nips nips2009 nips2009-13 knowledge-graph by maker-knowledge-mining

13 nips-2009-A Neural Implementation of the Kalman Filter


Source: pdf

Author: Robert Wilson, Leif Finkel

Abstract: Recent experimental evidence suggests that the brain is capable of approximating Bayesian inference in the face of noisy input stimuli. Despite this progress, the neural underpinnings of this computation are still poorly understood. In this paper we focus on the Bayesian filtering of stochastic time series and introduce a novel neural network, derived from a line attractor architecture, whose dynamics map directly onto those of the Kalman filter in the limit of small prediction error. When the prediction error is large we show that the network responds robustly to changepoints in a way that is qualitatively compatible with the optimal Bayesian model. The model suggests ways in which probability distributions are encoded in the brain and makes a number of testable experimental predictions. 1

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 edu Abstract Recent experimental evidence suggests that the brain is capable of approximating Bayesian inference in the face of noisy input stimuli. [sent-4, score-0.15]

2 In this paper we focus on the Bayesian filtering of stochastic time series and introduce a novel neural network, derived from a line attractor architecture, whose dynamics map directly onto those of the Kalman filter in the limit of small prediction error. [sent-6, score-0.603]

3 When the prediction error is large we show that the network responds robustly to changepoints in a way that is qualitatively compatible with the optimal Bayesian model. [sent-7, score-0.433]

4 We develop a neural network whose dynamics can be shown to approximate those of a one-dimensional Kalman filter, the Bayesian model when all the distributions are Gaussian. [sent-14, score-0.424]

5 Where the approximation breaks down, for large prediction errors, the network performs something akin to outlier or change detection and this ‘failure’ suggests ways in which the network can be extended to deal with more complex, non-Gaussian distributions over multiple dimensions. [sent-15, score-0.682]

6 Our approach rests on the modification of the line attractor network of Zhang [26]. [sent-16, score-0.612]

7 In particular, we make three changes to Zhang’s network, modifying the activation rule, the weights and the inputs in such a way that the network’s dynamics map exactly onto those of the Kalman filter when the prediction error is small. [sent-17, score-0.256]

8 Crucially, these modifications result in a network that is no longer a line attractor and thus no longer suffers from many of the limitations of these networks. [sent-18, score-0.612]

9 ˆ ¯ 3 The neural network The network is a modification of Zhang’s line attractor model of head direction cells [26]. [sent-22, score-1.088]

10 We use rate neurons and describe the state of the network at time t with the membrane potential vector, u(t), where each component of u(t) denotes the membrane potential of a single neuron. [sent-23, score-0.623]

11 In discrete time, the update equation is then u(t + 1) = wJf [u(t)] + I(t + 1) (4) where w scales the strength of the weights, J is the connectivity matrix, f [·] is the activation rule that maps membrane potential onto firing rate, and I(t + 1) is the input at time t + 1. [sent-24, score-0.396]

12 When w = 1, γ(t) = 0 and I(t) = 0, the network is a line attractor over a wide range of Kw , σw , c, S and µ, having a continuum of fixed points (as N → ∞). [sent-28, score-0.612]

13 Each fixed point has the same shape, taking the form of a smooth membrane potential profile, U(x) = Jsym f [U(x)], centered at location, x, in the network. [sent-29, score-0.146]

14 When γ(t) = 0, the bump of activity can be made to move over time (without losing its shape) [26] and hence, so long as γ(t) = v(t), implement the prediction step of the Kalman filter (equation 1). [sent-30, score-0.73]

15 That is, if the bump at time t is centered at x(t), i. [sent-31, score-0.48]

16 Thus, in ¯ ˆ x x this configuration, the network can already implement the first step of the Kalman filter through its recurrent connectivity. [sent-36, score-0.419]

17 The next two steps, equations 2 and 3, however, remain inaccessible as the network has no way of encoding uncertainty and it is unclear how it will deal with external inputs. [sent-37, score-0.432]

18 4 Relation to Kalman filter - small prediction error case In this section we outline how the neural network dynamics can be mapped onto those of a Kalman filter. [sent-38, score-0.533]

19 2 Our approach is to analyze the network in terms of U, which, for clarity, we define here to be the fixed point membrane potential profile of the network when w = 1, γ(t) = 0, I(t) = 0, S = S0 and µ = µ0 . [sent-40, score-0.693]

20 We begin by making the assumption that both the input, I(t), and the network membrane potential, u(t), take the form of scaled versions U, with the former encoding the noisy observations, z(t), and the latter encoding the network’s estimate of position, x(t), i. [sent-42, score-0.487]

21 w= S S0 + µ0 I (15) are identical to equations 2 and 3 so long as (a) α(t) ∝ 1 σx (t)2 ˆ (b) A(t) ∝ 1 σz (t)2 (c) µI ∝ σv (t)2 S (16) Thus the network dynamics, when the prediction error is small, map directly onto the Kalman filter equations. [sent-50, score-0.429]

22 1 Implications Reciprocal code for uncertainty in input and estimate Equation 16a provides a link between the strength of activity in the network and the overall uncertainty in the estimate of the Kalman filter, σx (t), with uncertainty decreasing as the activity increases. [sent-54, score-0.844]

23 A similar relation is also implied for the ˆ uncertainty in the observations, σz (t), where equation 16b suggests that this should be reflected in the magnitude of the input, A(t). [sent-55, score-0.13]

24 Code for velocity signal As with Zhang’s line attractor network [26], the mean of the velocity signal, v(t) is encoded into the recurrent connections of the network, with the degree of asymmetry in the weights, γ(t), proportional to the speed. [sent-57, score-0.902]

25 Such hard coding of the velocity signal represents a limitation of the model, as we would like to be able to deal with arbitrary, time varying speeds. [sent-58, score-0.178]

26 However, this kind of change could be implemented by pre-synaptic inhibition [24] or by using a ‘double-ring’ network similar to [25]. [sent-59, score-0.384]

27 Equation 16c implies that the variance of the velocity signal, σv (t), is encoded in the strength of the divisive feedback, µ (assuming constant S). [sent-60, score-0.255]

28 This is very different from Zhang’s model, that has no concept of uncertainty and is also very different from the traditional view of divisive inhibition that sees it as a mechanism for gain control [14, 13]. [sent-61, score-0.232]

29 The network is no longer a line attractor This can be seen by considering the fixed point values of the scale factor, α(t), when the input current, I(t) = 0. [sent-62, score-0.68]

30 This is a key result as it removes all of the constraints required for line attractor dynamics such as infinite precision in the weights and lack of noise in the network and thus the network is much more biologically plausible. [sent-64, score-1.118]

31 2 An example In figure 1 we demonstrate the ability of the network to approximate the dynamics of a onedimensional Kalman filter. [sent-66, score-0.429]

32 The input, shown in figure 1A, is a noiseless bump of current centered 4 B 100 80 80 neuron # 100 neuron # A 60 40 20 40 20 0 C 60 20 40 60 time step 80 100 0 D 60 20 40 60 time step 80 100 0 20 40 60 time step 80 100 2 σx(t) position 50 40 30 1. [sent-67, score-0.903]

33 5 20 0 0 20 40 60 time step 80 100 Figure 2: Response of the network when presented with a noisy moving bump input. [sent-69, score-0.856]

34 The observation noise has standard deviation σz (t) = 5, the speed v(t) = 0. [sent-71, score-0.136]

35 In accordance with equation 16b, the height of each bump is scaled by 1/σz (t)2 . [sent-75, score-0.508]

36 In figure 1B we plot the output activity of the network over time. [sent-76, score-0.454]

37 We assume that the network gets the correct velocity signal, i. [sent-78, score-0.362]

38 As can be seen from the plot, the amount of activity in the network steadily grows from zero over time to an asymptotic value, corresponding to the network’s increasing certainty in its predictions. [sent-85, score-0.456]

39 The position of the bump of activity in the network is also much less jittery than the input bump, reflecting a certain amount of smoothing. [sent-86, score-1.053]

40 In figure 1C we compare the positions of the input bumps (gray dots) with the position of the network bump (black line) and the output of the equivalent Kalman filter (red line). [sent-87, score-0.969]

41 The network clearly tracks the Kalman filter estimate extremely well. [sent-88, score-0.312]

42 The same is true for the network’s estimate of the uncertainty, computed as 1/ α(t) and shown as the black line in figure 1D, which tracks the Kalman filter uncertainty (red line) almost exactly. [sent-89, score-0.218]

43 5 Effect of input noise We now consider the effect of noise on the ability of the network to implement a Kalman filter. [sent-90, score-0.608]

44 In particular we consider noise in the input signal, which for this simple one layer network is equivalent to having noise in the update equation. [sent-91, score-0.569]

