nips nips2004 nips2004-76 knowledge-graph by maker-knowledge-mining
Source: pdf
Author: Rajesh P. Rao
Abstract: There is growing evidence from psychophysical and neurophysiological studies that the brain utilizes Bayesian principles for inference and decision making. An important open question is how Bayesian inference for arbitrary graphical models can be implemented in networks of spiking neurons. In this paper, we show that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for dynamic and hierarchical graphical models. The membrane potential dynamics of neurons is used to implement belief propagation in the log domain. The spiking probability of a neuron is shown to approximate the posterior probability of the preferred state encoded by the neuron, given past inputs. We illustrate the model using two examples: (1) a motion detection network in which the spiking probability of a direction-selective neuron becomes proportional to the posterior probability of motion in a preferred direction, and (2) a two-level hierarchical network that produces attentional effects similar to those observed in visual cortical areas V2 and V4. The hierarchical model offers a new Bayesian interpretation of attentional modulation in V2 and V4. 1
Reference: text
sentIndex sentText sentNum sentScore
1 edu Abstract There is growing evidence from psychophysical and neurophysiological studies that the brain utilizes Bayesian principles for inference and decision making. [sent-5, score-0.179]
2 An important open question is how Bayesian inference for arbitrary graphical models can be implemented in networks of spiking neurons. [sent-6, score-0.591]
3 In this paper, we show that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for dynamic and hierarchical graphical models. [sent-7, score-1.126]
4 The membrane potential dynamics of neurons is used to implement belief propagation in the log domain. [sent-8, score-0.824]
5 The spiking probability of a neuron is shown to approximate the posterior probability of the preferred state encoded by the neuron, given past inputs. [sent-9, score-1.069]
6 The hierarchical model offers a new Bayesian interpretation of attentional modulation in V2 and V4. [sent-11, score-0.223]
7 1 Introduction A wide range of psychophysical results have recently been successfully explained using Bayesian models [7, 8, 16, 19]. [sent-12, score-0.032]
8 These models have been able to account for human responses in tasks ranging from 3D shape perception to visuomotor control. [sent-13, score-0.125]
9 Simultaneously, there is accumulating evidence from human and monkey experiments that Bayesian mechanisms are at work during visual decision making [2, 5]. [sent-14, score-0.051]
10 An important question that has only recently received attention is how networks of cortical neurons can implement algorithms for Bayesian inference. [sent-16, score-0.598]
11 One powerful approach has been to build on the known properties of population coding models that represent information using a set of neural tuning curves or kernel functions [1, 20]. [sent-17, score-0.068]
12 Several proposals have been made regarding how a probability distribution could be encoded using population codes ([3, 18]; see [14] for an excellent review). [sent-18, score-0.209]
13 However, the problem of implementing general inference algorithms for arbitrary graphical models using population codes remains unresolved (some encouraging initial results are reported in Zemel et al. [sent-19, score-0.384]
14 For example, a neural implementation of approximate Bayesian inference for a hidden Markov model was investigated in [15]. [sent-22, score-0.15]
15 The question of how such an approach could be generalized to spiking neurons and arbitrary graphical models remained open. [sent-23, score-0.747]
16 In this paper, we propose a method for implementing Bayesian belief propagation in networks of spiking neurons. [sent-24, score-0.588]
17 We show that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for dynamic and hierarchical graphical models. [sent-25, score-1.126]
18 In the model, the dynamics of the membrane potential is used to implement on-line belief propagation in the log domain [15]. [sent-26, score-0.503]
19 A neuron’s spiking probability is shown to approximate the posterior probability of the preferred state encoded by the neuron, given past inputs. [sent-27, score-0.72]
20 We first show that for a visual motion detection task, the spiking probability of a direction-selective neuron becomes proportional to the posterior probability of motion in the neuron’s preferred direction. [sent-28, score-1.384]
21 We then show that in a two-level network, hierarchical Bayesian inference [9] produces responses that mimic the attentional effects seen in visual cortical areas V2 and V4. [sent-29, score-0.657]
22 1 Integrate-and-Fire Model of Spiking Neurons We begin with a recurrently-connected network of integrate-and-fire (IF) neurons receiving feedforward inputs denoted by the vector I. [sent-31, score-0.569]
23 If vi crosses a threshold T , the neuron fires a spike and vi is reset to the potential vreset . [sent-33, score-0.684]
24 Equation 1 can be rewritten in discrete form as: vi (t + 1) = vi (t) + (−vi (t) + wij Ij (t)) + j i. [sent-34, score-0.334]
25 vi (t + 1) uij vj (t) wij Ij (t) + = uij vj (t)) (2) j (3) j j where is the integration rate, uii = 1 + (uii − 1) and for i = j, uij = uij . [sent-36, score-1.091]
26 2 Stochastic Spiking in Noisy IF Neurons To model the effects of background inputs and the random openings of membrane channels, one can add a Gaussian white noise term to the right hand side of Equations 3 and 4. [sent-39, score-0.272]
27 This makes the spiking of neurons in the recurrent network stochastic. [sent-40, score-0.928]
28 Of particular interest to the present paper is the following exponential function for spiking probability suggested in [4] for noisy integrate-and-fire networks: P (neuron i spikes at time t) = ke(vi (t)−T )/c (5) where k and c are arbitrary constants. [sent-43, score-0.399]
29 1 Inference in a Single-Level Model We first consider on-line belief propagation in a single-level dynamic graphical model and show how it can be implemented in spiking networks. [sent-46, score-0.648]
30 The graphical model is shown in Figure 1A and corresponds to a classical hidden Markov model. [sent-47, score-0.117]
31 Let θ(t) represent the hidden state of a Markov model at time t with transition probabilities given by P (θ(t) = t t−1 θi |θ(t − 1) = θj ) = P (θi |θj ) for i, j = 1 . [sent-48, score-0.199]
32 Let I(t) be the observable output governed by the probabilities P (I(t)|θ(t)). [sent-52, score-0.07]
33 2 Neural Implementation of the Inference Algorithm By comparing the membrane potential equation (Eq. [sent-61, score-0.229]
34 7), it is clear that the first equation can implement the second if belief propagation is performed in the log domain [15], i. [sent-63, score-0.356]
35 Normalization by nt−1,t is implemented by subtracting log nt−1,t using inhibition. [sent-66, score-0.035]
36 1 An alternative approach, which was also found to yield satisfactory results, is to approximate the log-sum with a linear weighted sum [15], the weights being chosen to minimize the approximation error. [sent-67, score-0.07]
37 Each circle represents a node denoting the state variable θ t which can take on values θ1 , . [sent-70, score-0.071]
38 (B) Recurrent network for implementing on-line belief propagation for the graphical model in (A). [sent-74, score-0.456]
39 Each circle represents a neuron encoding a state θ i . [sent-75, score-0.42]
40 The probability distribution over state values at each time step is represented by the entire population. [sent-77, score-0.103]
41 (D) Two-level network for implementing online belief propagation for the graphical model in (C). [sent-79, score-0.456]
42 Arrows represent synaptic connections in the direction pointed by the arrow heads. [sent-80, score-0.234]
43 Finally, since the membrane potential vi (t + 1) is assumed to be proportional to log mt,t+1 i (Equation 9), we have: vi (t + 1) = c log mt,t+1 + T (12) i for some constants c and T . [sent-82, score-0.524]
44 Figure 1B illustrates the singlelevel recurrent network model that implements the on-line belief propagation equation 7. [sent-84, score-0.617]
45 3 Hierarchical Inference The model described above can be extended to perform on-line belief propagation and inference for arbitrary graphical models. [sent-86, score-0.449]
46 As an example, we describe the implementation for the two-level hierarchical graphical model in Figure 1C. [sent-87, score-0.243]
47 The equations above can be implemented in a 2-level hierarchical recurrent network of integrate-and-fire neurons in a manner similar to the 1-level case. [sent-90, score-0.818]
48 We assume that neuron i in level 1 encodes θ1,i as its preferred state while neuron i in level 2 encodes θ2,i . [sent-91, score-1.029]
49 We also assume specific feedforward and feedback neurons for computing and t conveying mt 1→2,i and m2→1,i respectively. [sent-92, score-0.703]
50 Taking the logarithm of both sides of Equations 17 and 18, we obtain equations that can be computed using the membrane potential dynamics of integrate-and-fire neurons (Equation 4). [sent-93, score-0.565]
51 , [10]) and potential implementations based on active dendritic interactions have been explored. [sent-97, score-0.195]
52 The model suggested here utilizes these multiplicative interactions within dendritic branches, in addition to a possible logarithmic transform of the signal before it sums with other signals at the soma. [sent-98, score-0.178]
53 Such a model is comparable to recent models of dendritic computation (see [6] for more details). [sent-99, score-0.102]
54 1 Single-Level Network: Probabilistic Motion Detection and Direction Selectivity We first tested the model in a 1D visual motion detection task [15]. [sent-101, score-0.295]
55 A single-level recurrent network of 30 neurons was used (see Figure 1B). [sent-102, score-0.658]
56 Figure 2A shows the feedforward weights for neurons 1, . [sent-103, score-0.5]
57 , 15: these were recurrently connected to encode transition probabilities biased for rightward motion as shown in Figure 2B. [sent-106, score-0.484]
58 , 30 were identical to Figure 2A but their recurrent connections encoded transition probabilities for leftward motion (see Figure 2B). [sent-110, score-0.872]
59 As seen in Figure 2C, neurons in the network exhibited direction selectivity. [sent-111, score-0.437]
60 Furthermore, the spiking probability of neurons reflected the posterior probabilities over time of motion direction at a given location (Figure 2D), suggesting a probabilistic interpretation of direction selective spiking responses in visual cortical areas such as V1 and MT. [sent-112, score-1.66]
61 2 Two-Level Network: Spatial Attention as Hierarchical Bayesian Inference We tested the two-level network implementation (Figure 1D) of hierarchical Bayesian inference using a simple attention task previously used in primate studies [17]. [sent-114, score-0.431]
62 In an input image, a vertical or horizontal bar could occur either on the left side, right side, or both sides (see Figure 3). [sent-115, score-0.183]
63 The corresponding 2-level generative model consisted of two states at level 2 (left or right side) and four states at level 1: vertical left, horizontal left, vertical right, horizontal right. [sent-116, score-0.489]
64 Each of these states was encoded by a neuron at the respective level. [sent-117, score-0.469]
65 The feedforward connections at level 1 were chosen to be vertically or horizontally oriented Gabor filters localized to the left or right side of the image. [sent-118, score-0.36]
66 Since the experiment used static images, the recurrent connections at each level implemented transition probabilities close to 1 for the same state and small random values for other states. [sent-119, score-0.598]
67 The transition probabilities t−1 t t t P (θ1,k |θ2,i , θ1,j ) were chosen such that for θ2 = left side, the transition probabilities for θt w15 w1 1 0. [sent-120, score-0.256]
