nips nips2012 nips2012-333 knowledge-graph by maker-knowledge-mining

333 nips-2012-Synchronization can Control Regularization in Neural Systems via Correlated Noise Processes


Source: pdf

Author: Jake Bouvrie, Jean-jeacques Slotine

Abstract: To learn reliable rules that can generalize to novel situations, the brain must be capable of imposing some form of regularization. Here we suggest, through theoretical and computational arguments, that the combination of noise with synchronization provides a plausible mechanism for regularization in the nervous system. The functional role of regularization is considered in a general context in which coupled computational systems receive inputs corrupted by correlated noise. Noise on the inputs is shown to impose regularization, and when synchronization upstream induces time-varying correlations across noise variables, the degree of regularization can be calibrated over time. The resulting qualitative behavior matches experimental data from visual cortex. 1

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 Here we suggest, through theoretical and computational arguments, that the combination of noise with synchronization provides a plausible mechanism for regularization in the nervous system. [sent-5, score-0.972]

2 The functional role of regularization is considered in a general context in which coupled computational systems receive inputs corrupted by correlated noise. [sent-6, score-0.46]

3 Noise on the inputs is shown to impose regularization, and when synchronization upstream induces time-varying correlations across noise variables, the degree of regularization can be calibrated over time. [sent-7, score-0.95]

4 How the nervous system imposes regularization is not entirely clear, however. [sent-12, score-0.356]

5 Intuitively, if noise is allowed to “smear” observations presented to a learning apparatus, overfitting may be mitigated – a well known phenomenon in artificial neural networks [1]. [sent-17, score-0.301]

6 In this paper we argue that noise provides an appealing, plausible mechanism for regularization in the nervous system. [sent-18, score-0.647]

7 We consider a general context in which coupled computational circuits subject to independent noise receive common inputs corrupted by spatially correlated noise. [sent-19, score-0.632]

8 Independent noise impacting higher-level computational elements may arise from more intrinsic, ambient neuronal noise sources, and may be roughly independent due to broader spatial distribution [6]. [sent-23, score-0.736]

9 To help understand the functional role of noise in inducing regularization, we propose a high-level model that can explain quantitatively how noise translates into regularization, and how regularization may be calibrated over time. [sent-24, score-0.802]

10 The ability to adjust regularization is key: as an organism accumulates 1 experience, its models of the world should be able to adjust to the complexity of the relationships and phenomena it encounters, as well as reconcile new information with prior probabilities. [sent-25, score-0.345]

11 Our point of view is complementary to Bayesian theories of learning; the representation and integration of sensory uncertainty is closely related to a regularization interpretation of learning in ill-posed settings. [sent-26, score-0.228]

12 We postulate that regularization may be plausibly controlled by one of the most ubiquitous mechanisms in the brain: synchronization. [sent-27, score-0.26]

13 A simple, one-dimensional regression (association) problem in the presence of both independent ambient noise and correlated measurement noise suffices to illustrate the core ideas. [sent-28, score-0.755]

14 When a learner is presented with a collection of noisy observations, we show that synchronization may be used to adjust the dependence between observational noise variables, and that this in turn leads to a quantifiable change in the degree of regularization imposed upon the learning task. [sent-29, score-0.985]

15 Regularization is further shown to both improve the convergence rate towards the solution to the regression problem, and reduce the negative impact of ambient noise. [sent-30, score-0.227]

16 The model’s qualitative behavior coincides with experimental data from visual tracking tasks [10] (area MT) and from anesthetized animals [24] (area V1), in which correlated noise impacts sensory measurements and correlations increase over short time scales. [sent-31, score-0.541]

17 Other experiments involving perceptual learning tasks have shown that noise correlations decrease with long-term training [8]. [sent-32, score-0.422]

18 The mechanism we propose suggests that changes in noise correlations arising from feedback synchronization can calibrate regularization, possibly leading to improved convergence properties or better solutions. [sent-33, score-0.859]

19 Collectively, the experimental evidence lends credence to the hypothesis that, at a high level, the brain may be optimizing its learning processes by adapting dependence among noise variables, with regularization an underlying computational theme. [sent-34, score-0.574]

20 In addition to supporting regularization, noise can be harnessed to facilitate distributed computation of the gradients needed to implement a dynamic optimization process. [sent-36, score-0.346]

21 This distributed approach to the implementation of dynamic learning processes further highlights a connection between parallel stochastic gradient descent algorithms [25, 15, 28], and neural computation. [sent-39, score-0.29]

22 2 Learning as noisy gradient descent on a network The learning process we will consider is that of a one-dimensional linear fitting problem described by a dynamic gradient based minimization of a square loss objective, in the spirit of Rao & Ballard [21]. [sent-40, score-0.309]

23 The gradient of E m with respect to the slope parameter is given by w E = − i=1 (yi − wxi )xi , and generates the continuous-time, noise-free gradient dynamics w = − w E(w). [sent-49, score-0.26]

24 ˙ (2) The learning dynamics we will consider, however, are assumed to be corrupted by two distinct kinds of noise: 2 (N1) Sensory observations (xi )i are corrupted by time-varying, correlated noise processes. [sent-50, score-0.557]

25 The noise (N2) may be significant (we do not take small noise limits) and can be attributed to some or all of: error in computing and sensing a gradient, intrinsic neuronal noise [6] (aggregated or localized), or interference between large assemblies of neurons or circuits. [sent-54, score-1.031]

26 Synchronization among circuits and/or populations will be modeled by considering multiple coupled dynamical systems, each receiving the same noisy observations. [sent-55, score-0.254]

27 The collective enhancement of precision hypothesis suggests that the nervous system copes with noise by averaging over collections of signals in order to reduce variation in behavior and improve computational accuracy [23, 13, 26, 3]. [sent-57, score-0.483]

28 If the coupling is strong enough, then the variance of the consensus trajectory decreases as O(1/n) after transients, if there are n signals or circuits [23, 17, 19, 3]. [sent-59, score-0.386]

29 We will consider regularization in the context of networks of coupled SDEs, and investigate the impact of coupling, redundancy (n) and regularization upon the convergence behavior of the system. [sent-60, score-0.591]

30 Formally, the noise-free flow (2) can be modified to include noise sources (N1) and (N2) as follows. [sent-62, score-0.301]

31 Noise (N1) may be modeled as a white-noise limit of Ornstein-Uhlenbeck (OU) processes (Zt )i , and (N2) as an additive diffusive noise term. [sent-63, score-0.419]

32 In differential form, we have dwt = − wt x + Zt 2 − x + Zt , y dt + σdBt √ i 2γ Zt i i i = 1, . [sent-64, score-0.347]

33 dZt = − dt + √ dBt , ε ε (3a) (3b) Here, Bt denotes the standard 1-dimensional Brownian motion and captures noise source (N2). [sent-68, score-0.431]

34 The i observations (x)i = xi are corrupted by the noise processes (Zt )i = Zt , following (N1). [sent-69, score-0.428]

35 The parameter 0 < ε i controls the correlation time of a given noise process. [sent-71, score-0.301]

36 Characterizing the noise Zt as (3b) with ε → 0 serves as both a modeling approximation/idealization and an analytical tool. [sent-73, score-0.301]

37 1 Homogenization The system (3a)-(3b) above is a classic “fast-slow” system: the gradient descent trajectory wt evolves on a timescale much longer than the O(ε) stochastic perturbations Zt . [sent-75, score-0.449]

38 Averaging over the fast variable Zt appearing in (3a) with respect to this distribution gives dwt = − wt ( x 2 + mγ 2 ) − x, y dt + σdBt , (5) and by Theorem 2. [sent-92, score-0.347]

39 2 Network structure Now consider n ≥ 1 diffusively coupled neural systems implementing the dynamics (5), with associated parameters w(t) = w1 (t), . [sent-95, score-0.246]

40 If Wij ≥ 0 is the coupling strength between systems i and j, L = diag(W 1) − W is the network Laplacian [16]. [sent-99, score-0.351]

41 Letting α := x 2 + mγ 2 , β := x, y and µ := (β/α)1, the coupled system can be written concisely as dwt = −(L + αI)wt dt + β1dt + σdBt = (L + αI)(µ − wt )dt + σdBt , (6) with Bt an n-dimensional Brownian motion. [sent-101, score-0.503]

42 In such a general setting, diffusive coupling is a natural and mathematically tractable choice that can capture the key, aggregate aspects of communication among neural systems. [sent-103, score-0.279]

43 Note that one can equivalently consider n systems (3a) and then homogenize assuming n copies of the (i) same noise process Zt , or n independent noise processes {Zt }i ; either choice also leads to (6). [sent-104, score-0.661]

44 3 Learning with noisy data imposes regularization Equation (6) is seen by inspection to be an OU process, and has solution (see e. [sent-105, score-0.286]

45 , n}, = e−(L+αI)t E[w(0)w(0) ]e−(L+αI)t + t→∞ E[wi (t)] − − (µ)i = −→ x, y x + mγ 2 2 (9) which is exactly the solution to the regularized regression problem min y − wx w∈R 2 + λw2 with regularization parameter λ := mγ 2 . [sent-113, score-0.205]

46 To summarize, we have considered a network of coupled, noisy gradient flows implementing unregularized linear regression. [sent-114, score-0.219]

47 How does coupling and redundancy (number of circuits n) impact convergence? [sent-119, score-0.387]

48 The first term of (11) estimates the transient part of the fluctuations term in (10), and we find that the rate of convergence to the synchronization subspace is 2(λ + α). [sent-130, score-0.414]

49 The second term term estimates the transient part of the centroid’s trajectory, and we see that the rate of convergence of the mean trajectory to equilibrium is 2α. [sent-131, score-0.22]

50 This term quantifies the steady-state interaction between: gradient noise (σ); regularization (α, via the observation noise γ); network topology (via λ), coupling strength (via λ), and redundancy (n; possibly λ). [sent-134, score-1.327]

51 Regularization improves both the synchronization rate and the rate of convergence to equilibrium. [sent-137, score-0.384]

52 Regularization contributes towards reducing the effect of the gradient noise σ: (N1) counteracts (N2). [sent-139, score-0.384]

53 Regularization changes the solution, so we cannot view regularization as a “free-parameter” that can be used solely to improve convergence or reduce noise. [sent-141, score-0.227]

54 Faster convergence rates and noise reduction should be viewed as beneficial side-effects, while the appropriate degree of regularization primarily depends on the learning problem at hand. [sent-142, score-0.528]

55 The number of circuits n and the coupling strength contribute towards reducing the effect of the gradient noise (N2) (that is, the variance of the error) and improve the synchronization rate, but do not affect the rate of convergence toward equilibrium. [sent-144, score-1.168]

56 Coupling strength and redundancy cannot be used to control the degree of regularization, since the equilibrium solution µ does not depend on n or the spectrum of L. [sent-146, score-0.24]

57 This is true no matter how the coupling weights Wij are chosen, since constants will always be in the null space of L and µ is a constant vector. [sent-147, score-0.22]

58 i In the next section we will show that if the noise processes {Zt }i are themselves trajectories of a coupled network, then synchronization can be a mechanism for controlling the regularization imposed on a learning process. [sent-148, score-1.005]

59 A collection of dependent observational noise processes is perhaps most conveniently modeled by coupling the OU dynamics (3b) introduced 5 before through another (symmetric) network Laplacian Lz : √ 1 2γ dZt = − (Lz + ηI)Zt dt + √ dBt , (12) ε ε for some η > 0. [sent-150, score-0.923]

60 We now have two networks: the first network of gradient systems is the same as before, but the observational noise process Zt is now generated by another network. [sent-151, score-0.537]

61 For purposes of analysis, this model suffices to capture generalized correlated noise sources. [sent-152, score-0.361]

62 In the actual biology, however, correlations may arise in a number of possible ways, which may or may not include diffusively coupled dynamic noise processes. [sent-153, score-0.534]

63 To analyze what happens when a network of learning systems (3a) is driven by observation noise of the form (12), we take an approach similar to that of the previous Section. [sent-154, score-0.392]

64 We can now dt + σdBt + γ 2 tr(Lz + ηI)−1 − x, y dt + σdBt where we have used that E[ Zt 2 ] = γ 2 tr(Lz + ηI)−1 . [sent-158, score-0.26]

65 An expression identical to (6), dwt = (L + αI)(µ − wt )dt + σdBt (14) is obtained by redefining α := x 2 + γ 2 tr(Lz + ηI)−1 and µ := ( x, y /α)1. [sent-160, score-0.217]

66 However now the covariance of Zt is a function of the network Laplacian Lz = Lz (t), which is defined by the topology and potentially time-varying coupling strengths of the noise network. [sent-165, score-0.607]

67 By adjusting the coupling in (12), we adjust the regularizai tion λ imposed upon (14). [sent-166, score-0.275]

68 When coupling increases, the dependence among the Zt increases and −1 tr(Lz + ηI) (and therefore α) decreases. [sent-167, score-0.22]

69 Thus, increased correlation among observational noise variables implies decreased regularization. [sent-168, score-0.398]

70 In the case of all-to-all coupling with uniform strength κ ≥ 0, for example, Lz has eigenvalues 0 = λ0 < λ1 = · · · = λm = mκ. [sent-169, score-0.295]

71 The regularization may in this case range over the interval λ m 1 = sup tr(Lz + ηI)−1 inf tr(Lz + ηI)−1 = < 2 ≤ κ η γ η κ by adjusting the coupling strength κ ∈ [0, ∞). [sent-170, score-0.463]

72 Note that all-to-all coupling may be plausibly implemented with O(n) connections using mechanisms such as quorum sensing (see [3, §2. [sent-171, score-0.388]

73 5 Distributed computation with noise We have argued that noise can serve as a mechanism for regularization. [sent-173, score-0.654]

74 The development is closely related to stochastic gradient descent (SGD) ideas appearing in stochastic approximation [25, 15] and adaptive optics [28]. [sent-176, score-0.261]

75 1 Parallel stochastic gradient descent Let J(u) : Rd → R be a locally Lipschitz Lyapunov cost functional we wish to minimize with respect to some set of control signals u(t) ∈ Rd . [sent-178, score-0.211]

76 2 Stochastic gradient model The parallel finite difference approximation (15) suggests a more biologically plausible mechanism for implementing gradient dynamics. [sent-222, score-0.347]

77 random variables, we can model parallel stochastic gradient descent as an Ito process: dut = −γ J(ut + Zt ) − J(ut ) Zt dt, σ 1 dZt = − Zt dt + √ dBt , ε ε u(0) = u0 (16a) Z(0) = z0 (16b) where Bt is a standard d-dimensional Brownian motion. [sent-226, score-0.406]

78 Additive noise affecting the gradient has been omitted from (16a) for simplicity, and does not change the fundamental results discussed in this section. [sent-227, score-0.384]

79 The perturbation noise Zt has again been modeled as a white-noise limit of OrnsteinUhlenbeck processes (16b). [sent-228, score-0.404]