45 Specifically, we consider input signals where the only source of noise is in the input current i. [sent-93, score-0.243]

46 there is no additional jitter in the position of the bump as there was in the noiseless case, thus we write I(t) = A(t)U (x(t)) + (t) (18) where (t) is some noise vector. [sent-95, score-0.74]

47 The main effect of the noise is that it perturbs the effective position of the input bump. [sent-96, score-0.315]

48 This can be modeled by extracting the maximum likelihood estimate of the input position given the noisy input and then using this position as the input to the equivalent Kalman filter. [sent-97, score-0.535]

49 Because of the noise, this extracted position is not, in general, the same as the noiseless input position and for zero mean Gaussian noise with covariance Σ, the variance of the perturbation, σz (t), 5 B 10 6 ERMS vs KF A α 4 2 0 8 6 4 2 0 0 0. [sent-98, score-0.497]

50 is approximately given by σz (t) ≈ 1 A(t) 2 U T Σ−1 U (19) Now, for the network to approximate a Kalman filter, equation 16b must hold which means that we require the magnitude of the covariance matrix to scale in proportion to the strength of the input signal, A(t), i. [sent-105, score-0.464]

51 Interestingly this relation is true for Poisson noise, the type of noise that is found all over the brain. [sent-108, score-0.131]

52 In figure 2 we demonstrate the ability of the network to approximate a Kalman filter. [sent-109, score-0.287]

53 In panel A we show the input current which is a moving bump of activity corrupted by independent Gaussian noise of standard deviation σnoise = 0. [sent-110, score-0.802]

54 This is a high noise setting and it is hard to see the bump location by eye. [sent-112, score-0.528]

55 The network dramatically cleans up this input signal (figure 2B) and the output activity, although still noisy, reflects the position of the underlying stimulus much more faithfully than the input. [sent-113, score-0.562]

56 In panel C we compare the position of the output bump in the network (black line) with that of the equivalent Kalman filter. [sent-115, score-0.848]

57 To do this we first fit the noisy input bump at each time to obtain input positions z(t) shown as gray dots. [sent-116, score-0.64]

58 which closely match those of the network (black line). [sent-119, score-0.287]

59 Similarly, there is good agreement between the two estimates of the uncertainty, σx (t), panel D (black line - network, red line - Kalman filter). [sent-120, score-0.209]

60 1 Performance of the network as a function of noise magnitude The noise not only affects the position of the input bump but also, in a slightly more subtle manner, causes a gradual decline in the ability of the network to emulate a Kalman filter. [sent-122, score-1.446]

61 The reason for this (outlined in more detail in the supplementary material) is that the output bump scale factor, α, decreases as a function of the noise variance, σnoise . [sent-123, score-0.587]

62 The average results of simulations on 100 neurons are shown as the red dots, while the black line represents the results of the theory in the supplementary material. [sent-125, score-0.331]

63 The reason for the decline in α as σnoise goes up is that, because of the rectifying non-linearity in the activation rule, increasing σnoise increases the amount of noisy activity in the network. [sent-126, score-0.252]

64 Because of inhibition (both divisive and subtractive) in the network, this ‘noisy activity’ competes with the bump activity and decreases it - thus reducing α. [sent-127, score-0.713]

65 We quantify this difference in figure 3B where we plot the root mean squared error (in units of neural position) between the network and the equivalent Kalman filter as a function of σnoise . [sent-129, score-0.342]

66 As before, the results of simulations are shown as red dots and the theory (outlined in the supplementary material) is the black line. [sent-130, score-0.229]

67 To give some sense for the scale on this plot, the horizontal blue line corresponds to the maximum height of the (noise free) input bump. [sent-131, score-0.186]

68 Thus we may conclude that the performance of the network and the theory are robust up to fairly large values of σnoise . [sent-132, score-0.287]

69 6 Response to changepoints (and outliers) - large prediction error case We now consider the dynamics of the network when the prediction error is large. [sent-134, score-0.621]

70 By large we mean that the prediction error is greater than the width of the bump of activity in the network. [sent-135, score-0.634]

71 a sustained large and abrupt change in the input position at a random time. [sent-138, score-0.24]

72 In the interests of space we focus only on the latter case and such an input, with a changepoint at t = 50, is shown in figure 4A. [sent-139, score-0.167]

73 As before, prior to the change, there is a single bump of activity whose position approximates that of a Kalman filter. [sent-141, score-0.698]

74 However, after the changepoint, the network maintains two bumps of activity for several time steps. [sent-142, score-0.509]

75 One at the original position, that shrinks over time and essentially predicts where the input would be if the change had not occurred, and a second, that grows over time, at the location of the input after the changepoint. [sent-143, score-0.2]

76 Thus in the period immediately after the changepoint, the network can be thought of as encoding two separate and competing hypotheses about the position of the stimulus, one corresponding to the case where no change has occurred, and the other, the case where a change occurred at t = 50. [sent-144, score-0.548]

77 In figure 4C we compare the position of the bump(s) in the network (black dots whose size reflects the size of each bump) to the output from the Kalman filter (red line). [sent-145, score-0.475]

78 Finally, in figure 4D we plot the scale factor, αi (t), of each bump as computed from the simulations (black dots) and from the approximate analytic solution described in the supplementary material (red line for bump at 30, blue line for bump at 80). [sent-148, score-1.579]

79 Thus, when confronted with a changepoint, the network no longer approximates a Kalman filter and instead maintains two competing hypotheses in a way that is qualitatively similar to that of the run-length distribution in [1]. [sent-150, score-0.287]

80 Although this work, at first glance, seems similar to what is presented here, there are several major differences, the main one being that our network is not a line attractor at all, while the results in [8] rely on this property. [sent-155, score-0.612]

81 Combined with divisive normalization, these networks can implement a Kalman filter exactly, while the model presented here can ‘only’ approximate one. [sent-158, score-0.129]

82 While this may seem like a limitation of our network, we see it as an advantage as the breakdown of the approximation leads to a more robust response to outliers and changepoints than a pure Kalman filter. [sent-159, score-0.156]

83 One possible way around this limitation is hinted at by the response of the network in the changepoint case where we saw two, largely independent bumps of activity in the network. [sent-162, score-0.703]

84 This ability to encode multiple ‘particles’ in the network may allow networks of this kind to implement something like the dynamics of a particle filter [12] that can approximate the inference process for non-linear and non-Gaussian systems. [sent-163, score-0.438]

85 3 Experimental predictions The model makes at least two easily testable predictions about the response of head direction cells [21, 22, 23] in rats. [sent-166, score-0.265]

86 The first comes by considering the response of the neurons in the ‘dark’. [sent-167, score-0.119]

87 Assuming that all bearing cues can indeed be eliminated, by setting A(t) = 0 in equation 13, we expect the activity of the neurons to fall off as 1/t and that the shape of the tuning curves will remain approximately constant. [sent-168, score-0.256]

88 Note that this prediction is vastly different from the behaviour of a line attractor, where we would not expect the level of activity to fall off at all in the dark. [sent-169, score-0.297]

89 In particular, one could imagine a training phase, where the position of one landmark is jittered over time, such that each time the rat encounters it it is at a slightly different heading. [sent-171, score-0.247]

90 In the test case, all other, reliable, landmark cues would be removed and the response of head direction cells measured in response to presentation of the unreliable cue alone. [sent-172, score-0.375]

91 The prediction of the model is that this would reduce the strength of the input, A, which in turn reduces the level of activity in the head direction cells, α. [sent-173, score-0.372]

92 This prediction is very different from that of a line attractor which would predict a constant level of activity regardless of the reliability of the landmark cues. [sent-175, score-0.638]

93 8 Conclusions In this paper we have introduced a novel neural network model whose dynamics map directly onto those of a one-dimensional Kalman filter when the prediction error is small. [sent-176, score-0.533]

94 This property is robust to noise and when the prediction error is large, such as for changepoints, the output of the network diverges from that of the Kalman filter, but in a way that is both interesting and useful. [sent-177, score-0.47]

95 Finally, the model makes two easily testable experimental predictions about head direction cells. [sent-178, score-0.151]

96 The contribution of noise to contrast invariance of orientation tuning in cat visual cortex. [sent-195, score-0.138]

97 Optimal sensorimotor integration in recurrent cortical networks: a neural implementation of kalman filters. [sent-239, score-0.641]