68 (B) Recurrent weights encoding the transition t+1 t probabilities P (θi |θj ) for i, j = 1, . [sent-165, score-0.163]
69 (C) Spiking responses of three of the first 15 neurons in the recurrent network (neurons 8, 10, and 12). [sent-170, score-0.75]
70 As is evident, these neurons have become selective for rightward motion as a consequence of the recurrent connections specified in (B). [sent-171, score-1.038]
71 (D) Posterior probabilities over time of motion direction (at a given location) encoded by the three neurons for rightward and leftward motion. [sent-172, score-1.015]
72 t t states θ1 coding for the right side were set to values close to zero (and vice versa, for θ 2 = right side). [sent-173, score-0.158]
73 As shown in Figure 3, the response of a neuron at level 1 that, for example, prefers a vertical edge on the right mimics the response of a V4 neuron with and without attention (see figure caption for more details). [sent-174, score-0.943]
74 The initial setting of the priors at level 2 is the crucial determinant of attentional modulation in level 1 neurons, suggesting that feedback from higher cortical areas may convey task-specific priors that are integrated into V4 responses. [sent-175, score-0.441]
75 5 Discussion and Conclusions We have shown that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for single- and multi-level dynamic graphical models. [sent-176, score-1.0]
76 The model suggests a new interpretation of the spiking probability of a neuron in terms of the posterior probability of the preferred state encoded by the neuron, given past inputs. [sent-177, score-1.034]
77 We illustrated the model using two problems: inference of motion direction in a single-level network and hierarchical inference of object identity at an attended visual location in a twolevel network. [sent-178, score-0.73]
78 In the first case, neurons generated direction-selective spikes encoding the probability of motion in a particular direction. [sent-179, score-0.605]
79 In the second case, attentional effects similar to those observed in primate cortical areas V2 and V4 emerged as a result of imposing appropriate priors at the highest level. [sent-180, score-0.331]
80 How does the approach scale to more realistic graphical models? [sent-182, score-0.117]
81 The two-level model explored in this paper assumed stationary objects, resulting in simplified dynamics for the two levels in our recurrent network. [sent-183, score-0.31]
82 Experiments are currently underway to test the robustness of the proposed model when richer classes of dynamics are introduced at the different levels. [sent-184, score-0.044]
83 (A) Top panel: Input image (lasting the first 15 time steps) containing a vertical bar (“Reference”) on the right side. [sent-186, score-0.151]
84 Middle: Three sample spike trains from the 1st level neuron whose preferred stimulus was a vertical bar on the right side. [sent-188, score-0.709]
85 Bottom: Posterior probability of a vertical bar (= spiking probability or instantaneous firing rate of the neuron) plotted over time. [sent-189, score-0.493]
86 Below: Sample spike trains and posterior probability for the same neuron as in (A). [sent-191, score-0.488]
87 (D) Responses from a neuron in primate area V4 without attention (top panel, Ref Att Away and Pair Att Away; compare with (A) and (B)) and with attention (bottom panel, Pair Att Ref; compare with (C)) (from [17]). [sent-193, score-0.529]
88 other open question is how active dendritic processes could support probabilistic integration of messages from local, lower-level, and higher-level neurons, as suggested in Section 3. [sent-195, score-0.18]
89 We intend to investigate this question using biophysical (compartmental) models of cortical neurons. [sent-196, score-0.125]
90 Finally, how can the feedforward, feedback, and recurrent synaptic weights in the networks be learned directly from input data? [sent-197, score-0.46]
91 We hope to investigate this question using biologically-plausible approximations to the expectation-maximization (EM) algorithm. [sent-198, score-0.039]
92 Neural computation of log likelihood in control of saccadic eye movements. [sent-217, score-0.035]
93 Population dynamics of spiking neurons: Fast transients, asynchronous states, and locking. [sent-228, score-0.314]
94 Loopy belief propagation for approximate inference: An empirical study. [sent-269, score-0.252]
95 Noise in integrate-and-fire neurons: From stochastic input to escape rates. [sent-280, score-0.045]
96 Competitive mechanisms subserve attention in macaque areas V2 and V4. [sent-310, score-0.112]
97 Doubly distributional population codes: Simultaneous representation of uncertainty and multiplicity. [sent-315, score-0.068]
wordName wordTfidf (topN-words)
[('neuron', 0.349), ('neurons', 0.321), ('spiking', 0.27), ('recurrent', 0.266), ('motion', 0.207), ('mt', 0.201), ('uij', 0.163), ('rightward', 0.149), ('leftward', 0.147), ('feedforward', 0.144), ('bayesian', 0.129), ('att', 0.128), ('hierarchical', 0.126), ('vi', 0.126), ('membrane', 0.117), ('graphical', 0.117), ('inference', 0.115), ('propagation', 0.111), ('synaptic', 0.109), ('belief', 0.106), ('dendritic', 0.102), ('attentional', 0.097), ('vj', 0.094), ('responses', 0.092), ('preferred', 0.09), ('message', 0.086), ('cortical', 0.086), ('level', 0.085), ('wij', 0.082), ('ref', 0.078), ('encoded', 0.076), ('posterior', 0.073), ('network', 0.071), ('state', 0.071), ('probabilities', 0.07), ('population', 0.068), ('vertical', 0.068), ('stim', 0.064), ('equation', 0.063), ('attention', 0.061), ('transition', 0.058), ('primate', 0.058), ('ij', 0.057), ('side', 0.052), ('bar', 0.052), ('noisy', 0.052), ('visual', 0.051), ('gerstner', 0.051), ('implementing', 0.051), ('areas', 0.051), ('networks', 0.05), ('potential', 0.049), ('plesser', 0.049), ('prescribes', 0.049), ('onset', 0.049), ('connections', 0.048), ('selective', 0.047), ('nt', 0.047), ('rao', 0.047), ('direction', 0.045), ('spikes', 0.045), ('panel', 0.045), ('escape', 0.045), ('dynamics', 0.044), ('interactions', 0.044), ('dynamic', 0.044), ('states', 0.044), ('ring', 0.043), ('uii', 0.043), ('past', 0.041), ('implement', 0.041), ('re', 0.041), ('effects', 0.039), ('instantaneous', 0.039), ('weiss', 0.039), ('messages', 0.039), ('question', 0.039), ('feedback', 0.037), ('detection', 0.037), ('dayan', 0.037), ('deneve', 0.036), ('dendrites', 0.036), ('away', 0.036), ('proportional', 0.036), ('weights', 0.035), ('approximate', 0.035), ('log', 0.035), ('spike', 0.034), ('equations', 0.034), ('neuroscience', 0.034), ('inputs', 0.033), ('codes', 0.033), ('perception', 0.033), ('arrow', 0.032), ('psychophysical', 0.032), ('zemel', 0.032), ('utilizes', 0.032), ('horizontal', 0.032), ('probability', 0.032), ('right', 0.031)]
simIndex simValue paperId paperTitle
same-paper 1 0.99999934 76 nips-2004-Hierarchical Bayesian Inference in Networks of Spiking Neurons
Author: Rajesh P. Rao
Abstract: There is growing evidence from psychophysical and neurophysiological studies that the brain utilizes Bayesian principles for inference and decision making. An important open question is how Bayesian inference for arbitrary graphical models can be implemented in networks of spiking neurons. In this paper, we show that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for dynamic and hierarchical graphical models. The membrane potential dynamics of neurons is used to implement belief propagation in the log domain. The spiking probability of a neuron is shown to approximate the posterior probability of the preferred state encoded by the neuron, given past inputs. We illustrate the model using two examples: (1) a motion detection network in which the spiking probability of a direction-selective neuron becomes proportional to the posterior probability of motion in a preferred direction, and (2) a two-level hierarchical network that produces attentional effects similar to those observed in visual cortical areas V2 and V4. The hierarchical model offers a new Bayesian interpretation of attentional modulation in V2 and V4. 1
2 0.34397829 28 nips-2004-Bayesian inference in spiking neurons
Author: Sophie Deneve
Abstract: We propose a new interpretation of spiking neurons as Bayesian integrators accumulating evidence over time about events in the external world or the body, and communicating to other neurons their certainties about these events. In this model, spikes signal the occurrence of new information, i.e. what cannot be predicted from the past activity. As a result, firing statistics are close to Poisson, albeit providing a deterministic representation of probabilities. We proceed to develop a theory of Bayesian inference in spiking neural networks, recurrent interactions implementing a variant of belief propagation. Many perceptual and motor tasks performed by the central nervous system are probabilistic, and can be described in a Bayesian framework [4, 3]. A few important but hidden properties, such as direction of motion, or appropriate motor commands, are inferred from many noisy, local and ambiguous sensory cues. These evidences are combined with priors about the sensory world and body. Importantly, because most of these inferences should lead to quick and irreversible decisions in a perpetually changing world, noisy cues have to be integrated on-line, but in a way that takes into account unpredictable events, such as a sudden change in motion direction or the appearance of a new stimulus. This raises the question of how this temporal integration can be performed at the neural level. It has been proposed that single neurons in sensory cortices represent and compute the log probability that a sensory variable takes on a certain value (eg Is visual motion in the neuron’s preferred direction?) [9, 7]. Alternatively, to avoid normalization issues and provide an appropriate signal for decision making, neurons could represent the log probability ratio of a particular hypothesis (eg is motion more likely to be towards the right than towards the left) [7, 6]. Log probabilities are convenient here, since under some assumptions, independent noisy cues simply combine linearly. Moreover, there are physiological evidence for the neural representation of log probabilities and log probability ratios [9, 6, 7]. However, these models assume that neurons represent probabilities in their firing rates. We argue that it is important to study how probabilistic information are encoded in spikes. Indeed, it seems spurious to marry the idea of an exquisite on-line integration of noisy cues with an underlying rate code that requires averaging on large populations of noisy neurons and long periods of time. In particular, most natural tasks require this integration to take place on the time scale of inter-spike intervals. Spikes are more efficiently signaling events ∗ Institute of Cognitive Science, 69645 Bron, France than analog quantities. In addition, a neural theory of inference with spikes will bring us closer to the physiological level and generate more easily testable predictions. Thus, we propose a new theory of neural processing in which spike trains provide a deterministic, online representation of a log-probability ratio. Spikes signals events, eg that the log-probability ratio has exceeded what could be predicted from previous spikes. This form of coding was loosely inspired by the idea of ”energy landscape” coding proposed by Hinton and Brown [2]. However, contrary to [2] and other theories using rate-based representation of probabilities, this model is self-consistent and does not require different models for encoding and decoding: As output spikes provide new, unpredictable, temporally independent evidence, they can be used directly as an input to other Bayesian neurons. Finally, we show that these neurons can be used as building blocks in a theory of approximate Bayesian inference in recurrent spiking networks. Connections between neurons implement an underlying Bayesian network, consisting of coupled hidden Markov models. Propagation of spikes is a form of belief propagation in this underlying graphical model. Our theory provides computational explanations of some general physiological properties of cortical neurons, such as spike frequency adaptation, Poisson statistics of spike trains, the existence of strong local inhibition in cortical columns, and the maintenance of a tight balance between excitation and inhibition. Finally, we discuss the implications of this model for the debate about temporal versus rate-based neural coding. 