80 (17) Convergence of continuous-time PSGD with quadratic cost We turn to studying the convergence behavior of (17) and the precise role of the stochastic perturbations Zt used to estimate the gradients. [sent-238, score-0.181]

81 However, one may also expect that the noise will play an important role in determining convergence properties since it is the noise that ultimately kicks the system “downhill” towards equilibrium. [sent-240, score-0.717]

82 Our earlier intuition that the perturbation noise σ should play a role in the rate of convergence is also confirmed: greater noise amplitudes lead to faster convergence. [sent-246, score-0.705]

83 5 Figure 1: (Left stack) Increased observation noise imposes greater regularization, and leads to a reduction in ambient noise. [sent-283, score-0.471]

84 (Right stack) Stronger coupling/correlation between observation noise processes decreases regularization. [sent-284, score-0.395]

85 6 Simulations We first simulated a network of gradient dynamics with uncoupled observation noise processes obeying (3). [sent-286, score-0.645]

86 To illustrate the effect of increasing observation noise variance, the parameter γ in (3b) was increased from 0. [sent-287, score-0.336]

87 The top plot shows the sample paths w(t) and time course of the observational noise deviation γ(t) (grey labeled trace). [sent-295, score-0.428]

88 ¯ Because the observation noise is increasing, the regularization λ = mγ 2 increases and the solution µ(t) to the regularized problem decreases in magnitude following (9). [sent-299, score-0.541]

89 A second experiment, described by the right-hand stack of plots in Figure 1, shows how synchronization can function to adjust regularization over time. [sent-302, score-0.582]

90 This simulation is inspired by the experimental study of noise correlations in cortical area MT due to [10], where it was suggested that time-varying correlations between pairs of neurons play a significant role in explaining behavioral variation in smooth-pursuit eye movements. [sent-303, score-0.533]

91 In particular, the findings in [10] and [4] suggest that short-term increases in noise correlations are likely to occur after feedback arrives and neurons within and upstream from MT synchronize. [sent-304, score-0.515]

92 We simulated a collection of correlated observation noise processes obeying (12) (ε = 10−3 , η = 3) with all-to-all topology and uniform coupling strength κz (t) increasing from 0 to 2 along the profile shown in Figure 1 (top-right plot, labeled gray trace). [sent-305, score-0.831]

93 This noise process Zt was then fed to a population of n = 5 units obeying (3a), with ambient noise σ = 1 and all-to-all coupling at fixed strength Wij = κ = 2. [sent-306, score-1.041]

94 The middle plot on the right-hand side shows the effect of increasing synchronization among the observation noise processes. [sent-308, score-0.691]

95 As the coupling increases, the noise becomes more correlated and regularization decreases. [sent-309, score-0.749]

96 With decreased regularization, the ambient noise is more pronounced. [sent-311, score-0.394]

97 An increase in the noise variance is apparent following the increase in observational noise correlation. [sent-313, score-0.733]

98 8 Acknowledgments The authors are grateful to Rodolfo Llinas for pointing out the plausible analogy between gradient search in adaptive optics and learning mechanisms in the brain. [sent-316, score-0.2]

99 Model for synchronization of pancreatic beta-cells by gap junction coupling. [sent-491, score-0.37]

100 Dynamical quorum sensing and synchronization in large populations of chemical oscillators. [sent-519, score-0.445]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('lz', 0.355), ('zt', 0.332), ('synchronization', 0.325), ('noise', 0.301), ('coupling', 0.22), ('dbt', 0.217), ('regularization', 0.168), ('ui', 0.141), ('dwt', 0.138), ('dt', 0.13), ('coupled', 0.1), ('observational', 0.097), ('ambient', 0.093), ('nervous', 0.09), ('psgd', 0.089), ('correlations', 0.088), ('ud', 0.087), ('gradient', 0.083), ('wt', 0.079), ('brownian', 0.078), ('strength', 0.075), ('perturbations', 0.074), ('circuits', 0.071), ('equilibrium', 0.07), ('corrupted', 0.068), ('dzt', 0.067), ('homogenization', 0.067), ('organism', 0.067), ('uctuations', 0.062), ('trajectory', 0.061), ('correlated', 0.06), ('sensory', 0.06), ('ou', 0.06), ('dynamics', 0.06), ('processes', 0.059), ('transients', 0.059), ('diffusive', 0.059), ('sde', 0.059), ('convergence', 0.059), ('redundancy', 0.058), ('neurons', 0.056), ('system', 0.056), ('network', 0.056), ('adjust', 0.055), ('parallel', 0.052), ('mechanism', 0.052), ('obeying', 0.051), ('descent', 0.048), ('stochastic', 0.048), ('mechanisms', 0.047), ('tr', 0.047), ('fw', 0.047), ('brain', 0.046), ('azt', 0.045), ('bouvrie', 0.045), ('diffusively', 0.045), ('dut', 0.045), ('harnessed', 0.045), ('pancreatic', 0.045), ('plausibly', 0.045), ('quorum', 0.045), ('sdes', 0.045), ('sherman', 0.045), ('tabareau', 0.045), ('perturbation', 0.044), ('populations', 0.044), ('au', 0.043), ('mt', 0.043), ('imposes', 0.042), ('implementing', 0.041), ('neuronal', 0.041), ('bt', 0.04), ('ergodic', 0.04), ('slotine', 0.039), ('fluctuations', 0.039), ('noisy', 0.039), ('impact', 0.038), ('solution', 0.037), ('plausible', 0.036), ('biophysical', 0.036), ('upstream', 0.036), ('averaging', 0.036), ('observation', 0.035), ('stack', 0.034), ('wij', 0.034), ('optics', 0.034), ('kiani', 0.034), ('wxi', 0.034), ('variance', 0.034), ('feedback', 0.034), ('perceptual', 0.033), ('functional', 0.032), ('visual', 0.032), ('inputs', 0.032), ('lyapunov', 0.031), ('sensing', 0.031), ('topology', 0.03), ('decision', 0.03), ('plot', 0.03), ('transient', 0.03)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.99999905 333 nips-2012-Synchronization can Control Regularization in Neural Systems via Correlated Noise Processes

Author: Jake Bouvrie, Jean-jeacques Slotine

Abstract: To learn reliable rules that can generalize to novel situations, the brain must be capable of imposing some form of regularization. Here we suggest, through theoretical and computational arguments, that the combination of noise with synchronization provides a plausible mechanism for regularization in the nervous system. The functional role of regularization is considered in a general context in which coupled computational systems receive inputs corrupted by correlated noise. Noise on the inputs is shown to impose regularization, and when synchronization upstream induces time-varying correlations across noise variables, the degree of regularization can be calibrated over time. The resulting qualitative behavior matches experimental data from visual cortex. 1

2 0.11594381 336 nips-2012-The Coloured Noise Expansion and Parameter Estimation of Diffusion Processes

Author: Simon Lyons, Amos J. Storkey, Simo Särkkä

Abstract: Stochastic differential equations (SDE) are a natural tool for modelling systems that are inherently noisy or contain uncertainties that can be modelled as stochastic processes. Crucial to the process of using SDE to build mathematical models is the ability to estimate parameters of those models from observed data. Over the past few decades, significant progress has been made on this problem, but we are still far from having a definitive solution. We describe a novel method of approximating a diffusion process that we show to be useful in Markov chain Monte-Carlo (MCMC) inference algorithms. We take the ‘white’ noise that drives a diffusion process and decompose it into two terms. The first is a ‘coloured noise’ term that can be deterministically controlled by a set of auxilliary variables. The second term is small and enables us to form a linear Gaussian ‘small noise’ approximation. The decomposition allows us to take a diffusion process of interest and cast it in a form that is amenable to sampling by MCMC methods. We explain why many state-of-the-art inference methods fail on highly nonlinear inference problems, and we demonstrate experimentally that our method performs well in such situations. Our results show that this method is a promising new tool for use in inference and parameter estimation problems. 1

3 0.097048774 13 nips-2012-A Nonparametric Conjugate Prior Distribution for the Maximizing Argument of a Noisy Function

Author: Pedro Ortega, Jordi Grau-moya, Tim Genewein, David Balduzzi, Daniel Braun

Abstract: We propose a novel Bayesian approach to solve stochastic optimization problems that involve finding extrema of noisy, nonlinear functions. Previous work has focused on representing possible functions explicitly, which leads to a two-step procedure of first, doing inference over the function space and second, finding the extrema of these functions. Here we skip the representation step and directly model the distribution over extrema. To this end, we devise a non-parametric conjugate prior based on a kernel regressor. The resulting posterior distribution directly captures the uncertainty over the maximum of the unknown function. Given t observations of the function, the posterior can be evaluated efficiently in time O(t2 ) up to a multiplicative constant. Finally, we show how to apply our model to optimize a noisy, non-convex, high-dimensional objective function.

4 0.093466543 195 nips-2012-Learning visual motion in recurrent neural networks

Author: Marius Pachitariu, Maneesh Sahani

Abstract: We present a dynamic nonlinear generative model for visual motion based on a latent representation of binary-gated Gaussian variables. Trained on sequences of images, the model learns to represent different movement directions in different variables. We use an online approximate inference scheme that can be mapped to the dynamics of networks of neurons. Probed with drifting grating stimuli and moving bars of light, neurons in the model show patterns of responses analogous to those of direction-selective simple cells in primary visual cortex. Most model neurons also show speed tuning and respond equally well to a range of motion directions and speeds aligned to the constraint line of their respective preferred speed. We show how these computations are enabled by a specific pattern of recurrent connections learned by the model. 1

5 0.093273506 114 nips-2012-Efficient coding provides a direct link between prior and likelihood in perceptual Bayesian inference