98 Head-direction cells recorded from postsubiculum in freely moving rats. [sent-329, score-0.167]

99 Head-direction cells recorded from postsubiculum in freely moving rats. [sent-340, score-0.167]

100 Representation of spatial orientation by the intrinsic dynamics of the head-direction cell ensemble: a theory. [sent-367, score-0.137]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('kalman', 0.548), ('bump', 0.421), ('network', 0.287), ('attractor', 0.241), ('lter', 0.187), ('changepoint', 0.14), ('position', 0.14), ('activity', 0.137), ('dynamics', 0.112), ('noise', 0.107), ('jsym', 0.1), ('membrane', 0.091), ('divisive', 0.09), ('line', 0.084), ('prediction', 0.076), ('velocity', 0.075), ('landmark', 0.075), ('head', 0.071), ('changepoints', 0.07), ('deneve', 0.07), ('recurrent', 0.068), ('input', 0.068), ('latham', 0.068), ('neurons', 0.066), ('inhibition', 0.065), ('gure', 0.062), ('neuroscience', 0.062), ('cells', 0.061), ('kw', 0.06), ('rhs', 0.06), ('ansatz', 0.06), ('jasym', 0.06), ('supplementary', 0.059), ('black', 0.056), ('strength', 0.056), ('uncertainty', 0.053), ('response', 0.053), ('bumps', 0.053), ('equation', 0.053), ('noisy', 0.051), ('neuron', 0.051), ('testable', 0.048), ('dots', 0.048), ('population', 0.043), ('noiseless', 0.042), ('red', 0.041), ('jij', 0.041), ('moving', 0.04), ('postsubiculum', 0.04), ('taube', 0.04), ('wjf', 0.04), ('implement', 0.039), ('signal', 0.038), ('bayesian', 0.035), ('activation', 0.035), ('height', 0.034), ('encoded', 0.034), ('material', 0.034), ('onto', 0.033), ('limitation', 0.033), ('equations', 0.033), ('change', 0.032), ('spiking', 0.032), ('nz', 0.032), ('sym', 0.032), ('direction', 0.032), ('time', 0.032), ('zhang', 0.032), ('brain', 0.031), ('cat', 0.031), ('jitter', 0.03), ('onedimensional', 0.03), ('unreliable', 0.03), ('plot', 0.03), ('external', 0.03), ('deviation', 0.029), ('encoding', 0.029), ('stimulus', 0.029), ('decline', 0.029), ('tracking', 0.028), ('occurred', 0.028), ('potential', 0.028), ('nv', 0.027), ('interests', 0.027), ('muller', 0.027), ('striate', 0.027), ('centered', 0.027), ('pro', 0.027), ('codes', 0.027), ('freely', 0.026), ('neural', 0.025), ('filter', 0.025), ('beck', 0.025), ('reliability', 0.025), ('tracks', 0.025), ('step', 0.025), ('simulations', 0.025), ('cell', 0.025), ('relation', 0.024), ('gain', 0.024)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 1.0000005 13 nips-2009-A Neural Implementation of the Kalman Filter

Author: Robert Wilson, Leif Finkel

Abstract: Recent experimental evidence suggests that the brain is capable of approximating Bayesian inference in the face of noisy input stimuli. Despite this progress, the neural underpinnings of this computation are still poorly understood. In this paper we focus on the Bayesian filtering of stochastic time series and introduce a novel neural network, derived from a line attractor architecture, whose dynamics map directly onto those of the Kalman filter in the limit of small prediction error. When the prediction error is large we show that the network responds robustly to changepoints in a way that is qualitatively compatible with the optimal Bayesian model. The model suggests ways in which probability distributions are encoded in the brain and makes a number of testable experimental predictions. 1

2 0.12834914 121 nips-2009-Know Thy Neighbour: A Normative Theory of Synaptic Depression

Author: Jean-pascal Pfister, Peter Dayan, Máté Lengyel

Abstract: Synapses exhibit an extraordinary degree of short-term malleability, with release probabilities and effective synaptic strengths changing markedly over multiple timescales. From the perspective of a fixed computational operation in a network, this seems like a most unacceptable degree of added variability. We suggest an alternative theory according to which short-term synaptic plasticity plays a normatively-justifiable role. This theory starts from the commonplace observation that the spiking of a neuron is an incomplete, digital, report of the analog quantity that contains all the critical information, namely its membrane potential. We suggest that a synapse solves the inverse problem of estimating the pre-synaptic membrane potential from the spikes it receives, acting as a recursive filter. We show that the dynamics of short-term synaptic depression closely resemble those required for optimal filtering, and that they indeed support high quality estimation. Under this account, the local postsynaptic potential and the level of synaptic resources track the (scaled) mean and variance of the estimated presynaptic membrane potential. We make experimentally testable predictions for how the statistics of subthreshold membrane potential fluctuations and the form of spiking non-linearity should be related to the properties of short-term plasticity in any particular cell type. 1

3 0.1186744 85 nips-2009-Explaining human multiple object tracking as resource-constrained approximate inference in a dynamic probabilistic model

Author: Ed Vul, George Alvarez, Joshua B. Tenenbaum, Michael J. Black

Abstract: Multiple object tracking is a task commonly used to investigate the architecture of human visual attention. Human participants show a distinctive pattern of successes and failures in tracking experiments that is often attributed to limits on an object system, a tracking module, or other specialized cognitive structures. Here we use a computational analysis of the task of object tracking to ask which human failures arise from cognitive limitations and which are consequences of inevitable perceptual uncertainty in the tracking task. We find that many human performance phenomena, measured through novel behavioral experiments, are naturally produced by the operation of our ideal observer model (a Rao-Blackwelized particle filter). The tradeoff between the speed and number of objects being tracked, however, can only arise from the allocation of a flexible cognitive resource, which can be formalized as either memory or attention. 1

4 0.10967527 99 nips-2009-Functional network reorganization in motor cortex can be explained by reward-modulated Hebbian learning

Author: Steven Chase, Andrew Schwartz, Wolfgang Maass, Robert A. Legenstein

Abstract: The control of neuroprosthetic devices from the activity of motor cortex neurons benefits from learning effects where the function of these neurons is adapted to the control task. It was recently shown that tuning properties of neurons in monkey motor cortex are adapted selectively in order to compensate for an erroneous interpretation of their activity. In particular, it was shown that the tuning curves of those neurons whose preferred directions had been misinterpreted changed more than those of other neurons. In this article, we show that the experimentally observed self-tuning properties of the system can be explained on the basis of a simple learning rule. This learning rule utilizes neuronal noise for exploration and performs Hebbian weight updates that are modulated by a global reward signal. In contrast to most previously proposed reward-modulated Hebbian learning rules, this rule does not require extraneous knowledge about what is noise and what is signal. The learning rule is able to optimize the performance of the model system within biologically realistic periods of time and under high noise levels. When the neuronal noise is fitted to experimental data, the model produces learning effects similar to those found in monkey experiments.

5 0.10809349 162 nips-2009-Neural Implementation of Hierarchical Bayesian Inference by Importance Sampling

Author: Lei Shi, Thomas L. Griffiths

Abstract: The goal of perception is to infer the hidden states in the hierarchical process by which sensory data are generated. Human behavior is consistent with the optimal statistical solution to this problem in many tasks, including cue combination and orientation detection. Understanding the neural mechanisms underlying this behavior is of particular importance, since probabilistic computations are notoriously challenging. Here we propose a simple mechanism for Bayesian inference which involves averaging over a few feature detection neurons which fire at a rate determined by their similarity to a sensory stimulus. This mechanism is based on a Monte Carlo method known as importance sampling, commonly used in computer science and statistics. Moreover, a simple extension to recursive importance sampling can be used to perform hierarchical Bayesian inference. We identify a scheme for implementing importance sampling with spiking neurons, and show that this scheme can account for human behavior in cue combination and the oblique effect. 1

6 0.098589137 164 nips-2009-No evidence for active sparsification in the visual cortex

7 0.094221808 215 nips-2009-Sensitivity analysis in HMMs with application to likelihood maximization

8 0.090823472 200 nips-2009-Reconstruction of Sparse Circuits Using Multi-neuronal Excitation (RESCUME)

9 0.087659411 165 nips-2009-Noise Characterization, Modeling, and Reduction for In Vivo Neural Recording

10 0.0843492 19 nips-2009-A joint maximum-entropy model for binary neural population patterns and continuous signals

11 0.082703106 210 nips-2009-STDP enables spiking neurons to detect hidden causes of their inputs

12 0.082543567 168 nips-2009-Non-stationary continuous dynamic Bayesian networks

13 0.081616007 163 nips-2009-Neurometric function analysis of population codes

14 0.077845417 219 nips-2009-Slow, Decorrelated Features for Pretraining Complex Cell-like Networks