1 Spikes and log posterior odds 1.1 Synaptic integration seen as inference in a hidden Markov chain We propose that each neuron codes for an underlying ”hidden” binary variable, xt , whose state evolves over time. We assume that xt depends only on the state at the previous time step, xt−dt , and is conditionally independent of other past states. The state xt can switch 1 from 0 to 1 with a constant rate ron = dt limdt→0 P (xt = 1|xt−dt = 0), and from 1 to 0 with a constant rate roff . For example, these transition rates could represent how often motion in a preferred direction appears the receptive field and how long it is likely to stay there. The neuron infers the state of its hidden variable from N noisy synaptic inputs, considered to be observations of the hidden state. In this initial version of the model, we assume that these inputs are conditionally independent homogeneous Poisson processes, synapse i i emitting a spike between time t and t + dt (si = 1) with constant probability qon dt if t i xt = 1, and another constant probability qoff dt if xt = 0. The synaptic spikes are assumed to be otherwise independent of previous synaptic spikes, previous states and spikes at other synapses. The resulting generative model is a hidden Markov chain (figure 1-A). However, rather than estimating the state of its hidden variable and communicating this estimate to other neurons (for example by emitting a spike when sensory evidence for xt = 1 goes above a threshold) the neuron reports and communicates its certainty that the current state is 1. This certainty takes the form of the log of the ratio of the probability that the hidden state is 1, and the probability that the state is 0, given all the synaptic inputs P (xt =1|s0→t ) received so far: Lt = log P (xt =0|s0→t ) . We use s0→t as a short hand notation for the N synaptic inputs received at present and in the past. We will refer to it as the log odds ratio. Thanks to the conditional independencies assumed in the generative model, we can compute this Log odds ratio iteratively. Taking the limit as dt goes to zero, we get the following differential equation: ˙ L = ron 1 + e−L − roff 1 + eL + i wi δ(si − 1) − θ t B. A. xt ron .roff dt qon , qoff st xt ron .roff i t st dt s qon , qoff qon , qoff st dt xt j st Ot It Gt Ot Lt t t dt C. E. 2 0 -2 -4 D. 500 1000 1500 2000 2500 2 3000 Count Log odds 4 20 Lt 0 -2 0 500 1000 1500 2000 2500 Time Ot 3000 0 200 400 600 ISI Figure 1: A. Generative model for the synaptic input. B. Schematic representation of log odds ratio encoding and decoding. The dashed circle represents both eventual downstream elements and the self-prediction taking place inside the model neuron. A spike is fired only when Lt exceeds Gt . C. One example trial, where the state switches from 0 to 1 (shaded area) and back to 0. plain: Lt , dotted: Gt . Black stripes at the top: corresponding spikes train. D. Mean Log odds ratio (dark line) and mean output firing rate (clear line). E. Output spike raster plot (1 line per trial) and ISI distribution for the neuron shown is C. and D. Clear line: ISI distribution for a poisson neuron with the same rate. wi , the synaptic weight, describe how informative synapse i is about the state of the hidden i qon variable, e.g. wi = log qi . Each synaptic spike (si = 1) gives an impulse to the log t off odds ratio, which is positive if this synapse is more active when the hidden state if 1 (i.e it increases the neuron’s confidence that the state is 1), and negative if this synapse is more active when xt = 0 (i.e it decreases the neuron’s confidence that the state is 1). The bias, θ, is determined by how informative it is not to receive any spike, e.g. θ = i i i qon − qoff . By convention, we will consider that the ”bias” is positive or zero (if not, we need simply to invert the status of the state x). 1.2 Generation of output spikes The spike train should convey a sparse representation of Lt , so that each spike reports new information about the state xt that is not redundant with that reported by other, preceding, spikes. This proposition is based on three arguments: First, spikes, being metabolically expensive, should be kept to a minimum. Second, spikes conveying redundant information would require a decoding of the entire spike train, whereas independent spike can be taken into account individually. And finally, we seek a self consistent model, with the spiking output having a similar semantics to its spiking input. To maximize the independence of the spikes (conditioned on xt ), we propose that the neuron fires only when the difference between its log odds ratio Lt and a prediction Gt of this log odds ratio based on the output spikes emitted so far reaches a certain threshold. Indeed, supposing that downstream elements predicts Lt as best as they can, the neuron only needs to fire when it expects that prediction to be too inaccurate (figure 1-B). In practice, this will happen when the neuron receives new evidence for xt = 1. Gt should thereby follow the same dynamics as Lt when spikes are not received. The equation for Gt and the output Ot (Ot = 1 when an output spike is fired) are given by: ˙ G = Ot = ron 1 + e−L − roff 1 + eL + go δ(Ot − 1) go 1. when Lt > Gt + , 0 otherwise, 2 (1) (2) Here go , a positive constant, is the only free parameter, the other parameters being constrained by the statistics of the synaptic input. 1.3 Results Figure 1-C plots a typical trial, showing the behavior of L, G and O before, during and after presentation of the stimulus. As random synaptic inputs are integrated, L fluctuates and eventually exceeds G + 0.5, leading to an output spike. Immediately after a spike, G jumps to G + go , which prevents (except in very rare cases) a second spike from immediately following the first. Thus, this ”jump” implements a relative refractory period. However, ron G decays as it tends to converge back to its stable level gstable = log roff . Thus L eventually exceeds G again, leading to a new spike. This threshold crossing happens more often during stimulation (xt = 1) as the net synaptic input alters to create a higher overall level of certainty, Lt . Mean Log odds ratio and output firing rate ¯ The mean firing rate Ot of the Bayesian neuron during presentation of its preferred stimulus (i.e. when xt switches from 0 to 1 and back to 0) is plotted in figure 1-D, together with the ¯ mean log posterior ratio Lt , both averaged over trials. Not surprisingly, the log-posterior ratio reflects the leaky integration of synaptic evidence, with an effective time constant that depends on the transition probabilities ron , roff . If the state is very stable (ron = roff ∼ 0), synaptic evidence is integrated over almost infinite time periods, the mean log posterior ratio tending to either increase or decrease linearly with time. In the example in figure 1D, the state is less stable, so ”old” synaptic evidence are discounted and Lt saturates. ¯ In contrast, the mean output firing rate Ot tracks the state of xt almost perfectly. This is because, as a form of predictive coding, the output spikes reflect the new synaptic i evidence, It = i δ(st − 1) − θ, rather than the log posterior ratio itself. In particular, the mean output firing rate is a rectified linear function of the mean input, e. g. + ¯ ¯ wi q i −θ . O= 1I= go i on(off) Analogy with a leaky integrate and fire neuron We can get an interesting insight into the computation performed by this neuron by linearizing L and G around their mean levels over trials. Here we reduce the analysis to prolonged, statistically stable periods when the state is constant (either ON or OFF). In this case, the ¯ ¯ mean level of certainty L and its output prediction G are also constant over time. We make the rough approximation that the post spike jump, go , and the input fluctuations are small ¯ compared to the mean level of certainty L. Rewriting Vt = Lt − Gt + go 2 as the ”membrane potential” of the Bayesian neuron: ˙ V = −kL V + It − ∆go − go Ot ¯ ¯ ¯ where kL = ron e−L + roff eL , the ”leak” of the membrane potential, depends on the overall ¯ level of certainty. ∆go is positive and a monotonic increasing function of go . A. s t1 dt s t1 s t1 dt B. C. x t1 x t3 dt x t3 x t3 dt x t1 x t1 x t1 x t2 x t3 x t1 … x tn x t3 x t2 … x tn … dt dt Lx2 D. x t2 dt s t2 dt x t2 s t2 x t2 dt s t2 dt Log odds 10 No inh -0.5 -1 -1 -1.5 -2 5 Feedback 500 1000 1500 2000 Tiger Stripes 0 -5 -10 500 1000 1500 2000 2500 Time Figure 2: A. Bayesian causal network for yt (tiger), x1 (stripes) and x2 (paws). B. A nett t work feedforward computing the log posterior for x1 . C. A recurrent network computing t the log posterior odds for all variables. D. Log odds ratio in a simulated trial with the net2 1 1 work in C (see text). Thick line: Lx , thin line: Lx , dash-dotted: Lx without inhibition. t t t 2 Insert: Lx averaged over trials, showing the effect of feedback. t The linearized Bayesian neuron thus acts in its stable regime as a leaky integrate and fire (LIF) neuron. The membrane potential Vt integrates its input, Jt = It − ∆go , with a leak kL . The neuron fires when its membrane potential reaches a constant threshold go . After ¯ each spikes, Vt is reset to 0. Interestingly, for appropriately chosen compression factor go , the mean input to the lin¯ ¯ earized neuron J = I − ∆go ≈ 0 1 . This means that the membrane potential is purely driven to its threshold by input fluctuations, or a random walk in membrane potential. As a consequence, the neuron’s firing will be memoryless, and close to a Poisson process. In particular, we found Fano factor close to 1 and quasi-exponential ISI distribution (figure 1E) on the entire range of parameters tested. Indeed, LIF neurons with balanced inputs have been proposed as a model to reproduce the statistics of real cortical neurons [8]. This balance is implemented in our model by the neuron’s effective self-inhibition, even when the synaptic input itself is not balanced. Decoding As we previously said, downstream elements could predict the log odds ratio Lt by computing Gt from the output spikes (Eq 1, fig 1-B). Of course, this requires an estimate of the transition probabilities ron , roff , that could be learned from the observed spike trains. However, we show next that explicit decoding is not necessary to perform bayesian inference in spiking networks. Intuitively, this is because the quantity that our model neurons receive and transmit, eg new information, is exactly what probabilistic inference algorithm propagate between connected statistical elements. 1 ¯ Even if go is not chosen optimally, the influence of the drift J is usually negligible compared to the large fluctuations in membrane potential. 2 Bayesian inference in cortical networks The model neurons, having the same input and output semantics, can be used as building blocks to implement more complex generative models consisting of coupled Markov chains. Consider, for example, the example in figure 2-A. Here, a ”parent” variable x1 t (the presence of a tiger) can cause the state of n other ”children” variables ([xk ]k=2...n ), t of whom two are represented (the presence of stripes,x2 , and motion, x3 ). The ”chilt t dren” variables are Bayesian neurons identical to those described previously. The resulting bayesian network consist of n + 1 coupled hidden Markov chains. Inference in this architecture corresponds to computing the log posterior odds ratio for the tiger, x1 , and the log t posterior of observing stripes or motion, ([xk ]k=2...n ), given the synaptic inputs received t by the entire network so far, i.e. s2 , . . . , sk . 0→t 0→t Unfortunately, inference and learning in this network (and in general in coupled Markov chains) requires very expensive computations, and cannot be performed by simply propagating messages over time and among the variable nodes. In particular, the state of a child k variable xt depends on xk , sk , x1 and the state of all other children at the previous t t t−dt time step, [xj ]2