Author: Xue-xin Wei, Alan Stocker

Abstract: A common challenge for Bayesian models of perception is the fact that the two fundamental Bayesian components, the prior distribution and the likelihood function, are formally unconstrained. Here we argue that a neural system that emulates Bayesian inference is naturally constrained by the way it represents sensory information in populations of neurons. More specifically, we show that an efficient coding principle creates a direct link between prior and likelihood based on the underlying stimulus distribution. The resulting Bayesian estimates can show biases away from the peaks of the prior distribution, a behavior seemingly at odds with the traditional view of Bayesian estimation, yet one that has been reported in human perception. We demonstrate that our framework correctly accounts for the repulsive biases previously reported for the perception of visual orientation, and show that the predicted tuning characteristics of the model neurons match the reported orientation tuning properties of neurons in primary visual cortex. Our results suggest that efficient coding is a promising hypothesis in constraining Bayesian models of perceptual inference. 1 Motivation Human perception is not perfect. Biases have been observed in a large number of perceptual tasks and modalities, of which the most salient ones constitute many well-known perceptual illusions. It has been suggested, however, that these biases do not reflect a failure of perception but rather an observer’s attempt to optimally combine the inherently noisy and ambiguous sensory information with appropriate prior knowledge about the world [13, 4, 14]. This hypothesis, which we will refer to as the Bayesian hypothesis, has indeed proven quite successful in providing a normative explanation of perception at a qualitative and, more recently, quantitative level (see e.g. [15]). A major challenge in forming models based on the Bayesian hypothesis is the correct selection of two main components: the prior distribution (belief) and the likelihood function. This has encouraged some to criticize the Bayesian hypothesis altogether, claiming that arbitrary choices for these components always allow for unjustified post-hoc explanations of the data [1]. We do not share this criticism, referring to a number of successful attempts to constrain prior beliefs and likelihood functions based on principled grounds. For example, prior beliefs have been defined as the relative distribution of the sensory variable in the environment in cases where these statistics are relatively easy to measure (e.g. local visual orientations [16]), or where it can be assumed that subjects have learned them over the course of the experiment (e.g. time perception [17]). Other studies have constrained the likelihood function according to known noise characteristics of neurons that are crucially involved in the specific perceptual process (e.g motion tuned neurons in visual cor∗ http://www.sas.upenn.edu/ astocker/lab 1 world neural representation efficient encoding percept Bayesian decoding Figure 1: Encoding-decoding framework. A stimulus representing a sensory variable θ elicits a firing rate response R = {r1 , r2 , ..., rN } in a population of N neurons. The perceptual task is to generate a ˆ good estimate θ(R) of the presented value of the sensory variable based on this population response. Our framework assumes that encoding is efficient, and decoding is Bayesian based on the likelihood p(R|θ), the prior p(θ), and a squared-error loss function. tex [18]). However, we agree that finding appropriate constraints is generally difficult and that prior beliefs and likelihood functions have been often selected on the basis of mathematical convenience. Here, we propose that the efficient coding hypothesis [19] offers a joint constraint on the prior and likelihood function in neural implementations of Bayesian inference. Efficient coding provides a normative description of how neurons encode sensory information, and suggests a direct link between measured perceptual discriminability, neural tuning characteristics, and environmental statistics [11]. We show how this link can be extended to a full Bayesian account of perception that includes perceptual biases. We validate our model framework against behavioral as well as neural data characterizing the perception of visual orientation. We demonstrate that we can account not only for the reported perceptual biases away from the cardinal orientations, but also for the specific response characteristics of orientation-tuned neurons in primary visual cortex. Our work is a novel proposal of how two important normative hypotheses in perception science, namely efficient (en)coding and Bayesian decoding, might be linked. 2 Encoding-decoding framework We consider perception as an inference process that takes place along the simplified neural encodingdecoding cascade illustrated in Fig. 11 . 2.1 Efficient encoding Efficient encoding proposes that the tuning characteristics of a neural population are adapted to the prior distribution p(θ) of the sensory variable such that the population optimally represents the sensory variable [19]. Different definitions of “optimally” are possible, and may lead to different results. Here, we assume an efficient representation that maximizes the mutual information between the sensory variable and the population response. With this definition and an upper limit on the total firing activity, the square-root of the Fisher Information must be proportional to the prior distribution [12, 21]. In order to constrain the tuning curves of individual neurons in the population we also impose a homogeneity constraint, requiring that there exists a one-to-one mapping F (θ) that transforms the ˜ physical space with units θ to a homogeneous space with units θ = F (θ) in which the stimulus distribution becomes uniform. This defines the mapping as θ F (θ) = p(χ)dχ , (1) −∞ which is the cumulative of the prior distribution p(θ). We then assume a neural population with identical tuning curves that evenly tiles the stimulus range in this homogeneous space. The population provides an efficient representation of the sensory variable θ according to the above constraints [11]. ˜ The tuning curves in the physical space are obtained by applying the inverse mapping F −1 (θ). Fig. 2 1 In the context of this paper, we consider ‘inferring’, ‘decoding’, and ‘estimating’ as synonymous. 2 stimulus distribution d samples # a Fisher information discriminability and average firing rates and b firing rate [ Hz] efficient encoding F likelihood function F -1 likelihood c symmetric asymmetric homogeneous space physical space Figure 2: Efficient encoding constrains the likelihood function. a) Prior distribution p(θ) derived from stimulus statistics. b) Efficient coding defines the shape of the tuning curves in the physical space by transforming a set of homogeneous neurons using a mapping F −1 that is the inverse of the cumulative of the prior p(θ) (see Eq. (1)). c) As a result, the likelihood shape is constrained by the prior distribution showing heavier tails on the side of lower prior density. d) Fisher information, discrimination threshold, and average firing rates are all uniform in the homogeneous space. illustrates the applied efficient encoding scheme, the mapping, and the concept of the homogeneous space for the example of a symmetric, exponentially decaying prior distribution p(θ). The key idea here is that by assuming efficient encoding, the prior (i.e. the stimulus distribution in the world) directly constrains the likelihood function. In particular, the shape of the likelihood is determined by the cumulative distribution of the prior. As a result, the likelihood is generally asymmetric, as shown in Fig. 2, exhibiting heavier tails on the side of the prior with lower density. 2.2 Bayesian decoding Let us consider a population of N sensory neurons that efficiently represents a stimulus variable θ as described above. A stimulus θ0 elicits a specific population response that is characterized by the vector R = [r1 , r2 , ..., rN ] where ri is the spike-count of the ith neuron over a given time-window τ . Under the assumption that the variability in the individual firing rates is governed by a Poisson process, we can write the likelihood function over θ as N p(R|θ) = (τ fi (θ))ri −τ fi (θ) e , ri ! i=1 (2) ˆ with fi (θ) describing the tuning curve of neuron i. We then define a Bayesian decoder θLSE as the estimator that minimizes the expected squared-error between the estimate and the true stimulus value, thus θp(R|θ)p(θ)dθ ˆ θLSE (R) = , (3) p(R|θ)p(θ)dθ where we use Bayes’ rule to appropriately combine the sensory evidence with the stimulus prior p(θ). 3 Bayesian estimates can be biased away from prior peaks Bayesian models of perception typically predict perceptual biases toward the peaks of the prior density, a characteristic often considered a hallmark of Bayesian inference. This originates from the 3 a b prior attraction prior prior attraction likelihood repulsion! likelihood c prior prior repulsive bias likelihood likelihood mean posterior mean posterior mean Figure 3: Bayesian estimates biased away from the prior. a) If the likelihood function is symmetric, then the estimate (posterior mean) is, on average, shifted away from the actual value of the sensory variable θ0 towards the prior peak. b) Efficient encoding typically leads to an asymmetric likelihood function whose normalized mean is away from the peak of the prior (relative to θ0 ). The estimate is determined by a combination of prior attraction and shifted likelihood mean, and can exhibit an overall repulsive bias. c) If p(θ0 ) < 0 and the likelihood is relatively narrow, then (1/p(θ)2 ) > 0 (blue line) and the estimate is biased away from the prior peak (see Eq. (6)). common approach of choosing a parametric description of the likelihood function that is computationally convenient (e.g. Gaussian). As a consequence, likelihood functions are typically assumed to be symmetric (but see [23, 24]), leaving the bias of the Bayesian estimator to be mainly determined by the shape of the prior density, i.e. leading to biases toward the peak of the prior (Fig. 3a). In our model framework, the shape of the likelihood function is constrained by the stimulus prior via efficient neural encoding, and is generally not symmetric for non-flat priors. It has a heavier tail on the side with lower prior density (Fig. 3b). The intuition is that due to the efficient allocation of neural resources, the side with smaller prior density will be encoded less accurately, leading to a broader likelihood function on that side. The likelihood asymmetry pulls the Bayes’ least-squares estimate away from the peak of the prior while at the same time the prior pulls it toward its peak. Thus, the resulting estimation bias is the combination of these two counter-acting forces - and both are determined by the prior! 3.1 General derivation of the estimation bias In the following, we will formally derive the mean estimation bias b(θ) of the proposed encodingdecoding framework. Specifically, we will study the conditions for which the bias is repulsive i.e. away from the peak of the prior density. ˆ We first re-write the estimator θLSE (3) by replacing θ with the inverse of its mapping to the homo−1 ˜ geneous space, i.e., θ = F (θ). The motivation for this is that the likelihood in the homogeneous space is symmetric (Fig. 2). Given a value θ0 and the elicited population response R, we can write the estimator as ˜ ˜ ˜ ˜ θp(R|θ)p(θ)dθ F −1 (θ)p(R|F −1 (θ))p(F −1 (θ))dF −1 (θ) ˆ θLSE (R) = = . ˜ ˜ ˜ p(R|θ)p(θ)dθ p(R|F −1 (θ))p(F −1 (θ))dF −1 (θ) Calculating the derivative of the inverse function and noting that F is the cumulative of the prior density, we get 1 1 1 ˜ ˜ ˜ ˜ ˜ ˜ dθ = dθ. dF −1 (θ) = (F −1 (θ)) dθ = dθ = −1 (θ)) ˜ F (θ) p(θ) p(F ˆ Hence, we can simplify θLSE (R) as ˆ θLSE (R) = ˜ ˜ ˜ F −1 (θ)p(R|F −1 (θ))dθ . ˜ ˜ p(R|F −1 (θ))dθ With ˜ K(R, θ) = ˜ p(R|F −1 (θ)) ˜ ˜ p(R|F −1 (θ))dθ 4 we can further simplify the notation and get ˆ θLSE (R) = ˜ ˜ ˜ F −1 (θ)K(R, θ)dθ . (4) ˆ ˜ In order to get the expected value of the estimate, θLSE (θ), we marginalize (4) over the population response space S, ˆ ˜ ˜ ˜ ˜ θLSE (θ) = p(R)F −1 (θ)K(R, θ)dθdR S = F −1 ˜ (θ)( ˜ ˜ p(R)K(R, θ)dR)dθ = ˜ ˜ ˜ F −1 (θ)L(θ)dθ, S where we define ˜ L(θ) = ˜ p(R)K(R, θ)dR. S ˜ ˜ ˜ It follows that L(θ)dθ = 1. Due to the symmetry in this space, it can be shown that L(θ) is ˜0 . Intuitively, L(θ) can be thought as the normalized ˜ symmetric around the true stimulus value θ average likelihood in the homogeneous space. We can then compute the expected bias at θ0 as b(θ0 ) = ˜ ˜ ˜ ˜ F −1 (θ)L(θ)dθ − F −1 (θ0 ) (5) ˜ This is expression is general where F −1 (θ) is defined as the inverse of the cumulative of an arbitrary ˜ prior density p(θ) (see Eq. (1)) and the dispersion of L(θ) is determined by the internal noise level. ˜ ˜ Assuming the prior density to be smooth, we expand F −1 in a neighborhood (θ0 − h, θ0 + h) that is larger than the support of the likelihood function. Using Taylor’s theorem with mean-value forms of the remainder, we get 1 ˜ ˜ ˜ ˜ ˜ ˜ ˜ ˜ F −1 (θ) = F −1 (θ0 ) + F −1 (θ0 ) (θ − θ0 ) + F −1 (θx ) (θ − θ0 )2 , 2 ˜ ˜ ˜ with θx lying between θ0 and θ. By applying this expression to (5), we find ˜ θ0 +h b(θ0 ) = = 1 2 ˜ θ0 −h 1 −1 ˜ ˜ ˜ ˜ ˜ 1 F (θx )θ (θ − θ0 )2 L(θ)dθ = ˜ 2 2 ˜ θ0 +h −( ˜ θ0 −h p(θx )θ ˜ ˜ 2 ˜ ˜ 1 )(θ − θ0 ) L(θ)dθ = p(θx )3 4 ˜ θ0 +h 1 ˜ − θ0 )2 L(θ)dθ ˜ ˜ ˜ ( ) ˜(θ ˜ p(F −1 (θx )) θ ( 1 ˜ ˜ ˜ ˜ ) (θ − θ0 )2 L(θ)dθ. p(θx )2 θ ˜ θ0 −h ˜ θ0 +h ˜ θ0 −h In general, there is no simple rule to judge the sign of b(θ0 ). However, if the prior is monotonic ˜ ˜ on the interval F −1 ((θ0 − h, θ0 + h)), then the sign of ( p(θ1 )2 ) is always the same as the sign of x 1 1 ( p(θ0 )2 ) . Also, if the likelihood is sufficiently narrow we can approximate ( p(θ1 )2 ) by ( p(θ0 )2 ) , x and therefore approximate the bias as b(θ0 ) ≈ C( 1 ) , p(θ0 )2 (6) where C is a positive constant. The result is quite surprising because it states that as long as the prior is monotonic over the support of the likelihood function, the expected estimation bias is always away from the peaks of the prior! 3.2 Internal (neural) versus external (stimulus) noise The above derivation of estimation bias is based on the assumption that all uncertainty about the sensory variable is caused by neural response variability. This level of internal noise depends on the response magnitude, and thus can be modulated e.g. by changing stimulus contrast. This contrastcontrolled noise modulation is commonly exploited in perceptual studies (e.g. [18]). Internal noise will always lead to repulsive biases in our framework if the prior is monotonic. If internal noise is low, the likelihood is narrow and thus the bias is small. Increasing internal noise leads to increasingly 5 larger biases up to the point where the likelihood becomes wide enough such that monotonicity of the prior over the support of the likelihood is potentially violated. Stimulus noise is another way to modulate the noise level in perception (e.g. random-dot motion stimuli). Such external noise, however, has a different effect on the shape of the likelihood function as compared to internal noise. It modifies the likelihood function (2) by convolving it with the noise kernel. External noise is frequently chosen as additive and symmetric (e.g. zero-mean Gaussian). It is straightforward to prove that such symmetric external noise does not lead to a change in the mean of the likelihood, and thus does not alter the repulsive effect induced by its asymmetry. However, by increasing the overall width of the likelihood, the attractive influence of the prior increases, resulting in an estimate that is closer to the prior peak than without external noise2 . 4 Perception of visual orientation We tested our framework by modelling the perception of visual orientation. Our choice was based on the fact that i) we have pretty good estimates of the prior distribution of local orientations in natural images, ii) tuning characteristics of orientation selective neurons in visual cortex are wellstudied (monkey/cat), and iii) biases in perceived stimulus orientation have been well characterized. We start by creating an efficient neural population based on measured prior distributions of local visual orientation, and then compare the resulting tuning characteristics of the population and the predicted perceptual biases with reported data in the literature. 4.1 Efficient neural model population for visual orientation Previous studies measured the statistics of the local orientation in large sets of natural images and consistently found that the orientation distribution is multimodal, peaking at the two cardinal orientations as shown in Fig. 4a [16, 20]. We assumed that the visual system’s prior belief over orientation p(θ) follows this distribution and approximate it formally as p(θ) ∝ 2 − | sin(θ)| (black line in Fig. 4b) . (7) Based on this prior distribution we defined an efficient neural representation for orientation. We assumed a population of model neurons (N = 30) with tuning curves that follow a von-Mises distribution in the homogeneous space on top of a constant spontaneous firing rate (5 Hz). We then ˜ applied the inverse transformation F −1 (θ) to all these tuning curves to get the corresponding tuning curves in the physical space (Fig. 4b - red curves), where F (θ) is the cumulative of the prior (7). The concentration parameter for the von-Mises tuning curves was set to κ ≈ 1.6 in the homogeneous space in order to match the measured average tuning width (∼ 32 deg) of neurons in area V1 of the macaque [9]. 4.2 Predicted tuning characteristics of neurons in primary visual cortex The orientation tuning characteristics of our model population well match neurophysiological data of neurons in primary visual cortex (V1). Efficient encoding predicts that the distribution of neurons’ preferred orientation follows the prior, with more neurons tuned to cardinal than oblique orientations by a factor of approximately 1.5. A similar ratio has been found for neurons in area V1 of monkey/cat [9, 10]. Also, the tuning widths of the model neurons vary between 25-42 deg depending on their preferred tuning (see Fig. 4c), matching the measured tuning width ratio of 0.6 between neurons tuned to the cardinal versus oblique orientations [9]. An important prediction of our model is that most of the tuning curves should be asymmetric. Such asymmetries have indeed been reported for the orientation tuning of neurons in area V1 [6, 7, 8]. We computed the asymmetry index for our model population as defined in previous studies [6, 7], and plotted it as a function of the preferred tuning of each neuron (Fig. 4d). The overall asymmetry index in our model population is 1.24 ± 0.11, which approximately matches the measured values for neurons in area V1 of the cat (1.26 ± 0.06) [6]. It also predicts that neurons tuned to the cardinal and oblique orientations should show less symmetry than those tuned to orientations in between. Finally, 2 Note, that these predictions are likely to change if the external noise is not symmetric. 6 a b 25 firing rate(Hz) 0 orientation(deg) asymmetry vs. tuning width 1.0 2.0 90 2.0 e asymmetry 1.0 0 asymmetry index 50 30 width (deg) 10 90 preferred tuning(deg) -90 0 d 0 0 90 asymmetry index 0 orientation(deg) tuning width -90 0 0 probability 0 -90 c efficient representation 0.01 0.01 image statistics -90 0 90 preferred tuning(deg) 25 30 35 40 tuning width (deg) Figure 4: Tuning characteristics of model neurons. a) Distribution of local orientations in natural images, replotted from [16]. b) Prior used in the model (black) and predicted tuning curves according to efficient coding (red). c) Tuning width as a function of preferred orientation. d) Tuning curves of cardinal and oblique neurons are more symmetric than those tuned to orientations in between. e) Both narrowly and broadly tuned neurons neurons show less asymmetry than neurons with tuning widths in between. neurons with tuning widths at the lower and upper end of the range are predicted to exhibit less asymmetry than those neurons whose widths lie in between these extremes (illustrated in Fig. 4e). These last two predictions have not been tested yet. 4.3 Predicted perceptual biases Our model framework also provides specific predictions for the expected perceptual biases. Humans show systematic biases in perceived orientation of visual stimuli such as e.g. arrays of Gabor patches (Fig. 5a,d). Two types of biases can be distinguished: First, perceived orientations show an absolute bias away from the cardinal orientations, thus away from the peaks of the orientation prior [2, 3]. We refer to these biases as absolute because they are typically measured by adjusting a noise-free reference until it matched the orientation of the test stimulus. Interestingly, these repulsive absolute biases are the larger the smaller the external stimulus noise is (see Fig. 5b). Second, the relative bias between the perceived overall orientations of a high-noise and a low-noise stimulus is toward the cardinal orientations as shown in Fig. 5c, and thus toward the peak of the prior distribution [3, 16]. The predicted perceptual biases of our model are shown Fig. 5e,f. We computed the likelihood function according to (2) and used the prior in (7). External noise was modeled by convolving the stimulus likelihood function with a Gaussian (different widths for different noise levels). The predictions well match both, the reported absolute bias away as well as the relative biases toward the cardinal orientations. Note, that our model framework correctly accounts for the fact that less external noise leads to larger absolute biases (see also discussion in section 3.2). 5 Discussion We have presented a modeling framework for perception that combines efficient (en)coding and Bayesian decoding. Efficient coding imposes constraints on the tuning characteristics of a population of neurons according to the stimulus distribution (prior). It thus establishes a direct link between prior and likelihood, and provides clear constraints on the latter for a Bayesian observer model of perception. We have shown that the resulting likelihoods are in general asymmetric, with 7 absolute bias (data) b c relative bias (data) -4 0 bias(deg) 4 a low-noise stimulus -90 e 90 absolute bias (model) low external noise high external noise 3 high-noise stimulus -90 f 0 90 relative bias (model) 0 bias(deg) d 0 attraction -3 repulsion -90 0 orientation (deg) 90 -90 0 orientation (deg) 90 Figure 5: Biases in perceived orientation: Human data vs. Model prediction. a,d) Low- and highnoise orientation stimuli of the type used in [3, 16]. b) Humans show absolute biases in perceived orientation that are away from the cardinal orientations. Data replotted from [2] (pink squares) and [3] (green (black) triangles: bias for low (high) external noise). c) Relative bias between stimuli with different external noise level (high minus low). Data replotted from [3] (blue triangles) and [16] (red circles). e,f) Model predictions for absolute and relative bias. heavier tails away from the prior peaks. We demonstrated that such asymmetric likelihoods can lead to the counter-intuitive prediction that a Bayesian estimator is biased away from the peaks of the prior distribution. Interestingly, such repulsive biases have been reported for human perception of visual orientation, yet a principled and consistent explanation of their existence has been missing so far. Here, we suggest that these counter-intuitive biases directly follow from the asymmetries in the likelihood function induced by efficient neural encoding of the stimulus. The good match between our model predictions and the measured perceptual biases and orientation tuning characteristics of neurons in primary visual cortex provides further support of our framework. Previous work has suggested that there might be a link between stimulus statistics, neuronal tuning characteristics, and perceptual behavior based on efficient coding principles, yet none of these studies has recognized the importance of the resulting likelihood asymmetries [16, 11]. We have demonstrated here that such asymmetries can be crucial in explaining perceptual data, even though the resulting estimates appear “anti-Bayesian” at first sight (see also models of sensory adaptation [23]). Note, that we do not provide a neural implementation of the Bayesian inference step. However, we and others have proposed various neural decoding schemes that can approximate Bayes’ leastsquares estimation using efficient coding [26, 25, 22]. It is also worth pointing out that our estimator is set to minimize total squared-error, and that other choices of the loss function (e.g. MAP estimator) could lead to different predictions. Our framework is general and should be directly applicable to other modalities. In particular, it might provide a new explanation for perceptual biases that are hard to reconcile with traditional Bayesian approaches [5]. Acknowledgments We thank M. Jogan and A. Tank for helpful comments on the manuscript. This work was partially supported by grant ONR N000141110744. 8 References [1] M. Jones, and B. C. Love. Bayesian fundamentalism or enlightenment? On the explanatory status and theoretical contributions of Bayesian models of cognition. Behavioral and Brain Sciences, 34, 169–231,2011. [2] D. P. Andrews. Perception of contours in the central fovea. Nature, 205:1218- 1220, 1965. [3] A. Tomassini, M. J.Morgam. and J. A. Solomon. Orientation uncertainty reduces perceived obliquity. Vision Res, 50, 541–547, 2010. [4] W. S. Geisler, D. Kersten. Illusions, perception and Bayes. Nature Neuroscience, 5(6):508- 510, 2002. [5] M. O. Ernst Perceptual learning: inverting the size-weight illusion. Current Biology, 19:R23- R25, 2009. [6] G. H. Henry, B. Dreher, P. O. Bishop. Orientation specificity of cells in cat striate cortex. J Neurophysiol, 37(6):1394-409,1974. [7] D. Rose, C. Blakemore An analysis of orientation selectivity in the cat’s visual cortex. Exp Brain Res., Apr 30;20(1):1-17, 1974. [8] N. V. Swindale. Orientation tuning curves: empirical description and estimation of parameters. Biol Cybern., 78(1):45-56, 1998. [9] R. L. De Valois, E. W. Yund, N. Hepler. The orientation and direction selectivity of cells in macaque visual cortex. Vision Res.,22, 531544,1982. [10] B. Li, M. R. Peterson, R. D. Freeman. The oblique effect: a neural basis in the visual cortex. J. Neurophysiol., 90, 204217, 2003. [11] D. Ganguli and E.P. Simoncelli. Implicit encoding of prior probabilities in optimal neural populations. In Adv. Neural Information Processing Systems NIPS 23, vol. 23:658–666, 2011. [12] M. D. McDonnell, N. G. Stocks. Maximally Informative Stimuli and Tuning Curves for Sigmoidal RateCoding Neurons and Populations. Phys Rev Lett., 101(5):058103, 2008. [13] H Helmholtz. Treatise on Physiological Optics (transl.). Thoemmes Press, Bristol, U.K., 2000. Original publication 1867. [14] Y. Weiss, E. Simoncelli, and E. Adelson. Motion illusions as optimal percept. Nature Neuroscience, 5(6):598–604, June 2002. [15] D.C. Knill and W. Richards, editors. Perception as Bayesian Inference. Cambridge University Press, 1996. [16] A R Girshick, M S Landy, and E P Simoncelli. Cardinal rules: visual orientation perception reflects knowledge of environmental statistics. Nat Neurosci, 14(7):926–932, Jul 2011. [17] M. Jazayeri and M.N. Shadlen. Temporal context calibrates interval timing. Nature Neuroscience, 13(8):914–916, 2010. [18] A.A. Stocker and E.P. Simoncelli. Noise characteristics and prior expectations in human visual speed perception. Nature Neuroscience, pages 578–585, April 2006. [19] H.B. Barlow. Possible principles underlying the transformation of sensory messages. In W.A. Rosenblith, editor, Sensory Communication, pages 217–234. MIT Press, Cambridge, MA, 1961. [20] D.M. Coppola, H.R. Purves, A.N. McCoy, and D. Purves The distribution of oriented contours in the real world. Proc Natl Acad Sci U S A., 95(7): 4002–4006, 1998. [21] N. Brunel and J.-P. Nadal. Mutual information, Fisher information and population coding. Neural Computation, 10, 7, 1731–1757, 1998. [22] X-X. Wei and A.A. Stocker. Bayesian inference with efficient neural population codes. In Lecture Notes in Computer Science, Artificial Neural Networks and Machine Learning - ICANN 2012, Lausanne, Switzerland, volume 7552, pages 523–530, 2012. [23] A.A. Stocker and E.P. Simoncelli. Sensory adaptation within a Bayesian framework for perception. In Y. Weiss, B. Sch¨ lkopf, and J. Platt, editors, Advances in Neural Information Processing Systems 18, pages o 1291–1298. MIT Press, Cambridge, MA, 2006. Oral presentation. [24] D.C. Knill. Robust cue integration: A Bayesian model and evidence from cue-conflict studies with stereoscopic and figure cues to slant. Journal of Vision, 7(7):1–24, 2007. [25] Deep Ganguli. Efficient coding and Bayesian inference with neural populations. PhD thesis, Center for Neural Science, New York University, New York, NY, September 2012. [26] B. Fischer. Bayesian estimates from heterogeneous population codes. In Proc. IEEE Intl. Joint Conf. on Neural Networks. IEEE, 2010. 9

