.

Second, in light of the considerable volatility of synaptic connections observed in the central nervous system [1720], it is hard to imagine that thalamocortical connectivity remains unchanged throughout the lifetime of an animal. Third, if activity independent plasticity alone underlies synaptic volatility, thalamocortical synaptic weights would be expected to be random. In this case, thalamic whisking input to layer 4 should be characterized by an extremely narrow distribution centered around the delayed mean thalamic preferred phase, cf. Fig 4b. Consequently, neglecting the possible contributions due to recurrent interactions with layer 4, a non-trivial distribution of L4I phases, with κL4I ∼ 1, can be obtained either via STDP or by pooling the whisking signal from an extremely small VPM population of N < 10 neurons. However, in the latter scenario (without STDP) the mean preferred phase of L4I neurons is expected to be determined by the (delayed) mean phase of VPM neurons (see e.g. Fig 4b), thus raising serious doubts as to the viability of the latter solution.

An alternative hypothesis to our proposed mechanism is that the recurrent dynamics within layer 4 generate the empirically observed wide distribution of preferred phases centered on κL4IψVPM − 2.2 [rad]. As layer 4 excitatory neurons have been reported to exhibit very weak rhythmic activity [13, 16], such a mechanism would mainly be based on recurrent interaction within the L4I population. The putative ways in which inhibitory interactions widen the distribution remain unclear. Nevertheless, this hypothesis should be explored further.

Our hypothesis views STDP as a continuous process. As a result of STDP, the preferred phase of a L4I neuron, in our model, is a stochastic process with a non-uniform drift, Fig 3b, which, in turn, induces a distribution over time for the phase, Fig 3e. As our model lacks recurrent interactions, phases of different L4I neurons will be i.i.d. random processes, albeit with different random initial conditions. Consequently, simultaneous sampling of the distribution of preferred phases of a population of L4I neurons will yield the same distribution as obtained by sampling the preferred phase of a single neuron over time. Thus, functionality, in terms of transmission of the whisking signal and retaining the stationary distribution of the preferred phases in the downstream population is obtained as a result of continuous remodelling of the entire population of synaptic weights.

Several key features of our hypothesis provide clear empirical predictions. Our theory links two empirical observations. One is the distribution of preferred phases in up- and downstream populations. The other is the STDP rule. The literature can provide a good assessment of the distribution of preferred phases of thalamic and of cortical whisking neurons. Further effort is required to estimate the STDP rule of the thalamocortical connections to L4I neurons. Thus, we can reject our hypothesis if the predicted distribution, based on our theory and the estimated learning rule, is not consistent with the empirically observed distribution.

Our theory also posits that the distribution of preferred phases results from their stochastic dynamics. Our theory also incorporates a direct relationship between the drift velocity and the distribution of preferred phases. Essentially, the system will spend more time around phases characterized by slower velocity; hence, a higher probability of preferred phases should correlate with slower drift velocity. To the best of our knowledge, the dynamics of the preferred phases of whisking neurons has not been studied. Monitoring the preferred phases of single L4I neurons over long periods of time (days and weeks) could serve as additional tests of our theory. Specifically, it would address the question of whether the preferred phases drift in time, and if so, whether the drift velocity and the distribution of preferred phases are consistent with our theory.

Further empirical exploration can test the degree in which our approximation of a purely feed-forward architecture (i.e., neglecting recurrent interactions) holds. Simultaneous recordings of a large population of fast spiking layer 4 whisking neurons can be harnessed to compare the distribution of preferred phases over space and time, which are expected to be identical in the absence of recurrent interactions. In addition, in the visual system, Lien and Scanziani combined intracellular recordings with optogenetic methods to assess the contribution of thalamic inputs to the selectivity of cortical neurons [49]. In the context of the current work, we explored the thalamic inputs to whisking neurons in the barrel cortex. Thus, applying ideas and approaches from the work of Lien and Scanziani could (i) provide a measurement of the thalamic input, and (ii) determine the relative contribution of feed-forward and recurrent connections to the selectivity to the phase of whisking in L4I neurons.