3 0.28826159 140 nips-2004-Optimal Information Decoding from Neuronal Populations with Specific Stimulus Selectivity
Author: Marcelo A. Montemurro, Stefano Panzeri
Abstract: A typical neuron in visual cortex receives most inputs from other cortical neurons with a roughly similar stimulus preference. Does this arrangement of inputs allow efficient readout of sensory information by the target cortical neuron? We address this issue by using simple modelling of neuronal population activity and information theoretic tools. We find that efficient synaptic information transmission requires that the tuning curve of the afferent neurons is approximately as wide as the spread of stimulus preferences of the afferent neurons reaching the target neuron. By meta analysis of neurophysiological data we found that this is the case for cortico-cortical inputs to neurons in visual cortex. We suggest that the organization of V1 cortico-cortical synaptic inputs allows optimal information transmission. 1
4 0.24510072 153 nips-2004-Reducing Spike Train Variability: A Computational Theory Of Spike-Timing Dependent Plasticity
Author: Sander M. Bohte, Michael C. Mozer
Abstract: Experimental studies have observed synaptic potentiation when a presynaptic neuron fires shortly before a postsynaptic neuron, and synaptic depression when the presynaptic neuron fires shortly after. The dependence of synaptic modulation on the precise timing of the two action potentials is known as spike-timing dependent plasticity or STDP. We derive STDP from a simple computational principle: synapses adapt so as to minimize the postsynaptic neuron’s variability to a given presynaptic input, causing the neuron’s output to become more reliable in the face of noise. Using an entropy-minimization objective function and the biophysically realistic spike-response model of Gerstner (2001), we simulate neurophysiological experiments and obtain the characteristic STDP curve along with other phenomena including the reduction in synaptic plasticity as synaptic efficacy increases. We compare our account to other efforts to derive STDP from computational principles, and argue that our account provides the most comprehensive coverage of the phenomena. Thus, reliability of neural response in the face of noise may be a key goal of cortical adaptation. 1
5 0.23691992 194 nips-2004-Theory of localized synfire chain: characteristic propagation speed of stable spike pattern
Author: Kosuke Hamaguchi, Masato Okada, Kazuyuki Aihara
Abstract: Repeated spike patterns have often been taken as evidence for the synfire chain, a phenomenon that a stable spike synchrony propagates through a feedforward network. Inter-spike intervals which represent a repeated spike pattern are influenced by the propagation speed of a spike packet. However, the relation between the propagation speed and network structure is not well understood. While it is apparent that the propagation speed depends on the excitatory synapse strength, it might also be related to spike patterns. We analyze a feedforward network with Mexican-Hattype connectivity (FMH) using the Fokker-Planck equation. We show that both a uniform and a localized spike packet are stable in the FMH in a certain parameter region. We also demonstrate that the propagation speed depends on the distinct firing patterns in the same network.
6 0.21050945 157 nips-2004-Saliency-Driven Image Acuity Modulation on a Reconfigurable Array of Spiking Silicon Neurons
7 0.20002542 181 nips-2004-Synergies between Intrinsic and Synaptic Plasticity in Individual Model Neurons
8 0.18472949 151 nips-2004-Rate- and Phase-coded Autoassociative Memory
9 0.17511015 112 nips-2004-Maximising Sensitivity in a Spiking Network
10 0.17269127 84 nips-2004-Inference, Attention, and Decision in a Bayesian Neural Architecture
11 0.16786973 148 nips-2004-Probabilistic Computation in Spiking Populations
12 0.1439144 173 nips-2004-Spike-timing Dependent Plasticity and Mutual Information Maximization for a Spiking Neuron Model
13 0.1199697 118 nips-2004-Methods for Estimating the Computational Power and Generalization Capability of Neural Microcircuits
14 0.11580093 26 nips-2004-At the Edge of Chaos: Real-time Computations and Self-Organized Criticality in Recurrent Neural Networks
15 0.11443482 58 nips-2004-Edge of Chaos Computation in Mixed-Mode VLSI - A Hard Liquid
16 0.11348259 116 nips-2004-Message Errors in Belief Propagation
17 0.096009031 35 nips-2004-Chemosensory Processing in a Spiking Model of the Olfactory Bulb: Chemotopic Convergence and Center Surround Inhibition
18 0.091048352 56 nips-2004-Dynamic Bayesian Networks for Brain-Computer Interfaces
19 0.090970382 46 nips-2004-Constraining a Bayesian Model of Human Visual Speed Perception
20 0.083513305 13 nips-2004-A Three Tiered Approach for Articulated Object Action Modeling and Recognition
topicId topicWeight
[(0, -0.241), (1, -0.469), (2, -0.085), (3, 0.032), (4, 0.064), (5, 0.061), (6, -0.026), (7, 0.023), (8, -0.06), (9, -0.117), (10, 0.016), (11, -0.05), (12, 0.044), (13, 0.001), (14, 0.062), (15, -0.024), (16, -0.06), (17, -0.034), (18, 0.01), (19, 0.033), (20, 0.062), (21, 0.019), (22, 0.104), (23, -0.087), (24, -0.009), (25, -0.052), (26, 0.061), (27, -0.095), (28, 0.059), (29, 0.087), (30, 0.051), (31, -0.017), (32, 0.011), (33, -0.09), (34, -0.081), (35, 0.007), (36, -0.063), (37, 0.037), (38, -0.034), (39, 0.095), (40, -0.114), (41, 0.081), (42, 0.05), (43, 0.018), (44, 0.049), (45, 0.002), (46, -0.183), (47, -0.034), (48, -0.073), (49, 0.024)]
simIndex simValue paperId paperTitle
same-paper 1 0.97938216 76 nips-2004-Hierarchical Bayesian Inference in Networks of Spiking Neurons
Author: Rajesh P. Rao
Abstract: There is growing evidence from psychophysical and neurophysiological studies that the brain utilizes Bayesian principles for inference and decision making. An important open question is how Bayesian inference for arbitrary graphical models can be implemented in networks of spiking neurons. In this paper, we show that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for dynamic and hierarchical graphical models. The membrane potential dynamics of neurons is used to implement belief propagation in the log domain. The spiking probability of a neuron is shown to approximate the posterior probability of the preferred state encoded by the neuron, given past inputs. We illustrate the model using two examples: (1) a motion detection network in which the spiking probability of a direction-selective neuron becomes proportional to the posterior probability of motion in a preferred direction, and (2) a two-level hierarchical network that produces attentional effects similar to those observed in visual cortical areas V2 and V4. The hierarchical model offers a new Bayesian interpretation of attentional modulation in V2 and V4. 1
2 0.85303348 140 nips-2004-Optimal Information Decoding from Neuronal Populations with Specific Stimulus Selectivity
Author: Marcelo A. Montemurro, Stefano Panzeri
Abstract: A typical neuron in visual cortex receives most inputs from other cortical neurons with a roughly similar stimulus preference. Does this arrangement of inputs allow efficient readout of sensory information by the target cortical neuron? We address this issue by using simple modelling of neuronal population activity and information theoretic tools. We find that efficient synaptic information transmission requires that the tuning curve of the afferent neurons is approximately as wide as the spread of stimulus preferences of the afferent neurons reaching the target neuron. By meta analysis of neurophysiological data we found that this is the case for cortico-cortical inputs to neurons in visual cortex. We suggest that the organization of V1 cortico-cortical synaptic inputs allows optimal information transmission. 1
3 0.78000426 28 nips-2004-Bayesian inference in spiking neurons
Author: Sophie Deneve
Abstract: We propose a new interpretation of spiking neurons as Bayesian integrators accumulating evidence over time about events in the external world or the body, and communicating to other neurons their certainties about these events. In this model, spikes signal the occurrence of new information, i.e. what cannot be predicted from the past activity. As a result, firing statistics are close to Poisson, albeit providing a deterministic representation of probabilities. We proceed to develop a theory of Bayesian inference in spiking neural networks, recurrent interactions implementing a variant of belief propagation. Many perceptual and motor tasks performed by the central nervous system are probabilistic, and can be described in a Bayesian framework [4, 3]. A few important but hidden properties, such as direction of motion, or appropriate motor commands, are inferred from many noisy, local and ambiguous sensory cues. These evidences are combined with priors about the sensory world and body. Importantly, because most of these inferences should lead to quick and irreversible decisions in a perpetually changing world, noisy cues have to be integrated on-line, but in a way that takes into account unpredictable events, such as a sudden change in motion direction or the appearance of a new stimulus. This raises the question of how this temporal integration can be performed at the neural level. It has been proposed that single neurons in sensory cortices represent and compute the log probability that a sensory variable takes on a certain value (eg Is visual motion in the neuron’s preferred direction?) [9, 7]. Alternatively, to avoid normalization issues and provide an appropriate signal for decision making, neurons could represent the log probability ratio of a particular hypothesis (eg is motion more likely to be towards the right than towards the left) [7, 6]. Log probabilities are convenient here, since under some assumptions, independent noisy cues simply combine linearly. Moreover, there are physiological evidence for the neural representation of log probabilities and log probability ratios [9, 6, 7]. However, these models assume that neurons represent probabilities in their firing rates. We argue that it is important to study how probabilistic information are encoded in spikes. Indeed, it seems spurious to marry the idea of an exquisite on-line integration of noisy cues with an underlying rate code that requires averaging on large populations of noisy neurons and long periods of time. In particular, most natural tasks require this integration to take place on the time scale of inter-spike intervals. Spikes are more efficiently signaling events ∗ Institute of Cognitive Science, 69645 Bron, France than analog quantities. In addition, a neural theory of inference with spikes will bring us closer to the physiological level and generate more easily testable predictions. Thus, we propose a new theory of neural processing in which spike trains provide a deterministic, online representation of a log-probability ratio. Spikes signals events, eg that the log-probability ratio has exceeded what could be predicted from previous spikes. This form of coding was loosely inspired by the idea of ”energy landscape” coding proposed by Hinton and Brown [2]. However, contrary to [2] and other theories using rate-based representation of probabilities, this model is self-consistent and does not require different models for encoding and decoding: As output spikes provide new, unpredictable, temporally independent evidence, they can be used directly as an input to other Bayesian neurons. Finally, we show that these neurons can be used as building blocks in a theory of approximate Bayesian inference in recurrent spiking networks. Connections between neurons implement an underlying Bayesian network, consisting of coupled hidden Markov models. Propagation of spikes is a form of belief propagation in this underlying graphical model. Our theory provides computational explanations of some general physiological properties of cortical neurons, such as spike frequency adaptation, Poisson statistics of spike trains, the existence of strong local inhibition in cortical columns, and the maintenance of a tight balance between excitation and inhibition. Finally, we discuss the implications of this model for the debate about temporal versus rate-based neural coding. 