nips nips2005 nips2005-164 knowledge-graph by maker-knowledge-mining
Source: pdf
Author: Viren Jain, Valentin Zhigulin, H. S. Seung
Abstract: There is little consensus about the computational function of top-down synaptic connections in the visual system. Here we explore the hypothesis that top-down connections, like bottom-up connections, reflect partwhole relationships. We analyze a recurrent network with bidirectional synaptic interactions between a layer of neurons representing parts and a layer of neurons representing wholes. Within each layer, there is lateral inhibition. When the network detects a whole, it can rigorously enforce part-whole relationships by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the theory of permitted and forbidden sets [3, 4]. The network behaviors are illustrated by recreating Rumelhart and McClelland’s “interactive activation” model [7]. In neural network models of visual object recognition [2, 6, 8], patterns of synaptic connectivity often reflect part-whole relationships between the features that are represented by neurons. For example, the connections of Figure 1 reflect the fact that feature B both contains simpler features A1, A2, and A3, and is contained in more complex features C1, C2, and C3. Such connectivity allows neurons to follow the rule that existence of the part is evidence for existence of the whole. By combining synaptic input from multiple sources of evidence for a feature, a neuron can “decide” whether that feature is present. 1 The synapses shown in Figure 1 are purely bottom-up, directed from simple to complex features. However, there are also top-down connections in the visual system, and there is little consensus about their function. One possibility is that top-down connections also reflect part-whole relationships. They allow feature detectors to make decisions using the rule that existence of the whole is evidence for existence of its parts. In this paper, we analyze the dynamics of a recurrent network in which part-whole relationships are stored as bidirectional synaptic interactions, rather than the unidirectional interactions of Figure 1. The network has a number of interesting computational capabilities. When the network detects a whole, it can rigorously enforce part-whole relationships 1 Synaptic connectivity may reflect other relationships besides part-whole. For example, invariances can be implemented by connecting detectors of several instances of the same feature to the same target, which is consequently an invariant detector of the feature. C1 C2 C3 B A1 A2 A3 Figure 1: The synaptic connections (arrows) of neuron B represent part-whole relationships. Feature B both contains simpler features and is contained in more complex features. The synaptic interactions are drawn one-way, as in most models of visual object recognition. Existence of the part is regarded as evidence for existence of the whole. This paper makes the interactions bidirectional, allowing the existence of the whole to be evidence for the existence of its parts. by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the recently developed theory of permitted and forbidden sets [3, 4]. Our model is closely related to the interactive activation model of word recognition, which was proposed by McClelland and Rumelhart to explain the word superiority effect studied by visual psychologists [7]. Here our concern is not to model a psychological effect, but to characterize mathematically how computations involving part-whole relationships can be carried out by a recurrent network. 1 Network model Suppose that we are given a set of part-whole relationships specified by a ξi = 1, if part i is contained in whole a 0, otherwise We assume that every whole contains at least one part, and every part is contained in at least one whole. The stimulus drives a layer of neurons that detect parts. These neurons also interact with a layer of neurons that detect wholes. We will refer to part-detectors as “P-neurons” and whole-detectors as “W-neurons.” The part-whole relationships are directly stored in the synaptic connections between P and a W neurons. If ξi = 1, the ith neuron in the P layer and the ath neuron in the W layer have a an excitatory interaction of strength γ. If ξi = 0, the neurons have an inhibitory interaction of strength σ. Furthermore, the P-neurons inhibit each other with strength β, and the Wneurons inhibit each other with strength α. All of these interactions are symmetric, and all activation functions are the rectification nonlinearity [z]+ = max{z, 0}. Then the dynamics of the network takes the form ˙ Wa + Wa a Pi ξ i − σ = γ i + a (1 − ξi )Pi − α i Wb , + ˙ Pi + Pi (1) b=a a Wa ξ i − σ = γ a a (1 − ξi )Wa − β a Pj + B i . j=i (2) where Bi is the input to the P layer from the stimulus. Figure 2 shows an example of a network with two wholes. Each whole contains two parts. One of the parts is contained in both wholes. -α Wa excitation γ -σ inhibition P1 B1 -β } W layer Wb -σ P2 -β B2 P3 } P layer B3 Figure 2: Model in example configuration: ξ = {(1, 1, 0), (0, 1, 1)}. When a stimulus is presented, it activates some of the P-neurons, which activate some of the W-neurons. The network eventually converges to a stable steady state. We will assume that α > 1. In the Appendix, we prove that this leads to unconditional winner-take-all behavior in the W layer. In other words, no more than one W-neuron can be active at a stable steady state. If a single W-neuron is active, then a whole has been detected. Potentially there are also many P-neurons active, indicating detection of parts. This representation may have different properties, depending on the choice of parameters β, γ, and σ. As discussed below, these include rigorous enforcement of part-whole relationships, completion of wholes by “filling in” missing parts, and non-recognition of parts that do not conform to a whole. 2 Enforcement of part-whole relationships Suppose that a single W-neuron is active at a stable steady state, so that a whole has been detected. Part-whole relationships are said to be enforced if the network always ignores parts that are not contained in the detected whole, despite potentially strong bottom-up evidence for them. It can be shown that enforcement follows from the inequality σ 2 + β 2 + γ 2 + 2σβγ > 1. (3) which guarantees that neuron i in the P layer is inactive, if neuron a in the W layer is a active and ξi = 0. When part-whole relations are enforced, prior knowledge about legal combinations of parts strictly constrains what may be perceived. This result is proven in the Appendix, and only an intuitive explanation is given here. Enforcement is easiest to understand when there is interlayer inhibition (σ > 0). In this case, the active W-neuron directly inhibits the forbidden P-neurons. The case of σ = 0 is more subtle. Then enforcement is mediated by lateral inhibition in the P layer. Excitatory feedback from the W-neuron has the effect of counteracting the lateral inhibition between the P-neurons that belong to the whole. As a result, these P-neurons become strongly activated enough to inhibit the rest of the P layer. 3 Completion of wholes by filling in missing parts If a W-neuron is active, it excites the P-neurons that belong to the whole. As a result, even if one of these P-neurons receives no bottom-up input (Bi = 0), it is still active. We call this phenomenon “completion,” and it is guaranteed to happen when (4) γ> β The network may thus “imagine” parts that are consistent with the recognized whole, but are not actually present in the stimulus. As with enforcement, this condition depends on top-down connections. √ In the special case γ = β, the interlayer excitation between a W-neuron and its P-neurons exactly cancels out the lateral inhibition between the P-neurons at a steady state. So the recurrent connections effectively vanish, letting the activity of the P-neurons be determined by their feedforward inputs. When the interlayer excitation is stronger than this, the inequality (4) holds, and completion occurs. 4 Non-recognition of a whole If there is no interlayer inhibition (σ = 0), then a single W-neuron is always active, assuming that there is some activity in the P layer. To see this, suppose for the sake of contradiction that all the W-neurons are inactive. Then they receive no inhibition to counteract the excitation from the P layer. This means some of them must be active, which contradicts our assumption. This means that the network always recognizes a whole, even if the stimulus is very different from any part-whole combination that is stored in the network. However, if interlayer inhibition is sufficiently strong (large σ), the network may refuse to recognize a whole. Neurons in the P layer are activated, but there is no activity in the W layer. Formal conditions on σ can be derived, but are not given here because of space limitations. In case of non-recognition, constraints on the P-layer are not enforced. It is possible for the network to detect a configuration of parts that is not consistent with any stored whole. 5 Example: Interactive Activation model To illustrate the computational capabilities of our network, we use it to recreate the interactive activation (IA) model of McClelland and Rumelhart. Figure 3 shows numerical simulations of a network containing three layers of neurons representing strokes, letters, and words, respectively. There are 16 possible strokes in each of four letter positions. For each stroke, there are two neurons, one signaling the presence of the stroke and the other signaling its absence. Letter neurons represent each letter of the alphabet in each of four positions. Word neurons represent each of 1200 common four letter words. The letter and word layers correspond to the P and W layers that were introduced previously. There are bidirectional interactions between the letter and word layers, and lateral inhibition within the layers. The letter neurons also receive input from the stroke neurons, but this interaction is unidirectional. Our network differs in two ways from the original IA model. First, all interactions involving letter and word neurons are symmetric. In the original model, the interactions between the letter and word layers were asymmetric. In particular, inhibitory connections only ran from letter neurons to word neurons, and not vice versa. Second, the only nonlinearity in our model is rectification. These two aspects allow us to apply the full machinery of the theory of permitted and forbidden sets. Figure 3 shows the result of presenting the stimulus “MO M” for four different settings of parameters. In each of the four cases, the word layer of the network converges to the same result, detecting the word “MOON”, which is the closest stored word to the stimulus. However, the activity in the letter layer is different in the four cases. input: P layer reconstruction W layer P layer reconstruction W layer completion noncompletion enforcement non-enforcement Figure 3: Simulation of 4 different parameter regimes in a letter- word recognition network. Within each panel, the middle column presents a feature- layer reconstruction based on the letter activity shown in the left column. W layer activity is shown in the right column. The top row shows the network state after 10 iterations of the dynamics. The bottom row shows the steady state. In the left column, the parameters obey the inequality (3), so that part- whole relationships are enforced. The activity of the letter layer is visualized by activating the strokes corresponding to each active letter neuron. The activated letters are part of the word “MOON”. In the top left, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom left, completion does not occur. In the simulations of the right column, parameters are such that part- whole relationships are not enforced. Consequently, the word layer is much more active. Bottom- up input provides evidence for several other letters, which is not suppressed. In the top right, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom right, the “O” neuron is not activated in the third position, so there is no completion. However, some letter neurons for the third position are activated, due to the input from neurons that indicate the absence of strokes. input: non-recognition event multi-stability Figure 4: Simulation of a non- recognition event and example of multistability. Figure 4 shows simulations for large σ, deep in the enforcement regime where non- recognition is a possibility. From one initial condition, the network converges to a state in which no W neurons are active, a non- recognition. From another initial condition, the network detects the word “NORM”. Deep in the enforcement regime, the top- down feedback can be so strong that the network has multiple stable states, many of which bear little resemblance to the stimulus at all. This is a problematic aspect of this network. It can be prevented by setting parameters at the edge of the enforcement regime. 