More broadly, our findings raise the question of whether the constant remodelling of synaptic efficacies is an artefact of the ‘biological hardware’ the brain must overcome, or whether it reflects an underlying principle of the central nervous system. Though we cannot provide an answer to this question at this point in time, we may speculate on the possible advantages of a dynamic solution. One interesting hypothesis to consider is that the constant remodelling makes the neuronal network more flexible and consequently enables a faster response to a changing environment.

In the current study, we made several simplifying assumptions. The spiking activity of the thalamic population was modeled as a rhythmic activity with a well-defined frequency. However, whisking activity spans a frequency band of several Hertz. Moreover, the thalamic input relays additional signals, such as touch and texture. These signals modify the cross-correlation structure and add ‘noise’ to the dynamics of the preferred phase of the downstream neuron. As a result, the distribution of preferred phases in the downstream population is likely to be wider. In addition, our analysis used a purely feed-forward architecture ignoring recurrent connections in layer 4, which may also affect the preferred phases in layer 4. A quantitative application of our theory to the whisker system should consider all these effects. Nevertheless, the theory presented provides the basic foundation to address these effects.

Methods

Temporal correlations

The cross-correlation between pre-synaptic neurons j and k at time difference Δt is given by:

Γ(j,k)(Δt)=ρj(t)ρk(t+Δt)=D2(1+γ22cos[νΔt+ϕjϕk])+δjkDδ(Δt). (13)

In the linear Poisson model, Eq (3), the cross-correlation between a pre-synaptic neuron and the post-synaptic neuron can be written as a linear combination of the cross-correlations in the upstream population; hence, the cross-correlation between the jth VPM neuron and the post-synaptic neuron is

Γj,post(Δt)=DNδ(Δtd)wj+D2(w¯+γ22w˜cos[ν(Δtd)+ϕjψ]). (14)

Where w and w˜eiψ are order parameters characterizing the synaptic weights profile, as defined in Eqs (5) and (6).

The mean field Fokker-Planck dynamics

For large N we obtain the continuum limit from Eq (11):

w˙(ϕ,t)λ=Fd(ϕ,t)+w¯(t)F0(ϕ,t)+w˜(t)F1(ϕ,t), (15)

where

Fd(ϕ,t)=w(ϕ,t)DN(f+(w(ϕ,t))K+(d)f(w(ϕ,t))K(d)), (16a)
F0(ϕ,t)=D2(K¯+f+(w(ϕ,t))K¯f(w(ϕ,t))), (16b)
F1(ϕ,t)=D2γ22(K˜+f+(w(ϕ,t))cos[ϕΩ+νdψ]K˜f(w(ϕ,t))cos[ϕΩνdψ]), (16c)

and K±, K˜±eiΩ±η are the Fourier transforms of the STDP kernels

K¯±=K±(Δ)dΔ, (17)
K˜±eiΩ±=K±(Δ)eiνΔdΔ. (18)

Note that in our specific choice of kernels, K±=1, by construction.

Fixed points of the mean field dynamics

The fixed point solution of Eq (15) is given by

w(ϕ)*=(1+α1/μ(1+X1+X+)1/μ)1, (19)

where

X±w˜w¯γ22K˜±cos(ϕνdΩ±ψ). (20)

Note that, from Eqs (19) and (20), the fixed point solution, w(ϕ)*, will depend on ϕ, for κVPM > 0. As μ grows to 1 the fixed point solution will become more uniform, see [22].

Details of the numerical simulations & statistical analysis

Scripts of the numerical simulations were written in Matlab. The numerical results presented in this paper were obtained by solving Eq (15) with the Euler method with a time step Δt = 0.1 and λ = 0.01. The cftool with the Nonlinear Least Squares method and a confidence level of 95% for all parameters was used for fitting the data presented in Figs 3, 5b, 6b and 7b.