6 0.089818567 187 nips-2012-Learning curves for multi-task Gaussian process regression

7 0.086080857 136 nips-2012-Forward-Backward Activation Algorithm for Hierarchical Hidden Markov Models

8 0.080661133 24 nips-2012-A mechanistic model of early sensory processing based on subtracting sparse representations

9 0.078688018 199 nips-2012-Link Prediction in Graphs with Autoregressive Features

10 0.076907195 77 nips-2012-Complex Inference in Neural Circuits with Probabilistic Population Codes and Topic Models

11 0.076593228 134 nips-2012-Finite Sample Convergence Rates of Zero-Order Stochastic Optimization Methods

12 0.07653635 151 nips-2012-High-Order Multi-Task Feature Learning to Identify Longitudinal Phenotypic Markers for Alzheimer's Disease Progression Prediction

13 0.07521107 312 nips-2012-Simultaneously Leveraging Output and Task Structures for Multiple-Output Regression

14 0.074620605 321 nips-2012-Spectral learning of linear dynamics from generalised-linear observations with application to neural population data

15 0.0741411 153 nips-2012-How Prior Probability Influences Decision Making: A Unifying Probabilistic Model

16 0.071935959 138 nips-2012-Fully Bayesian inference for neural models with negative-binomial spiking

17 0.071311198 287 nips-2012-Random function priors for exchangeable arrays with applications to graphs and relational data

18 0.071148314 190 nips-2012-Learning optimal spike-based representations

19 0.070865549 94 nips-2012-Delay Compensation with Dynamical Synapses

20 0.069336101 325 nips-2012-Stochastic optimization and sparse statistical recovery: Optimal algorithms for high dimensions


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, 0.21), (1, 0.014), (2, 0.007), (3, 0.089), (4, -0.012), (5, 0.148), (6, -0.001), (7, 0.039), (8, -0.015), (9, -0.017), (10, -0.025), (11, -0.034), (12, -0.011), (13, -0.026), (14, 0.033), (15, 0.001), (16, 0.01), (17, -0.04), (18, 0.046), (19, -0.039), (20, 0.019), (21, 0.059), (22, -0.024), (23, -0.029), (24, -0.038), (25, -0.068), (26, -0.063), (27, 0.04), (28, 0.039), (29, -0.012), (30, 0.021), (31, 0.057), (32, -0.006), (33, 0.039), (34, 0.061), (35, 0.005), (36, 0.034), (37, -0.015), (38, 0.075), (39, 0.086), (40, 0.042), (41, -0.057), (42, 0.019), (43, -0.082), (44, 0.043), (45, 0.03), (46, -0.016), (47, 0.018), (48, -0.083), (49, -0.002)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.95449996 333 nips-2012-Synchronization can Control Regularization in Neural Systems via Correlated Noise Processes

Author: Jake Bouvrie, Jean-jeacques Slotine

Abstract: To learn reliable rules that can generalize to novel situations, the brain must be capable of imposing some form of regularization. Here we suggest, through theoretical and computational arguments, that the combination of noise with synchronization provides a plausible mechanism for regularization in the nervous system. The functional role of regularization is considered in a general context in which coupled computational systems receive inputs corrupted by correlated noise. Noise on the inputs is shown to impose regularization, and when synchronization upstream induces time-varying correlations across noise variables, the degree of regularization can be calibrated over time. The resulting qualitative behavior matches experimental data from visual cortex. 1

2 0.69994521 94 nips-2012-Delay Compensation with Dynamical Synapses

Author: Chi Fung, K. Wong, Si Wu

Abstract: Time delay is pervasive in neural information processing. To achieve real-time tracking, it is critical to compensate the transmission and processing delays in a neural system. In the present study we show that dynamical synapses with shortterm depression can enhance the mobility of a continuous attractor network to the extent that the system tracks time-varying stimuli in a timely manner. The state of the network can either track the instantaneous position of a moving stimulus perfectly (with zero-lag) or lead it with an effectively constant time, in agreement with experiments on the head-direction systems in rodents. The parameter regions for delayed, perfect and anticipative tracking correspond to network states that are static, ready-to-move and spontaneously moving, respectively, demonstrating the strong correlation between tracking performance and the intrinsic dynamics of the network. We also find that when the speed of the stimulus coincides with the natural speed of the network state, the delay becomes effectively independent of the stimulus amplitude.

3 0.65877497 24 nips-2012-A mechanistic model of early sensory processing based on subtracting sparse representations

Author: Shaul Druckmann, Tao Hu, Dmitri B. Chklovskii

Abstract: Early stages of sensory systems face the challenge of compressing information from numerous receptors onto a much smaller number of projection neurons, a so called communication bottleneck. To make more efficient use of limited bandwidth, compression may be achieved using predictive coding, whereby predictable, or redundant, components of the stimulus are removed. In the case of the retina, Srinivasan et al. (1982) suggested that feedforward inhibitory connections subtracting a linear prediction generated from nearby receptors implement such compression, resulting in biphasic center-surround receptive fields. However, feedback inhibitory circuits are common in early sensory circuits and furthermore their dynamics may be nonlinear. Can such circuits implement predictive coding as well? Here, solving the transient dynamics of nonlinear reciprocal feedback circuits through analogy to a signal-processing algorithm called linearized Bregman iteration we show that nonlinear predictive coding can be implemented in an inhibitory feedback circuit. In response to a step stimulus, interneuron activity in time constructs progressively less sparse but more accurate representations of the stimulus, a temporally evolving prediction. This analysis provides a powerful theoretical framework to interpret and understand the dynamics of early sensory processing in a variety of physiological experiments and yields novel predictions regarding the relation between activity and stimulus statistics.