6 Discussion We have analyzed a recurrent network that performs computations involving part- whole relationships. The network can fill in missing parts and suppress parts that do not belong. These two computations are distinct and can be dissociated from each other, as shown in Figure 3. While these two computations can also be performed by associative memory models, they are not typically dissociable in these models. For example, in the Hopfield model pattern completion and noise suppression are both the result of recall of one of a finite number of stereotyped activity patterns. We believe that our model is more appropriate for perceptual systems, because its behavior is piecewise linear, due its reliance on rectification nonlinearity. Therefore, analog aspects of computation are able to coexist with the part-whole relationships. Furthermore, in our model the stimulus is encoded in maintained synaptic input to the network, rather than as an initial condition of the dynamics. A Appendix: Permitted and forbidden sets Our mathematical results depend on the theory of permitted and forbidden sets [3, 4], which is summarized briefly here. The theory is applicable to neural networks with rectification nonlinearity, of the form xi + xi = [bi + j Wij xj ]+ . Neuron i is said to be active when ˙ xi > 0. For a network of N neurons, there are 2N possible sets of active neurons. For each active set, consider the submatrix of Wij corresponding to the synapses between active neurons. If all eigenvalues of this submatrix have real parts less than or equal to unity, then the active set is said to be permitted. Otherwise the active set is said to be forbidden. A set is permitted if and only if there exists an input vector b such that those neurons are active at a stable steady state. Permitted sets can be regarded as memories stored in the synaptic connections Wij . If Wij is a symmetric matrix, the nesting property holds: every subset of a permitted set is permitted, and every superset of a forbidden set is forbidden. The present model can be seen as a general method for storing permitted sets in a recurrent network. This method introduces a neuron for each permitted set, relying on a unary or “grandmother cell” representation. In contrast, Xie et al.[9] used lateral inhibition in a single layer of neurons to store permitted sets. By introducing extra neurons, the present model achieves superior storage capacity, much as unary models of associative memory [1] surpass distributed models [5]. A.1 Unconditional winner-take-all in the W layer The synapses between two W-neurons have strengths 0 −α −α 0 The eigenvalues of this matrix are ±α. Therefore two W-neurons constitute a forbidden set if α > 1. By the nesting property, it follows more than two W-neurons is also a forbidden set, and that the W layer has the unconditional winner-take-all property. A.2 Part-whole combinations as permitted sets Theorem 1. Suppose that β < 1. If γ 2 < β + (1 − β)/k then any combination of k ≥ 1 parts consistent with a whole corresponds to a permitted set. Proof. Consider k parts belonging to a whole. They are represented by one W-neuron and k P-neurons, with synaptic connections given by the (k + 1) × (k + 1) matrix M= −β(11T − I) γ1 , γ1T 0 (5) where 1 is the k- dimensional vector whose elements are all equal to one. Two eigenvectors of M are of the form (1T c), and have the same eigenvalues as the 2 × 2 matrix −β(k − 1) γk γ 0 This matrix has eigenvalues less than one when γ 2 < β + (1 − β)/k and β(k − 1) + 2 > 0. The other k − 1 eigenvectors are of the form (dT , 0), where dT 1 = 0. These have eigenvalues β. Therefore all eigenvalues of W are less than one if the condition of the theorem is satisfied. A.3 Constraints on combining parts Here, we derive conditions under which the network can enforce the constraint that steady state activity be confined to parts that constitute a whole. Theorem 2. Suppose that β > 0 and σ 2 +β 2 +γ 2 +2σβγ > 1 If a W- neuron is active, then only P- neurons corresponding to parts contained in the relevant whole can be active at a stable steady state. Proof. Consider P- neurons Pi , Pj , and W- neuron Wa . Supa a pose that ξi = 1 but ξj = 0. As shown in Figure 5, the matrix of connections is given by: W = 0 −β γ −β 0 −σ γ −σ 0 (6) Wa γ Pi -σ -β Pj Figure 5: A set of one W- neuron and two P- neurons is forbidden if one part belongs to the whole and the other does not. This set is permitted if all eigenvalues of W − I have negative real parts. The characteristic equation of I − W is λ3 + b1 λ2 + b2 λ + b3 = 0, where b1 = 3, b2 = 3 − σ 2 − β 2 − γ 2 and b3 = 1−2σβγ−σ 2 −β 2 −γ 2 . According to the Routh- Hurwitz theorem, all the eigenvalues have negative real parts if and only if b1 > 0, b3 > 0 and b1 b2 > b3 . Clearly, the first condition is always satisfied. The second condition is more restrictive than the third. It is satisfied only when σ 2 + β 2 + γ 2 + 2σβγ < 1. Hence, one of the eigenvalues has a positive real part when this condition is broken, i.e., when σ 2 +β 2 +γ 2 +2σβγ > 1. By the nesting property, any larger set of P- neurons inconsistent with the W- neuron is also forbidden. A.4 Completion of wholes √ Theorem 3. If γ > β and a single W- neuron a is active at a steady state, then Pi > 0 a for all i such that ξi = 1. Proof. Suppose that the detected whole has k parts. At the steady state Pi = a ξi Bi − (β − γ 2 )Ptot 1−β + where Ptot = Pi = i 1 1 − β + (β − γ 2 )k k a B i ξi i=1 (7) A.5 Preventing runaway If feedback loops cause the network activity to diverge, then the preceding analyses are not relevant. Here we give a sufficient condition guaranteeing that runaway instability does not happen. It is not a necessary condition. Interestingly, the condition implies the condition of Theorem 1. Theorem 4. Suppose that P and W obey the dynamics of Eqs. (1) and (2), and define the objective function E 1−α 2 = − 2 Wa a α + 2 Wa a 1−β + 2 a Pi Wa ξi + σ Bi Pi − γ i 2 ia Pi2 i 2 β + 2 Pi i a (1 − ξi )Pi Wa . (8) ia Then E is a Lyapunov like function that, given β > γ 2 − dynamics to a stable steady state. 1−γ 2 N −1 , ensures convergence of the Proof. (sketch) Differentiation of E with respect to time shows that that E is nonincreasing in the nonnegative orthant and constant only at steady states of the network dynamics. We must also show that E is radially unbounded, which is true if the quadratic part of E is copositive definite. Note that the last term of E is lower-bounded by zero and the previous term is upper bounded by γ ia Pi Wa . We assume α > 1. Thus, we can use Cauchy’s 2 2 inequality, i Pi2 ≥ ( i Pi ) /N , and the fact that a Wa ≤ ( a Wa )2 for Wa ≥ 0, to derive E≥ 1 2 Wa )2 + ( If β > γ 2 − unbounded. a 1−γ 2 N −1 , 1 − β + βN ( N Pi )2 − 2γ( i Wa a Pi ) i − Bi Pi . (9) i the quadratic form in the inequality is positive definite and E is radially References [1] E. B. Baum, J. Moody, and F. Wilczek. Internal representations for associative memory. Biol. Cybern., 59:217–228, 1988. [2] K. Fukushima. Neocognitron: a self organizing neural network model for a mechanism of pattern recognition unaffected by shift in position. Biol Cybern, 36(4):193–202, 1980. [3] R.H. Hahnloser, R. Sarpeshkar, M.A. Mahowald, R.J. Douglas, and H.S. Seung. Digital selection and analogue amplification coexist in a cortex-inspired silicon circuit. Nature, 405(6789):947– 51, Jun 22 2000. [4] R.H. Hahnloser, H.S. Seung, and J.-J. Slotine. Permitted and forbidden sets in symmetric threshold-linear networks. Neural Computation, 15:621–638, 2003. [5] J.J. Hopfield. Neural networks and physical systems with emergent collective computational abilities. Proc Natl Acad Sci U S A, 79(8):2554–8, Apr 1982. [6] Y. LeCun, B. Boser, J. S. Denker, D. Henderson, R. E. Howard, W. Hubbard, and L. D. Jackel. Backpropagation applied to handwritten zip code recognition. Neural Comput., 1:541–551, 1989. [7] J. L. McClelland and D. E. Rumelhart. An interactive activation model of context effects in letter perception: Part i. an account of basic findings. Psychological Review, 88(5):375–407, Sep 1981. [8] M Riesenhuber and T Poggio. Hierarchical models of object recognition in cortex. Nat Neurosci, 2(11):1019–25, Nov 1999. [9] X. Xie, R.H. Hahnloser, and H. S. Seung. Selectively grouping neurons in recurrent networks of lateral inhibition. Neural Computation, 14:2627–2646, 2002.
Reference: text
sentIndex sentText sentNum sentScore
1 edu Abstract There is little consensus about the computational function of top-down synaptic connections in the visual system. [sent-8, score-0.265]
2 We analyze a recurrent network with bidirectional synaptic interactions between a layer of neurons representing parts and a layer of neurons representing wholes. [sent-10, score-1.633]
3 When the network detects a whole, it can rigorously enforce part-whole relationships by ignoring parts that do not belong. [sent-12, score-0.528]
4 The network can complete the whole by filling in missing parts. [sent-13, score-0.376]
5 The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. [sent-14, score-0.642]
6 Parameter regimes in which these behaviors happen are identified using the theory of permitted and forbidden sets [3, 4]. [sent-15, score-0.674]
7 In neural network models of visual object recognition [2, 6, 8], patterns of synaptic connectivity often reflect part-whole relationships between the features that are represented by neurons. [sent-17, score-0.455]
8 Such connectivity allows neurons to follow the rule that existence of the part is evidence for existence of the whole. [sent-19, score-0.485]
9 By combining synaptic input from multiple sources of evidence for a feature, a neuron can “decide” whether that feature is present. [sent-20, score-0.286]
10 They allow feature detectors to make decisions using the rule that existence of the whole is evidence for existence of its parts. [sent-24, score-0.386]
11 In this paper, we analyze the dynamics of a recurrent network in which part-whole relationships are stored as bidirectional synaptic interactions, rather than the unidirectional interactions of Figure 1. [sent-25, score-0.727]
12 When the network detects a whole, it can rigorously enforce part-whole relationships 1 Synaptic connectivity may reflect other relationships besides part-whole. [sent-27, score-0.521]
13 C1 C2 C3 B A1 A2 A3 Figure 1: The synaptic connections (arrows) of neuron B represent part-whole relationships. [sent-29, score-0.35]
14 This paper makes the interactions bidirectional, allowing the existence of the whole to be evidence for the existence of its parts. [sent-33, score-0.476]
15 The network can complete the whole by filling in missing parts. [sent-35, score-0.376]
16 The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. [sent-36, score-0.642]
17 Parameter regimes in which these behaviors happen are identified using the recently developed theory of permitted and forbidden sets [3, 4]. [sent-37, score-0.674]
18 Our model is closely related to the interactive activation model of word recognition, which was proposed by McClelland and Rumelhart to explain the word superiority effect studied by visual psychologists [7]. [sent-38, score-0.501]
19 Here our concern is not to model a psychological effect, but to characterize mathematically how computations involving part-whole relationships can be carried out by a recurrent network. [sent-39, score-0.258]
20 1 Network model Suppose that we are given a set of part-whole relationships specified by a ξi = 1, if part i is contained in whole a 0, otherwise We assume that every whole contains at least one part, and every part is contained in at least one whole. [sent-40, score-0.651]
21 The stimulus drives a layer of neurons that detect parts. [sent-41, score-0.55]
22 These neurons also interact with a layer of neurons that detect wholes. [sent-42, score-0.681]
23 ” The part-whole relationships are directly stored in the synaptic connections between P and a W neurons. [sent-44, score-0.432]
24 If ξi = 1, the ith neuron in the P layer and the ath neuron in the W layer have a an excitatory interaction of strength γ. [sent-45, score-0.81]
25 If ξi = 0, the neurons have an inhibitory interaction of strength σ. [sent-46, score-0.249]
26 j=i (2) where Bi is the input to the P layer from the stimulus. [sent-50, score-0.267]
27 -α Wa excitation γ -σ inhibition P1 B1 -β } W layer Wb -σ P2 -β B2 P3 } P layer B3 Figure 2: Model in example configuration: ξ = {(1, 1, 0), (0, 1, 1)}. [sent-54, score-0.737]
28 The network eventually converges to a stable steady state. [sent-56, score-0.397]
29 In other words, no more than one W-neuron can be active at a stable steady state. [sent-59, score-0.399]
30 As discussed below, these include rigorous enforcement of part-whole relationships, completion of wholes by “filling in” missing parts, and non-recognition of parts that do not conform to a whole. [sent-63, score-0.664]