Modeling pre-synaptic phase distributions

Unless stated otherwise, STDP dynamics in the mean field limit was simulated without quenched disorder. To this end, the preferred phase, ϕk, of the kth neuron in a population of N pre-synaptic VPM neurons was set by the condition πφkPr(φ)dϕ=k/N. In Fig 4b and 4c we used the accept-reject method [50, 51] to sample the phases.

Data Availability

All relevant data are within the manuscript and its Supporting Information files.

Funding Statement

Maoz Shamir received the following grant: The Israel Science Foundation (ISF) grant number 300/16. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Isett BR, Feldman DE. Cortical Coding of Whisking Phase during Surface Whisking. Current Biology. 2020;. doi: 10.1016/j.cub.2020.05.064 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Kleinfeld D, Deschênes M. Neuronal basis for object location in the vibrissa scanning sensorimotor system. Neuron. 2011;72(3):455–468. doi: 10.1016/j.neuron.2011.10.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Ahissar E, Knutsen PM. Object localization with whiskers. Biological cybernetics. 2008;98(6):449–458. doi: 10.1007/s00422-008-0214-4 [DOI] [PubMed] [Google Scholar]
  • 4.Diamond ME, Von Heimendahl M, Knutsen PM, Kleinfeld D, Ahissar E. ‘Where’ and ‘what’in the whisker sensorimotor system. Nature Reviews Neuroscience. 2008;9(8):601–612. doi: 10.1038/nrn2411 [DOI] [PubMed] [Google Scholar]
  • 5.Severson KS, Xu D, Van de Loo M, Bai L, Ginty DD, O’Connor DH. Active touch and self-motion encoding by merkel cell-associated afferents. Neuron. 2017;94(3):666–676. doi: 10.1016/j.neuron.2017.03.045 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Moore JD, Lindsay NM, Deschênes M, Kleinfeld D. Vibrissa self-motion and touch are reliably encoded along the same somatosensory pathway from brainstem through thalamus. PLoS Biol. 2015;13(9):e1002253. doi: 10.1371/journal.pbio.1002253 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Campagner D, Evans MH, Bale MR, Erskine A, Petersen RS. Prediction of primary somatosensory neuron activity during active tactile exploration. Elife. 2016;5:e10696. doi: 10.7554/eLife.10696 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Urbain N, Salin PA, Libourel PA, Comte JC, Gentet LJ, Petersen CC. Whisking-related changes in neuronal firing and membrane potential dynamics in the somatosensory thalamus of awake mice. Cell reports. 2015;13(4):647–656. doi: 10.1016/j.celrep.2015.09.029 [DOI] [PubMed] [Google Scholar]
  • 9.Wallach A, Bagdasarian K, Ahissar E. On-going computation of whisking phase by mechanoreceptors. Nature neuroscience. 2016;19(3):487–493. doi: 10.1038/nn.4221 [DOI] [PubMed] [Google Scholar]
  • 10.Szwed M, Bagdasarian K, Ahissar E. Encoding of vibrissal active touch. Neuron. 2003;40(3):621–630. doi: 10.1016/S0896-6273(03)00671-8 [DOI] [PubMed] [Google Scholar]
  • 11.Yu C, Derdikman D, Haidarliu S, Ahissar E. Parallel thalamic pathways for whisking and touch signals in the rat. PLoS Biol. 2006;4(5):e124. doi: 10.1371/journal.pbio.0040124 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ego-Stengel V, Le Cam J, Shulz DE. Coding of apparent motion in the thalamic nucleus of the rat vibrissal somatosensory system. Journal of Neuroscience. 2012;32(10):3339–3351. doi: 10.1523/JNEUROSCI.3890-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Yu J, Gutnisky DA, Hires SA, Svoboda K. Layer 4 fast-spiking interneurons filter thalamocortical signals during active somatosensation. Nature neuroscience. 2016;19(12):1647–1657. doi: 10.1038/nn.4412 [DOI] [PubMed] [Google Scholar]