1 Spikes and log posterior odds 1.1 Synaptic integration seen as inference in a hidden Markov chain We propose that each neuron codes for an underlying ”hidden” binary variable, xt , whose state evolves over time. We assume that xt depends only on the state at the previous time step, xt−dt , and is conditionally independent of other past states. The state xt can switch 1 from 0 to 1 with a constant rate ron = dt limdt→0 P (xt = 1|xt−dt = 0), and from 1 to 0 with a constant rate roff . For example, these transition rates could represent how often motion in a preferred direction appears the receptive field and how long it is likely to stay there. The neuron infers the state of its hidden variable from N noisy synaptic inputs, considered to be observations of the hidden state. In this initial version of the model, we assume that these inputs are conditionally independent homogeneous Poisson processes, synapse i i emitting a spike between time t and t + dt (si = 1) with constant probability qon dt if t i xt = 1, and another constant probability qoff dt if xt = 0. The synaptic spikes are assumed to be otherwise independent of previous synaptic spikes, previous states and spikes at other synapses. The resulting generative model is a hidden Markov chain (figure 1-A). However, rather than estimating the state of its hidden variable and communicating this estimate to other neurons (for example by emitting a spike when sensory evidence for xt = 1 goes above a threshold) the neuron reports and communicates its certainty that the current state is 1. This certainty takes the form of the log of the ratio of the probability that the hidden state is 1, and the probability that the state is 0, given all the synaptic inputs P (xt =1|s0→t ) received so far: Lt = log P (xt =0|s0→t ) . We use s0→t as a short hand notation for the N synaptic inputs received at present and in the past. We will refer to it as the log odds ratio. Thanks to the conditional independencies assumed in the generative model, we can compute this Log odds ratio iteratively. Taking the limit as dt goes to zero, we get the following differential equation: ˙ L = ron 1 + e−L − roff 1 + eL + i wi δ(si − 1) − θ t B. A. xt ron .roff dt qon , qoff st xt ron .roff i t st dt s qon , qoff qon , qoff st dt xt j st Ot It Gt Ot Lt t t dt C. E. 2 0 -2 -4 D. 500 1000 1500 2000 2500 2 3000 Count Log odds 4 20 Lt 0 -2 0 500 1000 1500 2000 2500 Time Ot 3000 0 200 400 600 ISI Figure 1: A. Generative model for the synaptic input. B. Schematic representation of log odds ratio encoding and decoding. The dashed circle represents both eventual downstream elements and the self-prediction taking place inside the model neuron. A spike is fired only when Lt exceeds Gt . C. One example trial, where the state switches from 0 to 1 (shaded area) and back to 0. plain: Lt , dotted: Gt . Black stripes at the top: corresponding spikes train. D. Mean Log odds ratio (dark line) and mean output firing rate (clear line). E. Output spike raster plot (1 line per trial) and ISI distribution for the neuron shown is C. and D. Clear line: ISI distribution for a poisson neuron with the same rate. wi , the synaptic weight, describe how informative synapse i is about the state of the hidden i qon variable, e.g. wi = log qi . Each synaptic spike (si = 1) gives an impulse to the log t off odds ratio, which is positive if this synapse is more active when the hidden state if 1 (i.e it increases the neuron’s confidence that the state is 1), and negative if this synapse is more active when xt = 0 (i.e it decreases the neuron’s confidence that the state is 1). The bias, θ, is determined by how informative it is not to receive any spike, e.g. θ = i i i qon − qoff . By convention, we will consider that the ”bias” is positive or zero (if not, we need simply to invert the status of the state x). 1.2 Generation of output spikes The spike train should convey a sparse representation of Lt , so that each spike reports new information about the state xt that is not redundant with that reported by other, preceding, spikes. This proposition is based on three arguments: First, spikes, being metabolically expensive, should be kept to a minimum. Second, spikes conveying redundant information would require a decoding of the entire spike train, whereas independent spike can be taken into account individually. And finally, we seek a self consistent model, with the spiking output having a similar semantics to its spiking input. To maximize the independence of the spikes (conditioned on xt ), we propose that the neuron fires only when the difference between its log odds ratio Lt and a prediction Gt of this log odds ratio based on the output spikes emitted so far reaches a certain threshold. Indeed, supposing that downstream elements predicts Lt as best as they can, the neuron only needs to fire when it expects that prediction to be too inaccurate (figure 1-B). In practice, this will happen when the neuron receives new evidence for xt = 1. Gt should thereby follow the same dynamics as Lt when spikes are not received. The equation for Gt and the output Ot (Ot = 1 when an output spike is fired) are given by: ˙ G = Ot = ron 1 + e−L − roff 1 + eL + go δ(Ot − 1) go 1. when Lt > Gt + , 0 otherwise, 2 (1) (2) Here go , a positive constant, is the only free parameter, the other parameters being constrained by the statistics of the synaptic input. 1.3 Results Figure 1-C plots a typical trial, showing the behavior of L, G and O before, during and after presentation of the stimulus. As random synaptic inputs are integrated, L fluctuates and eventually exceeds G + 0.5, leading to an output spike. Immediately after a spike, G jumps to G + go , which prevents (except in very rare cases) a second spike from immediately following the first. Thus, this ”jump” implements a relative refractory period. However, ron G decays as it tends to converge back to its stable level gstable = log roff . Thus L eventually exceeds G again, leading to a new spike. This threshold crossing happens more often during stimulation (xt = 1) as the net synaptic input alters to create a higher overall level of certainty, Lt . Mean Log odds ratio and output firing rate ¯ The mean firing rate Ot of the Bayesian neuron during presentation of its preferred stimulus (i.e. when xt switches from 0 to 1 and back to 0) is plotted in figure 1-D, together with the ¯ mean log posterior ratio Lt , both averaged over trials. Not surprisingly, the log-posterior ratio reflects the leaky integration of synaptic evidence, with an effective time constant that depends on the transition probabilities ron , roff . If the state is very stable (ron = roff ∼ 0), synaptic evidence is integrated over almost infinite time periods, the mean log posterior ratio tending to either increase or decrease linearly with time. In the example in figure 1D, the state is less stable, so ”old” synaptic evidence are discounted and Lt saturates. ¯ In contrast, the mean output firing rate Ot tracks the state of xt almost perfectly. This is because, as a form of predictive coding, the output spikes reflect the new synaptic i evidence, It = i δ(st − 1) − θ, rather than the log posterior ratio itself. In particular, the mean output firing rate is a rectified linear function of the mean input, e. g. + ¯ ¯ wi q i −θ . O= 1I= go i on(off) Analogy with a leaky integrate and fire neuron We can get an interesting insight into the computation performed by this neuron by linearizing L and G around their mean levels over trials. Here we reduce the analysis to prolonged, statistically stable periods when the state is constant (either ON or OFF). In this case, the ¯ ¯ mean level of certainty L and its output prediction G are also constant over time. We make the rough approximation that the post spike jump, go , and the input fluctuations are small ¯ compared to the mean level of certainty L. Rewriting Vt = Lt − Gt + go 2 as the ”membrane potential” of the Bayesian neuron: ˙ V = −kL V + It − ∆go − go Ot ¯ ¯ ¯ where kL = ron e−L + roff eL , the ”leak” of the membrane potential, depends on the overall ¯ level of certainty. ∆go is positive and a monotonic increasing function of go . A. s t1 dt s t1 s t1 dt B. C. x t1 x t3 dt x t3 x t3 dt x t1 x t1 x t1 x t2 x t3 x t1 … x tn x t3 x t2 … x tn … dt dt Lx2 D. x t2 dt s t2 dt x t2 s t2 x t2 dt s t2 dt Log odds 10 No inh -0.5 -1 -1 -1.5 -2 5 Feedback 500 1000 1500 2000 Tiger Stripes 0 -5 -10 500 1000 1500 2000 2500 Time Figure 2: A. Bayesian causal network for yt (tiger), x1 (stripes) and x2 (paws). B. A nett t work feedforward computing the log posterior for x1 . C. A recurrent network computing t the log posterior odds for all variables. D. Log odds ratio in a simulated trial with the net2 1 1 work in C (see text). Thick line: Lx , thin line: Lx , dash-dotted: Lx without inhibition. t t t 2 Insert: Lx averaged over trials, showing the effect of feedback. t The linearized Bayesian neuron thus acts in its stable regime as a leaky integrate and fire (LIF) neuron. The membrane potential Vt integrates its input, Jt = It − ∆go , with a leak kL . The neuron fires when its membrane potential reaches a constant threshold go . After ¯ each spikes, Vt is reset to 0. Interestingly, for appropriately chosen compression factor go , the mean input to the lin¯ ¯ earized neuron J = I − ∆go ≈ 0 1 . This means that the membrane potential is purely driven to its threshold by input fluctuations, or a random walk in membrane potential. As a consequence, the neuron’s firing will be memoryless, and close to a Poisson process. In particular, we found Fano factor close to 1 and quasi-exponential ISI distribution (figure 1E) on the entire range of parameters tested. Indeed, LIF neurons with balanced inputs have been proposed as a model to reproduce the statistics of real cortical neurons [8]. This balance is implemented in our model by the neuron’s effective self-inhibition, even when the synaptic input itself is not balanced. Decoding As we previously said, downstream elements could predict the log odds ratio Lt by computing Gt from the output spikes (Eq 1, fig 1-B). Of course, this requires an estimate of the transition probabilities ron , roff , that could be learned from the observed spike trains. However, we show next that explicit decoding is not necessary to perform bayesian inference in spiking networks. Intuitively, this is because the quantity that our model neurons receive and transmit, eg new information, is exactly what probabilistic inference algorithm propagate between connected statistical elements. 1 ¯ Even if go is not chosen optimally, the influence of the drift J is usually negligible compared to the large fluctuations in membrane potential. 2 Bayesian inference in cortical networks The model neurons, having the same input and output semantics, can be used as building blocks to implement more complex generative models consisting of coupled Markov chains. Consider, for example, the example in figure 2-A. Here, a ”parent” variable x1 t (the presence of a tiger) can cause the state of n other ”children” variables ([xk ]k=2...n ), t of whom two are represented (the presence of stripes,x2 , and motion, x3 ). The ”chilt t dren” variables are Bayesian neurons identical to those described previously. The resulting bayesian network consist of n + 1 coupled hidden Markov chains. Inference in this architecture corresponds to computing the log posterior odds ratio for the tiger, x1 , and the log t posterior of observing stripes or motion, ([xk ]k=2...n ), given the synaptic inputs received t by the entire network so far, i.e. s2 , . . . , sk . 0→t 0→t Unfortunately, inference and learning in this network (and in general in coupled Markov chains) requires very expensive computations, and cannot be performed by simply propagating messages over time and among the variable nodes. In particular, the state of a child k variable xt depends on xk , sk , x1 and the state of all other children at the previous t t t−dt time step, [xj ]2