31 2 Enforcement of part-whole relationships Suppose that a single W-neuron is active at a stable steady state, so that a whole has been detected. [sent-64, score-0.689]
32 Part-whole relationships are said to be enforced if the network always ignores parts that are not contained in the detected whole, despite potentially strong bottom-up evidence for them. [sent-65, score-0.564]
33 It can be shown that enforcement follows from the inequality σ 2 + β 2 + γ 2 + 2σβγ > 1. [sent-66, score-0.249]
34 (3) which guarantees that neuron i in the P layer is inactive, if neuron a in the W layer is a active and ξi = 0. [sent-67, score-0.909]
35 In this case, the active W-neuron directly inhibits the forbidden P-neurons. [sent-71, score-0.405]
36 Then enforcement is mediated by lateral inhibition in the P layer. [sent-73, score-0.421]
37 Excitatory feedback from the W-neuron has the effect of counteracting the lateral inhibition between the P-neurons that belong to the whole. [sent-74, score-0.262]
38 3 Completion of wholes by filling in missing parts If a W-neuron is active, it excites the P-neurons that belong to the whole. [sent-76, score-0.295]
39 We call this phenomenon “completion,” and it is guaranteed to happen when (4) γ> β The network may thus “imagine” parts that are consistent with the recognized whole, but are not actually present in the stimulus. [sent-78, score-0.329]
40 √ In the special case γ = β, the interlayer excitation between a W-neuron and its P-neurons exactly cancels out the lateral inhibition between the P-neurons at a steady state. [sent-80, score-0.603]
41 So the recurrent connections effectively vanish, letting the activity of the P-neurons be determined by their feedforward inputs. [sent-81, score-0.281]
42 When the interlayer excitation is stronger than this, the inequality (4) holds, and completion occurs. [sent-82, score-0.352]
43 4 Non-recognition of a whole If there is no interlayer inhibition (σ = 0), then a single W-neuron is always active, assuming that there is some activity in the P layer. [sent-83, score-0.483]
44 This means that the network always recognizes a whole, even if the stimulus is very different from any part-whole combination that is stored in the network. [sent-87, score-0.295]
45 However, if interlayer inhibition is sufficiently strong (large σ), the network may refuse to recognize a whole. [sent-88, score-0.489]
46 Neurons in the P layer are activated, but there is no activity in the W layer. [sent-89, score-0.339]
47 It is possible for the network to detect a configuration of parts that is not consistent with any stored whole. [sent-92, score-0.376]
48 Figure 3 shows numerical simulations of a network containing three layers of neurons representing strokes, letters, and words, respectively. [sent-94, score-0.412]
49 There are 16 possible strokes in each of four letter positions. [sent-95, score-0.275]
50 Letter neurons represent each letter of the alphabet in each of four positions. [sent-97, score-0.425]
51 Word neurons represent each of 1200 common four letter words. [sent-98, score-0.425]
52 The letter and word layers correspond to the P and W layers that were introduced previously. [sent-99, score-0.527]
53 There are bidirectional interactions between the letter and word layers, and lateral inhibition within the layers. [sent-100, score-0.787]
54 The letter neurons also receive input from the stroke neurons, but this interaction is unidirectional. [sent-101, score-0.482]
55 First, all interactions involving letter and word neurons are symmetric. [sent-103, score-0.692]
56 In the original model, the interactions between the letter and word layers were asymmetric. [sent-104, score-0.551]
57 In particular, inhibitory connections only ran from letter neurons to word neurons, and not vice versa. [sent-105, score-0.709]
58 These two aspects allow us to apply the full machinery of the theory of permitted and forbidden sets. [sent-107, score-0.519]
59 In each of the four cases, the word layer of the network converges to the same result, detecting the word “MOON”, which is the closest stored word to the stimulus. [sent-109, score-1.017]
60 However, the activity in the letter layer is different in the four cases. [sent-110, score-0.557]
61 input: P layer reconstruction W layer P layer reconstruction W layer completion noncompletion enforcement non-enforcement Figure 3: Simulation of 4 different parameter regimes in a letter- word recognition network. [sent-111, score-1.705]
62 Within each panel, the middle column presents a feature- layer reconstruction based on the letter activity shown in the left column. [sent-112, score-0.587]
63 In the left column, the parameters obey the inequality (3), so that part- whole relationships are enforced. [sent-116, score-0.382]
64 The activity of the letter layer is visualized by activating the strokes corresponding to each active letter neuron. [sent-117, score-0.973]
65 The activated letters are part of the word “MOON”. [sent-118, score-0.308]
66 In the simulations of the right column, parameters are such that part- whole relationships are not enforced. [sent-121, score-0.29]
67 However, some letter neurons for the third position are activated, due to the input from neurons that indicate the absence of strokes. [sent-126, score-0.632]
68 Figure 4 shows simulations for large σ, deep in the enforcement regime where non- recognition is a possibility. [sent-128, score-0.265]
69 From one initial condition, the network converges to a state in which no W neurons are active, a non- recognition. [sent-129, score-0.346]
70 From another initial condition, the network detects the word “NORM”. [sent-130, score-0.354]
71 Deep in the enforcement regime, the top- down feedback can be so strong that the network has multiple stable states, many of which bear little resemblance to the stimulus at all. [sent-131, score-0.493]
72 6 Discussion We have analyzed a recurrent network that performs computations involving part- whole relationships. [sent-134, score-0.449]
73 The network can fill in missing parts and suppress parts that do not belong. [sent-135, score-0.519]
74 Furthermore, in our model the stimulus is encoded in maintained synaptic input to the network, rather than as an initial condition of the dynamics. [sent-141, score-0.257]
75 A Appendix: Permitted and forbidden sets Our mathematical results depend on the theory of permitted and forbidden sets [3, 4], which is summarized briefly here. [sent-142, score-0.849]
76 For a network of N neurons, there are 2N possible sets of active neurons. [sent-145, score-0.313]
77 For each active set, consider the submatrix of Wij corresponding to the synapses between active neurons. [sent-146, score-0.349]
78 If all eigenvalues of this submatrix have real parts less than or equal to unity, then the active set is said to be permitted. [sent-147, score-0.45]
79 A set is permitted if and only if there exists an input vector b such that those neurons are active at a stable steady state. [sent-149, score-0.861]
80 Permitted sets can be regarded as memories stored in the synaptic connections Wij . [sent-150, score-0.346]
81 If Wij is a symmetric matrix, the nesting property holds: every subset of a permitted set is permitted, and every superset of a forbidden set is forbidden. [sent-151, score-0.591]
82 The present model can be seen as a general method for storing permitted sets in a recurrent network. [sent-152, score-0.39]
83 This method introduces a neuron for each permitted set, relying on a unary or “grandmother cell” representation. [sent-153, score-0.414]
84 [9] used lateral inhibition in a single layer of neurons to store permitted sets. [sent-155, score-0.96]
85 1 Unconditional winner-take-all in the W layer The synapses between two W-neurons have strengths 0 −α −α 0 The eigenvalues of this matrix are ±α. [sent-158, score-0.379]
86 Therefore two W-neurons constitute a forbidden set if α > 1. [sent-159, score-0.299]
87 By the nesting property, it follows more than two W-neurons is also a forbidden set, and that the W layer has the unconditional winner-take-all property. [sent-160, score-0.653]
88 2 Part-whole combinations as permitted sets Theorem 1. [sent-162, score-0.288]
89 If γ 2 < β + (1 − β)/k then any combination of k ≥ 1 parts consistent with a whole corresponds to a permitted set. [sent-164, score-0.583]
90 3 Constraints on combining parts Here, we derive conditions under which the network can enforce the constraint that steady state activity be confined to parts that constitute a whole. [sent-173, score-0.801]
91 Suppose that β > 0 and σ 2 +β 2 +γ 2 +2σβγ > 1 If a W- neuron is active, then only P- neurons corresponding to parts contained in the relevant whole can be active at a stable steady state. [sent-175, score-1.114]
92 As shown in Figure 5, the matrix of connections is given by: W = 0 −β γ −β 0 −σ γ −σ 0 (6) Wa γ Pi -σ -β Pj Figure 5: A set of one W- neuron and two P- neurons is forbidden if one part belongs to the whole and the other does not. [sent-179, score-0.898]
93 This set is permitted if all eigenvalues of W − I have negative real parts. [sent-180, score-0.332]
94 By the nesting property, any larger set of P- neurons inconsistent with the W- neuron is also forbidden. [sent-189, score-0.396]
95 If γ > β and a single W- neuron a is active at a steady state, then Pi > 0 a for all i such that ξi = 1. [sent-192, score-0.459]
96 5 Preventing runaway If feedback loops cause the network activity to diverge, then the preceding analyses are not relevant. [sent-196, score-0.284]
97 (8) ia Then E is a Lyapunov like function that, given β > γ 2 − dynamics to a stable steady state. [sent-203, score-0.328]
98 (sketch) Differentiation of E with respect to time shows that that E is nonincreasing in the nonnegative orthant and constant only at steady states of the network dynamics. [sent-205, score-0.34]
99 An interactive activation model of context effects in letter perception: Part i. [sent-269, score-0.365]
100 Selectively grouping neurons in recurrent networks of lateral inhibition. [sent-281, score-0.404]
wordName wordTfidf (topN-words)
[('layer', 0.267), ('forbidden', 0.264), ('permitted', 0.255), ('wa', 0.224), ('letter', 0.218), ('pi', 0.212), ('neurons', 0.207), ('steady', 0.201), ('enforcement', 0.19), ('word', 0.177), ('whole', 0.171), ('parts', 0.157), ('active', 0.141), ('network', 0.139), ('inhibition', 0.136), ('synaptic', 0.126), ('completion', 0.122), ('relationships', 0.119), ('neuron', 0.117), ('connections', 0.107), ('interlayer', 0.104), ('recurrent', 0.102), ('activated', 0.099), ('lateral', 0.095), ('interactions', 0.09), ('interactive', 0.089), ('existence', 0.086), ('mcclelland', 0.083), ('stored', 0.08), ('eigenvalues', 0.077), ('stimulus', 0.076), ('activity', 0.072), ('nesting', 0.072), ('refuse', 0.072), ('wholes', 0.072), ('bidirectional', 0.071), ('lling', 0.071), ('ia', 0.07), ('excitation', 0.067), ('missing', 0.066), ('layers', 0.066), ('contained', 0.063), ('bi', 0.062), ('inequality', 0.059), ('recti', 0.058), ('activation', 0.058), ('hahnloser', 0.057), ('strokes', 0.057), ('stroke', 0.057), ('conform', 0.057), ('stable', 0.057), ('condition', 0.055), ('pj', 0.053), ('inhibit', 0.053), ('unconditional', 0.05), ('wij', 0.05), ('ptot', 0.048), ('valentin', 0.048), ('viren', 0.048), ('xie', 0.048), ('regimes', 0.048), ('said', 0.043), ('evidence', 0.043), ('ect', 0.043), ('strength', 0.042), ('unary', 0.042), ('runaway', 0.042), ('suppose', 0.041), ('behaviors', 0.041), ('recognition', 0.04), ('associative', 0.04), ('enforce', 0.04), ('detects', 0.038), ('nonlinearity', 0.038), ('recognize', 0.038), ('moon', 0.038), ('radially', 0.038), ('wb', 0.038), ('hop', 0.038), ('coexist', 0.038), ('computations', 0.037), ('appendix', 0.036), ('rigorously', 0.035), ('constitute', 0.035), ('deep', 0.035), ('synapses', 0.035), ('theorem', 0.034), ('rumelhart', 0.033), ('obey', 0.033), ('sets', 0.033), ('happen', 0.033), ('part', 0.032), ('submatrix', 0.032), ('consensus', 0.032), ('howard', 0.032), ('connectivity', 0.031), ('feedback', 0.031), ('signaling', 0.03), ('reconstruction', 0.03), ('lled', 0.029)]
simIndex simValue paperId paperTitle
same-paper 1 1.0000004 164 nips-2005-Representing Part-Whole Relationships in Recurrent Neural Networks
Author: Viren Jain, Valentin Zhigulin, H. S. Seung