  • 14.Grion N, Akrami A, Zuo Y, Stella F, Diamond ME. Coherence between rat sensorimotor system and hippocampus is enhanced during tactile discrimination. PLoS biology. 2016;14(2):e1002384. doi: 10.1371/journal.pbio.1002384 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Yu J, Hu H, Agmon A, Svoboda K. Recruitment of GABAergic interneurons in the barrel cortex during active tactile behavior. Neuron. 2019;104(2):412–427. doi: 10.1016/j.neuron.2019.07.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Gutnisky DA, Yu J, Hires SA, To MS, Bale M, Svoboda K, et al. Mechanisms underlying a thalamocortical transformation during active tactile sensation. PLoS computational biology. 2017;13(6):e1005576. doi: 10.1371/journal.pcbi.1005576 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Loewenstein Y, Kuras A, Rumpel S. Multiplicative dynamics underlie the emergence of the log-normal distribution of spine sizes in the neocortex in vivo. Journal of Neuroscience. 2011;31(26):9481–9488. doi: 10.1523/JNEUROSCI.6130-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Mongillo G, Rumpel S, Loewenstein Y. Intrinsic volatility of synaptic connections—a challenge to the synaptic trace theory of memory. Current opinion in neurobiology. 2017;46:7–13. doi: 10.1016/j.conb.2017.06.006 [DOI] [PubMed] [Google Scholar]
  • 19.Mongillo G, Rumpel S, Loewenstein Y. Inhibitory connectivity defines the realm of excitatory plasticity. Nature neuroscience. 2018;21(10):1463–1470. doi: 10.1038/s41593-018-0226-x [DOI] [PubMed] [Google Scholar]
  • 20.Ziv NE, Brenner N. Synaptic tenacity or lack thereof: spontaneous remodeling of synapses. Trends in neurosciences. 2018;41(2):89–99. doi: 10.1016/j.tins.2017.12.003 [DOI] [PubMed] [Google Scholar]
  • 21.Cruikshank SJ, Lewis TJ, Connors BW. Synaptic basis for intense thalamocortical activation of feedforward inhibitory cells in neocortex. Nature neuroscience. 2007;10(4):462–468. doi: 10.1038/nn1861 [DOI] [PubMed] [Google Scholar]
  • 22.Luz Y, Shamir M. Oscillations via Spike-Timing Dependent Plasticity in a Feed-Forward Model. PLoS computational biology. 2016;12(4):e1004878. doi: 10.1371/journal.pcbi.1004878 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Sherf N, Maoz S. Multiplexing rhythmic information by spike timing dependent plasticity. PLoS computational biology. 2020;16(6):e1008000. doi: 10.1371/journal.pcbi.1008000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Softky WR, Koch C. The highly irregular firing of cortical cells is inconsistent with temporal integration of random EPSPs. Journal of Neuroscience. 1993;13(1):334–350. doi: 10.1523/JNEUROSCI.13-01-00334.1993 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Shadlen MN, Newsome WT. The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. Journal of neuroscience. 1998;18(10):3870–3896. doi: 10.1523/JNEUROSCI.18-10-03870.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Masquelier T, Hugues E, Deco G, Thorpe SJ. Oscillations, phase-of-firing coding, and spike timing-dependent plasticity: an efficient learning scheme. Journal of Neuroscience. 2009;29(43):13484–13493. doi: 10.1523/JNEUROSCI.2207-09.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Morrison A, Diesmann M, Gerstner W. Phenomenological models of synaptic plasticity based on spike timing. Biological cybernetics. 2008;98(6):459–478. doi: 10.1007/s00422-008-0233-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Song S, Miller KD, Abbott LF. Competitive Hebbian learning through spike-timing-dependent synaptic plasticity. Nature neuroscience. 2000;3(9):919. doi: 10.1038/78829 [DOI] [PubMed] [Google Scholar]
  • 29.Kempter R, Gerstner W, Hemmen JLv. Intrinsic stabilization of output rates by spike-based Hebbian learning. Neural computation. 2001;13(12):2709–2741. doi: 10.1162/089976601317098501 [DOI] [PubMed] [Google Scholar]
  • 30.Kempter R, Gerstner W, Van Hemmen JL. Hebbian learning and spiking neurons. Physical Review E. 1999;59(4):4498. doi: 10.1103/PhysRevE.59.4498 [DOI] [Google Scholar]
  • 31.Luz Y, Shamir M. Balancing feed-forward excitation and inhibition via Hebbian inhibitory synaptic plasticity. PLoS computational biology. 2012;8(1):e1002334. doi: 10.1371/journal.pcbi.1002334 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Luz Y, Shamir M. The effect of STDP temporal kernel structure on the learning dynamics of single excitatory and inhibitory synapses. PloS one. 2014;9(7):e101109. doi: 10.1371/journal.pone.0101109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Gütig R, Aharonov R, Rotter S, Sompolinsky H. Learning input correlations through nonlinear temporally asymmetric Hebbian plasticity. Journal of Neuroscience. 2003;23(9):3697–3714. doi: 10.1523/JNEUROSCI.23-09-03697.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Markram H, Lübke J, Frotscher M, Sakmann B. Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science. 1997;275(5297):213–215. doi: 10.1126/science.275.5297.213 [DOI] [PubMed] [Google Scholar]