4 0.59537178 224 nips-2012-Multi-scale Hyper-time Hardware Emulation of Human Motor Nervous System Based on Spiking Neurons using FPGA

Author: C. M. Niu, Sirish Nandyala, Won J. Sohn, Terence Sanger

Abstract: Our central goal is to quantify the long-term progression of pediatric neurological diseases, such as a typical 10-15 years progression of child dystonia. To this purpose, quantitative models are convincing only if they can provide multi-scale details ranging from neuron spikes to limb biomechanics. The models also need to be evaluated in hyper-time, i.e. significantly faster than real-time, for producing useful predictions. We designed a platform with digital VLSI hardware for multiscale hyper-time emulations of human motor nervous systems. The platform is constructed on a scalable, distributed array of Field Programmable Gate Array (FPGA) devices. All devices operate asynchronously with 1 millisecond time granularity, and the overall system is accelerated to 365x real-time. Each physiological component is implemented using models from well documented studies and can be flexibly modified. Thus the validity of emulation can be easily advised by neurophysiologists and clinicians. For maximizing the speed of emulation, all calculations are implemented in combinational logic instead of clocked iterative circuits. This paper presents the methodology of building FPGA modules emulating a monosynaptic spinal loop. Emulated activities are qualitatively similar to real human data. Also discussed is the rationale of approximating neural circuitry by organizing neurons with sparse interconnections. In conclusion, our platform allows emulating pathological abnormalities such that motor symptoms will emerge and can be analyzed. It compels us to test the origins of childhood motor disorders and predict their long-term progressions. 1 Challenges of studying developmental motor disorders There is currently no quantitative model of how a neuropathological condition, which mainly affects the function of neurons, ends up causing the functional abnormalities identified in clinical examinations. The gap in knowledge is particularly evident for disorders in developing human nervous systems, i.e. childhood neurological diseases. In these cases, the ultimate clinical effect of cellu1 lar injury is compounded by a complex interplay among the child’s injury, development, behavior, experience, plasticity, etc. Qualitative insight has been provided by clinical experiences into the association between particular types of injury and particular types of outcome. Their quantitative linkages, nevertheless, have yet to be created – neither in clinic nor in cellular physiological tests. This discrepancy is significantly more prominent for individual child patients, which makes it very difficult to estimate the efficacy of treatment plans. In order to understand the consequence of injury and discover new treatments, it is necessary to create a modeling toolset with certain design guidelines, such that child neurological diseases can be quantitatively analyzed. Perhaps more than any other organ, the brain necessarily operates on multiple spatial and temporal scales. On the one hand, it is the neurons that perform fundamental computations, but neurons have to interact with large-scale organs (ears, eyes, skeletal muscles, etc.) to achieve global functions. This multi-scale nature worths more attention in injuries, where the overall deficits depend on both the cellular effects of injuries and the propagated consequences. On the other hand, neural processes in developmental diseases usually operate on drastically different time scales, e.g. spinal reflex in milliseconds versus learning in years. Thus when studying motor nervous systems, mathematical modeling is convincing only if it can provide multi-scale details, ranging from neuron spikes to limb biomechanics; also the models should be evaluated with time granularity as small as 1 millisecond, meanwhile the evaluation needs to continue trillions of cycles in order to cover years of life. It is particularly challenging to describe the multi-scale nature of human nervous system when modeling childhood movement disorders. Note that for a child who suffered brain injury at birth, the full development of all motor symptoms may easily take more than 10 years. Therefore the millisecondbased model needs to be evaluated significantly faster than real-time, otherwise the model will fail to produce any useful predictions in time. We have implemented realistic models for spiking motoneurons, sensory neurons, neural circuitry, muscle fibers and proprioceptors using VLSI and programmable logic technologies. All models are computed in Field Programmable Gate Array (FPGA) hardware in 365 times real-time. Therefore one year of disease progression can be assessed after one day of emulation. This paper presents the methodology of building the emulation platform. The results demonstrate that our platform is capable of producing physiologically realistic multi-scale signals, which are usually scarce in experiments. Successful emulations enabled by this platform will be used to verify theories of neuropathology. New treatment mechanisms and drug effects can also be emulated before animal experiments or clinical trials. 2 Methodology of multi-scale neural emulation A. Human arm B. Monosynaptic spinal loop C. Inner structure of muscle spindle Gamma Secondary dynamic Gamma output input static Primary input output Bag 1 αMN Bag 2 Chain Figure 1: Illustration of the multi-scale nature of motor nervous system. The motor part of human nervous system is responsible for maintaining body postures and generating voluntary movements. The multi-scale nature of motor nervous system is demonstrated in Fig.1. When the elbow (Fig.1A) is maintaining a posture or performing a movement, a force is established by the involved muscle based on how much spiking excitation the muscle receives from its αmotoneurons (Fig.1B). The α-motoneurons are regulated by a variety of sensory input, part of which comes directly from the proprioceptors in the muscle. As the primary proprioceptor found in skeletal muscles, a muscle spindle is another complex system that has its own microscopic Multiple-InputMultiple-Output structure (Fig.1C). Spindles continuously provide information about the length and lengthening speed of the muscle fiber. A muscle with its regulating motoneurons, sensory neurons and proprioceptors constitutes a monosynaptic spinal loop. This minimalist neurophysiological 2 structure is used as an example for explaining the multi-scale hyper-time emulation in hardware. Additional structures can be added to the backbone set-up using similar methodologies. 2.1 Modularized architecture for multi-scale models Decades of studies on neurophysiology provided an abundance of models characterizing different components of the human motor nervous system. The informational characteristics of physiological components allowed us to model them as functional structures, i.e. each of which converting input signals to certain outputs. In particular, within a monosynaptic spinal loop illustrated in Fig.1B, stretching the muscle will elicit a chain of physiological activities in: muscle stretch ⇒ spindle ⇒ sensory neuron ⇒ synapse ⇒ motoneuron ⇒ muscle contraction. The adjacent components must have compatible interfaces, and the interfacing variables must also be physiologically realistic. In our design, each component is mathematically described in Table 1: Table 1: Functional definition of neural models COMPONENT Neuron Synapse Muscle Spindle MATHEMATICAL DEFINITION S(t) = fneuron (I, t) I(t) = fsynapse (S, t) ˙ T (t) = fmuscle (S, L, L, t) ˙ Γdynamic , Γstatic , t) A(t) = fspindle (L, L, all components are modeled as black-box functions that map the inputs to the outputs. The meanings of these mathematical definitions are explained below. This design allows existing physiological models to be easily inserted and switched. In all models the input signals are time-varying, e.g. I = I(t), L = L(t) , etc. The argument of t in input signals are omitted throughout this paper. 2.2 Selection of models for emulation Models were selected in consideration of their computational cost, physiological verisimilitude, and whether it can be adapted to the mathematical form defined in Table 1. Model of Neuron The informational process for a neuron is to take post-synaptic current I as the input, and produce a binary spike train S in the output. The neuron model adopted in the emulation was developed by Izhikevich [1]: = 0.04v 2 + 5v + 140 − u + I = a(bv − u) v u (1) (2) if v = 30 mV, then v ← c, u ← u + d where a, b, c, d are free parameters tuned to achieve certain firing patterns. Membrane potential v directly determines a binary spike train S(t) that S(t) = 1 if v ≥ 30, otherwise S(t) = 0. Note that v in Izhikevich model is in millivolts and time t is in milliseconds. Therefore the coefficients in eq.1 need to be adjusted in correspondence to SI units. Model of Synapse When a pre-synaptic neuron spikes, i.e. S(0) = 1, an excitatory synapse subsequently issues an Excitatory Post-Synaptic Current (EPSC) that drives the post-synaptic neuron. Neural recording of hair cells in rats [2] provided evidence that the time profile of EPSC can be well characterized using the equations below: I(t) = Vm × e t d Vm −τ 0 t − e− τr Vm if t ≥ 0 (3) otherwise The key parameters in a synapse model is the time constants for rising (τr ) and decaying (τd ). In our emulation τr = 0.001 s and τr = 0.003 s. 3 Model of Muscle force and electromyograph (EMG) The primary effect of skeletal muscle is to convert α-motoneuron spikes S into force T , depending ˙ on the muscle’s instantaneous length L and lengthening speed L. We used Hill’s muscle model in the emulation with parameter tuning described in [3]. Another measurable output of muscle is electromyograph (EMG). EMG is the small skin current polarized by motor unit action potential (MUAP) when it travels along muscle fibers. Models exist to describe the typical waveform picked by surface EMG electrodes. In this project we chose to implement the one described in [4]. Model of Proprioceptor Spindle is a sensory organ that provides the main source of proprioceptive information. As can be seen in Fig.1C, a spindle typically produces two afferent outputs (primary Ia and secondary II) ˙ according to its gamma fusimotor drives (Γdynamic and Γstatic ) and muscle states (L and L). There is currently no closed-form models describing spindle functions due to spindle’s significant nonlinearity. On representative model that numerically approximates the spindle dynamics was developed by Mileusnic et al. [5]. The model used differential equations to characterize a typical cat soleus spindle. Eqs.4-10 present a subset of this model for one type of spindle fiber (bag1): Γdynamic − x0 /τ Γdynamic + Ω2 bag1 x0 ˙ = x1 ˙ = x2 1 = [TSR − TB − TP R − Γ1 x0 ] M x2 ˙ (4) (5) (6) where TSR TB TP R CSS = KSR (L − x1 − LSR0 ) (7) 0.3 = (B0 + B1 x0 ) · (x1 − R) · CSS · |x2 | = KP R (x1 − LP R0 ) 2 = −1 −1000x2 1+e (8) (9) (10) Eq.8 and 10 suggest that evaluating the spindle model requires multiplication, division as well as more complex arithmetics like polynomials and exponentials. The implementation details are described in Section 3. 2.3 Neuron connectivity with sparse interconnections Although the number of spinal neurons (~1 billion) is significantly less compared to that of cortical neurons (~100 billion), a fully connected spinal network still means approximately 2 trillion synaptic endings [6]. Implementing such a huge number of synapses imposes a major challenge, if not impossible, given limited hardware resource. In this platform we approximated the neural connectivity by sparsely connecting sensory neurons to motoneurons as parallel pathways. We do not attempt to introduce the full connectivity. The rationale is that in a neural control system, the effect of a single neuron can be considered as mapping current state x to change in state x through a band-limited channel. Therefore when a collection of ˙ neurons are firing stochastically, the probability of x depends on both x and the firing behavior s ˙ (s = 1 when spiking, otherwise s = 0) of each neuron, as such: p(x|x, s) = p(x|s = 1)p(s = 1|x) + p(x|s = 0)p(s = 0|x) ˙ ˙ ˙ (11) Eq.11 is a master equation that determines a probability flow on the state. From the Kramers-Moyal expansion we can associate this probability flow with a partial differential equation: ∂ p(x, t) ∂t ∞ − = i=1 ∂ ∂x i D(i) (x)p(x, t) (12) where D(i) (x) is a time-invariant term that modifies the change of probability density based on its i-th gradient. 4 Under certain conditions [7, 8], D(i) (x) for i > 2 all vanish and therefore the probability flow can be described deterministically using a linear operator L: ∂ ∂ ∂ 2 (2) D (x) p(x, t) = Lp(x, t) (13) p(x, t) = − D(1) (x) + ∂t ∂x ∂x2 This means that various Ls can be superimposed to achieve complex system dynamics (illustrated in Fig.2A). B. Equivalent network with sparse interconnections A. Neuron function as superimposed linear operators SN Sensory Input + SN SN SN αMN αMN αMN Motor Output αMN Figure 2: Functions of neuron population can be described as the combination of linear operators (A). Therefore the original neural function can be equivalently produced by sparsely connected neurons formalizing parallel pathways (B). As a consequence, the statistical effect of two fully connected neuron populations is equivalent to ones that are only sparsely connected, as long as the probability flow can be described by the same L. For a movement task, in particular, it is the statistical effect from the neuron ensemble onto skeletal muscles that determines the global behavior. Therefore we argue that it is feasible to approximate the spinal cord connectivity by sparsely interconnecting sensory and motor neurons (Fig.2B). Here a pool of homogenous sensory neurons projects to another pool of homogeneous α-motoneurons. Pseudorandom noise is added to the input of all homogeneous neurons within a population. It is worth noting that this approximation significantly reduces the number of synapses that need to be implemented in hardware. 3 Hardware implementation on FPGA We select FPGA as the implementation device due to its inherent parallelism that resembles the nervous system. FPGA is favored over GPU or clustered CPUs because it is relatively easy to network hundreds of nodes under flexible protocols. The platform is distributed on multiple nodes of Xilinx Spartan-6 devices. The interfacing among FPGAs and computers is created using OpalKelly development board XEM6010. The dynamic range of variables is tight in models of Izhikevich neuron, synapse and EMG. This helps maintaining the accuracy of models even when they are evaluated in 32-bit fixed-point arithmetics. The spindle model, in contrast, requires floating-point arithmetics due to its wide dynamic range and complex calculations (see eq.4-10). Hyper-time computations with floating-point numbers are resource consuming and therefore need to be implemented with special attentions. 3.1 Floating-point arithmetics in combinational logic Our arithmetic implementations are compatible with IEEE-754 standard. Typical floating-point arithmetic IP cores are either pipe-lined or based on iterative algorithms such as CORDIC, all of which require clocks to schedule the calculation. In our platform, no clock is provided for model evaluations thus all arithmetics need to be executed in pure combinational logic. Taking advantage of combinational logic allows all model evaluations to be 1) fast, the evaluation time depends entirely on the propagating and settling time of signals, which is on the order of microseconds, and 2) parallel, each model is evaluated on its own circuit without waiting for any other results. Our implementations of adder and multiplier are inspired by the open source project “Free FloatingPoint Madness”, available at http://www.hmc.edu/chips/. Please contact the authors of this paper if the modified code is needed. 5 Fast combinational floating-point division Floating-point division is even more resource demanding than multiplications. We avoided directly implementing the dividing algorithm by approximating it with additions and multiplications. Our approach is inspired by an algorithm described in [9], which provides a good approximation of the inverse square root for any positive number x within one Newton-Raphson iteration: 1 x Q(x) = √ ≈ x(1.5 − · x2 ) 2 x (x > 0) (14) Q(x) can be implemented only using floating-point adders and multipliers. Thereby any division with a positive divisor can be achieved if two blocks of Q(x) are concatenated: a a (15) = √ √ = a · Q(b) · Q(b) (b > 0) b b· b This algorithm has been adjusted to also work with negative divisors (b < 0). Numerical integrators for differential equations Evaluating the instantaneous states of differential equation models require a fixed-step numerical integrator. Backward Euler’s Method was chosen to balance the numerical error and FPGA usage: x ˙ xn+1 = f (x, t) = xn + T f (xn+1 , tn+1 ) (16) (17) where T is the sampling interval. f (x, t) is the derivative function for state variable x. 3.2 Asynchronous spike-based communication between FPGA chips Clock Spike clean count Counter 1 1 2 1 2 3 Figure 3: Timing diagram of asynchronous spike-based communication FPGA nodes are networked by transferring 1-bit binary spikes to each other. Our design allowed the sender and the receiver to operate on independent clocks without having to synchronize. The timing diagram of the spike-based communication is shown in Fig.3. The sender issues Spike with a pulse width of 1/(365 × Femu ) second. Each Spike then triggers a counting event on the receiver, meanwhile each Clock first reads the accumulated spike count and subsequently cleans the counter. Note that the phase difference between Spike and Clock is not predictable due to asynchronicity. 3.3 Serialize neuron evaluations within a homogeneous population Different neuron populations are instantiated as standalone circuits. Within in each population, however, homogeneous neurons mentioned in Section 2.3 are evaluated in series in order to optimize FPGA usage. Within each FPGA node all modules operate with a central clock, which is the only source allowed to trigger any updating event. Therefore the maximal number of neurons that can be serialized (Nserial ) is restrained by the following relationship: Ffpga = C × Nserial × 365 × Femu (18) Here Ffpga is the fastest clock rate that a FPGA can operate on; C = 4 is the minimal clock cycles needed for updating each state variable in the on-chip memory; Femu = 1 kHz is the time granularity of emulation (1 millisecond), and 365 × Femu represents 365x real-time. Consider that Xilinx 6 Spartan-6 FPGA devices peaks at 200MHz central clock frequency, the theoretical maximum of neurons that can be serialized is Nserial 200 MHz/(4 × 365 × 1 kHz) ≈ 137 (19) In the current design we choose Nserial = 128. 4 Results: emulated activities of motor nervous system Figure 4 shows the implemented monosynaptic spinal loop in schematics and in operation. Each FPGA node is able to emulate monosynaptic spinal loops consisting of 1,024 sensory and 1,024 motor neurons, i.e. 2,048 neurons in total. The spike-based asynchronous communication is successful between two FPGA nodes. Note that the emulation has to be significantly slowed down for on-line plotting. When the emulation is at full speed (365x real-time) the software front-end is not able to visualize the signals due to limited data throughput. 128 SNs 128 αMNs SN αMN 128 SNs 128 αMNs SN αMN ... 8 parallel pathways 2,048 neurons Figure 4: The neural emulation platform in operation. Left: Neural circuits implemented for each FPGA node including 2,048 neurons. SN = Sensory Neuron; αMN = α-motoneuron. Center: One working FPGA node. Right: Two FPGA nodes networked using asynchronous spiking protocol. The emulation platform successfully created multi-scale information when the muscle is externally stretched (Fig.5A). We also tested if our emulated motor system is able to produce the recruitment order and size principles observed in real physiological data. It has been well known that when a voluntary motor command is sent to the α-motoneuron pool, the motor units are recruited in an order that small ones get recruited first, followed by the big ones [10]. The comparison between our results and real data are shown in Fig.5B, where the top panel shows 20 motor unit activities emulated using our platform, and the bottom panel shows decoded motor unit activities from real human EMG [11]. No qualitative difference was found. 5 Discussion and future work We designed a hardware platform for emulating the multi-scale motor nervous activities in hypertime. We managed to use one node of single Xilinx Spartan-6 FPGA to emulate monosynaptic spinal loops consisting of 2,048 neurons, associated muscles and proprioceptors. The neurons are organized as parallel pathways with sparse interconnections. The emulation is successfully accelerated to 365x real-time. The platform can be scaled by networking multiple FPGA nodes, which is enabled by an asynchronous spike-based communication protocol. The emulated monosynaptic spinal loops are capable of producing reflex-like activities in response to muscle stretch. Our results of motor unit recruitment order are compatible with the physiological data collected in real human subjects. There is a question of whether this stochastic system turns out chaotic, especially with accumulated errors from Backward Euler’s integrator. Note that the firing property of a neuron population is usually stable even with explicit noise [8], and spindle inputs are measured from real robots so the integrator errors are corrected at every iteration. To our knowledge, the system is not critically sensitive to the initial conditions or integrator errors. This question, however, is both interesting and important for in-depth investigations in the future. 7 It has been shown [12] that replicating classic types of spinal interneurons (propriospinal, Iaexcitatory, Ia-inhibitory, Renshaw, etc.) is sufficient to produce stabilizing responses and rapid reaching movement in a wrist. Our platform will introduce those interneurons to describe the known spinal circuitry in further details. Physiological models will also be refined as needed. For the purpose of modeling movement behavior or diseases, Izhikevich model is a good balance between verisimilitude and computational cost. Nevertheless when testing drug effects along disease progression, neuron models are expected to cover sufficient molecular details including how neurotransmitters affect various ion channels. With the advancing of programmable semiconductor technology, it is expected to upgrade our neuron model to Hodgkin-Huxley’s. For the muscle models, Hill’s type of model does not fit the muscle properties accurately enough when the muscle is being shortened. Alternative models will be tested. Other studies showed that the functional dexterity of human limbs – especially in the hands – is critically enabled by the tendon configurations and joint geometry [13]. As a result, if our platform is used to understand whether known neurophysiology and biomechanics are sufficient to produce able and pathological movements, it will be necessary to use this platform to control human-like limbs. Since the emulation speed can be flexibly adjusted from arbitrarily slow to 365x real-time, when speeded to exactly 1x real-time the platform will function as a digital controller with 1kHz refresh rate. The main purpose of the emulation is to learn how certain motor disorders progress during childhood development. This first requires the platform to reproduce motor symptoms that are compatible with clinical observations. For example it has been suggested that muscle spasticity in rats is associated with decreased soma size of α-motoneurons [14], which presumably reduced the firing threshold of neurons. Thus when lower firing threshold is introduced to the emulated motoneuron pool, similar EMG patterns as in [15] should be observed. It is also necessary for the symptoms to evolve with neural plasticity. In the current version we presume that the structure of each component remains time invariant. In the future work Spike Timing Dependent Plasticity (STDP) will be introduced such that all components are subject to temporal modifications. B. Verify motor unit recruitment pattern A. Multi-scale activities from emulation Emulation 1s Stretch Spindle Ia Sensory post-synaptic current Real Data Motoneurons Muscle Force EMG Figure 5: A) Physiological activity emulated by each model when the muscle is sinusoidally stretched. B) Comparing the emulated motor unit recruitment order with real experimental data. Acknowledgments The authors thank Dr. Gerald Loeb for helping set up the emulation of spindle models. This project is supported by NIH NINDS grant R01NS069214-02. 8 References [1] Izhikevich, E. M. Simple model of spiking neurons. IEEE transactions on neural networks / a publication of the IEEE Neural Networks Council 14, 1569–1572 (2003). [2] Glowatzki, E. & Fuchs, P. A. Transmitter release at the hair cell ribbon synapse. Nature neuroscience 5, 147–154 (2002). [3] Shadmehr, R. & Wise, S. P. A Mathematical Muscle Model. In Supplementary documents for “Computational Neurobiology of Reaching and Pointing”, 1–18 (MIT Press, Cambridge, MA, 2005). [4] Fuglevand, A. J., Winter, D. A. & Patla, A. E. Models of recruitment and rate coding organization in motor-unit pools. Journal of neurophysiology 70, 2470–2488 (1993). [5] Mileusnic, M. P., Brown, I. E., Lan, N. & Loeb, G. E. Mathematical models of proprioceptors. I. Control and transduction in the muscle spindle. Journal of neurophysiology 96, 1772–1788 (2006). [6] Gelfan, S., Kao, G. & Ruchkin, D. S. The dendritic tree of spinal neurons. The Journal of comparative neurology 139, 385–411 (1970). [7] Sanger, T. D. Neuro-mechanical control using differential stochastic operators. In Engineering in Medicine and Biology Society (EMBC), 2010 Annual International Conference of the IEEE, 4494–4497 (2010). [8] Sanger, T. D. Distributed control of uncertain systems using superpositions of linear operators. Neural computation 23, 1911–1934 (2011). [9] Lomont, C. Fast inverse square root (2003). URL http://www.lomont.org/Math/Papers/ 2003/InvSqrt.pdf. [10] Henneman, E. Relation between size of neurons and their susceptibility to discharge. Science (New York, N.Y.) 126, 1345–1347 (1957). [11] De Luca, C. J. & Hostage, E. C. Relationship between firing rate and recruitment threshold of motoneurons in voluntary isometric contractions. Journal of neurophysiology 104, 1034–1046 (2010). [12] Raphael, G., Tsianos, G. A. & Loeb, G. E. Spinal-like regulator facilitates control of a two-degree-offreedom wrist. The Journal of neuroscience : the official journal of the Society for Neuroscience 30, 9431–9444 (2010). [13] Valero-Cuevas, F. J. et al. The tendon network of the fingers performs anatomical computation at a macroscopic scale. IEEE transactions on bio-medical engineering 54, 1161–1166 (2007). [14] Brashear, A. & Elovic, E. Spasticity: Diagnosis and Management (Demos Medical, 2010), 1 edn. [15] Levin, M. F. & Feldman, A. G. The role of stretch reflex threshold regulation in normal and impaired motor control. Brain research 657, 23–30 (1994). 9