Abstract: There is little consensus about the computational function of top-down synaptic connections in the visual system. Here we explore the hypothesis that top-down connections, like bottom-up connections, reflect partwhole relationships. We analyze a recurrent network with bidirectional synaptic interactions between a layer of neurons representing parts and a layer of neurons representing wholes. Within each layer, there is lateral inhibition. When the network detects a whole, it can rigorously enforce part-whole relationships by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the theory of permitted and forbidden sets [3, 4]. The network behaviors are illustrated by recreating Rumelhart and McClelland’s “interactive activation” model [7]. In neural network models of visual object recognition [2, 6, 8], patterns of synaptic connectivity often reflect part-whole relationships between the features that are represented by neurons. For example, the connections of Figure 1 reflect the fact that feature B both contains simpler features A1, A2, and A3, and is contained in more complex features C1, C2, and C3. Such connectivity allows neurons to follow the rule that existence of the part is evidence for existence of the whole. By combining synaptic input from multiple sources of evidence for a feature, a neuron can “decide” whether that feature is present. 1 The synapses shown in Figure 1 are purely bottom-up, directed from simple to complex features. However, there are also top-down connections in the visual system, and there is little consensus about their function. One possibility is that top-down connections also reflect part-whole relationships. They allow feature detectors to make decisions using the rule that existence of the whole is evidence for existence of its parts. In this paper, we analyze the dynamics of a recurrent network in which part-whole relationships are stored as bidirectional synaptic interactions, rather than the unidirectional interactions of Figure 1. The network has a number of interesting computational capabilities. When the network detects a whole, it can rigorously enforce part-whole relationships 1 Synaptic connectivity may reflect other relationships besides part-whole. For example, invariances can be implemented by connecting detectors of several instances of the same feature to the same target, which is consequently an invariant detector of the feature. C1 C2 C3 B A1 A2 A3 Figure 1: The synaptic connections (arrows) of neuron B represent part-whole relationships. Feature B both contains simpler features and is contained in more complex features. The synaptic interactions are drawn one-way, as in most models of visual object recognition. Existence of the part is regarded as evidence for existence of the whole. This paper makes the interactions bidirectional, allowing the existence of the whole to be evidence for the existence of its parts. by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the recently developed theory of permitted and forbidden sets [3, 4]. Our model is closely related to the interactive activation model of word recognition, which was proposed by McClelland and Rumelhart to explain the word superiority effect studied by visual psychologists [7]. Here our concern is not to model a psychological effect, but to characterize mathematically how computations involving part-whole relationships can be carried out by a recurrent network. 1 Network model Suppose that we are given a set of part-whole relationships specified by a ξi = 1, if part i is contained in whole a 0, otherwise We assume that every whole contains at least one part, and every part is contained in at least one whole. The stimulus drives a layer of neurons that detect parts. These neurons also interact with a layer of neurons that detect wholes. We will refer to part-detectors as “P-neurons” and whole-detectors as “W-neurons.” The part-whole relationships are directly stored in the synaptic connections between P and a W neurons. If ξi = 1, the ith neuron in the P layer and the ath neuron in the W layer have a an excitatory interaction of strength γ. If ξi = 0, the neurons have an inhibitory interaction of strength σ. Furthermore, the P-neurons inhibit each other with strength β, and the Wneurons inhibit each other with strength α. All of these interactions are symmetric, and all activation functions are the rectification nonlinearity [z]+ = max{z, 0}. Then the dynamics of the network takes the form ˙ Wa + Wa a Pi ξ i − σ = γ i + a (1 − ξi )Pi − α i Wb , + ˙ Pi + Pi (1) b=a a Wa ξ i − σ = γ a a (1 − ξi )Wa − β a Pj + B i . j=i (2) where Bi is the input to the P layer from the stimulus. Figure 2 shows an example of a network with two wholes. Each whole contains two parts. One of the parts is contained in both wholes. -α Wa excitation γ -σ inhibition P1 B1 -β } W layer Wb -σ P2 -β B2 P3 } P layer B3 Figure 2: Model in example configuration: ξ = {(1, 1, 0), (0, 1, 1)}. When a stimulus is presented, it activates some of the P-neurons, which activate some of the W-neurons. The network eventually converges to a stable steady state. We will assume that α > 1. In the Appendix, we prove that this leads to unconditional winner-take-all behavior in the W layer. In other words, no more than one W-neuron can be active at a stable steady state. If a single W-neuron is active, then a whole has been detected. Potentially there are also many P-neurons active, indicating detection of parts. This representation may have different properties, depending on the choice of parameters β, γ, and σ. As discussed below, these include rigorous enforcement of part-whole relationships, completion of wholes by “filling in” missing parts, and non-recognition of parts that do not conform to a whole. 2 Enforcement of part-whole relationships Suppose that a single W-neuron is active at a stable steady state, so that a whole has been detected. Part-whole relationships are said to be enforced if the network always ignores parts that are not contained in the detected whole, despite potentially strong bottom-up evidence for them. It can be shown that enforcement follows from the inequality σ 2 + β 2 + γ 2 + 2σβγ > 1. (3) which guarantees that neuron i in the P layer is inactive, if neuron a in the W layer is a active and ξi = 0. When part-whole relations are enforced, prior knowledge about legal combinations of parts strictly constrains what may be perceived. This result is proven in the Appendix, and only an intuitive explanation is given here. Enforcement is easiest to understand when there is interlayer inhibition (σ > 0). In this case, the active W-neuron directly inhibits the forbidden P-neurons. The case of σ = 0 is more subtle. Then enforcement is mediated by lateral inhibition in the P layer. Excitatory feedback from the W-neuron has the effect of counteracting the lateral inhibition between the P-neurons that belong to the whole. As a result, these P-neurons become strongly activated enough to inhibit the rest of the P layer. 3 Completion of wholes by filling in missing parts If a W-neuron is active, it excites the P-neurons that belong to the whole. As a result, even if one of these P-neurons receives no bottom-up input (Bi = 0), it is still active. We call this phenomenon “completion,” and it is guaranteed to happen when (4) γ> β The network may thus “imagine” parts that are consistent with the recognized whole, but are not actually present in the stimulus. As with enforcement, this condition depends on top-down connections. √ In the special case γ = β, the interlayer excitation between a W-neuron and its P-neurons exactly cancels out the lateral inhibition between the P-neurons at a steady state. So the recurrent connections effectively vanish, letting the activity of the P-neurons be determined by their feedforward inputs. When the interlayer excitation is stronger than this, the inequality (4) holds, and completion occurs. 4 Non-recognition of a whole If there is no interlayer inhibition (σ = 0), then a single W-neuron is always active, assuming that there is some activity in the P layer. To see this, suppose for the sake of contradiction that all the W-neurons are inactive. Then they receive no inhibition to counteract the excitation from the P layer. This means some of them must be active, which contradicts our assumption. This means that the network always recognizes a whole, even if the stimulus is very different from any part-whole combination that is stored in the network. However, if interlayer inhibition is sufficiently strong (large σ), the network may refuse to recognize a whole. Neurons in the P layer are activated, but there is no activity in the W layer. Formal conditions on σ can be derived, but are not given here because of space limitations. In case of non-recognition, constraints on the P-layer are not enforced. It is possible for the network to detect a configuration of parts that is not consistent with any stored whole. 5 Example: Interactive Activation model To illustrate the computational capabilities of our network, we use it to recreate the interactive activation (IA) model of McClelland and Rumelhart. Figure 3 shows numerical simulations of a network containing three layers of neurons representing strokes, letters, and words, respectively. There are 16 possible strokes in each of four letter positions. For each stroke, there are two neurons, one signaling the presence of the stroke and the other signaling its absence. Letter neurons represent each letter of the alphabet in each of four positions. Word neurons represent each of 1200 common four letter words. The letter and word layers correspond to the P and W layers that were introduced previously. There are bidirectional interactions between the letter and word layers, and lateral inhibition within the layers. The letter neurons also receive input from the stroke neurons, but this interaction is unidirectional. Our network differs in two ways from the original IA model. First, all interactions involving letter and word neurons are symmetric. In the original model, the interactions between the letter and word layers were asymmetric. In particular, inhibitory connections only ran from letter neurons to word neurons, and not vice versa. Second, the only nonlinearity in our model is rectification. These two aspects allow us to apply the full machinery of the theory of permitted and forbidden sets. Figure 3 shows the result of presenting the stimulus “MO M” for four different settings of parameters. In each of the four cases, the word layer of the network converges to the same result, detecting the word “MOON”, which is the closest stored word to the stimulus. However, the activity in the letter layer is different in the four cases. input: P layer reconstruction W layer P layer reconstruction W layer completion noncompletion enforcement non-enforcement Figure 3: Simulation of 4 different parameter regimes in a letter- word recognition network. Within each panel, the middle column presents a feature- layer reconstruction based on the letter activity shown in the left column. W layer activity is shown in the right column. The top row shows the network state after 10 iterations of the dynamics. The bottom row shows the steady state. In the left column, the parameters obey the inequality (3), so that part- whole relationships are enforced. The activity of the letter layer is visualized by activating the strokes corresponding to each active letter neuron. The activated letters are part of the word “MOON”. In the top left, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom left, completion does not occur. In the simulations of the right column, parameters are such that part- whole relationships are not enforced. Consequently, the word layer is much more active. Bottom- up input provides evidence for several other letters, which is not suppressed. In the top right, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom right, the “O” neuron is not activated in the third position, so there is no completion. However, some letter neurons for the third position are activated, due to the input from neurons that indicate the absence of strokes. input: non-recognition event multi-stability Figure 4: Simulation of a non- recognition event and example of multistability. Figure 4 shows simulations for large σ, deep in the enforcement regime where non- recognition is a possibility. From one initial condition, the network converges to a state in which no W neurons are active, a non- recognition. From another initial condition, the network detects the word “NORM”. Deep in the enforcement regime, the top- down feedback can be so strong that the network has multiple stable states, many of which bear little resemblance to the stimulus at all. This is a problematic aspect of this network. It can be prevented by setting parameters at the edge of the enforcement regime. 