  • 35.Bi Gq, Poo Mm. Synaptic modifications in cultured hippocampal neurons: dependence on spike timing, synaptic strength, and postsynaptic cell type. Journal of neuroscience. 1998;18(24):10464–10472. doi: 10.1523/JNEUROSCI.18-24-10464.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Sjöström PJ, Turrigiano GG, Nelson SB. Rate, timing, and cooperativity jointly determine cortical synaptic plasticity. Neuron. 2001;32(6):1149–1164. doi: 10.1016/S0896-6273(01)00542-6 [DOI] [PubMed] [Google Scholar]
  • 37.Zhang LI, Tao HW, Holt CE, Harris WA, Poo Mm. A critical window for cooperation and competition among developing retinotectal synapses. Nature. 1998;395(6697):37. doi: 10.1038/25665 [DOI] [PubMed] [Google Scholar]
  • 38.Abbott LF, Nelson SB. Synaptic plasticity: taming the beast. Nature Neuroscience. 2000;3(11):1178–1183. doi: 10.1038/81453 [DOI] [PubMed] [Google Scholar]
  • 39.Froemke RC, Tsay IA, Raad M, Long JD, Dan Y. Contribution of individual spikes in burst-induced long-term synaptic modification. Journal of neurophysiology. 2006;95(3):1620–1629. doi: 10.1152/jn.00910.2005 [DOI] [PubMed] [Google Scholar]
  • 40.Nishiyama M, Hong K, Mikoshiba K, Poo MM, Kato K. Calcium stores regulate the polarity and input specificity of synaptic modification. Nature. 2000;408:584–588. doi: 10.1038/35046067 [DOI] [PubMed] [Google Scholar]
  • 41.Shouval HZ, Bear MF, Cooper LN. A unified model of NMDA receptor-dependent bidirectional synaptic plasticity. Proceedings of the National Academy of Sciences. 2002;99(16):10831–10836. doi: 10.1073/pnas.152343099 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Woodin MA, Ganguly K, ming Poo M. Coincident Pre- and Postsynaptic Activity Modifies GABAergic Synapses by Postsynaptic Changes in Cl− Transporter Activity. Neuron. 2003;39(5):807—820. doi: 10.1016/S0896-6273(03)00507-5 [DOI] [PubMed] [Google Scholar]
  • 43.Itami C, Kimura F. Developmental switch in spike timing-dependent plasticity at layers 4–2/3 in the rodent barrel cortex. Journal of Neuroscience. 2012;32(43):15000–15011. doi: 10.1523/JNEUROSCI.2506-12.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Itami C, Huang JY, Yamasaki M, Watanabe M, Lu HC, Kimura F. Developmental switch in spike timing-dependent plasticity and cannabinoid-dependent reorganization of the thalamocortical projection in the barrel cortex. Journal of Neuroscience. 2016;36(26):7039–7054. doi: 10.1523/JNEUROSCI.4280-15.2016 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Kimura F, Itami C. A hypothetical model concerning how spike-timing-dependent plasticity contributes to neural circuit formation and initiation of the critical period in barrel cortex. Journal of Neuroscience. 2019;39(20):3784–3791. doi: 10.1523/JNEUROSCI.1684-18.2019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Morrison A, Diesmann M, Gerstner W. Phenomenological models of synaptic plasticity based on spike timing. Biological cybernetics. 2008;98(6):459–478. doi: 10.1007/s00422-008-0233-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Ni J, Chen JL. Long-range cortical dynamics: a perspective from the mouse sensorimotor whisker system. European Journal of Neuroscience. 2017;46(8):2315–2324. doi: 10.1111/ejn.13698 [DOI] [PubMed] [Google Scholar]
  • 48.Inglebert Y, Aljadeff J, Brunel N, Debanne D. Synaptic plasticity rules with physiological calcium levels. Proceedings of the National Academy of Sciences. 2020;117(52):33639–33648. doi: 10.1073/pnas.2013663117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lien AD, Scanziani M. Tuned thalamic excitation is amplified by visual cortical circuits. Nature neuroscience. 2013;16(9):1315–1323. doi: 10.1038/nn.3488 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Chib S, Greenberg E. Understanding the metropolis-hastings algorithm. The american statistician. 1995;49(4):327–335. doi: 10.1080/00031305.1995.10476177 [DOI] [Google Scholar]
  • 51.Robert C, Casella G. Monte Carlo statistical methods. Springer Science & Business Media; 2013. [Google Scholar]
PLoS Comput Biol. doi: 10.1371/journal.pcbi.1009353.r001