4 0.75467795 157 nips-2004-Saliency-Driven Image Acuity Modulation on a Reconfigurable Array of Spiking Silicon Neurons
Author: R. J. Vogelstein, Udayan Mallik, Eugenio Culurciello, Gert Cauwenberghs, Ralph Etienne-Cummings
Abstract: We have constructed a system that uses an array of 9,600 spiking silicon neurons, a fast microcontroller, and digital memory, to implement a reconfigurable network of integrate-and-fire neurons. The system is designed for rapid prototyping of spiking neural networks that require high-throughput communication with external address-event hardware. Arbitrary network topologies can be implemented by selectively routing address-events to specific internal or external targets according to a memory-based projective field mapping. The utility and versatility of the system is demonstrated by configuring it as a three-stage network that accepts input from an address-event imager, detects salient regions of the image, and performs spatial acuity modulation around a high-resolution fovea that is centered on the location of highest salience. 1
Author: Baranidharan Raman, Ricardo Gutierrez-osuna
Abstract: This paper presents a neuromorphic model of two olfactory signalprocessing primitives: chemotopic convergence of olfactory receptor neurons, and center on-off surround lateral inhibition in the olfactory bulb. A self-organizing model of receptor convergence onto glomeruli is used to generate a spatially organized map, an olfactory image. This map serves as input to a lattice of spiking neurons with lateral connections. The dynamics of this recurrent network transforms the initial olfactory image into a spatio-temporal pattern that evolves and stabilizes into odor- and intensity-coding attractors. The model is validated using experimental data from an array of temperature-modulated gas sensors. Our results are consistent with recent neurobiological findings on the antennal lobe of the honeybee and the locust. 1 In trod u ction An artificial olfactory system comprises of an array of cross-selective chemical sensors followed by a pattern recognition engine. An elegant alternative for the processing of sensor-array signals, normally performed with statistical pattern recognition techniques [1], involves adopting solutions from the biological olfactory system. The use of neuromorphic approaches provides an opportunity for formulating new computational problems in machine olfaction, including mixture segmentation, background suppression, olfactory habituation, and odor-memory associations. A biologically inspired approach to machine olfaction involves (1) identifying key signal processing primitives in the olfactory pathway, (2) adapting these primitives to account for the unique properties of chemical sensor signals, and (3) applying the models to solving specific computational problems. The biological olfactory pathway can be divided into three general stages: (i) olfactory epithelium, where primary reception takes place, (ii) olfactory bulb (OB), where the bulk of signal processing is performed and, (iii) olfactory cortex, where odor associations are stored. A review of literature on olfactory signal processing reveals six key primitives in the olfactory pathway that can be adapted for use in machine olfaction. These primitives are: (a) chemical transduction into a combinatorial code by a large population of olfactory receptor neurons (ORN), (b) chemotopic convergence of ORN axons onto glomeruli (GL), (c) logarithmic compression through lateral inhibition at the GL level by periglomerular interneurons, (d) contrast enhancement through lateral inhibition of mitral (M) projection neurons by granule interneurons, (e) storage and association of odor memories in the piriform cortex, and (f) bulbar modulation through cortical feedback [2, 3]. This article presents a model that captures the first three abovementioned primitives: population coding, chemotopic convergence and contrast enhancement. The model operates as follows. First, a large population of cross-selective pseudosensors is generated from an array of metal-oxide (MOS) gas sensors by means of temperature modulation. Next, a self-organizing model of convergence is used to cluster these pseudo-sensors according to their relative selectivity. This clustering generates an initial spatial odor map at the GL layer. Finally, a lattice of spiking neurons with center on-off surround lateral connections is used to transform the GL map into identity- and intensity-specific attractors. The model is validated using a database of temperature-modulated sensor patterns from three analytes at three concentration levels. The model is shown to address the first problem in biologically-inspired machine olfaction: intensity and identity coding of a chemical stimulus in a manner consistent with neurobiology [4, 5]. 2 M o d e l i n g c h e m o t opi c c o n v e r g e n c e The projection of sensory signals onto the olfactory bulb is organized such that ORNs expressing the same receptor gene converge onto one or a few GLs [3]. This convergence transforms the initial combinatorial code into an organized spatial pattern (i.e., an olfactory image). In addition, massive convergence improves the signal to noise ratio by integrating signals from multiple receptor neurons [6]. When incorporating this principle into machine olfaction, a fundamental difference between the artificial and biological counterparts must be overcome: the input dimensionality at the receptor/sensor level. The biological olfactory system employs a large population of ORNs (over 100 million in humans, replicated from 1,000 primary receptor types), whereas its artificial analogue uses a few chemical sensors (commonly one replica of up to 32 different sensor types). To bridge this gap, we employ a sensor excitation technique known as temperature modulation [7]. MOS sensors are conventionally driven in an isothermal fashion by maintaining a constant temperature. However, the selectivity of these devices is a function of the operating temperature. Thus, capturing the sensor response at multiple temperatures generates a wealth of additional information as compared to the isothermal mode of operation. If the temperature is modulated slow enough (e.g., mHz), the behavior of the sensor at each point in the temperature cycle can then be treated as a pseudo-sensor, and thus used to simulate a large population of cross-selective ORNs (refer to Figure 1(a)). To model chemotopic convergence, these temperature-modulated pseudo-sensors (referred to as ORNs in what follows) must be clustered according to their selectivity [8]. As a first approximation, each ORN can be modeled by an affinity vector [9] consisting of the responses across a set of C analytes: r K i = K i1 , K i2 ,..., K iC (1) [ ] where K ia is the response of the ith ORN to analyte a. The selectivity of this ORN r is then defined by the orientation of the affinity vector Κ i . A close look at the OB also shows that neighboring GLs respond to similar odors [10]. Therefore, we model the ORN-GL projection with a Kohonen self-organizing map (SOM) [11]. In our model, the SOM is trained to model the distribution of r ORNs in chemical sensitivity space, defined by the affinity vector Κ i . Once the training of the SOM is completed, each ORN is assigned to the closest SOM node (a simulated GL) in affinity space, thereby forming a convergence map. The response of each GL can then be computed as Ga = σ j (∑ N i =1 Wij ⋅ ORN ia ) (2) where ORN ia is the response of pseudo-sensor i to analyte a, Wij=1 if pseudo-sensor i converges to GL j and zero otherwise, and σ (⋅) is a squashing sigmoidal function that models saturation. This convergence model works well under the assumption that the different sensory inputs are reasonably uncorrelated. Unfortunately, most gas sensors are extremely collinear. As a result, this convergence model degenerates into a few dominant GLs that capture most of the sensory activity, and a large number of dormant GLs that do not receive any projections. To address this issue, we employ a form of competition known as conscience learning [12], which incorporates a habituation mechanism to prevent certain SOM nodes from dominating the competition. In this scheme, the fraction of times that a particular SOM node wins the competition is used as a bias to favor non-winning nodes. This results in a spreading of the ORN projections to neighboring units and, therefore, significantly reduces the number of dormant units. We measure the performance of the convergence mapping with the entropy across the lattice, H = −∑ Pi log Pi , where Pi is the fraction of ORNs that project to SOM node i [13]. To compare Kohonen and conscience learning, we built convergence mappings with 3,000 pseudo-sensors and 400 GL units (refer to section 4 for details). The theoretical maximum of the entropy for this network, which corresponds to a uniform distribution, is 8.6439. When trained with Kohonen’s algorithm, the entropy of the SOM is 7.3555. With conscience learning, the entropy increases to 8.2280. Thus, conscience is an effective mechanism to improve the spreading of ORN projections across the GL lattice. 3 M o d e l i n g t h e o l f a c t o r y b u l b n e t wo r k Mitral cells, which synapse ORNs at the GL level, transform the initial olfactory image into a spatio-temporal code by means of lateral inhibition. Two roles have been suggested for this lateral inhibition: (a) sharpening of the molecular tuning range of individual M cells with respect to that of their corresponding ORNs [10], and (b) global redistribution of activity, such that the bulb-wide representation of an odorant, rather than the individual tuning ranges, becomes specific and concise over time [3]. More recently, center on-off surround inhibitory connections have been found in the OB [14]. These circuits have been suggested to perform pattern normalization, noise reduction and contrast enhancement of the spatial patterns. We model each M cell using a leaky integrate-and-fire spiking neuron [15]. The input current I(t) and change in membrane potential u(t) of a neuron are given by: I (t ) = du u (t ) +C dt R (3) du τ = −u (t ) + R ⋅ I (t ) [τ = RC ] dt Each M cell receives current Iinput from ORNs and current Ilateral from lateral connections with other M cells: I input ( j ) = ∑Wij ⋅ ORNi i (4) I lateral ( j , t ) = ∑ Lkj ⋅ α (k , t − 1) k where Wij indicates the presence/absence of a synapse between ORNi and Mj, as determined by the chemotopic mapping, Lkj is the efficacy of the lateral connection between Mk and Mj, and α(k,t-1) is the post-synaptic current generated by a spike at Mk: α (k , t − 1) = − g (k , t − 1) ⋅ [u ( j, t − 1) + − Esyn ] (5) g(k,t-1) is the conductance of the synapse between Mk and Mj at time t-1, u(j,t-1) is the membrane potential of Mj at time t-1 and the + subscript indicates this value becomes zero if negative, and Esyn is the reverse synaptic potential. The change in conductance of post-synaptic membrane is: & g (k , t ) = dg (k , t ) − g (k , t ) = + z (k , t ) dt τ syn & z (k , t ) = dz (k , t ) − z ( k , t ) = + g norm ⋅ spk ( k , t ) dt τ syn (6) where z(.) and g(.) are low pass filters of the form exp(-t/τsyn) and t ⋅ exp(−t / τ syn ) , respectively, τsyn is the synaptic time constant, gnorm is a normalization constant, and spk(j,t) marks the occurrence of a spike in neuron i at time t: 1 u ( j , t ) = Vspike spk ( j , t ) = 0 u ( j , t ) ≠ Vspike (7) Combining equations (3) and (4), the membrane potential can be expressed as: du ( j , t ) − u ( j, t ) I lateral ( j, t ) I input ( j ) = + + dt RC C C & u ( j , t − 1) + u ( j , t − 1) ⋅ dt u ( j, t ) < Vthreshold u ( j, t ) = Vspike u ( j, t ) ≥ Vthreshold & u ( j, t ) = (8) When the membrane potential reaches Vthreshold, a spike is generated, and the membrane potential is reset to Vrest. Any further inputs to the neuron are ignored during the subsequent refractory period. Following [14], lateral interactions are modeled with a center on-off surround matrix Lij. Each M cell makes excitatory synapses to nearby M cells (d