6 Discussion We have analyzed a recurrent network that performs computations involving part- whole relationships. The network can fill in missing parts and suppress parts that do not belong. These two computations are distinct and can be dissociated from each other, as shown in Figure 3. While these two computations can also be performed by associative memory models, they are not typically dissociable in these models. For example, in the Hopfield model pattern completion and noise suppression are both the result of recall of one of a finite number of stereotyped activity patterns. We believe that our model is more appropriate for perceptual systems, because its behavior is piecewise linear, due its reliance on rectification nonlinearity. Therefore, analog aspects of computation are able to coexist with the part-whole relationships. Furthermore, in our model the stimulus is encoded in maintained synaptic input to the network, rather than as an initial condition of the dynamics. A Appendix: Permitted and forbidden sets Our mathematical results depend on the theory of permitted and forbidden sets [3, 4], which is summarized briefly here. The theory is applicable to neural networks with rectification nonlinearity, of the form xi + xi = [bi + j Wij xj ]+ . Neuron i is said to be active when ˙ xi > 0. For a network of N neurons, there are 2N possible sets of active neurons. For each active set, consider the submatrix of Wij corresponding to the synapses between active neurons. If all eigenvalues of this submatrix have real parts less than or equal to unity, then the active set is said to be permitted. Otherwise the active set is said to be forbidden. A set is permitted if and only if there exists an input vector b such that those neurons are active at a stable steady state. Permitted sets can be regarded as memories stored in the synaptic connections Wij . If Wij is a symmetric matrix, the nesting property holds: every subset of a permitted set is permitted, and every superset of a forbidden set is forbidden. The present model can be seen as a general method for storing permitted sets in a recurrent network. This method introduces a neuron for each permitted set, relying on a unary or “grandmother cell” representation. In contrast, Xie et al.[9] used lateral inhibition in a single layer of neurons to store permitted sets. By introducing extra neurons, the present model achieves superior storage capacity, much as unary models of associative memory [1] surpass distributed models [5]. A.1 Unconditional winner-take-all in the W layer The synapses between two W-neurons have strengths 0 −α −α 0 The eigenvalues of this matrix are ±α. Therefore two W-neurons constitute a forbidden set if α > 1. By the nesting property, it follows more than two W-neurons is also a forbidden set, and that the W layer has the unconditional winner-take-all property. A.2 Part-whole combinations as permitted sets Theorem 1. Suppose that β < 1. If γ 2 < β + (1 − β)/k then any combination of k ≥ 1 parts consistent with a whole corresponds to a permitted set. Proof. Consider k parts belonging to a whole. They are represented by one W-neuron and k P-neurons, with synaptic connections given by the (k + 1) × (k + 1) matrix M= −β(11T − I) γ1 , γ1T 0 (5) where 1 is the k- dimensional vector whose elements are all equal to one. Two eigenvectors of M are of the form (1T c), and have the same eigenvalues as the 2 × 2 matrix −β(k − 1) γk γ 0 This matrix has eigenvalues less than one when γ 2 < β + (1 − β)/k and β(k − 1) + 2 > 0. The other k − 1 eigenvectors are of the form (dT , 0), where dT 1 = 0. These have eigenvalues β. Therefore all eigenvalues of W are less than one if the condition of the theorem is satisfied. A.3 Constraints on combining parts Here, we derive conditions under which the network can enforce the constraint that steady state activity be confined to parts that constitute a whole. Theorem 2. Suppose that β > 0 and σ 2 +β 2 +γ 2 +2σβγ > 1 If a W- neuron is active, then only P- neurons corresponding to parts contained in the relevant whole can be active at a stable steady state. Proof. Consider P- neurons Pi , Pj , and W- neuron Wa . Supa a pose that ξi = 1 but ξj = 0. As shown in Figure 5, the matrix of connections is given by: W = 0 −β γ −β 0 −σ γ −σ 0 (6) Wa γ Pi -σ -β Pj Figure 5: A set of one W- neuron and two P- neurons is forbidden if one part belongs to the whole and the other does not. This set is permitted if all eigenvalues of W − I have negative real parts. The characteristic equation of I − W is λ3 + b1 λ2 + b2 λ + b3 = 0, where b1 = 3, b2 = 3 − σ 2 − β 2 − γ 2 and b3 = 1−2σβγ−σ 2 −β 2 −γ 2 . According to the Routh- Hurwitz theorem, all the eigenvalues have negative real parts if and only if b1 > 0, b3 > 0 and b1 b2 > b3 . Clearly, the first condition is always satisfied. The second condition is more restrictive than the third. It is satisfied only when σ 2 + β 2 + γ 2 + 2σβγ < 1. Hence, one of the eigenvalues has a positive real part when this condition is broken, i.e., when σ 2 +β 2 +γ 2 +2σβγ > 1. By the nesting property, any larger set of P- neurons inconsistent with the W- neuron is also forbidden. A.4 Completion of wholes √ Theorem 3. If γ > β and a single W- neuron a is active at a steady state, then Pi > 0 a for all i such that ξi = 1. Proof. Suppose that the detected whole has k parts. At the steady state Pi = a ξi Bi − (β − γ 2 )Ptot 1−β + where Ptot = Pi = i 1 1 − β + (β − γ 2 )k k a B i ξi i=1 (7) A.5 Preventing runaway If feedback loops cause the network activity to diverge, then the preceding analyses are not relevant. Here we give a sufficient condition guaranteeing that runaway instability does not happen. It is not a necessary condition. Interestingly, the condition implies the condition of Theorem 1. Theorem 4. Suppose that P and W obey the dynamics of Eqs. (1) and (2), and define the objective function E 1−α 2 = − 2 Wa a α + 2 Wa a 1−β + 2 a Pi Wa ξi + σ Bi Pi − γ i 2 ia Pi2 i 2 β + 2 Pi i a (1 − ξi )Pi Wa . (8) ia Then E is a Lyapunov like function that, given β > γ 2 − dynamics to a stable steady state. 1−γ 2 N −1 , ensures convergence of the Proof. (sketch) Differentiation of E with respect to time shows that that E is nonincreasing in the nonnegative orthant and constant only at steady states of the network dynamics. We must also show that E is radially unbounded, which is true if the quadratic part of E is copositive definite. Note that the last term of E is lower-bounded by zero and the previous term is upper bounded by γ ia Pi Wa . We assume α > 1. Thus, we can use Cauchy’s 2 2 inequality, i Pi2 ≥ ( i Pi ) /N , and the fact that a Wa ≤ ( a Wa )2 for Wa ≥ 0, to derive E≥ 1 2 Wa )2 + ( If β > γ 2 − unbounded. a 1−γ 2 N −1 , 1 − β + βN ( N Pi )2 − 2γ( i Wa a Pi ) i − Bi Pi . (9) i the quadratic form in the inequality is positive definite and E is radially References [1] E. B. Baum, J. Moody, and F. Wilczek. Internal representations for associative memory. Biol. Cybern., 59:217–228, 1988. [2] K. Fukushima. Neocognitron: a self organizing neural network model for a mechanism of pattern recognition unaffected by shift in position. Biol Cybern, 36(4):193–202, 1980. [3] R.H. Hahnloser, R. Sarpeshkar, M.A. Mahowald, R.J. Douglas, and H.S. Seung. Digital selection and analogue amplification coexist in a cortex-inspired silicon circuit. Nature, 405(6789):947– 51, Jun 22 2000. [4] R.H. Hahnloser, H.S. Seung, and J.-J. Slotine. Permitted and forbidden sets in symmetric threshold-linear networks. Neural Computation, 15:621–638, 2003. [5] J.J. Hopfield. Neural networks and physical systems with emergent collective computational abilities. Proc Natl Acad Sci U S A, 79(8):2554–8, Apr 1982. [6] Y. LeCun, B. Boser, J. S. Denker, D. Henderson, R. E. Howard, W. Hubbard, and L. D. Jackel. Backpropagation applied to handwritten zip code recognition. Neural Comput., 1:541–551, 1989. [7] J. L. McClelland and D. E. Rumelhart. An interactive activation model of context effects in letter perception: Part i. an account of basic findings. Psychological Review, 88(5):375–407, Sep 1981. [8] M Riesenhuber and T Poggio. Hierarchical models of object recognition in cortex. Nat Neurosci, 2(11):1019–25, Nov 1999. [9] X. Xie, R.H. Hahnloser, and H. S. Seung. Selectively grouping neurons in recurrent networks of lateral inhibition. Neural Computation, 14:2627–2646, 2002.
2 0.19843222 181 nips-2005-Spiking Inputs to a Winner-take-all Network
Author: Matthias Oster, Shih-Chii Liu
Abstract: Recurrent networks that perform a winner-take-all computation have been studied extensively. Although some of these studies include spiking networks, they consider only analog input rates. We present results of this winner-take-all computation on a network of integrate-and-fire neurons which receives spike trains as inputs. We show how we can configure the connectivity in the network so that the winner is selected after a pre-determined number of input spikes. We discuss spiking inputs with both regular frequencies and Poisson-distributed rates. The robustness of the computation was tested by implementing the winner-take-all network on an analog VLSI array of 64 integrate-and-fire neurons which have an innate variance in their operating parameters. 1
3 0.12406213 118 nips-2005-Learning in Silicon: Timing is Everything
Author: John V. Arthur, Kwabena Boahen
Abstract: We describe a neuromorphic chip that uses binary synapses with spike timing-dependent plasticity (STDP) to learn stimulated patterns of activity and to compensate for variability in excitability. Specifically, STDP preferentially potentiates (turns on) synapses that project from excitable neurons, which spike early, to lethargic neurons, which spike late. The additional excitatory synaptic current makes lethargic neurons spike earlier, thereby causing neurons that belong to the same pattern to spike in synchrony. Once learned, an entire pattern can be recalled by stimulating a subset. 1 Variability in Neural Systems Evidence suggests precise spike timing is important in neural coding, specifically, in the hippocampus. The hippocampus uses timing in the spike activity of place cells (in addition to rate) to encode location in space [1]. Place cells employ a phase code: the timing at which a neuron spikes relative to the phase of the inhibitory theta rhythm (5-12Hz) conveys information. As an animal approaches a place cell’s preferred location, the place cell not only increases its spike rate, but also spikes at earlier phases in the theta cycle. To implement a phase code, the theta rhythm is thought to prevent spiking until the input synaptic current exceeds the sum of the neuron threshold and the decreasing inhibition on the downward phase of the cycle [2]. However, even with identical inputs and common theta inhibition, neurons do not spike in synchrony. Variability in excitability spreads the activity in phase. Lethargic neurons (such as those with high thresholds) spike late in the theta cycle, since their input exceeds the sum of the neuron threshold and theta inhibition only after the theta inhibition has had time to decrease. Conversely, excitable neurons (such as those with low thresholds) spike early in the theta cycle. Consequently, variability in excitability translates into variability in timing. We hypothesize that the hippocampus achieves its precise spike timing (about 10ms) through plasticity enhanced phase-coding (PEP). The source of hippocampal timing precision in the presence of variability (and noise) remains unexplained. Synaptic plasticity can compensate for variability in excitability if it increases excitatory synaptic input to neurons in inverse proportion to their excitabilities. Recasting this in a phase-coding framework, we desire a learning rule that increases excitatory synaptic input to neurons directly related to their phases. Neurons that lag require additional synaptic input, whereas neurons that lead 120µm 190µm A B Figure 1: STDP Chip. A The chip has a 16-by-16 array of microcircuits; one microcircuit includes four principal neurons, each with 21 STDP circuits. B The STDP Chip is embedded in a circuit board including DACs, a CPLD, a RAM chip, and a USB chip, which communicates with a PC. require none. The spike timing-dependent plasticity (STDP) observed in the hippocampus satisfies this requirement [3]. It requires repeated pre-before-post spike pairings (within a time window) to potentiate and repeated post-before-pre pairings to depress a synapse. Here we validate our hypothesis with a model implemented in silicon, where variability is as ubiquitous as it is in biology [4]. Section 2 presents our silicon system, including the STDP Chip. Section 3 describes and characterizes the STDP circuit. Section 4 demonstrates that PEP compensates for variability and provides evidence that STDP is the compensation mechanism. Section 5 explores a desirable consequence of PEP: unconventional associative pattern recall. Section 6 discusses the implications of the PEP model, including its benefits and applications in the engineering of neuromorphic systems and in the study of neurobiology. 