15 0.072151624 246 nips-2009-Time-Varying Dynamic Bayesian Networks

16 0.071460813 228 nips-2009-Speeding up Magnetic Resonance Image Acquisition by Bayesian Multi-Slice Adaptive Compressed Sensing

17 0.071334295 43 nips-2009-Bayesian estimation of orientation preference maps

18 0.070082836 38 nips-2009-Augmenting Feature-driven fMRI Analyses: Semi-supervised learning and resting state activity

19 0.069431953 231 nips-2009-Statistical Models of Linear and Nonlinear Contextual Interactions in Early Visual Processing

20 0.068259858 52 nips-2009-Code-specific policy gradient rules for spiking neurons


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, -0.16), (1, -0.151), (2, 0.168), (3, 0.091), (4, 0.073), (5, 0.042), (6, -0.017), (7, 0.008), (8, -0.027), (9, -0.063), (10, -0.005), (11, -0.011), (12, -0.023), (13, -0.006), (14, 0.025), (15, -0.006), (16, -0.079), (17, 0.0), (18, -0.072), (19, 0.061), (20, 0.052), (21, 0.043), (22, 0.099), (23, 0.003), (24, 0.019), (25, 0.052), (26, -0.058), (27, 0.028), (28, -0.02), (29, -0.073), (30, -0.058), (31, -0.064), (32, 0.095), (33, -0.04), (34, -0.072), (35, -0.039), (36, 0.034), (37, -0.034), (38, 0.069), (39, -0.025), (40, -0.085), (41, 0.013), (42, -0.044), (43, -0.01), (44, -0.055), (45, -0.016), (46, 0.063), (47, -0.016), (48, 0.022), (49, -0.111)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.96464813 13 nips-2009-A Neural Implementation of the Kalman Filter

Author: Robert Wilson, Leif Finkel

Abstract: Recent experimental evidence suggests that the brain is capable of approximating Bayesian inference in the face of noisy input stimuli. Despite this progress, the neural underpinnings of this computation are still poorly understood. In this paper we focus on the Bayesian filtering of stochastic time series and introduce a novel neural network, derived from a line attractor architecture, whose dynamics map directly onto those of the Kalman filter in the limit of small prediction error. When the prediction error is large we show that the network responds robustly to changepoints in a way that is qualitatively compatible with the optimal Bayesian model. The model suggests ways in which probability distributions are encoded in the brain and makes a number of testable experimental predictions. 1

2 0.74573505 99 nips-2009-Functional network reorganization in motor cortex can be explained by reward-modulated Hebbian learning

Author: Steven Chase, Andrew Schwartz, Wolfgang Maass, Robert A. Legenstein

Abstract: The control of neuroprosthetic devices from the activity of motor cortex neurons benefits from learning effects where the function of these neurons is adapted to the control task. It was recently shown that tuning properties of neurons in monkey motor cortex are adapted selectively in order to compensate for an erroneous interpretation of their activity. In particular, it was shown that the tuning curves of those neurons whose preferred directions had been misinterpreted changed more than those of other neurons. In this article, we show that the experimentally observed self-tuning properties of the system can be explained on the basis of a simple learning rule. This learning rule utilizes neuronal noise for exploration and performs Hebbian weight updates that are modulated by a global reward signal. In contrast to most previously proposed reward-modulated Hebbian learning rules, this rule does not require extraneous knowledge about what is noise and what is signal. The learning rule is able to optimize the performance of the model system within biologically realistic periods of time and under high noise levels. When the neuronal noise is fitted to experimental data, the model produces learning effects similar to those found in monkey experiments.

3 0.66026652 200 nips-2009-Reconstruction of Sparse Circuits Using Multi-neuronal Excitation (RESCUME)