Decision Letter 0

Lyle J Graham, Abigail Morrison

Transfer Alert

This paper was transferred from another journal. As a result, its full editorial history (including decision letters, peer reviews and author responses) may not be present.

12 Jul 2021

Dear Mr. Sherf,

Thank you very much for submitting your manuscript "STDP and the distribution of preferred phases in the whisker system" for consideration at PLOS Computational Biology. As with all papers reviewed by the journal, your manuscript was reviewed by members of the editorial board and by several independent reviewers. The reviewers appreciated the attention to an important topic. Based on the reviews, we are likely to accept this manuscript for publication, providing that you modify the manuscript according to the review recommendations.

Please prepare and submit your revised manuscript within 30 days. If you anticipate any delay, please let us know the expected resubmission date by replying to this email.

When you are ready to resubmit, please upload the following:

[1] A letter containing a detailed list of your responses to all review comments, and a description of the changes you have made in the manuscript. Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out

[2] Two versions of the revised manuscript: one with either highlights or tracked changes denoting where the text has been changed; the other a clean version (uploaded as the manuscript file).

Important additional instructions are given below your reviewer comments.

Thank you again for your submission to our journal. We hope that our editorial process has been constructive so far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Abigail Morrison

Associate Editor

PLOS Computational Biology

Lyle Graham

Deputy Editor

PLOS Computational Biology

***********************

A link appears below if there are any accompanying review attachments. If you believe any reviews to be missing, please contact ploscompbiol@plos.org immediately:

[LINK]

Reviewer's Responses to Questions

Comments to the Authors:

Please note here if the review is uploaded as an attachment.

Reviewer #1: In the manuscript the authors provide a theoretical account for a puzzling experimental observation in the whisker system, i.e., the distribution of preferred phases in the cortex is not significantly sharper than that observed in the thalamus (contrary to what one would expect according to a simple "pooling" model). The non-sharpening is, according to the theory, a result of synaptic volatility induced by spike-timing dependent plasticity. The theory also makes experimental predictions that could be tested.