5 0.59148896 336 nips-2012-The Coloured Noise Expansion and Parameter Estimation of Diffusion Processes

Author: Simon Lyons, Amos J. Storkey, Simo Särkkä

Abstract: Stochastic differential equations (SDE) are a natural tool for modelling systems that are inherently noisy or contain uncertainties that can be modelled as stochastic processes. Crucial to the process of using SDE to build mathematical models is the ability to estimate parameters of those models from observed data. Over the past few decades, significant progress has been made on this problem, but we are still far from having a definitive solution. We describe a novel method of approximating a diffusion process that we show to be useful in Markov chain Monte-Carlo (MCMC) inference algorithms. We take the ‘white’ noise that drives a diffusion process and decompose it into two terms. The first is a ‘coloured noise’ term that can be deterministically controlled by a set of auxilliary variables. The second term is small and enables us to form a linear Gaussian ‘small noise’ approximation. The decomposition allows us to take a diffusion process of interest and cast it in a form that is amenable to sampling by MCMC methods. We explain why many state-of-the-art inference methods fail on highly nonlinear inference problems, and we demonstrate experimentally that our method performs well in such situations. Our results show that this method is a promising new tool for use in inference and parameter estimation problems. 1

6 0.58711439 321 nips-2012-Spectral learning of linear dynamics from generalised-linear observations with application to neural population data

7 0.58531052 262 nips-2012-Optimal Neural Tuning Curves for Arbitrary Stimulus Distributions: Discrimax, Infomax and Minimum $L p$ Loss

8 0.5785529 177 nips-2012-Learning Invariant Representations of Molecules for Atomization Energy Prediction

9 0.57349187 113 nips-2012-Efficient and direct estimation of a neural subunit model for sensory coding

10 0.57163507 114 nips-2012-Efficient coding provides a direct link between prior and likelihood in perceptual Bayesian inference

11 0.56874818 195 nips-2012-Learning visual motion in recurrent neural networks

12 0.56539154 110 nips-2012-Efficient Reinforcement Learning for High Dimensional Linear Quadratic Systems

13 0.56478244 268 nips-2012-Perfect Dimensionality Recovery by Variational Bayesian PCA

14 0.5588721 66 nips-2012-Causal discovery with scale-mixture model for spatiotemporal variance dependencies

15 0.5553714 39 nips-2012-Analog readout for optical reservoir computers

16 0.54596484 46 nips-2012-Assessing Blinding in Clinical Trials

17 0.54310876 161 nips-2012-Interpreting prediction markets: a stochastic approach

18 0.53464651 77 nips-2012-Complex Inference in Neural Circuits with Probabilistic Population Codes and Topic Models

19 0.53023785 153 nips-2012-How Prior Probability Influences Decision Making: A Unifying Probabilistic Model

20 0.52935982 199 nips-2012-Link Prediction in Graphs with Autoregressive Features


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(0, 0.037), (17, 0.036), (21, 0.058), (25, 0.178), (36, 0.017), (38, 0.126), (39, 0.01), (42, 0.044), (54, 0.028), (55, 0.034), (74, 0.049), (76, 0.123), (77, 0.011), (80, 0.098), (92, 0.076)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.85374153 333 nips-2012-Synchronization can Control Regularization in Neural Systems via Correlated Noise Processes

Author: Jake Bouvrie, Jean-jeacques Slotine

Abstract: To learn reliable rules that can generalize to novel situations, the brain must be capable of imposing some form of regularization. Here we suggest, through theoretical and computational arguments, that the combination of noise with synchronization provides a plausible mechanism for regularization in the nervous system. The functional role of regularization is considered in a general context in which coupled computational systems receive inputs corrupted by correlated noise. Noise on the inputs is shown to impose regularization, and when synchronization upstream induces time-varying correlations across noise variables, the degree of regularization can be calibrated over time. The resulting qualitative behavior matches experimental data from visual cortex. 1

2 0.77175075 77 nips-2012-Complex Inference in Neural Circuits with Probabilistic Population Codes and Topic Models