Author: Tao Hu, Anthony Leonardo, Dmitri B. Chklovskii

Abstract: One of the central problems in neuroscience is reconstructing synaptic connectivity in neural circuits. Synapses onto a neuron can be probed by sequentially stimulating potentially pre-synaptic neurons while monitoring the membrane voltage of the post-synaptic neuron. Reconstructing a large neural circuit using such a “brute force” approach is rather time-consuming and inefficient because the connectivity in neural circuits is sparse. Instead, we propose to measure a post-synaptic neuron’s voltage while stimulating sequentially random subsets of multiple potentially pre-synaptic neurons. To reconstruct these synaptic connections from the recorded voltage we apply a decoding algorithm recently developed for compressive sensing. Compared to the brute force approach, our method promises significant time savings that grow with the size of the circuit. We use computer simulations to find optimal stimulation parameters and explore the feasibility of our reconstruction method under realistic experimental conditions including noise and non-linear synaptic integration. Multineuronal stimulation allows reconstructing synaptic connectivity just from the spiking activity of post-synaptic neurons, even when sub-threshold voltage is unavailable. By using calcium indicators, voltage-sensitive dyes, or multi-electrode arrays one could monitor activity of multiple postsynaptic neurons simultaneously, thus mapping their synaptic inputs in parallel, potentially reconstructing a complete neural circuit. 1 In tro d uc tio n Understanding information processing in neural circuits requires systematic characterization of synaptic connectivity [1, 2]. The most direct way to measure synapses between a pair of neurons is to stimulate potentially pre-synaptic neuron while recording intra-cellularly from the potentially post-synaptic neuron [3-8]. This method can be scaled to reconstruct multiple synaptic connections onto one neuron by combining intracellular recordings from the postsynaptic neuron with photo-activation of pre-synaptic neurons using glutamate uncaging [913] or channelrhodopsin [14, 15], or with multi-electrode arrays [16, 17]. Neurons are sequentially stimulated to fire action potentials by scanning a laser beam (or electrode voltage) over a brain slice, while synaptic weights are measured by recording post-synaptic voltage. Although sequential excitation of single potentially pre-synaptic neurons could reveal connectivity, such a “brute force” approach is inefficient because the connectivity among neurons is sparse. Even among nearby neurons in the cerebral cortex, the probability of connection is only about ten percent [3-8]. Connection probability decays rapidly with the 1 distance between neurons and falls below one percent on the scale of a cortical column [3, 8]. Thus, most single-neuron stimulation trials would result in zero response making the brute force approach slow, especially for larger circuits. Another drawback of the brute force approach is that single-neuron stimulation cannot be combined efficiently with methods allowing parallel recording of neural activity, such as calcium imaging [18-22], voltage-sensitive dyes [23-25] or multi-electrode arrays [17, 26]. As these techniques do not reliably measure sub-threshold potential but report only spiking activity, they would reveal only the strongest connections that can drive a neuron to fire [2730]. Therefore, such combination would reveal only a small fraction of the circuit. We propose to circumvent the above limitations of the brute force approach by stimulating multiple potentially pre-synaptic neurons simultaneously and reconstructing individual connections by using a recently developed method called compressive sensing (CS) [31-35]. In each trial, we stimulate F neurons randomly chosen out of N potentially pre-synaptic neurons and measure post-synaptic activity. Although each measurement yields only a combined response to stimulated neurons, if synaptic inputs sum linearly in a post-synaptic neuron, one can reconstruct the weights of individual connections by using an optimization algorithm. Moreover, if the synaptic connections are sparse, i.e. only K << N potentially pre-synaptic neurons make synaptic connections onto a post-synaptic neuron, the required number of trials M ~ K log(N/K), which is much less than N [31-35]. The proposed method can be used even if only spiking activity is available. Because multiple neurons are driven to fire simultaneously, if several of them synapse on the post-synaptic neuron, they can induce one or more spikes in that neuron. As quantized spike counts carry less information than analog sub-threshold voltage recordings, reconstruction requires a larger number of trials. Yet, the method can be used to reconstruct a complete feedforward circuit from spike recordings. Reconstructing neural circuit with multi-neuronal excitation may be compared with mapping retinal ganglion cell receptive fields. Typically, photoreceptors are stimulated by white-noise checkerboard stimulus and the receptive field is obtained by Reverse Correlation (RC) in case of sub-threshold measurements or Spike-Triggered Average (STA) of the stimulus [36, 37]. Although CS may use the same stimulation protocol, for a limited number of trials, the reconstruction quality is superior to RC or STA. 2 Ma pp i ng sy na pti c inp ut s o nto o n e ne uro n We start by formalizing the problem of mapping synaptic connections from a population of N potentially pre-synaptic neurons onto a single neuron, as exemplified by granule cells synapsing onto a Purkinje cell (Figure 1a). Our experimental protocol can be illustrated using linear algebra formalism, Figure 1b. We represent synaptic weights as components of a column vector x, where zeros represent non-existing connections. Each row in the stimulation matrix A represents a trial, ones indicating neurons driven to spike once and zeros indicating non-spiking neurons. The number of rows in the stimulation matrix A is equal to the number of trials M. The column vector y represents M measurements of membrane voltage obtained by an intra-cellular recording from the post-synaptic neuron: y = Ax. (1) In order to recover individual synaptic weights, Eq. (1) must be solved for x. RC (or STA) solution to this problem is x = (ATA)-1AT y, which minimizes (y-Ax)2 if M>N. In the case M << N for a sparse circuit. In this section we search computationally for the minimum number of trials required for exact reconstruction as a function of the number of non-zero synaptic weights K out of N potentially pre-synaptic neurons. First, note that the number of trials depends on the number of stimulated neurons F. If F = 1 we revert to the brute force approach and the number of measurements is N, while for F = N, the measurements are redundant and no finite number suffices. As the minimum number of measurements is expected to scale as K logN, there must be an optimal F which makes each measurement most informative about x. To determine the optimal number of stimulated neurons F for given K and N, we search for the minimum number of trials M, which allows a perfect reconstruction of the synaptic connectivity x. For each F, we generate 50 synaptic weight vectors and attempt reconstruction from sequentially increasing numbers of trials. The value of M, at which all 50 recoveries are successful (up to computer round-off error), estimates the number of trial needed for reconstruction with probability higher than 98%. By repeating this procedure 50 times for each F, we estimate the mean and standard deviation of M. We find that, for given N and K, the minimum number of trials, M, as a function of the number of stimulated neurons, F, has a shallow minimum. As K decreases, the minimum shifts towards larger F because more neurons should be stimulated simultaneously for sparser x. For the explored range of simulation parameters, the minimum is located close to 0.1N. Next, we set F = 0.1N and explore how the minimum number of measurements required for exact reconstruction depends on K and N. Results of the simulations following the recipe described above are shown in Figure 3a. As expected, when x is sparse, M grows approximately linearly with K (Figure 3b), and logarithmically with N (Figure 3c). N = 1000 180 25 160 20 140 120 15 100 10 80 5 150 250 400 650 1000 Number of potential connections (N) K = 30 220 200 (a) 200 220 Number of measurements (M) 30 Number of measurements (M) Number of actual connections (K) Number of necessary measurements (M) (b) 180 160 140 120 100 80 5 10 15 20 25 30 Number of actual connections (K) 210 (c) 200 190 180 170 160 150 140 130 120 2 10 10 3 Number of potential connections (N) Figure 3: a) Minimum number of measurements M required for reconstruction as a function of the number of actual synapses, K, and the number of potential synapses, N. b) For given N, we find M ~ K. c) For given K, we find M ~ logN (note semi-logarithmic scale in c). 4 R o b ust nes s o f re con st r uc t io n s t o noi se a n d v io la tio n o f si m pli fy in g a ss umpt io n s To make our simulation more realistic we now take into account three possible sources of noise: 1) In reality, post-synaptic voltage on a given synapse varies from trial to trial [4, 5, 46-52], an effect we call synaptic noise. Such noise detrimentally affects reconstructions because each row of A is multiplied by a different instantiation of vector x. 2) Stimulation of neurons may be imprecise exciting a slightly different subset of neurons than intended and/or firing intended neurons multiple times. We call this effect stimulation noise. Such noise detrimentally affects reconstructions because, in its presence, the actual measurement matrix A is different from the one used for recovery. 3) A synapse may fail to release neurotransmitter with some probability. Naturally, in the presence of noise, reconstructions cannot be exact. We quantify the 4 reconstruction x − xr l2 = error ∑ N i =1 by the normalized x − xr l2–error l2 / xl , where 2 ( xi − xri ) 2 . We plot normalized reconstruction error in brute force approach (M = N = 500 trials) as a function of noise, as well as CS and RC reconstruction errors (M = 200, 600 trials), Figure 4. 2 0.9 Normalized reconstruction error ||x-x|| /||x|| 1 r 2 For each noise source, the reconstruction error of the brute force approach can be achieved with 60% fewer trials by CS method for the above parameters (Figure 4). For the same number of trials, RC method performs worse. Naturally, the reconstruction error decreases with the number of trials. The reconstruction error is most sensitive to stimulation noise and least sensitive to synaptic noise. 1 (a) 1 (b) 0.9 0.8 (c) 0.9 0.8 0.8 0.7 0.7 0.6 0.6 0.5 0.5 0.4 0.4 0.4 0.3 0.3 0.3 0.2 0.2 0.2 0.1 0.1 0.1 0 0 0.7 RC: M=200 RC: M=600 Brute force method: M=500 CS: M=200 CS: M=600 0.6 0.5 0 0.05 0.1 0.15 Synaptic noise level 0.2 0.25 0 0.05 0.1 0.15 Stimulation noise level 0.2 0.25 0 0 0.05 0.1 0.15 0.2 0.25 Synaptic failure probability Figure 4: Impact of noise on the reconstruction quality for N = 500, K = 30, F = 50. a) Recovery error due to trial-to-trial variation in synaptic weight. The response y is calculated using the synaptic connectivity x perturbed by an additive Gaussian noise. The noise level is given by the coefficient of variation of synaptic weight. b) Recovery error due to stimulation noise. The matrix A used for recovery is obtained from the binary matrix used to calculate the measurement vector y by shifting, in each row, a fraction of ones specified by the noise level to random positions. c) Recovery error due to synaptic failures. The detrimental effect of the stimulation noise on the reconstruction can be eliminated by monitoring spiking activity of potentially pre-synaptic neurons. By using calcium imaging [18-22], voltage-sensitive dyes [23] or multi-electrode arrays [17, 26] one could record the actual stimulation matrix. Because most random matrices satisfy the reconstruction requirements [31, 34, 35], the actual stimulation matrix can be used for a successful recovery instead of the intended one. If neuronal activity can be monitored reliably, experiments can be done in a different mode altogether. Instead of stimulating designated neurons with high fidelity by using highly localized and intense light, one could stimulate all neurons with low probability. Random firing events can be detected and used in the recovery process. The light intensity can be tuned to stimulate the optimal number of neurons per trial. Next, we explore the sensitivity of the proposed reconstruction method to the violation of simplifying assumptions. First, whereas our simulation assumes that the actual number of connections, K, is known, in reality, connectivity sparseness is known a priori only approximately. Will this affect reconstruction results? In principle, CS does not require prior knowledge of K for reconstruction [31, 34, 35]. For the CoSaMP algorithm, however, it is important to provide value K larger than the actual value (Figure 5a). Then, the algorithm will find all the actual synaptic weights plus some extra non-zero weights, negligibly small when compared to actual ones. Thus, one can provide the algorithm with the value of K safely larger than the actual one and then threshold the reconstruction result according to the synaptic noise level. Second, whereas we assumed a linear summation of inputs [53], synaptic integration may be 2 non-linear [54]. We model non-linearity by setting y = yl + α yl , where yl represents linearly summed synaptic inputs. Results of simulations (Figure 5b) show that although nonlinearity can significantly degrade CS reconstruction quality, it still performs better than RC. 5 (b) 0.45 0.4 Actual K = 30 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 10 20 30 40 50 60 Normalized reconstruction error ||x-x||2/||x|| r 2 Normalized reconstrcution error ||x-x||2/||x|| r 2 (a) 0.9 0.8 0.7 CS RC 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.15-0.12-0.09-0.06-0.03 0 0.03 0.06 0.09 0.12 0.15 Relative strength of the non-linear term α × mean(y ) K fed to CoSaMP l Figure 5: Sensitivity of reconstruction error to the violation of simplifying assumptions for N = 500, K = 30, M = 200, F = 50. a) The quality of the reconstruction is not affected if the CoSaMP algorithm is fed with the value of K larger than actual. b) Reconstruction error computed in 100 realizations for each value of the quadratic term relative to the linear term. 5 Ma pp i ng sy na pti c inp ut s o nto a n e uro na l po pu la tio n Until now, we considered reconstruction of synaptic inputs onto one neuron using subthreshold measurements of its membrane potential. In this section, we apply CS to reconstructing synaptic connections onto a population of potentially post-synaptic neurons. Because in CS the choice of stimulated neurons is non-adaptive, by recording from all potentially post-synaptic neurons in response to one sequence of trials one can reconstruct a complete feedforward network (Figure 6). (a) x p(y=1) A Normalized reconstruction error ||x/||x||2−xr/||xr||2||2 y (b) (c) 100 (d) 500 700 900 1100 Number of spikes 1 STA CS 0.8 0.6 0.4 0.2 0 1000 0 300 3000 5000 7000 9000 Number of trials (M) Ax Figure 6: Mapping of a complete feedforward network. a) Each post-synaptic neuron (red) receives synapses from a sparse subset of potentially pre-synaptic neurons (blue). b) Linear algebra representation of the experimental protocol. c) Probability of firing as a function of synaptic current. d) Comparison of CS and STA reconstruction error using spike trains for N = 500, K = 30 and F = 50. Although attractive, such parallelization raises several issues. First, patching a large number of neurons is unrealistic and, therefore, monitoring membrane potential requires using different methods, such as calcium imaging [18-22], voltage sensitive dyes [23-25] or multielectrode arrays [17, 26]. As these methods can report reliably only spiking activity, the measurement is not analog but discrete. Depending on the strength of summed synaptic inputs compared to the firing threshold, the postsynaptic neuron may be silent, fire once or multiple times. As a result, the measured response y is quantized by the integer number of spikes. Such quantized measurements are less informative than analog measurements of the sub-threshold membrane potential. In the extreme case of only two quantization levels, spike or no spike, each measurement contains only 1 bit of information. Therefore, to achieve reasonable reconstruction quality using quantized measurements, a larger number of trials M>>N is required. We simulate circuit reconstruction from spike recordings in silico as follows. First, we draw synaptic weights from an experimentally motivated distribution. Second, we generate a 6 random stimulation matrix and calculate the product Ax. Third, we linear half-wave rectify this product and use the result as the instantaneous firing rate for the Poisson spike generator (Figure 6c). We used a rectifying threshold that results in 10% of spiking trials as typically observed in experiments. Fourth, we reconstruct synaptic weights using STA and CS and compare the results with the generated weights. We calculated mean error over 100 realizations of the simulation protocol (Figure 6d). Due to the non-linear spike generating procedure, x can be recovered only up to a scaling factor. We propose to calibrate x with a few brute-force measurements of synaptic weights. Thus, in calculating the reconstruction error using l2 norm, we normalize both the generated and recovered synaptic weights. Such definition is equivalent to the angular error, which is often used to evaluate the performance of STA in mapping receptive field [37, 55]. Why is CS superior to STA for a given number of trials (Figure 6d)? Note that spikeless trials, which typically constitute a majority, also carry information about connectivity. While STA discards these trials, CS takes them into account. In particular, CoSaMP starts with the STA solution as zeroth iteration and improves on it by using the results of all trials and the sparseness prior. 6 D i s c uss ion We have demonstrated that sparse feedforward networks can be reconstructed by stimulating multiple potentially pre-synaptic neurons simultaneously and monitoring either subthreshold or spiking response of potentially post-synaptic neurons. When sub-threshold voltage is recorded, significantly fewer measurements are required than in the brute force approach. Although our method is sensitive to noise (with stimulation noise worse than synapse noise), it is no less robust than the brute force approach or RC. The proposed reconstruction method can also recover inputs onto a neuron from spike counts, albeit with more trials than from sub-threshold potential measurements. This is particularly useful when intra-cellular recordings are not feasible and only spiking can be detected reliably, for example, when mapping synaptic inputs onto multiple neurons in parallel. For a given number of trials, our method yields smaller error than STA. The proposed reconstruction method assumes linear summation of synaptic inputs (both excitatory and inhibitory) and is sensitive to non-linearity of synaptic integration. Therefore, it is most useful for studying connections onto neurons, in which synaptic integration is close to linear. On the other hand, multi-neuron stimulation is closer than single-neuron stimulation to the intrinsic activity in the live brain and can be used to study synaptic integration under realistic conditions. In contrast to circuit reconstruction using intrinsic neuronal activity [56, 57], our method relies on extrinsic stimulation of neurons. Can our method use intrinsic neuronal activity instead? We see two major drawbacks of such approach. First, activity of non-monitored presynaptic neurons may significantly distort reconstruction results. Thus, successful reconstruction would require monitoring all active pre-synaptic neurons, which is rather challenging. Second, reliable reconstruction is possible only when the activity of presynaptic neurons is uncorrelated. Yet, their activity may be correlated, for example, due to common input. We thank Ashok Veeraraghavan for introducing us to CS, Anthony Leonardo for making a retina dataset available for the analysis, Lou Scheffer and Hong Young Noh for commenting on the manuscript and anonymous reviewers for helpful suggestions. References [1] Luo, L., Callaway, E.M. & Svoboda, K. (2008) Genetic dissection of neural circuits. Neuron 57(5):634-660. [2] Helmstaedter, M., Briggman, K.L. & Denk, W. (2008) 3D structural imaging of the brain with photons and electrons. Current opinion in neurobiology 18(6):633-641. [3] Holmgren, C., Harkany, T., Svennenfors, B. & Zilberter, Y. (2003) Pyramidal cell communication within local networks in layer 2/3 of rat neocortex. Journal of Physiology 551:139-153. [4] Markram, H. (1997) A network of tufted layer 5 pyramidal neurons. Cerebral Cortex 7(6):523-533. 7 [5] Markram, H., Lubke, J., Frotscher, M., Roth, A. & Sakmann, B. (1997) Physiology and anatomy of synaptic connections between thick tufted pyramidal neurones in the developing rat neocortex. Journal of Physiology 500(2):409-440. [6] Thomson, A.M. & Bannister, A.P. (2003) Interlaminar connections in the neocortex. Cerebral Cortex 13(1):5-14. [7] Thomson, A.M., West, D.C., Wang, Y. & Bannister, A.P. (2002) Synaptic connections and small circuits involving excitatory and inhibitory neurons in layers 2-5 of adult rat and cat neocortex: triple intracellular recordings and biocytin labelling in vitro. Cerebral Cortex 12(9):936-953. [8] Song, S., Sjostrom, P.J., Reigl, M., Nelson, S. & Chklovskii, D.B. (2005) Highly nonrandom features of synaptic connectivity in local cortical circuits. Plos Biology 3(3):e68. [9] Callaway, E.M. & Katz, L.C. (1993) Photostimulation using caged glutamate reveals functional circuitry in living brain slices. Proceedings of the National Academy of Sciences of the United States of America 90(16):7661-7665. [10] Dantzker, J.L. & Callaway, E.M. (2000) Laminar sources of synaptic input to cortical inhibitory interneurons and pyramidal neurons. Nature Neuroscience 3(7):701-707. [11] Shepherd, G.M. & Svoboda, K. (2005) Laminar and columnar organization of ascending excitatory projections to layer 2/3 pyramidal neurons in rat barrel cortex. Journal of Neuroscience 25(24):5670-5679. [12] Nikolenko, V., Poskanzer, K.E. & Yuste, R. (2007) Two-photon photostimulation and imaging of neural circuits. Nature Methods 4(11):943-950. [13] Shoham, S., O'connor, D.H., Sarkisov, D.V. & Wang, S.S. (2005) Rapid neurotransmitter uncaging in spatially defined patterns. Nature Methods 2(11):837-843. [14] Gradinaru, V., Thompson, K.R., Zhang, F., Mogri, M., Kay, K., Schneider, M.B. & Deisseroth, K. (2007) Targeting and readout strategies for fast optical neural control in vitro and in vivo. Journal of Neuroscience 27(52):14231-14238. [15] Petreanu, L., Huber, D., Sobczyk, A. & Svoboda, K. (2007) Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nature Neuroscience 10(5):663-668. [16] Na, L., Watson, B.O., Maclean, J.N., Yuste, R. & Shepard, K.L. (2008) A 256×256 CMOS Microelectrode Array for Extracellular Neural Stimulation of Acute Brain Slices. Solid-State Circuits Conference, 2008. ISSCC 2008. Digest of Technical Papers. IEEE International. [17] Fujisawa, S., Amarasingham, A., Harrison, M.T. & Buzsaki, G. (2008) Behavior-dependent shortterm assembly dynamics in the medial prefrontal cortex. Nature Neuroscience 11(7):823-833. [18] Ikegaya, Y., Aaron, G., Cossart, R., Aronov, D., Lampl, I., Ferster, D. & Yuste, R. (2004) Synfire chains and cortical songs: temporal modules of cortical activity. Science 304(5670):559-564. [19] Ohki, K., Chung, S., Ch'ng, Y.H., Kara, P. & Reid, R.C. (2005) Functional imaging with cellular resolution reveals precise micro-architecture in visual cortex. Nature 433(7026):597-603. [20] Stosiek, C., Garaschuk, O., Holthoff, K. & Konnerth, A. (2003) In vivo two-photon calcium imaging of neuronal networks. Proceedings of the National Academy of Sciences of the United States of America 100(12):7319-7324. [21] Svoboda, K., Denk, W., Kleinfeld, D. & Tank, D.W. (1997) In vivo dendritic calcium dynamics in neocortical pyramidal neurons. Nature 385(6612):161-165. [22] Sasaki, T., Minamisawa, G., Takahashi, N., Matsuki, N. & Ikegaya, Y. (2009) Reverse optical trawling for synaptic connections in situ. Journal of Neurophysiology 102(1):636-643. [23] Zecevic, D., Djurisic, M., Cohen, L.B., Antic, S., Wachowiak, M., Falk, C.X. & Zochowski, M.R. (2003) Imaging nervous system activity with voltage-sensitive dyes. Current Protocols in Neuroscience Chapter 6:Unit 6.17. [24] Cacciatore, T.W., Brodfuehrer, P.D., Gonzalez, J.E., Jiang, T., Adams, S.R., Tsien, R.Y., Kristan, W.B., Jr. & Kleinfeld, D. (1999) Identification of neural circuits by imaging coherent electrical activity with FRET-based dyes. Neuron 23(3):449-459. [25] Taylor, A.L., Cottrell, G.W., Kleinfeld, D. & Kristan, W.B., Jr. (2003) Imaging reveals synaptic targets of a swim-terminating neuron in the leech CNS. Journal of Neuroscience 23(36):11402-11410. [26] Hutzler, M., Lambacher, A., Eversmann, B., Jenkner, M., Thewes, R. & Fromherz, P. (2006) High-resolution multitransistor array recording of electrical field potentials in cultured brain slices. Journal of Neurophysiology 96(3):1638-1645. [27] Egger, V., Feldmeyer, D. & Sakmann, B. (1999) Coincidence detection and changes of synaptic efficacy in spiny stellate neurons in rat barrel cortex. Nature Neuroscience 2(12):1098-1105. [28] Feldmeyer, D., Egger, V., Lubke, J. & Sakmann, B. (1999) Reliable synaptic connections between pairs of excitatory layer 4 neurones within a single 'barrel' of developing rat somatosensory cortex. Journal of Physiology 521:169-190. [29] Peterlin, Z.A., Kozloski, J., Mao, B.Q., Tsiola, A. & Yuste, R. (2000) Optical probing of neuronal circuits with calcium indicators. Proceedings of the National Academy of Sciences of the United States of America 97(7):3619-3624. 8 [30] Thomson, A.M., Deuchars, J. & West, D.C. (1993) Large, deep layer pyramid-pyramid single axon EPSPs in slices of rat motor cortex display paired pulse and frequency-dependent depression, mediated presynaptically and self-facilitation, mediated postsynaptically. Journal of Neurophysiology 70(6):2354-2369. [31] Baraniuk, R.G. (2007) Compressive sensing. Ieee Signal Processing Magazine 24(4):118-120. [32] Candes, E.J. (2008) Compressed Sensing. Twenty-Second Annual Conference on Neural Information Processing Systems, Tutorials. [33] Candes, E.J., Romberg, J.K. & Tao, T. (2006) Stable signal recovery from incomplete and inaccurate measurements. Communications on Pure and Applied Mathematics 59(8):1207-1223. [34] Candes, E.J. & Tao, T. (2006) Near-optimal signal recovery from random projections: Universal encoding strategies? Ieee Transactions on Information Theory 52(12):5406-5425. [35] Donoho, D.L. (2006) Compressed sensing. Ieee Transactions on Information Theory 52(4):12891306. [36] Ringach, D. & Shapley, R. (2004) Reverse Correlation in Neurophysiology. Cognitive Science 28:147-166. [37] Schwartz, O., Pillow, J.W., Rust, N.C. & Simoncelli, E.P. (2006) Spike-triggered neural characterization. Journal of Vision 6(4):484-507. [38] Candes, E.J. & Tao, T. (2005) Decoding by linear programming. Ieee Transactions on Information Theory 51(12):4203-4215. [39] Needell, D. & Vershynin, R. (2009) Uniform Uncertainty Principle and Signal Recovery via Regularized Orthogonal Matching Pursuit. Foundations of Computational Mathematics 9(3):317-334. [40] Tropp, J.A. & Gilbert, A.C. (2007) Signal recovery from random measurements via orthogonal matching pursuit. Ieee Transactions on Information Theory 53(12):4655-4666. [41] Dai, W. & Milenkovic, O. (2009) Subspace Pursuit for Compressive Sensing Signal Reconstruction. Ieee Transactions on Information Theory 55(5):2230-2249. [42] Needell, D. & Tropp, J.A. (2009) CoSaMP: Iterative signal recovery from incomplete and inaccurate samples. Applied and Computational Harmonic Analysis 26(3):301-321. [43] Varshney, L.R., Sjostrom, P.J. & Chklovskii, D.B. (2006) Optimal information storage in noisy synapses under resource constraints. Neuron 52(3):409-423. [44] Brunel, N., Hakim, V., Isope, P., Nadal, J.P. & Barbour, B. (2004) Optimal information storage and the distribution of synaptic weights: perceptron versus Purkinje cell. Neuron 43(5):745-757. [45] Napoletani, D. & Sauer, T.D. (2008) Reconstructing the topology of sparsely connected dynamical networks. Physical Review E 77(2):026103. [46] Allen, C. & Stevens, C.F. (1994) An evaluation of causes for unreliability of synaptic transmission. Proceedings of the National Academy of Sciences of the United States of America 91(22):10380-10383. [47] Hessler, N.A., Shirke, A.M. & Malinow, R. (1993) The probability of transmitter release at a mammalian central synapse. Nature 366(6455):569-572. [48] Isope, P. & Barbour, B. (2002) Properties of unitary granule cell-->Purkinje cell synapses in adult rat cerebellar slices. Journal of Neuroscience 22(22):9668-9678. [49] Mason, A., Nicoll, A. & Stratford, K. (1991) Synaptic transmission between individual pyramidal neurons of the rat visual cortex in vitro. Journal of Neuroscience 11(1):72-84. [50] Raastad, M., Storm, J.F. & Andersen, P. (1992) Putative Single Quantum and Single Fibre Excitatory Postsynaptic Currents Show Similar Amplitude Range and Variability in Rat Hippocampal Slices. European Journal of Neuroscience 4(1):113-117. [51] Rosenmund, C., Clements, J.D. & Westbrook, G.L. (1993) Nonuniform probability of glutamate release at a hippocampal synapse. Science 262(5134):754-757. [52] Sayer, R.J., Friedlander, M.J. & Redman, S.J. (1990) The time course and amplitude of EPSPs evoked at synapses between pairs of CA3/CA1 neurons in the hippocampal slice. Journal of Neuroscience 10(3):826-836. [53] Cash, S. & Yuste, R. (1999) Linear summation of excitatory inputs by CA1 pyramidal neurons. Neuron 22(2):383-394. [54] Polsky, A., Mel, B.W. & Schiller, J. (2004) Computational subunits in thin dendrites of pyramidal cells. Nature Neuroscience 7(6):621-627. [55] Paninski, L. (2003) Convergence properties of three spike-triggered analysis techniques. Network: Computation in Neural Systems 14(3):437-464. [56] Okatan, M., Wilson, M.A. & Brown, E.N. (2005) Analyzing functional connectivity using a network likelihood model of ensemble neural spiking activity. Neural Computation 17(9):1927-1961. [57] Timme, M. (2007) Revealing network connectivity from response dynamics. Physical Review Letters 98(22):224101. 9