The results presented appear correct, as far as I can check, and novel. The assumptions underlying the modeling and its limitations are carefully discussed. The manuscript is well written and the figures informative. I found the results intriguing and certainly worth being disseminated. I have no major comment, just a curiosity: what could be the functional relevance of such a mechanism?

Reviewer #2: Neurons in the VPM and in the barrel cortex are characterized by a preferred phase. The distribution of preferred phases in the two regions is non-uniform and the widths of the two distributions is comparable. In this paper, Sherf and Shamir show that this seemingly innocuous observation is in fact highly non-trivial, and discuss a possible interpretation of this result in the framework of STDP.

I find the results of this paper both interesting and novel and I strongly recommend that it would be published in PCB. However, I think that the presentation of the results is far from optimal and be substantially improved, specifically with respect to the relationship with the biological observations.

Comments:

The important observation that the widths of the distributions are comparable is NOT supported by reference 17, as claimed in line 19. I do not doubt the observation and in fact, there is evidence for it in Yu et al., Nature Neuroscience, 2016 and Kleinfeld and Deschenes, Neuron 2011, references that are cited elsewhere in the manuscript. However, I did not find it in reference 17.

Another related point that should be clarified with respect to the relation with biology is the difference between measurements that are done simultaneously and measurements done in the same animal in different days – what can we learn from the existing literature.

In order to explain why the fact that the two distributions are comparable is surprising, the authors discuss random or uniform pooling. Explaining why such pooling is even relevant is done only in the Discussion, even this only to some extent in the framework of random synaptic changes.

I did not understand lines 244-247 and 251-252 of the Discussion section – they need to be clarified.

Reviewer #3: Rat and mice detect objects by rhythmically swiping there whiskers.

Thalamic neurons which track the phase of the whisking cycle encode

information about the location of the whisker. These neurons respond preferentially to

a particular phase which are narrowly distributed. Cortical neurons pool the rhythmic

excitation from several tens of thalamic neurons. Therefore, one may expect

a very narrow distribution of preferred phase of layer 4 cortical neuron narrow . However, this is not

observed experimentally. The paper argues that this issue can be solved if ihalamo-cortical synapses

exhibit short term dependent plasticity. To this end, the authors combine an analytical calculations

and simulations.

This is is very nice piece of work. The approach is elegant. The results are convincing.

The paper is very well written and I had pleasure to read it. I have only a couple of comments.

1) Has mentioned by the authors at the end of the discussion, they neglected the recurrent interaction

in layer 4. This should be emphasized much earlier.

Recurrent interactions - especially inhibition- may have a key role in broadening the

distribution of preferred phases in layer 4. This is because inhibition can de-correlate neuronal

activity. I suggest to the authors to elaborate a bit more about this alternative hypothesis.

2) In fact the paper determines the distribution of the preferred phase of the aggregate

thalamic input into cortical neurons. Optogenetic techniques combined with intracellular

recordings (i.e Lien and Scanziani, 2013, in the context of V1) makes it possible to test

the hypothesis of the paper. I think that commenting about that can be inspiring to the reader.

**********

Have the authors made all data and (if applicable) computational code underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data and code underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data and code should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data or code —e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #1: None

Reviewer #2: Yes

Reviewer #3: Yes

**********

PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: No

Reviewer #2: No

Reviewer #3: No

Figure Files:

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email us at figures@plos.org.

Data Requirements:

Please note that, as a condition of publication, PLOS' data policy requires that you make available all data used to draw the conclusions outlined in your manuscript. Data must be deposited in an appropriate repository, included within the body of the manuscript, or uploaded as supporting information. This includes all numerical values that were used to generate graphs, histograms etc.. For an example in PLOS Biology see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5.

Reproducibility:

To enhance the reproducibility of your results, we recommend that you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option to publish peer-reviewed clinical study protocols. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

References:

Review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the rebuttal letter that accompanies your revised manuscript.

If you need to cite a retracted article, indicate the article’s retracted status in the References list and also include a citation and full reference for the retraction notice.