Author: Jeff Beck, Alexandre Pouget, Katherine A. Heller

Abstract: Recent experiments have demonstrated that humans and animals typically reason probabilistically about their environment. This ability requires a neural code that represents probability distributions and neural circuits that are capable of implementing the operations of probabilistic inference. The proposed probabilistic population coding (PPC) framework provides a statistically efficient neural representation of probability distributions that is both broadly consistent with physiological measurements and capable of implementing some of the basic operations of probabilistic inference in a biologically plausible way. However, these experiments and the corresponding neural models have largely focused on simple (tractable) probabilistic computations such as cue combination, coordinate transformations, and decision making. As a result it remains unclear how to generalize this framework to more complex probabilistic computations. Here we address this short coming by showing that a very general approximate inference algorithm known as Variational Bayesian Expectation Maximization can be naturally implemented within the linear PPC framework. We apply this approach to a generic problem faced by any given layer of cortex, namely the identification of latent causes of complex mixtures of spikes. We identify a formal equivalent between this spike pattern demixing problem and topic models used for document classification, in particular Latent Dirichlet Allocation (LDA). We then construct a neural network implementation of variational inference and learning for LDA that utilizes a linear PPC. This network relies critically on two non-linear operations: divisive normalization and super-linear facilitation, both of which are ubiquitously observed in neural circuits. We also demonstrate how online learning can be achieved using a variation of Hebb’s rule and describe an extension of this work which allows us to deal with time varying and correlated latent causes. 1 Introduction to Probabilistic Inference in Cortex Probabilistic (Bayesian) reasoning provides a coherent and, in many ways, optimal framework for dealing with complex problems in an uncertain world. It is, therefore, somewhat reassuring that behavioural experiments reliably demonstrate that humans and animals behave in a manner consistent with optimal probabilistic reasoning when performing a wide variety of perceptual [1, 2, 3], motor [4, 5, 6], and cognitive tasks[7]. This remarkable ability requires a neural code that represents probability distribution functions of task relevant stimuli rather than just single values. While there 1 are many ways to represent functions, Bayes rule tells us that when it comes to probability distribution functions, there is only one statistically optimal way to do it. More precisely, Bayes Rule states that any pattern of activity, r, that efficiently represents a probability distribution over some task relevant quantity s, must satisfy the relationship p(s|r) ∝ p(r|s)p(s), where p(r|s) is the stimulus conditioned likelihood function that specifies the form of neural variability, p(s) gives the prior belief regarding the stimulus, and p(s|r) gives the posterior distribution over values of the stimulus, s given the representation r . Of course, it is unlikely that the nervous system consistently achieves this level of optimality. None-the-less, Bayes rule suggests the existence of a link between neural variability as characterized by the likelihood function p(r|s) and the state of belief of a mature statistical learning machine such as the brain. The so called Probabilistic Population Coding (or PPC) framework[8, 9, 10] takes this link seriously by proposing that the function encoded by a pattern of neural activity r is, in fact, the likelihood function p(r|s). When this is the case, the precise form of the neural variability informs the nature of the neural code. For example, the exponential family of statistical models with linear sufficient statistics has been shown to be flexible enough to model the first and second order statistics of in vivo recordings in awake behaving monkeys[9, 11, 12] and anesthetized cats[13]. When the likelihood function is modeled in this way, the log posterior probability over the stimulus is linearly encoded by neural activity, i.e. log p(s|r) = h(s) · r − log Z(r) (1) Here, the stimulus dependent kernel, h(s), is a vector of functions of s, the dot represents a standard dot product, and Z(r) is the partition function which serves to normalize the posterior. This log linear form for a posterior distribution is highly computationally convenient and allows for evidence integration to be implemented via linear operations on neural activity[14, 8]. Proponents of this kind of linear PPC have demonstrated how to build biologically plausible neural networks capable of implementing the operations of probabilistic inference that are needed to optimally perform the behavioural tasks listed above. This includes, linear PPC implementations of cue combination[8], evidence integration over time, maximum likelihood and maximum a posterior estimation[9], coordinate transformation/auditory localization[10], object tracking/Kalman filtering[10], explaining away[10], and visual search[15]. Moreover, each of these neural computations has required only a single recurrently connected layer of neurons that is capable of just two non-linear operations: coincidence detection and divisive normalization, both of which are widely observed in cortex[16, 17]. Unfortunately, this research program has been a piecemeal effort that has largely proceeded by building neural networks designed deal with particular problems. As a result, there have been no proposals for a general principle by which neural network implementations of linear PPCs might be generated and no suggestions regarding how to deal with complex (intractable) problems of probabilistic inference. In this work, we will partially address this short coming by showing that Variation Bayesian Expectation Maximization (VBEM) algorithm provides a general scheme for approximate inference and learning with linear PPCs. In section 2, we briefly review the VBEM algorithm and show how it naturally leads to a linear PPC representation of the posterior as well as constraints on the neural network dynamics which build that PPC representation. Because this section describes the VB-PPC approach rather abstractly, the remainder of the paper is dedicated to concrete applications. As a motivating example, we consider the problem of inferring the concentrations of odors in an olfactory scene from a complex pattern of spikes in a population of olfactory receptor neurons (ORNs). In section 3, we argue that this requires solving a spike pattern demixing problem which is indicative of the generic problem faced by many layers of cortex. We then show that this demixing problem is equivalent to the problem addressed by a class of models for text documents know as probabilistic topic models, in particular Latent Dirichlet Allocation or LDA[18]. In section 4, we apply the VB-PPC approach to build a neural network implementation of probabilistic inference and learning for LDA. This derivation shows that causal inference with linear PPC’s also critically relies on divisive normalization. This result suggests that this particular non-linearity may be involved in very general and fundamental probabilistic computation, rather than simply playing a role in gain modulation. In this section, we also show how this formulation allows for a probabilistic treatment of learning and show that a simple variation of Hebb’s rule can implement Bayesian learning in neural circuits. 2 We conclude this work by generalizing this approach to time varying inputs by introducing the Dynamic Document Model (DDM) which can infer short term fluctuations in the concentrations of individual topics/odors and can be used to model foraging and other tracking tasks. 2 Variational Bayesian Inference with linear Probabilistic Population Codes Variational Bayesian (VB) inference refers to a class of deterministic methods for approximating the intractable integrals which arise in the context of probabilistic reasoning. Properly implemented it can result a fast alternative to sampling based methods of inference such as MCMC[19] sampling. Generically, the goal of any Bayesian inference algorithm is to infer a posterior distribution over behaviourally relevant latent variables Z given observations X and a generative model which specifies the joint distribution p(X, Θ, Z). This task is confounded by the fact that the generative model includes latent parameters Θ which must be marginalized out, i.e. we wish to compute, p(Z|X) ∝ p(X, Θ, Z)dΘ (2) When the number of latent parameters is large this integral can be quite unwieldy. The VB algorithms simplify this marginalization by approximating the complex joint distribution over behaviourally relevant latents and parameters, p(Θ, Z|X), with a distribution q(Θ, Z) for which integrals of this form are easier to deal with in some sense. There is some art to choosing the particular form for the approximating distribution to make the above integral tractable, however, a factorized approximation is common, i.e. q(Θ, Z) = qΘ (Θ)qZ (Z). Regardless, for any given observation X, the approximate posterior is found by minimizing the Kullback-Leibler divergence between q(Θ, Z) and p(Θ, Z|X). When a factorized posterior is assumed, the Variational Bayesian Expectation Maximization (VBEM) algorithm finds a local minimum of the KL divergence by iteratively updating, qΘ (Θ) and qZ (Z) according to the scheme n log qΘ (Θ) ∼ log p(X, Θ, Z) n qZ (Z) and n+1 log qZ (Z) ∼ log p(X, Θ, Z) n qΘ (Θ) (3) Here the brackets indicate an expected value taken with respect to the subscripted probability distribution function and the tilde indicates equality up to a constant which is independent of Θ and Z. The key property to note here is that the approximate posterior which results from this procedure is in an exponential family form and is therefore representable by a linear PPC (Eq. 1). This feature allows for the straightforward construction of networks which implement the VBEM algorithm with linear PPC’s in the following way. If rn and rn are patterns of activity that use a linear PPC representation Θ Z of the relevant posteriors, then n log qΘ (Θ) ∼ hΘ (Θ) · rn Θ and n+1 log qZ (Z) ∼ hZ (Z) · rn+1 . Z (4) Here the stimulus dependent kernels hZ (Z) and hΘ (Θ) are chosen so that their outer product results in a basis that spans the function space on Z × Θ given by log p(X, Θ, Z) for every X. This choice guarantees that there exist functions fΘ (X, rn ) and fZ (X, rn ) such that Z Θ rn = fΘ (X, rn ) Θ Z and rn+1 = fZ (X, rn ) Θ Z (5) satisfy Eq. 3. When this is the case, simply iterating the discrete dynamical system described by Eq. 5 until convergence will find the VBEM approximation to the posterior. This is one way to build a neural network implementation of the VB algorithm. However, its not the only way. In general, any dynamical system which has stable fixed points in common with Eq. 5 can also be said to implement the VBEM algorithm. In the example below we will take advantage of this flexibility in order to build biologically plausible neural network implementations. 3 Response! to Mixture ! of Odors! Single Odor Response Cause Intensity Figure 1: (Left) Each cause (e.g. coffee) in isolation results in a pattern of neural activity (top). When multiple causes contribute to a scene this results in an overall pattern of neural activity which is a mixture of these patterns weighted by the intensities (bottom). (Right) The resulting pattern can be represented by a raster, where each spike is colored by its corresponding latent cause. 3 Probabilistic Topic Models for Spike Train Demixing Consider the problem of odor identification depicted in Fig. 1. A typical mammalian olfactory system consists of a few hundred different types of olfactory receptor neurons (ORNs), each of which responds to a wide range of volatile chemicals. This results in a highly distributed code for each odor. Since, a typical olfactory scene consists of many different odors at different concentrations, the pattern of ORN spike trains represents a complex mixture. Described in this way, it is easy to see that the problem faced by early olfactory cortex can be described as the task of demixing spike trains to infer latent causes (odor intensities). In many ways this olfactory problem is a generic problem faced by each cortical layer as it tries to make sense of the activity of the neurons in the layer below. The input patterns of activity consist of spikes (or spike counts) labeled by the axons which deliver them and summarized by a histogram which indicates how many spikes come from each input neuron. Of course, just because a spike came from a particular neuron does not mean that it had a particular cause, just as any particular ORN spike could have been caused by any one of a large number of volatile chemicals. Like olfactory codes, cortical codes are often distributed and multiple latent causes can be present at the same time. Regardless, this spike or histogram demixing problem is formally equivalent to a class of demixing problems which arise in the context of probabilistic topic models used for document modeling. A simple but successful example of this kind of topic model is called Latent Dirichlet Allocation (LDA) [18]. LDA assumes that word order in documents is irrelevant and, therefore, models documents as histograms of word counts. It also assumes that there are K topics and that each of these topics appears in different proportions in each document, e.g. 80% of the words in a document might be concerned with coffee and 20% with strawberries. Words from a given topic are themselves drawn from a distribution over words associated with that topic, e.g. when talking about coffee you have a 5% chance of using the word ’bitter’. The goal of LDA is to infer both the distribution over topics discussed in each document and the distribution of words associated with each topic. We can map the generative model for LDA onto the task of spike demixing in cortex by letting topics become latent causes or odors, words become neurons, word occurrences become spikes, word distributions associated with each topic become patterns of neural activity associated with each cause, and different documents become the observed patterns of neural activity on different trials. This equivalence is made explicit in Fig. 2 which describes the standard generative model for LDA applied to documents on the left and mixtures of spikes on the right. 4 LDA Inference and Network Implementation In this section we will apply the VB-PPC formulation to build a biologically plausible network capable of approximating probabilistic inference for spike pattern demixing. For simplicity, we will use the equivalent Gamma-Poisson formulation of LDA which directly models word and topic counts 4 1. For each topic k = 1, . . . , K, (a) Distribution over words βk ∼ Dirichlet(η0 ) 2. For document d = 1, . . . , D, (a) Distribution over topics θd ∼ Dirichlet(α0 ) (b) For word m = 1, . . . , Ωd i. Topic assignment zd,m ∼ Multinomial(θd ) ii. Word assignment ωd,m ∼ Multinomial(βzm ) 1. For latent cause k = 1, . . . , K, (a) Pattern of neural activity βk ∼ Dirichlet(η0 ) 2. For scene d = 1, . . . , D, (a) Relative intensity of each cause θd ∼ Dirichlet(α0 ) (b) For spike m = 1, . . . , Ωd i. Cause assignment zd,m ∼ Multinomial(θd ) ii. Neuron assignment ωd,m ∼ Multinomial(βzm ) Figure 2: (Left) The LDA generative model in the context of document modeling. (Right) The corresponding LDA generative model mapped onto the problem of spike demixing. Text related attributes on the left, in red, have been replaced with neural attributes on the right, in green. rather than topic assignments. Specifically, we define, Rd,j to be the number of times neuron j fires during trial d. Similarly, we let Nd,j,k to be the number of times a spike in neuron j comes from cause k in trial d. These new variables play the roles of the cause and neuron assignment variables, zd,m and ωd,m by simply counting them up. If we let cd,k be an un-normalized intensity of cause j such that θd,k = cd,k / k cd,k then the generative model, Rd,j = k Nd,j,k Nd,j,k ∼ Poisson(βj,k cd,k ) 0 cd,k ∼ Gamma(αk , C −1 ). (6) is equivalent to the topic models described above. Here the parameter C is a scale parameter which sets the expected total number of spikes from the population on each trial. Note that, the problem of inferring the wj,k and cd,k is a non-negative matrix factorization problem similar to that considered by Lee and Seung[20]. The primary difference is that, here, we are attempting to infer a probability distribution over these quantities rather than maximum likelihood estimates. See supplement for details. Following the prescription laid out in section 2, we approximate the posterior over latent variables given a set of input patterns, Rd , d = 1, . . . , D, with a factorized distribution of the form, qN (N)qc (c)qβ (β). This results in marginal posterior distributions q (β:,k |η:,k ), q cd,k |αd,k , C −1 + 1 ), and q (Nd,j,: | log pd,j,: , Rd,i ) which are Dirichlet, Gamma, and Multinomial respectively. Here, the parameters η:,k , αd,k , and log pd,j,: are the natural parameters of these distributions. The VBEM update algorithm yields update rules for these parameters which are summarized in Fig. 3 Algorithm1. Algorithm 1: Batch VB updates 1: while ηj,k not converged do 2: for d = 1, · · · , D do 3: while pd,j,k , αd,k not converged do 4: αd,k → α0 + j Rd,j pd,j,k 5: pd,j,k → Algorithm 2: Online VB updates 1: for d = 1, · · · , D do 2: reinitialize pj,k , αk ∀j, k 3: while pj,k , αk not converged do 4: αk → α0 + j Rd,j pj,k 5: pj,k → exp (ψ(ηj,k )−ψ(¯k )) exp ψ(αk ) η η i exp (ψ(ηj,i )−ψ(¯i )) exp ψ(αi ) exp (ψ(ηj,k )−ψ(¯k )) exp ψ(αd,k ) η η i exp (ψ(ηj,i )−ψ(¯i )) exp ψ(αd,i ) 6: end while 7: end for 8: ηj,k = η 0 + 9: end while end while ηj,k → (1 − dt)ηj,k + dt(η 0 + Rd,j pj,k ) 8: end for 6: 7: d Rd,j pd,j,k Figure 3: Here ηk = j ηj,k and ψ(x) is the digamma function so that exp ψ(x) is a smoothed ¯ threshold linear function. Before we move on to the neural network implementation, note that this standard formulation of variational inference for LDA utilizes a batch learning scheme that is not biologically plausible. Fortunately, an online version of this variational algorithm was recently proposed and shown to give 5 superior results when compared to the batch learning algorithm[21]. This algorithm replaces the sum over d in update equation for ηj,k with an incremental update based upon only the most recently observed pattern of spikes. See Fig. 3 Algorithm 2. 4.1 Neural Network Implementation Recall that the goal was to build a neural network that implements the VBEM algorithm for the underlying latent causes of a mixture of spikes using a neural code that represents the posterior distribution via a linear PPC. A linear PPC represents the natural parameters of a posterior distribution via a linear operation on neural activity. Since the primary quantity of interest here is the posterior distribution over odor concentrations, qc (c|α), this means that we need a pattern of activity rα which is linearly related to the αk ’s in the equations above. One way to accomplish this is to simply assume that the firing rates of output neurons are equal to the positive valued αk parameters. Fig. 4 depicts the overall network architecture. Input patterns of activity, R, are transmitted to the synapses of a population of output neurons which represent the αk ’s. The output activity is pooled to ¯ form an un-normalized prediction of the activity of each input neuron, Rj , given the output layer’s current state of belief about the latent causes of the Rj . The activity at each synapse targeted by input neuron j is then inhibited divisively by this prediction. This results in a dendrite that reports to the ¯ soma a quantity, Nj,k , which represents the fraction of unexplained spikes from input neuron j that could be explained by latent cause k. A continuous time dynamical system with this feature and the property that it shares its fixed points with the LDA algorithm is given by d ¯ Nj,k dt d αk dt ¯ ¯ = wj,k Rj − Rj Nj,k = (7) ¯ Nj,k exp (ψ (¯k )) (α0 − αk ) + exp (ψ (αk )) η (8) i ¯ where Rj = k wj,k exp (ψ (αk )), and wj,k = exp (ψ (ηj,k )). Note that, despite its form, it is Eq. 7 which implements the required divisive normalization operation since, in the steady state, ¯ ¯ Nj,k = wj,k Rj /Rj . Regardless, this network has a variety of interesting properties that align well with biology. It predicts that a balance of excitation and inhibition is maintained in the dendrites via divisive normalization and that the role of inhibitory neurons is to predict the input spikes which target individual dendrites. It also predicts superlinear facilitation. Specifically, the final term on the right of Eq. 8 indicates that more active cells will be more sensitive to their dendritic inputs. Alternatively, this could be implemented via recurrent excitation at the population level. In either case, this is the mechanism by which the network implements a sparse prior on topic concentrations and stands in stark contrast to the winner take all mechanisms which rely on competitive mutual inhibition mechanisms. Additionally, the ηj in Eq. 8 represents a cell wide ’leak’ parameter that indicates that the total leak should be ¯ roughly proportional to the sum total weight of the synapses which drive the neuron. This predicts that cells that are highly sensitive to input should also decay back to baseline more quickly. This implementation also predicts Hebbian learning of synaptic weights. To observe this fact, note that the online update rule for the ηj,k parameters can be implemented by simply correlating the activity at ¯ each synapse, Nj,k with activity at the soma αj via the equation: τL d ¯ wj,k = exp (ψ (¯k )) (η0 − 1/2 − wj,k ) + Nj,k exp ψ (αk ) η dt (9) where τL is a long time constant for learning and we have used the fact that exp (ψ (ηjk )) ≈ ηjk −1/2 for x > 1. For a detailed derivation see the supplementary material. 5 Dynamic Document Model LDA is a rather simple generative model that makes several unrealistic assumptions about mixtures of sensory and cortical spikes. In particular, it assumes both that there are no correlations between the 6 Targeted Divisive Normalization Targeted Divisive Normalization αj Ri Input Neurons Recurrent Connections ÷ ÷ -1 -1 Σ μj Nij Ri Synapses Output Neurons Figure 4: The LDA network model. Dendritically targeted inhibition is pooled from the activity of all neurons in the output layer and acts divisively. Σ jj' Nij Input Neurons Synapses Output Neurons Figure 5: DDM network model also includes recurrent connections which target the soma with both a linear excitatory signal and an inhibitory signal that also takes the form of a divisive normalization. intensities of latent causes and that there are no correlations between the intensities of latent causes in temporally adjacent trials or scenes. This makes LDA a rather poor computational model for a task like olfactory foraging which requires the animal to track the rise a fall of odor intensities as it navigates its environment. We can model this more complicated task by replacing the static cause or odor intensity parameters with dynamic odor intensity parameters whose behavior is governed by an exponentiated Ornstein-Uhlenbeck process with drift and diffusion matrices given by (Λ and ΣD ). We call this variant of LDA the Dynamic Document Model (DDM) as it could be used to model smooth changes in the distribution of topics over the course of a single document. 5.1 DDM Model Thus the generative model for the DDM is as follows: 1. For latent cause k = 1, . . . , K, (a) Cause distribution over spikes βk ∼ Dirichlet(η0 ) 2. For scene t = 1, . . . , T , (a) Log intensity of causes c(t) ∼ Normal(Λct−1 , ΣD ) (b) Number of spikes in neuron j resulting from cause k, Nj,k (t) ∼ Poisson(βj,k exp ck (t)) (c) Number of spikes in neuron j, Rj (t) = k Nj,k (t) This model bears many similarities to the Correlated and Dynamic topic models[22], but models dynamics over a short time scale, where the dynamic relationship (Λ, ΣD ) is important. 5.2 Network Implementation Once again the quantity of interest is the current distribution of latent causes, p(c(t)|R(τ ), τ = 0..T ). If no spikes occur then no evidence is presented and posterior inference over c(t) is simply given by an undriven Kalman filter with parameters (Λ, ΣD ). A recurrent neural network which uses a linear PPC to encode a posterior that evolves according to a Kalman filter has the property that neural responses are linearly related to the inverse covariance matrix of the posterior as well as that inverse covariance matrix times the posterior mean. In the absence of evidence, it is easy to show that these quantities must evolve according to recurrent dynamics which implement divisive normalization[10]. Thus, the patterns of neural activity which linearly encode them must do so as well. When a new spike arrives, optimal inference is no longer possible and a variational approximation must be utilized. As is shown in the supplement, this variational approximation is similar to the variational approximation used for LDA. As a result, a network which can divisively inhibit its synapses is able to implement approximate Bayesian inference. Curiously, this implies that the addition of spatial and temporal correlations to the latent causes adds very little complexity to the VB-PPC network implementation of probabilistic inference. All that is required is an additional inhibitory population which targets the somata in the output population. See Fig. 5. 7 Natural Parameters Natural Parameters (α) 0.4 200 450 180 0.3 Network Estimate Network Estimate 500 400 350 300 250 200 150 100 0.1 0 50 100 150 200 250 300 350 400 450 500 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 140 120 0.4 0.3 100 0.2 80 0.1 0 60 40 0.4 20 50 0 0 0.2 160 0 0 0.3 0.2 20 40 60 80 100 120 VBEM Estimate VBEM Estimate 140 160 180 200 0.1 0 Figure 6: (Left) Neural network approximation to the natural parameters of the posterior distribution over topics (the α’s) as a function of the VBEM estimate of those same parameters for a variety of ’documents’. (Center) Same as left, but for the natural parameters of the DDM (i.e the entries of the matrix Σ−1 (t) and Σ−1 µ(t) of the distribution over log topic intensities. (Right) Three example traces for cause intensity in the DDM. Black shows true concentration, blue and red (indistinguishable) show MAP estimates for the network and VBEM algorithms. 6 Experimental Results We compared the PPC neural network implementations of the variational inference with the standard VBEM algorithm. This comparison is necessary because the two algorithms are not guaranteed to converge to the same solution due to the fact that we only required that the neural network dynamics have the same fixed points as the standard VBEM algorithm. As a result, it is possible for the two algorithms to converge to different local minima of the KL divergence. For the network implementation of LDA we find good agreement between the neural network and VBEM estimates of the natural parameters of the posterior. See Fig. 6(left) which shows the two algorithms estimates of the shape parameter of the posterior distribution over topic (odor) concentrations (a quantity which is proportional to the expected concentration). This agreement, however, is not perfect, especially when posterior predicted concentrations are low. In part, this is due to the fact we are presenting the network with difficult inference problems for which the true posterior distribution over topics (odors) is highly correlated and multimodal. As a result, the objective function (KL divergence) is littered with local minima. Additionally, the discrete iterations of the VBEM algorithm can take very large steps in the space of natural parameters while the neural network implementation cannot. In contrast, the network implementation of the DDM is in much better agreement with the VBEM estimation. See Fig. 6(right). This is because the smooth temporal dynamics of the topics eliminate the need for the VBEM algorithm to take large steps. As a result, the smooth network dynamics are better able to accurately track the VBEM algorithms output. For simulation details please see the supplement. 7 Discussion and Conclusion In this work we presented a general framework for inference and learning with linear Probabilistic Population codes. This framework takes advantage of the fact that the Variational Bayesian Expectation Maximization algorithm generates approximate posterior distributions which are in an exponential family form. This is precisely the form needed in order to make probability distributions representable by a linear PPC. We then outlined a general means by which one can build a neural network implementation of the VB algorithm using this kind of neural code. We applied this VB-PPC framework to generate a biologically plausible neural network for spike train demixing. We chose this problem because it has many of the features of the canonical problem faced by nearly every layer of cortex, i.e. that of inferring the latent causes of complex mixtures of spike trains in the layer below. Curiously, this very complicated problem of probabilistic inference and learning ended up having a remarkably simple network solution, requiring only that neurons be capable of implementing divisive normalization via dendritically targeted inhibition and superlinear facilitation. Moreover, we showed that extending this approach to the more complex dynamic case in which latent causes change in intensity over time does not substantially increase the complexity of the neural circuit. Finally, we would like to note that, while we utilized a rate coding scheme for our linear PPC, the basic equations would still apply to any spike based log probability codes such as that considered Beorlin and Deneve[23]. 8 References [1] Daniel Kersten, Pascal Mamassian, and Alan Yuille. Object perception as Bayesian inference. Annual review of psychology, 55:271–304, January 2004. [2] Marc O Ernst and Martin S Banks. Humans integrate visual and haptic information in a statistically optimal fashion. Nature, 415(6870):429–33, 2002. [3] Yair Weiss, Eero P Simoncelli, and Edward H Adelson. Motion illusions as optimal percepts. Nature neuroscience, 5(6):598–604, 2002. [4] P N Sabes. The planning and control of reaching movements. Current opinion in neurobiology, 10(6): 740–6, 2000. o [5] Konrad P K¨ rding and Daniel M Wolpert. Bayesian integration in sensorimotor learning. Nature, 427 (6971):244–7, 2004. [6] Emanuel Todorov. Optimality principles in sensorimotor control. Nature neuroscience, 7(9):907–15, 2004. [7] Erno T´ gl´ s, Edward Vul, Vittorio Girotto, Michel Gonzalez, Joshua B Tenenbaum, and Luca L Bonatti. e a Pure reasoning in 12-month-old infants as probabilistic inference. Science (New York, N.Y.), 332(6033): 1054–9, 2011. [8] W.J. Ma, J.M. Beck, P.E. Latham, and A. Pouget. Bayesian inference with probabilistic population codes. Nature Neuroscience, 2006. [9] Jeffrey M Beck, Wei Ji Ma, Roozbeh Kiani, Tim Hanks, Anne K Churchland, Jamie Roitman, Michael N Shadlen, Peter E Latham, and Alexandre Pouget. Probabilistic population codes for Bayesian decision making. Neuron, 60(6):1142–52, 2008. [10] J. M. Beck, P. E. Latham, and a. Pouget. Marginalization in Neural Circuits with Divisive Normalization. Journal of Neuroscience, 31(43):15310–15319, 2011. [11] Tianming Yang and Michael N Shadlen. Probabilistic reasoning by neurons. Nature, 447(7148):1075–80, 2007. [12] RHS Carpenter and MLL Williams. Neural computation of log likelihood in control of saccadic eye movements. Nature, 1995. [13] Arnulf B a Graf, Adam Kohn, Mehrdad Jazayeri, and J Anthony Movshon. Decoding the activity of neuronal populations in macaque primary visual cortex. Nature neuroscience, 14(2):239–45, 2011. [14] HB Barlow. Pattern Recognition and the Responses of Sensory Neurons. Annals of the New York Academy of Sciences, 1969. [15] Wei Ji Ma, Vidhya Navalpakkam, Jeffrey M Beck, Ronald Van Den Berg, and Alexandre Pouget. Behavior and neural basis of near-optimal visual search. Nature Neuroscience, (May), 2011. [16] DJ Heeger. Normalization of cell responses in cat striate cortex. Visual Neuroscience, 9, 1992. [17] M Carandini, D J Heeger, and J a Movshon. Linearity and normalization in simple cells of the macaque primary visual cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience, 17(21):8621–44, 1997. [18] D. Blei, A. Ng, and M. Jordan. Latent Dirichlet Allocation. JMLR, 2003. [19] M. Beal. Variational Algorithms for Approximate Bayesian Inference. PhD thesis, Gatsby Unit, UCL, 2003. [20] D D Lee and H S Seung. Learning the parts of objects by non-negative matrix factorization. Nature, 401 (6755):788–91, 1999. [21] M. Hoffman, D. Blei, and F. Bach. Online learning for Latent Dirichlet Allocation. In NIPS, 2010. [22] D. Blei and J. Lafferty. Dynamic topic models. In ICML, 2006. [23] M. Boerlin and S. Deneve. Spike-based population coding and working memory. PLOS computational biology, 2011. 9