6 0.69266075 194 nips-2004-Theory of localized synfire chain: characteristic propagation speed of stable spike pattern
7 0.58671945 84 nips-2004-Inference, Attention, and Decision in a Bayesian Neural Architecture
8 0.58018661 153 nips-2004-Reducing Spike Train Variability: A Computational Theory Of Spike-Timing Dependent Plasticity
9 0.57209629 181 nips-2004-Synergies between Intrinsic and Synaptic Plasticity in Individual Model Neurons
10 0.56649768 151 nips-2004-Rate- and Phase-coded Autoassociative Memory
11 0.52234221 112 nips-2004-Maximising Sensitivity in a Spiking Network
12 0.47774014 118 nips-2004-Methods for Estimating the Computational Power and Generalization Capability of Neural Microcircuits
13 0.45379215 58 nips-2004-Edge of Chaos Computation in Mixed-Mode VLSI - A Hard Liquid
14 0.45226014 46 nips-2004-Constraining a Bayesian Model of Human Visual Speed Perception
15 0.43534055 173 nips-2004-Spike-timing Dependent Plasticity and Mutual Information Maximization for a Spiking Neuron Model
16 0.41240057 148 nips-2004-Probabilistic Computation in Spiking Populations
17 0.37970939 26 nips-2004-At the Edge of Chaos: Real-time Computations and Self-Organized Criticality in Recurrent Neural Networks
18 0.36434457 86 nips-2004-Instance-Specific Bayesian Model Averaging for Classification
19 0.34111336 128 nips-2004-Neural Network Computation by In Vitro Transcriptional Circuits
20 0.3404094 25 nips-2004-Assignment of Multiplicative Mixtures in Natural Images
topicId topicWeight
[(13, 0.086), (15, 0.152), (26, 0.067), (31, 0.025), (33, 0.149), (35, 0.083), (50, 0.027), (82, 0.304)]
simIndex simValue paperId paperTitle
1 0.87253338 122 nips-2004-Modelling Uncertainty in the Game of Go
Author: David H. Stern, Thore Graepel, David MacKay
Abstract: Go is an ancient oriental game whose complexity has defeated attempts to automate it. We suggest using probability in a Bayesian sense to model the uncertainty arising from the vast complexity of the game tree. We present a simple conditional Markov random field model for predicting the pointwise territory outcome of a game. The topology of the model reflects the spatial structure of the Go board. We describe a version of the Swendsen-Wang process for sampling from the model during learning and apply loopy belief propagation for rapid inference and prediction. The model is trained on several hundred records of professional games. Our experimental results indicate that the model successfully learns to predict territory despite its simplicity. 1
same-paper 2 0.84351039 76 nips-2004-Hierarchical Bayesian Inference in Networks of Spiking Neurons
Author: Rajesh P. Rao
Abstract: There is growing evidence from psychophysical and neurophysiological studies that the brain utilizes Bayesian principles for inference and decision making. An important open question is how Bayesian inference for arbitrary graphical models can be implemented in networks of spiking neurons. In this paper, we show that recurrent networks of noisy integrate-and-fire neurons can perform approximate Bayesian inference for dynamic and hierarchical graphical models. The membrane potential dynamics of neurons is used to implement belief propagation in the log domain. The spiking probability of a neuron is shown to approximate the posterior probability of the preferred state encoded by the neuron, given past inputs. We illustrate the model using two examples: (1) a motion detection network in which the spiking probability of a direction-selective neuron becomes proportional to the posterior probability of motion in a preferred direction, and (2) a two-level hierarchical network that produces attentional effects similar to those observed in visual cortical areas V2 and V4. The hierarchical model offers a new Bayesian interpretation of attentional modulation in V2 and V4. 1
3 0.791789 188 nips-2004-The Laplacian PDF Distance: A Cost Function for Clustering in a Kernel Feature Space
Author: Robert Jenssen, Deniz Erdogmus, Jose Principe, Torbjørn Eltoft
Abstract: A new distance measure between probability density functions (pdfs) is introduced, which we refer to as the Laplacian pdf distance. The Laplacian pdf distance exhibits a remarkable connection to Mercer kernel based learning theory via the Parzen window technique for density estimation. In a kernel feature space defined by the eigenspectrum of the Laplacian data matrix, this pdf distance is shown to measure the cosine of the angle between cluster mean vectors. The Laplacian data matrix, and hence its eigenspectrum, can be obtained automatically based on the data at hand, by optimal Parzen window selection. We show that the Laplacian pdf distance has an interesting interpretation as a risk function connected to the probability of error. 1
4 0.69410998 28 nips-2004-Bayesian inference in spiking neurons
Author: Sophie Deneve
Abstract: We propose a new interpretation of spiking neurons as Bayesian integrators accumulating evidence over time about events in the external world or the body, and communicating to other neurons their certainties about these events. In this model, spikes signal the occurrence of new information, i.e. what cannot be predicted from the past activity. As a result, firing statistics are close to Poisson, albeit providing a deterministic representation of probabilities. We proceed to develop a theory of Bayesian inference in spiking neural networks, recurrent interactions implementing a variant of belief propagation. Many perceptual and motor tasks performed by the central nervous system are probabilistic, and can be described in a Bayesian framework [4, 3]. A few important but hidden properties, such as direction of motion, or appropriate motor commands, are inferred from many noisy, local and ambiguous sensory cues. These evidences are combined with priors about the sensory world and body. Importantly, because most of these inferences should lead to quick and irreversible decisions in a perpetually changing world, noisy cues have to be integrated on-line, but in a way that takes into account unpredictable events, such as a sudden change in motion direction or the appearance of a new stimulus. This raises the question of how this temporal integration can be performed at the neural level. It has been proposed that single neurons in sensory cortices represent and compute the log probability that a sensory variable takes on a certain value (eg Is visual motion in the neuron’s preferred direction?) [9, 7]. Alternatively, to avoid normalization issues and provide an appropriate signal for decision making, neurons could represent the log probability ratio of a particular hypothesis (eg is motion more likely to be towards the right than towards the left) [7, 6]. Log probabilities are convenient here, since under some assumptions, independent noisy cues simply combine linearly. Moreover, there are physiological evidence for the neural representation of log probabilities and log probability ratios [9, 6, 7]. However, these models assume that neurons represent probabilities in their firing rates. We argue that it is important to study how probabilistic information are encoded in spikes. Indeed, it seems spurious to marry the idea of an exquisite on-line integration of noisy cues with an underlying rate code that requires averaging on large populations of noisy neurons and long periods of time. In particular, most natural tasks require this integration to take place on the time scale of inter-spike intervals. Spikes are more efficiently signaling events ∗ Institute of Cognitive Science, 69645 Bron, France than analog quantities. In addition, a neural theory of inference with spikes will bring us closer to the physiological level and generate more easily testable predictions. Thus, we propose a new theory of neural processing in which spike trains provide a deterministic, online representation of a log-probability ratio. Spikes signals events, eg that the log-probability ratio has exceeded what could be predicted from previous spikes. This form of coding was loosely inspired by the idea of ”energy landscape” coding proposed by Hinton and Brown [2]. However, contrary to [2] and other theories using rate-based representation of probabilities, this model is self-consistent and does not require different models for encoding and decoding: As output spikes provide new, unpredictable, temporally independent evidence, they can be used directly as an input to other Bayesian neurons. Finally, we show that these neurons can be used as building blocks in a theory of approximate Bayesian inference in recurrent spiking networks. Connections between neurons implement an underlying Bayesian network, consisting of coupled hidden Markov models. Propagation of spikes is a form of belief propagation in this underlying graphical model. Our theory provides computational explanations of some general physiological properties of cortical neurons, such as spike frequency adaptation, Poisson statistics of spike trains, the existence of strong local inhibition in cortical columns, and the maintenance of a tight balance between excitation and inhibition. Finally, we discuss the implications of this model for the debate about temporal versus rate-based neural coding. 1 Spikes and log posterior odds 1.1 Synaptic integration seen as inference in a hidden Markov chain We propose that each neuron codes for an underlying ”hidden” binary variable, xt , whose state evolves over time. We assume that xt depends only on the state at the previous time step, xt−dt , and is conditionally independent of other past states. The state xt can switch 1 from 0 to 1 with a constant rate ron = dt limdt→0 P (xt = 1|xt−dt = 0), and from 1 to 0 with a constant rate roff . For example, these transition rates could represent how often motion in a preferred direction appears the receptive field and how long it is likely to stay there. The neuron infers the state of its hidden variable from N noisy synaptic inputs, considered to be observations of the hidden state. In this initial version of the model, we assume that these inputs are conditionally independent homogeneous Poisson processes, synapse i i emitting a spike between time t and t + dt (si = 1) with constant probability qon dt if t i xt = 1, and another constant probability qoff dt if xt = 0. The synaptic spikes are assumed to be otherwise independent of previous synaptic spikes, previous states and spikes at other synapses. The resulting generative model is a hidden Markov chain (figure 1-A). However, rather than estimating the state of its hidden variable and communicating this estimate to other neurons (for example by emitting a spike when sensory evidence for xt = 1 goes above a threshold) the neuron reports and communicates its certainty that the current state is 1. This certainty takes the form of the log of the ratio of the probability that the hidden state is 1, and the probability that the state is 0, given all the synaptic inputs P (xt =1|s0→t ) received so far: Lt = log P (xt =0|s0→t ) . We use s0→t as a short hand notation for the N synaptic inputs received at present and in the past. We will refer to it as the log odds ratio. Thanks to the conditional independencies assumed in the generative model, we can compute this Log odds ratio iteratively. Taking the limit as dt goes to zero, we get the following differential equation: ˙ L = ron 1 + e−L − roff 1 + eL + i wi δ(si − 1) − θ t B. A. xt ron .roff dt qon , qoff st xt ron .roff i t st dt s qon , qoff qon , qoff st dt xt j st Ot It Gt Ot Lt t t dt C. E. 2 0 -2 -4 D. 500 1000 1500 2000 2500 2 3000 Count Log odds 4 20 Lt 0 -2 0 500 1000 1500 2000 2500 Time Ot 3000 0 200 400 600 ISI Figure 1: A. Generative model for the synaptic input. B. Schematic representation of log odds ratio encoding and decoding. The dashed circle represents both eventual downstream elements and the self-prediction taking place inside the model neuron. A spike is fired only when Lt exceeds Gt . C. One example trial, where the state switches from 0 to 1 (shaded area) and back to 0. plain: Lt , dotted: Gt . Black stripes at the top: corresponding spikes train. D. Mean Log odds ratio (dark line) and mean output firing rate (clear line). E. Output spike raster plot (1 line per trial) and ISI distribution for the neuron shown is C. and D. Clear line: ISI distribution for a poisson neuron with the same rate. wi , the synaptic weight, describe how informative synapse i is about the state of the hidden i qon variable, e.g. wi = log qi . Each synaptic spike (si = 1) gives an impulse to the log t off odds ratio, which is positive if this synapse is more active when the hidden state if 1 (i.e it increases the neuron’s confidence that the state is 1), and negative if this synapse is more active when xt = 0 (i.e it decreases the neuron’s confidence that the state is 1). The bias, θ, is determined by how informative it is not to receive any spike, e.g. θ = i i i qon − qoff . By convention, we will consider that the ”bias” is positive or zero (if not, we need simply to invert the status of the state x). 