2 Silicon System We have designed, submitted, and tested a silicon implementation of PEP. The STDP Chip was fabricated through MOSIS in a 1P5M 0.25µm CMOS process, with just under 750,000 transistors in just over 10mm2 of area. It has a 32 by 32 array of excitatory principal neurons commingled with a 16 by 16 array of inhibitory interneurons that are not used here (Figure 1A). Each principal neuron has 21 STDP synapses. The address-event representation (AER) [5] is used to transmit spikes off chip and to receive afferent and recurrent spike input. To configure the STDP Chip as a recurrent network, we embedded it in a circuit board (Figure 1B). The board has five primary components: a CPLD (complex programmable logic device), the STDP Chip, a RAM chip, a USB interface chip, and DACs (digital-to-analog converters). The central component in the system is the CPLD. The CPLD handles AER traffic, mediates communication between devices, and implements recurrent connections by accessing a lookup table, stored in the RAM chip. The USB interface chip provides a bidirectional link with a PC. The DACs control the analog biases in the system, including the leak current, which the PC varies in real-time to create the global inhibitory theta rhythm. The principal neuron consists of a refractory period and calcium-dependent potassium circuit (RCK), a synapse circuit, and a soma circuit (Figure 2A). RCK and the synapse are ISOMA Soma Synapse STDP Presyn. Spike PE LPF A Presyn. Spike Raster AH 0 0.1 Spike probability RCK Postsyn. Spike B 0.05 0.1 0.05 0.1 0.08 0.06 0.04 0.02 0 0 Time(s) Figure 2: Principal neuron. A A simplified schematic is shown, including: the synapse, refractory and calcium-dependent potassium channel (RCK), soma, and axon-hillock (AH) circuits, plus their constituent elements, the pulse extender (PE) and the low-pass filter (LPF). B Spikes (dots) from 81 principal neurons are temporally dispersed, when excited by poisson-like inputs (58Hz) and inhibited by the common 8.3Hz theta rhythm (solid line). The histogram includes spikes from five theta cycles. composed of two reusable blocks: the low-pass filter (LPF) and the pulse extender (PE). The soma is a modified version of the LPF, which receives additional input from an axonhillock circuit (AH). RCK is inhibitory to the neuron. It consists of a PE, which models calcium influx during a spike, and a LPF, which models calcium buffering. When AH fires a spike, a packet of charge is dumped onto a capacitor in the PE. The PE’s output activates until the charge decays away, which takes a few milliseconds. Also, while the PE is active, charge accumulates on the LPF’s capacitor, lowering the LPF’s output voltage. Once the PE deactivates, this charge leaks away as well, but this takes tens of milliseconds because the leak is smaller. The PE’s and the LPF’s inhibitory effects on the soma are both described below in terms of the sum (ISHUNT ) of the currents their output voltages produce in pMOS transistors whose sources are at Vdd (see Figure 2A). Note that, in the absence of spikes, these currents decay exponentially, with a time-constant determined by their respective leaks. The synapse circuit is excitatory to the neuron. It is composed of a PE, which represents the neurotransmitter released into the synaptic cleft, and a LPF, which represents the bound neurotransmitter. The synapse circuit is similar to RCK in structure but differs in function: It is activated not by the principal neuron itself but by the STDP circuits (or directly by afferent spikes that bypass these circuits, i.e., fixed synapses). The synapse’s effect on the soma is also described below in terms of the current (ISYN ) its output voltage produces in a pMOS transistor whose source is at Vdd. The soma circuit is a leaky integrator. It receives excitation from the synapse circuit and shunting inhibition from RCK and has a leak current as well. Its temporal behavior is described by: τ dISOMA ISYN I0 + ISOMA = dt ISHUNT where ISOMA is the current the capacitor’s voltage produces in a pMOS transistor whose source is at Vdd (see Figure 2A). ISHUNT is the sum of the leak, refractory, and calciumdependent potassium currents. These currents also determine the time constant: τ = C Ut κISHUNT , where I0 and κ are transistor parameters and Ut is the thermal voltage. STDP circuit ~LTP SRAM Presynaptic spike A ~LTD Inverse number of pairings Integrator Decay Postsynaptic spike Potentiation 0.1 0.05 0 0.05 0.1 Depression -80 -40 0 Presynaptic spike Postsynaptic spike 40 Spike timing: t pre - t post (ms) 80 B Figure 3: STDP circuit design and characterization. A The circuit is composed of three subcircuits: decay, integrator, and SRAM. B The circuit potentiates when the presynaptic spike precedes the postsynaptic spike and depresses when the postsynaptic spike precedes the presynaptic spike. The soma circuit is connected to an AH, the locus of spike generation. The AH consists of model voltage-dependent sodium and potassium channel populations (modified from [6] by Kai Hynna). It initiates the AER signaling process required to send a spike off chip. To characterize principal neuron variability, we excited 81 neurons with poisson-like 58Hz spike trains (Figure 2B). We made these spike trains poisson-like by starting with a regular 200Hz spike train and dropping spikes randomly, with probability of 0.71. Thus spikes were delivered to neurons that won the coin toss in synchrony every 5ms. However, neurons did not lock onto the input synchrony due to filtering by the synaptic time constant (see Figure 2B). They also received a common inhibitory input at the theta frequency (8.3Hz), via their leak current. Each neuron was prevented from firing more than one spike in a theta cycle by its model calcium-dependent potassium channel population. The principal neurons’ spike times were variable. To quantify the spike variability, we used timing precision, which we define as twice the standard deviation of spike times accumulated from five theta cycles. With an input rate of 58Hz the timing precision was 34ms. 3 STDP Circuit The STDP circuit (related to [7]-[8]), for which the STDP Chip is named, is the most abundant, with 21,504 copies on the chip. This circuit is built from three subcircuits: decay, integrator, and SRAM (Figure 3A). The decay and integrator are used to implement potentiation, and depression, in a symmetric fashion. The SRAM holds the current binary state of the synapse, either potentiated or depressed. For potentiation, the decay remembers the last presynaptic spike. Its capacitor is charged when that spike occurs and discharges linearly thereafter. A postsynaptic spike samples the charge remaining on the capacitor, passes it through an exponential function, and dumps the resultant charge into the integrator. This charge decays linearly thereafter. At the time of the postsynaptic spike, the SRAM, a cross-coupled inverter pair, reads the voltage on the integrator’s capacitor. If it exceeds a threshold, the SRAM switches state from depressed to potentiated (∼LTD goes high and ∼LTP goes low). The depression side of the STDP circuit is exactly symmetric, except that it responds to postsynaptic activation followed by presynaptic activation and switches the SRAM’s state from potentiated to depressed (∼LTP goes high and ∼LTD goes low). When the SRAM is in the potentiated state, the presynaptic 50 After STDP 83 92 100 Timing precision(ms) Before STDP 75 B Before STDP After STDP 40 30 20 10 0 50 60 70 80 90 Input rate(Hz) 100 50 58 67 text A 0.2 0.4 Time(s) 0.6 0.2 0.4 Time(s) 0.6 C Figure 4: Plasticity enhanced phase-coding. A Spike rasters of 81 neurons (9 by 9 cluster) display synchrony over a two-fold range of input rates after STDP. B The degree of enhancement is quantified by timing precision. C Each neuron (center box) sends synapses to (dark gray) and receives synapses from (light gray) twenty-one randomly chosen neighbors up to five nodes away (black indicates both connections). spike activates the principal neuron’s synapse; otherwise the spike has no effect. We characterized the STDP circuit by activating a plastic synapse and a fixed synapse– which elicits a spike at different relative times. We repeated this pairing at 16Hz. We counted the number of pairings required to potentiate (or depress) the synapse. Based on this count, we calculated the efficacy of each pairing as the inverse number of pairings required (Figure 3B). For example, if twenty pairings were required to potentiate the synapse, the efficacy of that pre-before-post time-interval was one twentieth. The efficacy of both potentiation and depression are fit by exponentials with time constants of 11.4ms and 94.9ms, respectively. This behavior is similar to that observed in the hippocampus: potentiation has a shorter time constant and higher maximum efficacy than depression [3]. 4 Recurrent Network We carried out an experiment designed to test the STDP circuit’s ability to compensate for variability in spike timing through PEP. Each neuron received recurrent connections from 21 randomly selected neurons within an 11 by 11 neighborhood centered on itself (see Figure 4C). Conversely, it made recurrent connections to randomly chosen neurons within the same neighborhood. These connections were mediated by STDP circuits, initialized to the depressed state. We chose a 9 by 9 cluster of neurons and delivered spikes at a mean rate of 50 to 100Hz to each one (dropping spikes with a probability of 0.75 to 0.5 from a regular 200Hz train) and provided common theta inhibition as before. We compared the variability in spike timing after five seconds of learning with the initial distribution. Phase coding was enhanced after STDP (Figure 4A). Before STDP, spike timing among neurons was highly variable (except for the very highest input rate). After STDP, variability was virtually eliminated (except for the very lowest input rate). Initially, the variability, characterized by timing precision, was inversely related to the input rate, decreasing from 34 to 13ms. After five seconds of STDP, variability decreased and was largely independent of input rate, remaining below 11ms. Potentiated synapses 25 A Synaptic state after STDP 20 15 10 5 0 B 50 100 150 200 Spiking order 250 Figure 5: Compensating for variability. A Some synapses (dots) become potentiated (light) while others remain depressed (dark) after STDP. B The number of potentiated synapses neurons make (pluses) and receive (circles) is negatively (r = -0.71) and positively (r = 0.76) correlated to their rank in the spiking order, respectively. Comparing the number of potentiated synapses each neuron made or received with its excitability confirmed the PEP hypothesis (i.e., leading neurons provide additional synaptic current to lagging neurons via potentiated recurrent synapses). In this experiment, to eliminate variability due to noise (as opposed to excitability), we provided a 17 by 17 cluster of neurons with a regular 200Hz excitatory input. Theta inhibition was present as before and all synapses were initialized to the depressed state. After 10 seconds of STDP, a large fraction of the synapses were potentiated (Figure 5A). When the number of potentiated synapses each neuron made or received was plotted versus its rank in spiking order (Figure 5B), a clear correlation emerged (r = -0.71 or 0.76, respectively). As expected, neurons that spiked early made more and received fewer potentiated synapses. In contrast, neurons that spiked late made fewer and received more potentiated synapses. 5 Pattern Completion After STDP, we found that the network could recall an entire pattern given a subset, thus the same mechanisms that compensated for variability and noise could also compensate for lack of information. We chose a 9 by 9 cluster of neurons as our pattern and delivered a poisson-like spike train with mean rate of 67Hz to each one as in the first experiment. Theta inhibition was present as before and all synapses were initialized to the depressed state. Before STDP, we stimulated a subset of the pattern and only neurons in that subset spiked (Figure 6A). After five seconds of STDP, we stimulated the same subset again. This time they recruited spikes from other neurons in the pattern, completing it (Figure 6B). Upon varying the fraction of the pattern presented, we found that the fraction recalled increased faster than the fraction presented. We selected subsets of the original pattern randomly, varying the fraction of neurons chosen from 0.1 to 1.0 (ten trials for each). We classified neurons as active if they spiked in the two second period over which we recorded. Thus, we characterized PEP’s pattern-recall performance as a function of the probability that the pattern in question’s neurons are activated (Figure 6C). At a fraction of 0.50 presented, nearly all of the neurons in the pattern are consistently activated (0.91±0.06), showing robust pattern completion. We fitted the recall performance with a sigmoid that reached 0.50 recall fraction with an input fraction of 0.30. No spurious neurons were activated during any trials. Rate(Hz) Rate(Hz) 8 7 7 6 6 5 5 0.6 0.4 2 0.2 0 0 3 3 2 1 1 A 0.8 4 4 Network activity before STDP 1 Fraction of pattern actived 8 0 B Network activity after STDP C 0 0.2 0.4 0.6 0.8 Fraction of pattern stimulated 1 Figure 6: Associative recall. A Before STDP, half of the neurons in a pattern are stimulated; only they are activated. B After STDP, half of the neurons in a pattern are stimulated, and all are activated. C The fraction of the pattern activated grows faster than the fraction stimulated. 