PLoS Comput Biol. doi: 10.1371/journal.pcbi.1009353.r003

Decision Letter 1

Lyle J Graham, Abigail Morrison

17 Aug 2021

Dear Mr. Sherf,

We are pleased to inform you that your manuscript 'STDP and the distribution of preferred phases in the whisker system' has been provisionally accepted for publication in PLOS Computational Biology.

Before your manuscript can be formally accepted you will need to complete some formatting changes, which you will receive in a follow up email. A member of our team will be in touch with a set of requests.

Please note that your manuscript will not be scheduled for publication until you have made the required changes, so a swift response is appreciated.

IMPORTANT: The editorial review process is now complete. PLOS will only permit corrections to spelling, formatting or significant scientific errors from this point onwards. Requests for major changes, or any which affect the scientific understanding of your work, will cause delays to the publication date of your manuscript.

Should you, your institution's press office or the journal office choose to press release your paper, you will automatically be opted out of early publication. We ask that you notify us now if you or your institution is planning to press release the article. All press must be co-ordinated with PLOS.

Thank you again for supporting Open Access publishing; we are looking forward to publishing your work in PLOS Computational Biology. 

Best regards,

Abigail Morrison

Associate Editor

PLOS Computational Biology

Lyle Graham

Deputy Editor

PLOS Computational Biology

***********************************************************

Reviewer's Responses to Questions

Comments to the Authors:

Please note here if the review is uploaded as an attachment.

Reviewer #2: The authors have successfully addressed all my concerns.

I have a single additional minor suggestion.

The motivation to this paper is the observation that synapses are volatile and hence are expected to converge, in the absence of activity-dependent synaptic plasticity, to random connections. By contrast, in Figure 2c, as well as throughout the text, the authors compare the experimental data to two baseline models, one in which pooling is uniform and the other in which pooling random. It is not clear to why the uniform model is biologically-relevant, as synaptic volatility is expected to result in random connectivity, not uniform.

I think that the clarity of the paper would be improved if the authors remove the uniform-model from the figure and text. However, I leave this decision to the authors.

Reviewer #3: The authors have taken into account all my comments.

**********

Have the authors made all data and (if applicable) computational code underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data and code underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data and code should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data or code —e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #2: Yes

Reviewer #3: None

**********

PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #2: No

Reviewer #3: No

PLoS Comput Biol. doi: 10.1371/journal.pcbi.1009353.r004

Acceptance letter

Lyle J Graham, Abigail Morrison

3 Sep 2021

PCOMPBIOL-D-21-00842R1

STDP and the distribution of preferred phases in the whisker system

Dear Dr Sherf,

I am pleased to inform you that your manuscript has been formally accepted for publication in PLOS Computational Biology. Your manuscript is now with our production department and you will be notified of the publication date in due course.

The corresponding author will soon be receiving a typeset proof for review, to ensure errors have not been introduced during production. Please review the PDF proof of your manuscript carefully, as this is the last chance to correct any errors. Please note that major changes, or those which affect the scientific understanding of the work, will likely cause delays to the publication date of your manuscript.

Soon after your final files are uploaded, unless you have opted out, the early version of your manuscript will be published online. The date of the early version will be your article's publication date. The final article will be published to the same URL, and all versions of the paper will be accessible to readers.

Thank you again for supporting PLOS Computational Biology and open-access publishing. We are looking forward to publishing your work!

With kind regards,

Katalin Szabo

PLOS Computational Biology | Carlyle House, Carlyle Road, Cambridge CB4 3DN | United Kingdom ploscompbiol@plos.org | Phone +44 (0) 1223-442824 | ploscompbiol.org | @PLOSCompBiol

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    Attachment

    Submitted filename: CoverLetter20210725.doc

    Data Availability Statement

    All relevant data are within the manuscript and its Supporting Information files.


    Articles from PLoS Computational Biology are provided here courtesy of PLOS

    RESOURCES