4 0.64094162 121 nips-2009-Know Thy Neighbour: A Normative Theory of Synaptic Depression

Author: Jean-pascal Pfister, Peter Dayan, Máté Lengyel

Abstract: Synapses exhibit an extraordinary degree of short-term malleability, with release probabilities and effective synaptic strengths changing markedly over multiple timescales. From the perspective of a fixed computational operation in a network, this seems like a most unacceptable degree of added variability. We suggest an alternative theory according to which short-term synaptic plasticity plays a normatively-justifiable role. This theory starts from the commonplace observation that the spiking of a neuron is an incomplete, digital, report of the analog quantity that contains all the critical information, namely its membrane potential. We suggest that a synapse solves the inverse problem of estimating the pre-synaptic membrane potential from the spikes it receives, acting as a recursive filter. We show that the dynamics of short-term synaptic depression closely resemble those required for optimal filtering, and that they indeed support high quality estimation. Under this account, the local postsynaptic potential and the level of synaptic resources track the (scaled) mean and variance of the estimated presynaptic membrane potential. We make experimentally testable predictions for how the statistics of subthreshold membrane potential fluctuations and the form of spiking non-linearity should be related to the properties of short-term plasticity in any particular cell type. 1

5 0.63751704 162 nips-2009-Neural Implementation of Hierarchical Bayesian Inference by Importance Sampling