1.2 Generation of output spikes The spike train should convey a sparse representation of Lt , so that each spike reports new information about the state xt that is not redundant with that reported by other, preceding, spikes. This proposition is based on three arguments: First, spikes, being metabolically expensive, should be kept to a minimum. Second, spikes conveying redundant information would require a decoding of the entire spike train, whereas independent spike can be taken into account individually. And finally, we seek a self consistent model, with the spiking output having a similar semantics to its spiking input. To maximize the independence of the spikes (conditioned on xt ), we propose that the neuron fires only when the difference between its log odds ratio Lt and a prediction Gt of this log odds ratio based on the output spikes emitted so far reaches a certain threshold. Indeed, supposing that downstream elements predicts Lt as best as they can, the neuron only needs to fire when it expects that prediction to be too inaccurate (figure 1-B). In practice, this will happen when the neuron receives new evidence for xt = 1. Gt should thereby follow the same dynamics as Lt when spikes are not received. The equation for Gt and the output Ot (Ot = 1 when an output spike is fired) are given by: ˙ G = Ot = ron 1 + e−L − roff 1 + eL + go δ(Ot − 1) go 1. when Lt > Gt + , 0 otherwise, 2 (1) (2) Here go , a positive constant, is the only free parameter, the other parameters being constrained by the statistics of the synaptic input. 1.3 Results Figure 1-C plots a typical trial, showing the behavior of L, G and O before, during and after presentation of the stimulus. As random synaptic inputs are integrated, L fluctuates and eventually exceeds G + 0.5, leading to an output spike. Immediately after a spike, G jumps to G + go , which prevents (except in very rare cases) a second spike from immediately following the first. Thus, this ”jump” implements a relative refractory period. However, ron G decays as it tends to converge back to its stable level gstable = log roff . Thus L eventually exceeds G again, leading to a new spike. This threshold crossing happens more often during stimulation (xt = 1) as the net synaptic input alters to create a higher overall level of certainty, Lt . Mean Log odds ratio and output firing rate ¯ The mean firing rate Ot of the Bayesian neuron during presentation of its preferred stimulus (i.e. when xt switches from 0 to 1 and back to 0) is plotted in figure 1-D, together with the ¯ mean log posterior ratio Lt , both averaged over trials. Not surprisingly, the log-posterior ratio reflects the leaky integration of synaptic evidence, with an effective time constant that depends on the transition probabilities ron , roff . If the state is very stable (ron = roff ∼ 0), synaptic evidence is integrated over almost infinite time periods, the mean log posterior ratio tending to either increase or decrease linearly with time. In the example in figure 1D, the state is less stable, so ”old” synaptic evidence are discounted and Lt saturates. ¯ In contrast, the mean output firing rate Ot tracks the state of xt almost perfectly. This is because, as a form of predictive coding, the output spikes reflect the new synaptic i evidence, It = i δ(st − 1) − θ, rather than the log posterior ratio itself. In particular, the mean output firing rate is a rectified linear function of the mean input, e. g. + ¯ ¯ wi q i −θ . O= 1I= go i on(off) Analogy with a leaky integrate and fire neuron We can get an interesting insight into the computation performed by this neuron by linearizing L and G around their mean levels over trials. Here we reduce the analysis to prolonged, statistically stable periods when the state is constant (either ON or OFF). In this case, the ¯ ¯ mean level of certainty L and its output prediction G are also constant over time. We make the rough approximation that the post spike jump, go , and the input fluctuations are small ¯ compared to the mean level of certainty L. Rewriting Vt = Lt − Gt + go 2 as the ”membrane potential” of the Bayesian neuron: ˙ V = −kL V + It − ∆go − go Ot ¯ ¯ ¯ where kL = ron e−L + roff eL , the ”leak” of the membrane potential, depends on the overall ¯ level of certainty. ∆go is positive and a monotonic increasing function of go . A. s t1 dt s t1 s t1 dt B. C. x t1 x t3 dt x t3 x t3 dt x t1 x t1 x t1 x t2 x t3 x t1 … x tn x t3 x t2 … x tn … dt dt Lx2 D. x t2 dt s t2 dt x t2 s t2 x t2 dt s t2 dt Log odds 10 No inh -0.5 -1 -1 -1.5 -2 5 Feedback 500 1000 1500 2000 Tiger Stripes 0 -5 -10 500 1000 1500 2000 2500 Time Figure 2: A. Bayesian causal network for yt (tiger), x1 (stripes) and x2 (paws). B. A nett t work feedforward computing the log posterior for x1 . C. A recurrent network computing t the log posterior odds for all variables. D. Log odds ratio in a simulated trial with the net2 1 1 work in C (see text). Thick line: Lx , thin line: Lx , dash-dotted: Lx without inhibition. t t t 2 Insert: Lx averaged over trials, showing the effect of feedback. t The linearized Bayesian neuron thus acts in its stable regime as a leaky integrate and fire (LIF) neuron. The membrane potential Vt integrates its input, Jt = It − ∆go , with a leak kL . The neuron fires when its membrane potential reaches a constant threshold go . After ¯ each spikes, Vt is reset to 0. Interestingly, for appropriately chosen compression factor go , the mean input to the lin¯ ¯ earized neuron J = I − ∆go ≈ 0 1 . This means that the membrane potential is purely driven to its threshold by input fluctuations, or a random walk in membrane potential. As a consequence, the neuron’s firing will be memoryless, and close to a Poisson process. In particular, we found Fano factor close to 1 and quasi-exponential ISI distribution (figure 1E) on the entire range of parameters tested. Indeed, LIF neurons with balanced inputs have been proposed as a model to reproduce the statistics of real cortical neurons [8]. This balance is implemented in our model by the neuron’s effective self-inhibition, even when the synaptic input itself is not balanced. Decoding As we previously said, downstream elements could predict the log odds ratio Lt by computing Gt from the output spikes (Eq 1, fig 1-B). Of course, this requires an estimate of the transition probabilities ron , roff , that could be learned from the observed spike trains. However, we show next that explicit decoding is not necessary to perform bayesian inference in spiking networks. Intuitively, this is because the quantity that our model neurons receive and transmit, eg new information, is exactly what probabilistic inference algorithm propagate between connected statistical elements. 1 ¯ Even if go is not chosen optimally, the influence of the drift J is usually negligible compared to the large fluctuations in membrane potential. 2 Bayesian inference in cortical networks The model neurons, having the same input and output semantics, can be used as building blocks to implement more complex generative models consisting of coupled Markov chains. Consider, for example, the example in figure 2-A. Here, a ”parent” variable x1 t (the presence of a tiger) can cause the state of n other ”children” variables ([xk ]k=2...n ), t of whom two are represented (the presence of stripes,x2 , and motion, x3 ). The ”chilt t dren” variables are Bayesian neurons identical to those described previously. The resulting bayesian network consist of n + 1 coupled hidden Markov chains. Inference in this architecture corresponds to computing the log posterior odds ratio for the tiger, x1 , and the log t posterior of observing stripes or motion, ([xk ]k=2...n ), given the synaptic inputs received t by the entire network so far, i.e. s2 , . . . , sk . 0→t 0→t Unfortunately, inference and learning in this network (and in general in coupled Markov chains) requires very expensive computations, and cannot be performed by simply propagating messages over time and among the variable nodes. In particular, the state of a child k variable xt depends on xk , sk , x1 and the state of all other children at the previous t t t−dt time step, [xj ]2
5 0.67754763 194 nips-2004-Theory of localized synfire chain: characteristic propagation speed of stable spike pattern
Author: Kosuke Hamaguchi, Masato Okada, Kazuyuki Aihara
Abstract: Repeated spike patterns have often been taken as evidence for the synfire chain, a phenomenon that a stable spike synchrony propagates through a feedforward network. Inter-spike intervals which represent a repeated spike pattern are influenced by the propagation speed of a spike packet. However, the relation between the propagation speed and network structure is not well understood. While it is apparent that the propagation speed depends on the excitatory synapse strength, it might also be related to spike patterns. We analyze a feedforward network with Mexican-Hattype connectivity (FMH) using the Fokker-Planck equation. We show that both a uniform and a localized spike packet are stable in the FMH in a certain parameter region. We also demonstrate that the propagation speed depends on the distinct firing patterns in the same network.
6 0.65719289 173 nips-2004-Spike-timing Dependent Plasticity and Mutual Information Maximization for a Spiking Neuron Model
7 0.6469878 148 nips-2004-Probabilistic Computation in Spiking Populations
8 0.64526927 181 nips-2004-Synergies between Intrinsic and Synaptic Plasticity in Individual Model Neurons
9 0.64115608 84 nips-2004-Inference, Attention, and Decision in a Bayesian Neural Architecture
10 0.63379109 151 nips-2004-Rate- and Phase-coded Autoassociative Memory
11 0.63310236 153 nips-2004-Reducing Spike Train Variability: A Computational Theory Of Spike-Timing Dependent Plasticity
12 0.63031149 112 nips-2004-Maximising Sensitivity in a Spiking Network
13 0.62649333 154 nips-2004-Resolving Perceptual Aliasing In The Presence Of Noisy Sensors
14 0.62430835 118 nips-2004-Methods for Estimating the Computational Power and Generalization Capability of Neural Microcircuits
15 0.61358333 1 nips-2004-A Cost-Shaping LP for Bellman Error Minimization with Performance Guarantees
16 0.61116689 131 nips-2004-Non-Local Manifold Tangent Learning
17 0.60866219 4 nips-2004-A Generalized Bradley-Terry Model: From Group Competition to Individual Skill
18 0.60810536 172 nips-2004-Sparse Coding of Natural Images Using an Overcomplete Set of Limited Capacity Units
19 0.60698664 189 nips-2004-The Power of Selective Memory: Self-Bounded Learning of Prediction Suffix Trees
20 0.60544157 16 nips-2004-Adaptive Discriminative Generative Model and Its Applications