6 Discussion Our results demonstrate that PEP successfully compensates for graded variations in our silicon recurrent network using binary (on–off) synapses (in contrast with [8], where weights are graded). While our chip results are encouraging, variability was not eliminated in every case. In the case of the lowest input (50Hz), we see virtually no change (Figure 4A). We suspect the timing remains imprecise because, with such low input, neurons do not spike every theta cycle and, consequently, provide fewer opportunities for the STDP synapses to potentiate. This shortfall illustrates the system’s limits; it can only compensate for variability within certain bounds, and only for activity appropriate to the PEP model. As expected, STDP is the mechanism responsible for PEP. STDP potentiated recurrent synapses from leading neurons to lagging neurons, reducing the disparity among the diverse population of neurons. Even though the STDP circuits are themselves variable, with different efficacies and time constants, when using timing the sign of the weight-change is always correct (data not shown). For this reason, we chose STDP over other more physiological implementations of plasticity, such as membrane-voltage-dependent plasticity (MVDP), which has the capability to learn with graded voltage signals [9], such as those found in active dendrites, providing more computational power [10]. Previously, we investigated a MVDP circuit, which modeled a voltage-dependent NMDAreceptor-gated synapse [11]. It potentiated when the calcium current analog exceeded a threshold, which was designed to occur only during a dendritic action potential. This circuit produced behavior similar to STDP, implying it could be used in PEP. However, it was sensitive to variability in the NMDA and potentiation thresholds, causing a fraction of the population to potentiate anytime the synapse received an input and another fraction to never potentiate, rendering both subpopulations useless. Therefore, the simpler, less biophysical STDP circuit won out over the MVDP circuit: In our system timing is everything. Associative storage and recall naturally emerge in the PEP network when synapses between neurons coactivated by a pattern are potentiated. These synapses allow neurons to recruit their peers when a subset of the pattern is presented, thereby completing the pattern. However, this form of pattern storage and completion differs from Hopfield’s attractor model [12] . Rather than forming symmetric, recurrent neuronal circuits, our recurrent network forms asymmetric circuits in which neurons make connections exclusively to less excitable neurons in the pattern. In both the poisson-like and regular cases (Figures 4 & 5), only about six percent of potentiated connections were reciprocated, as expected by chance. We plan to investigate the storage capacity of this asymmetric form of associative memory. Our system lends itself to modeling brain regions that use precise spike timing, such as the hippocampus. We plan to extend the work presented to store and recall sequences of patterns, as the hippocampus is hypothesized to do. Place cells that represent different locations spike at different phases of the theta cycle, in relation to the distance to their preferred locations. This sequential spiking will allow us to link patterns representing different locations in the order those locations are visited, thereby realizing episodic memory. We propose PEP as a candidate neural mechanism for information coding and storage in the hippocampal system. Observations from the CA1 region of the hippocampus suggest that basal dendrites (which primarily receive excitation from recurrent connections) support submillisecond timing precision, consistent with PEP [13]. We have shown, in a silicon model, PEP’s ability to exploit such fast recurrent connections to sharpen timing precision as well as to associatively store and recall patterns. Acknowledgments We thank Joe Lin for assistance with chip generation. The Office of Naval Research funded this work (Award No. N000140210468). References [1] O’Keefe J. & Recce M.L. (1993). Phase relationship between hippocampal place units and the EEG theta rhythm. Hippocampus 3(3):317-330. [2] Mehta M.R., Lee A.K. & Wilson M.A. (2002) Role of experience and oscillations in transforming a rate code into a temporal code. Nature 417(6890):741-746. [3] Bi G.Q. & Wang H.X. (2002) Temporal asymmetry in spike timing-dependent synaptic plasticity. Physiology & Behavior 77:551-555. [4] Rodriguez-Vazquez, A., Linan, G., Espejo S. & Dominguez-Castro R. (2003) Mismatch-induced trade-offs and scalability of analog preprocessing visual microprocessor chips. Analog Integrated Circuits and Signal Processing 37:73-83. [5] Boahen K.A. (2000) Point-to-point connectivity between neuromorphic chips using address events. IEEE Transactions on Circuits and Systems II 47:416-434. [6] Culurciello E.R., Etienne-Cummings R. & Boahen K.A. (2003) A biomorphic digital image sensor. IEEE Journal of Solid State Circuits 38:281-294. [7] Bofill A., Murray A.F & Thompson D.P. (2005) Citcuits for VLSI Implementation of Temporally Asymmetric Hebbian Learning. In: Advances in Neural Information Processing Systems 14, MIT Press, 2002. [8] Cameron K., Boonsobhak V., Murray A. & Renshaw D. (2005) Spike timing dependent plasticity (STDP) can ameliorate process variations in neuromorphic VLSI. IEEE Transactions on Neural Networks 16(6):1626-1627. [9] Chicca E., Badoni D., Dante V., D’Andreagiovanni M., Salina G., Carota L., Fusi S. & Del Giudice P. (2003) A VLSI recurrent network of integrate-and-fire neurons connected by plastic synapses with long-term memory. IEEE Transaction on Neural Networks 14(5):1297-1307. [10] Poirazi P., & Mel B.W. (2001) Impact of active dendrites and structural plasticity on the memory capacity of neural tissue. Neuron 29(3)779-796. [11] Arthur J.V. & Boahen K. (2004) Recurrently connected silicon neurons with active dendrites for one-shot learning. In: IEEE International Joint Conference on Neural Networks 3, pp.1699-1704. [12] Hopfield J.J. (1984) Neurons with graded response have collective computational properties like those of two-state neurons. Proceedings of the National Academy of Science 81(10):3088-3092. [13] Ariav G., Polsky A. & Schiller J. (2003) Submillisecond precision of the input-output transformation function mediated by fast sodium dendritic spikes in basal dendrites of CA1 pyramidal neurons. Journal of Neuroscience 23(21):7750-7758.
4 0.11701804 157 nips-2005-Principles of real-time computing with feedback applied to cortical microcircuit models
Author: Wolfgang Maass, Prashant Joshi, Eduardo D. Sontag
Abstract: The network topology of neurons in the brain exhibits an abundance of feedback connections, but the computational function of these feedback connections is largely unknown. We present a computational theory that characterizes the gain in computational power achieved through feedback in dynamical systems with fading memory. It implies that many such systems acquire through feedback universal computational capabilities for analog computing with a non-fading memory. In particular, we show that feedback enables such systems to process time-varying input streams in diverse ways according to rules that are implemented through internal states of the dynamical system. In contrast to previous attractor-based computational models for neural networks, these flexible internal states are high-dimensional attractors of the circuit dynamics, that still allow the circuit state to absorb new information from online input streams. In this way one arrives at novel models for working memory, integration of evidence, and reward expectation in cortical circuits. We show that they are applicable to circuits of conductance-based Hodgkin-Huxley (HH) neurons with high levels of noise that reflect experimental data on invivo conditions. 1
5 0.10610275 196 nips-2005-Two view learning: SVM-2K, Theory and Practice
Author: Jason Farquhar, David Hardoon, Hongying Meng, John S. Shawe-taylor, Sándor Szedmák
Abstract: Kernel methods make it relatively easy to define complex highdimensional feature spaces. This raises the question of how we can identify the relevant subspaces for a particular learning task. When two views of the same phenomenon are available kernel Canonical Correlation Analysis (KCCA) has been shown to be an effective preprocessing step that can improve the performance of classification algorithms such as the Support Vector Machine (SVM). This paper takes this observation to its logical conclusion and proposes a method that combines this two stage learning (KCCA followed by SVM) into a single optimisation termed SVM-2K. We present both experimental and theoretical analysis of the approach showing encouraging results and insights. 1
6 0.096206114 7 nips-2005-A Cortically-Plausible Inverse Problem Solving Method Applied to Recognizing Static and Kinematic 3D Objects
7 0.093886606 124 nips-2005-Measuring Shared Information and Coordinated Activity in Neuronal Networks
8 0.092101291 6 nips-2005-A Connectionist Model for Constructive Modal Reasoning
9 0.089114331 61 nips-2005-Dynamical Synapses Give Rise to a Power-Law Distribution of Neuronal Avalanches
10 0.08652515 64 nips-2005-Efficient estimation of hidden state dynamics from spike trains
11 0.082883537 99 nips-2005-Integrate-and-Fire models with adaptation are good enough
12 0.0801346 108 nips-2005-Layered Dynamic Textures
13 0.077473521 134 nips-2005-Neural mechanisms of contrast dependent receptive field size in V1
14 0.077331461 176 nips-2005-Silicon growth cones map silicon retina
15 0.075221255 106 nips-2005-Large-scale biophysical parameter estimation in single neurons via constrained linear regression
16 0.074738376 28 nips-2005-Analyzing Auditory Neurons by Learning Distance Functions
17 0.073519096 74 nips-2005-Faster Rates in Regression via Active Learning
18 0.071469218 141 nips-2005-Norepinephrine and Neural Interrupts
19 0.071149677 100 nips-2005-Interpolating between types and tokens by estimating power-law generators
20 0.069432087 8 nips-2005-A Criterion for the Convergence of Learning with Spike Timing Dependent Plasticity
topicId topicWeight
[(0, 0.187), (1, -0.204), (2, -0.058), (3, -0.024), (4, -0.009), (5, 0.053), (6, -0.017), (7, 0.088), (8, -0.01), (9, 0.03), (10, -0.028), (11, 0.077), (12, 0.005), (13, -0.101), (14, -0.146), (15, 0.1), (16, 0.058), (17, 0.007), (18, 0.01), (19, 0.097), (20, 0.098), (21, 0.038), (22, -0.052), (23, -0.015), (24, -0.119), (25, 0.022), (26, 0.011), (27, 0.068), (28, 0.107), (29, -0.079), (30, 0.117), (31, 0.148), (32, -0.237), (33, 0.128), (34, 0.099), (35, 0.022), (36, 0.061), (37, -0.036), (38, -0.023), (39, -0.067), (40, -0.014), (41, -0.21), (42, -0.143), (43, -0.063), (44, 0.037), (45, 0.123), (46, -0.037), (47, 0.182), (48, 0.056), (49, -0.084)]
simIndex simValue paperId paperTitle
same-paper 1 0.98263294 164 nips-2005-Representing Part-Whole Relationships in Recurrent Neural Networks
Author: Viren Jain, Valentin Zhigulin, H. S. Seung