Author: Lei Shi, Thomas L. Griffiths

Abstract: The goal of perception is to infer the hidden states in the hierarchical process by which sensory data are generated. Human behavior is consistent with the optimal statistical solution to this problem in many tasks, including cue combination and orientation detection. Understanding the neural mechanisms underlying this behavior is of particular importance, since probabilistic computations are notoriously challenging. Here we propose a simple mechanism for Bayesian inference which involves averaging over a few feature detection neurons which fire at a rate determined by their similarity to a sensory stimulus. This mechanism is based on a Monte Carlo method known as importance sampling, commonly used in computer science and statistics. Moreover, a simple extension to recursive importance sampling can be used to perform hierarchical Bayesian inference. We identify a scheme for implementing importance sampling with spiking neurons, and show that this scheme can account for human behavior in cue combination and the oblique effect. 1

6 0.60428756 246 nips-2009-Time-Varying Dynamic Bayesian Networks

7 0.55793029 210 nips-2009-STDP enables spiking neurons to detect hidden causes of their inputs

8 0.53635871 165 nips-2009-Noise Characterization, Modeling, and Reduction for In Vivo Neural Recording

9 0.5202592 168 nips-2009-Non-stationary continuous dynamic Bayesian networks

10 0.47743875 203 nips-2009-Replacing supervised classification learning by Slow Feature Analysis in spiking neural networks

11 0.47346529 231 nips-2009-Statistical Models of Linear and Nonlinear Contextual Interactions in Early Visual Processing

12 0.46950501 164 nips-2009-No evidence for active sparsification in the visual cortex

13 0.43572339 38 nips-2009-Augmenting Feature-driven fMRI Analyses: Semi-supervised learning and resting state activity

14 0.42407492 194 nips-2009-Predicting the Optimal Spacing of Study: A Multiscale Context Model of Memory

15 0.41813102 54 nips-2009-Compositionality of optimal control laws

16 0.41374162 188 nips-2009-Perceptual Multistability as Markov Chain Monte Carlo Inference

17 0.3970609 228 nips-2009-Speeding up Magnetic Resonance Image Acquisition by Bayesian Multi-Slice Adaptive Compressed Sensing

18 0.3936688 219 nips-2009-Slow, Decorrelated Features for Pretraining Complex Cell-like Networks

19 0.37016451 85 nips-2009-Explaining human multiple object tracking as resource-constrained approximate inference in a dynamic probabilistic model

20 0.36814782 52 nips-2009-Code-specific policy gradient rules for spiking neurons


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(7, 0.019), (24, 0.025), (25, 0.067), (35, 0.057), (36, 0.068), (39, 0.059), (58, 0.085), (61, 0.013), (62, 0.299), (71, 0.051), (81, 0.049), (86, 0.079), (91, 0.028)]

similar papers list:

simIndex simValue paperId paperTitle

1 0.82162607 164 nips-2009-No evidence for active sparsification in the visual cortex

Author: Pietro Berkes, Ben White, Jozsef Fiser

Abstract: The proposal that cortical activity in the visual cortex is optimized for sparse neural activity is one of the most established ideas in computational neuroscience. However, direct experimental evidence for optimal sparse coding remains inconclusive, mostly due to the lack of reference values on which to judge the measured sparseness. Here we analyze neural responses to natural movies in the primary visual cortex of ferrets at different stages of development and of rats while awake and under different levels of anesthesia. In contrast with prediction from a sparse coding model, our data shows that population and lifetime sparseness decrease with visual experience, and increase from the awake to anesthetized state. These results suggest that the representation in the primary visual cortex is not actively optimized to maximize sparseness. 1

same-paper 2 0.78988272 13 nips-2009-A Neural Implementation of the Kalman Filter

Author: Robert Wilson, Leif Finkel

Abstract: Recent experimental evidence suggests that the brain is capable of approximating Bayesian inference in the face of noisy input stimuli. Despite this progress, the neural underpinnings of this computation are still poorly understood. In this paper we focus on the Bayesian filtering of stochastic time series and introduce a novel neural network, derived from a line attractor architecture, whose dynamics map directly onto those of the Kalman filter in the limit of small prediction error. When the prediction error is large we show that the network responds robustly to changepoints in a way that is qualitatively compatible with the optimal Bayesian model. The model suggests ways in which probability distributions are encoded in the brain and makes a number of testable experimental predictions. 1

3 0.65588129 41 nips-2009-Bayesian Source Localization with the Multivariate Laplace Prior

Author: Marcel V. Gerven, Botond Cseke, Robert Oostenveld, Tom Heskes

Abstract: We introduce a novel multivariate Laplace (MVL) distribution as a sparsity promoting prior for Bayesian source localization that allows the specification of constraints between and within sources. We represent the MVL distribution as a scale mixture that induces a coupling between source variances instead of their means. Approximation of the posterior marginals using expectation propagation is shown to be very efficient due to properties of the scale mixture representation. The computational bottleneck amounts to computing the diagonal elements of a sparse matrix inverse. Our approach is illustrated using a mismatch negativity paradigm for which MEG data and a structural MRI have been acquired. We show that spatial coupling leads to sources which are active over larger cortical areas as compared with an uncoupled prior. 1

4 0.56920892 99 nips-2009-Functional network reorganization in motor cortex can be explained by reward-modulated Hebbian learning

Author: Steven Chase, Andrew Schwartz, Wolfgang Maass, Robert A. Legenstein

Abstract: The control of neuroprosthetic devices from the activity of motor cortex neurons benefits from learning effects where the function of these neurons is adapted to the control task. It was recently shown that tuning properties of neurons in monkey motor cortex are adapted selectively in order to compensate for an erroneous interpretation of their activity. In particular, it was shown that the tuning curves of those neurons whose preferred directions had been misinterpreted changed more than those of other neurons. In this article, we show that the experimentally observed self-tuning properties of the system can be explained on the basis of a simple learning rule. This learning rule utilizes neuronal noise for exploration and performs Hebbian weight updates that are modulated by a global reward signal. In contrast to most previously proposed reward-modulated Hebbian learning rules, this rule does not require extraneous knowledge about what is noise and what is signal. The learning rule is able to optimize the performance of the model system within biologically realistic periods of time and under high noise levels. When the neuronal noise is fitted to experimental data, the model produces learning effects similar to those found in monkey experiments.

5 0.5543744 162 nips-2009-Neural Implementation of Hierarchical Bayesian Inference by Importance Sampling

Author: Lei Shi, Thomas L. Griffiths

Abstract: The goal of perception is to infer the hidden states in the hierarchical process by which sensory data are generated. Human behavior is consistent with the optimal statistical solution to this problem in many tasks, including cue combination and orientation detection. Understanding the neural mechanisms underlying this behavior is of particular importance, since probabilistic computations are notoriously challenging. Here we propose a simple mechanism for Bayesian inference which involves averaging over a few feature detection neurons which fire at a rate determined by their similarity to a sensory stimulus. This mechanism is based on a Monte Carlo method known as importance sampling, commonly used in computer science and statistics. Moreover, a simple extension to recursive importance sampling can be used to perform hierarchical Bayesian inference. We identify a scheme for implementing importance sampling with spiking neurons, and show that this scheme can account for human behavior in cue combination and the oblique effect. 1

6 0.55410606 19 nips-2009-A joint maximum-entropy model for binary neural population patterns and continuous signals

7 0.53510749 62 nips-2009-Correlation Coefficients are Insufficient for Analyzing Spike Count Dependencies

8 0.53027451 210 nips-2009-STDP enables spiking neurons to detect hidden causes of their inputs

9 0.51412243 155 nips-2009-Modelling Relational Data using Bayesian Clustered Tensor Factorization

10 0.51250267 163 nips-2009-Neurometric function analysis of population codes

11 0.51167691 9 nips-2009-A Game-Theoretic Approach to Hypergraph Clustering

12 0.5069977 158 nips-2009-Multi-Label Prediction via Sparse Infinite CCA

13 0.50504583 70 nips-2009-Discriminative Network Models of Schizophrenia

14 0.502406 17 nips-2009-A Sparse Non-Parametric Approach for Single Channel Separation of Known Sounds

15 0.5016048 38 nips-2009-Augmenting Feature-driven fMRI Analyses: Semi-supervised learning and resting state activity

16 0.50063503 145 nips-2009-Manifold Embeddings for Model-Based Reinforcement Learning under Partial Observability

17 0.49848118 113 nips-2009-Improving Existing Fault Recovery Policies

18 0.49736366 254 nips-2009-Variational Gaussian-process factor analysis for modeling spatio-temporal data

19 0.49701467 174 nips-2009-Nonparametric Latent Feature Models for Link Prediction

20 0.49498621 28 nips-2009-An Additive Latent Feature Model for Transparent Object Recognition