3 0.76818812 65 nips-2012-Cardinality Restricted Boltzmann Machines

Author: Kevin Swersky, Ilya Sutskever, Daniel Tarlow, Richard S. Zemel, Ruslan Salakhutdinov, Ryan P. Adams

Abstract: The Restricted Boltzmann Machine (RBM) is a popular density model that is also good for extracting features. A main source of tractability in RBM models is that, given an input, the posterior distribution over hidden variables is factorizable and can be easily computed and sampled from. Sparsity and competition in the hidden representation is beneficial, and while an RBM with competition among its hidden units would acquire some of the attractive properties of sparse coding, such constraints are typically not added, as the resulting posterior over the hidden units seemingly becomes intractable. In this paper we show that a dynamic programming algorithm can be used to implement exact sparsity in the RBM’s hidden units. We also show how to pass derivatives through the resulting posterior marginals, which makes it possible to fine-tune a pre-trained neural network with sparse hidden layers. 1

4 0.76719594 113 nips-2012-Efficient and direct estimation of a neural subunit model for sensory coding

Author: Brett Vintch, Andrew Zaharia, J Movshon, Hhmi) Hhmi), Eero P. Simoncelli

Abstract: Many visual and auditory neurons have response properties that are well explained by pooling the rectified responses of a set of spatially shifted linear filters. These filters cannot be estimated using spike-triggered averaging (STA). Subspace methods such as spike-triggered covariance (STC) can recover multiple filters, but require substantial amounts of data, and recover an orthogonal basis for the subspace in which the filters reside rather than the filters themselves. Here, we assume a linear-nonlinear–linear-nonlinear (LN-LN) cascade model in which the first linear stage is a set of shifted (‘convolutional’) copies of a common filter, and the first nonlinear stage consists of rectifying scalar nonlinearities that are identical for all filter outputs. We refer to these initial LN elements as the ‘subunits’ of the receptive field. The second linear stage then computes a weighted sum of the responses of the rectified subunits. We present a method for directly fitting this model to spike data, and apply it to both simulated and real neuronal data from primate V1. The subunit model significantly outperforms STA and STC in terms of cross-validated accuracy and efficiency. 1

5 0.76261389 83 nips-2012-Controlled Recognition Bounds for Visual Learning and Exploration

Author: Vasiliy Karasev, Alessandro Chiuso, Stefano Soatto

Abstract: We describe the tradeoff between the performance in a visual recognition problem and the control authority that the agent can exercise on the sensing process. We focus on the problem of “visual search” of an object in an otherwise known and static scene, propose a measure of control authority, and relate it to the expected risk and its proxy (conditional entropy of the posterior density). We show this analytically, as well as empirically by simulation using the simplest known model that captures the phenomenology of image formation, including scaling and occlusions. We show that a “passive” agent given a training set can provide no guarantees on performance beyond what is afforded by the priors, and that an “omnipotent” agent, capable of infinite control authority, can achieve arbitrarily good performance (asymptotically). In between these limiting cases, the tradeoff can be characterized empirically. 1

6 0.75777149 229 nips-2012-Multimodal Learning with Deep Boltzmann Machines

7 0.75581557 48 nips-2012-Augmented-SVM: Automatic space partitioning for combining multiple non-linear dynamics

8 0.7557981 316 nips-2012-Small-Variance Asymptotics for Exponential Family Dirichlet Process Mixture Models

9 0.7548247 112 nips-2012-Efficient Spike-Coding with Multiplicative Adaptation in a Spike Response Model

10 0.75470662 292 nips-2012-Regularized Off-Policy TD-Learning

11 0.75414723 252 nips-2012-On Multilabel Classification and Ranking with Partial Feedback

12 0.7532742 197 nips-2012-Learning with Recursive Perceptual Representations

13 0.75227845 105 nips-2012-Dynamic Pruning of Factor Graphs for Maximum Marginal Prediction

14 0.75224537 148 nips-2012-Hamming Distance Metric Learning

15 0.75206864 216 nips-2012-Mirror Descent Meets Fixed Share (and feels no regret)

16 0.75114822 355 nips-2012-Truncation-free Online Variational Inference for Bayesian Nonparametric Models

17 0.75114471 186 nips-2012-Learning as MAP Inference in Discrete Graphical Models

18 0.74880505 365 nips-2012-Why MCA? Nonlinear sparse coding with spike-and-slab prior for neurally plausible image encoding

19 0.74705994 260 nips-2012-Online Sum-Product Computation Over Trees

20 0.74697828 302 nips-2012-Scaling MPE Inference for Constrained Continuous Markov Random Fields with Consensus Optimization