Abstract: There is little consensus about the computational function of top-down synaptic connections in the visual system. Here we explore the hypothesis that top-down connections, like bottom-up connections, reflect partwhole relationships. We analyze a recurrent network with bidirectional synaptic interactions between a layer of neurons representing parts and a layer of neurons representing wholes. Within each layer, there is lateral inhibition. When the network detects a whole, it can rigorously enforce part-whole relationships by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the theory of permitted and forbidden sets [3, 4]. The network behaviors are illustrated by recreating Rumelhart and McClelland’s “interactive activation” model [7]. In neural network models of visual object recognition [2, 6, 8], patterns of synaptic connectivity often reflect part-whole relationships between the features that are represented by neurons. For example, the connections of Figure 1 reflect the fact that feature B both contains simpler features A1, A2, and A3, and is contained in more complex features C1, C2, and C3. Such connectivity allows neurons to follow the rule that existence of the part is evidence for existence of the whole. By combining synaptic input from multiple sources of evidence for a feature, a neuron can “decide” whether that feature is present. 1 The synapses shown in Figure 1 are purely bottom-up, directed from simple to complex features. However, there are also top-down connections in the visual system, and there is little consensus about their function. One possibility is that top-down connections also reflect part-whole relationships. They allow feature detectors to make decisions using the rule that existence of the whole is evidence for existence of its parts. In this paper, we analyze the dynamics of a recurrent network in which part-whole relationships are stored as bidirectional synaptic interactions, rather than the unidirectional interactions of Figure 1. The network has a number of interesting computational capabilities. When the network detects a whole, it can rigorously enforce part-whole relationships 1 Synaptic connectivity may reflect other relationships besides part-whole. For example, invariances can be implemented by connecting detectors of several instances of the same feature to the same target, which is consequently an invariant detector of the feature. C1 C2 C3 B A1 A2 A3 Figure 1: The synaptic connections (arrows) of neuron B represent part-whole relationships. Feature B both contains simpler features and is contained in more complex features. The synaptic interactions are drawn one-way, as in most models of visual object recognition. Existence of the part is regarded as evidence for existence of the whole. This paper makes the interactions bidirectional, allowing the existence of the whole to be evidence for the existence of its parts. by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the recently developed theory of permitted and forbidden sets [3, 4]. Our model is closely related to the interactive activation model of word recognition, which was proposed by McClelland and Rumelhart to explain the word superiority effect studied by visual psychologists [7]. Here our concern is not to model a psychological effect, but to characterize mathematically how computations involving part-whole relationships can be carried out by a recurrent network. 1 Network model Suppose that we are given a set of part-whole relationships specified by a ξi = 1, if part i is contained in whole a 0, otherwise We assume that every whole contains at least one part, and every part is contained in at least one whole. The stimulus drives a layer of neurons that detect parts. These neurons also interact with a layer of neurons that detect wholes. We will refer to part-detectors as “P-neurons” and whole-detectors as “W-neurons.” The part-whole relationships are directly stored in the synaptic connections between P and a W neurons. If ξi = 1, the ith neuron in the P layer and the ath neuron in the W layer have a an excitatory interaction of strength γ. If ξi = 0, the neurons have an inhibitory interaction of strength σ. Furthermore, the P-neurons inhibit each other with strength β, and the Wneurons inhibit each other with strength α. All of these interactions are symmetric, and all activation functions are the rectification nonlinearity [z]+ = max{z, 0}. Then the dynamics of the network takes the form ˙ Wa + Wa a Pi ξ i − σ = γ i + a (1 − ξi )Pi − α i Wb , + ˙ Pi + Pi (1) b=a a Wa ξ i − σ = γ a a (1 − ξi )Wa − β a Pj + B i . j=i (2) where Bi is the input to the P layer from the stimulus. Figure 2 shows an example of a network with two wholes. Each whole contains two parts. One of the parts is contained in both wholes. -α Wa excitation γ -σ inhibition P1 B1 -β } W layer Wb -σ P2 -β B2 P3 } P layer B3 Figure 2: Model in example configuration: ξ = {(1, 1, 0), (0, 1, 1)}. When a stimulus is presented, it activates some of the P-neurons, which activate some of the W-neurons. The network eventually converges to a stable steady state. We will assume that α > 1. In the Appendix, we prove that this leads to unconditional winner-take-all behavior in the W layer. In other words, no more than one W-neuron can be active at a stable steady state. If a single W-neuron is active, then a whole has been detected. Potentially there are also many P-neurons active, indicating detection of parts. This representation may have different properties, depending on the choice of parameters β, γ, and σ. As discussed below, these include rigorous enforcement of part-whole relationships, completion of wholes by “filling in” missing parts, and non-recognition of parts that do not conform to a whole. 2 Enforcement of part-whole relationships Suppose that a single W-neuron is active at a stable steady state, so that a whole has been detected. Part-whole relationships are said to be enforced if the network always ignores parts that are not contained in the detected whole, despite potentially strong bottom-up evidence for them. It can be shown that enforcement follows from the inequality σ 2 + β 2 + γ 2 + 2σβγ > 1. (3) which guarantees that neuron i in the P layer is inactive, if neuron a in the W layer is a active and ξi = 0. When part-whole relations are enforced, prior knowledge about legal combinations of parts strictly constrains what may be perceived. This result is proven in the Appendix, and only an intuitive explanation is given here. Enforcement is easiest to understand when there is interlayer inhibition (σ > 0). In this case, the active W-neuron directly inhibits the forbidden P-neurons. The case of σ = 0 is more subtle. Then enforcement is mediated by lateral inhibition in the P layer. Excitatory feedback from the W-neuron has the effect of counteracting the lateral inhibition between the P-neurons that belong to the whole. As a result, these P-neurons become strongly activated enough to inhibit the rest of the P layer. 3 Completion of wholes by filling in missing parts If a W-neuron is active, it excites the P-neurons that belong to the whole. As a result, even if one of these P-neurons receives no bottom-up input (Bi = 0), it is still active. We call this phenomenon “completion,” and it is guaranteed to happen when (4) γ> β The network may thus “imagine” parts that are consistent with the recognized whole, but are not actually present in the stimulus. As with enforcement, this condition depends on top-down connections. √ In the special case γ = β, the interlayer excitation between a W-neuron and its P-neurons exactly cancels out the lateral inhibition between the P-neurons at a steady state. So the recurrent connections effectively vanish, letting the activity of the P-neurons be determined by their feedforward inputs. When the interlayer excitation is stronger than this, the inequality (4) holds, and completion occurs. 4 Non-recognition of a whole If there is no interlayer inhibition (σ = 0), then a single W-neuron is always active, assuming that there is some activity in the P layer. To see this, suppose for the sake of contradiction that all the W-neurons are inactive. Then they receive no inhibition to counteract the excitation from the P layer. This means some of them must be active, which contradicts our assumption. This means that the network always recognizes a whole, even if the stimulus is very different from any part-whole combination that is stored in the network. However, if interlayer inhibition is sufficiently strong (large σ), the network may refuse to recognize a whole. Neurons in the P layer are activated, but there is no activity in the W layer. Formal conditions on σ can be derived, but are not given here because of space limitations. In case of non-recognition, constraints on the P-layer are not enforced. It is possible for the network to detect a configuration of parts that is not consistent with any stored whole. 5 Example: Interactive Activation model To illustrate the computational capabilities of our network, we use it to recreate the interactive activation (IA) model of McClelland and Rumelhart. Figure 3 shows numerical simulations of a network containing three layers of neurons representing strokes, letters, and words, respectively. There are 16 possible strokes in each of four letter positions. For each stroke, there are two neurons, one signaling the presence of the stroke and the other signaling its absence. Letter neurons represent each letter of the alphabet in each of four positions. Word neurons represent each of 1200 common four letter words. The letter and word layers correspond to the P and W layers that were introduced previously. There are bidirectional interactions between the letter and word layers, and lateral inhibition within the layers. The letter neurons also receive input from the stroke neurons, but this interaction is unidirectional. Our network differs in two ways from the original IA model. First, all interactions involving letter and word neurons are symmetric. In the original model, the interactions between the letter and word layers were asymmetric. In particular, inhibitory connections only ran from letter neurons to word neurons, and not vice versa. Second, the only nonlinearity in our model is rectification. These two aspects allow us to apply the full machinery of the theory of permitted and forbidden sets. Figure 3 shows the result of presenting the stimulus “MO M” for four different settings of parameters. In each of the four cases, the word layer of the network converges to the same result, detecting the word “MOON”, which is the closest stored word to the stimulus. However, the activity in the letter layer is different in the four cases. input: P layer reconstruction W layer P layer reconstruction W layer completion noncompletion enforcement non-enforcement Figure 3: Simulation of 4 different parameter regimes in a letter- word recognition network. Within each panel, the middle column presents a feature- layer reconstruction based on the letter activity shown in the left column. W layer activity is shown in the right column. The top row shows the network state after 10 iterations of the dynamics. The bottom row shows the steady state. In the left column, the parameters obey the inequality (3), so that part- whole relationships are enforced. The activity of the letter layer is visualized by activating the strokes corresponding to each active letter neuron. The activated letters are part of the word “MOON”. In the top left, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom left, completion does not occur. In the simulations of the right column, parameters are such that part- whole relationships are not enforced. Consequently, the word layer is much more active. Bottom- up input provides evidence for several other letters, which is not suppressed. In the top right, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom right, the “O” neuron is not activated in the third position, so there is no completion. However, some letter neurons for the third position are activated, due to the input from neurons that indicate the absence of strokes. input: non-recognition event multi-stability Figure 4: Simulation of a non- recognition event and example of multistability. Figure 4 shows simulations for large σ, deep in the enforcement regime where non- recognition is a possibility. From one initial condition, the network converges to a state in which no W neurons are active, a non- recognition. From another initial condition, the network detects the word “NORM”. Deep in the enforcement regime, the top- down feedback can be so strong that the network has multiple stable states, many of which bear little resemblance to the stimulus at all. This is a problematic aspect of this network. It can be prevented by setting parameters at the edge of the enforcement regime. 6 Discussion We have analyzed a recurrent network that performs computations involving part- whole relationships. The network can fill in missing parts and suppress parts that do not belong. These two computations are distinct and can be dissociated from each other, as shown in Figure 3. While these two computations can also be performed by associative memory models, they are not typically dissociable in these models. For example, in the Hopfield model pattern completion and noise suppression are both the result of recall of one of a finite number of stereotyped activity patterns. We believe that our model is more appropriate for perceptual systems, because its behavior is piecewise linear, due its reliance on rectification nonlinearity. Therefore, analog aspects of computation are able to coexist with the part-whole relationships. Furthermore, in our model the stimulus is encoded in maintained synaptic input to the network, rather than as an initial condition of the dynamics. A Appendix: Permitted and forbidden sets Our mathematical results depend on the theory of permitted and forbidden sets [3, 4], which is summarized briefly here. The theory is applicable to neural networks with rectification nonlinearity, of the form xi + xi = [bi + j Wij xj ]+ . Neuron i is said to be active when ˙ xi > 0. For a network of N neurons, there are 2N possible sets of active neurons. For each active set, consider the submatrix of Wij corresponding to the synapses between active neurons. If all eigenvalues of this submatrix have real parts less than or equal to unity, then the active set is said to be permitted. Otherwise the active set is said to be forbidden. A set is permitted if and only if there exists an input vector b such that those neurons are active at a stable steady state. Permitted sets can be regarded as memories stored in the synaptic connections Wij . If Wij is a symmetric matrix, the nesting property holds: every subset of a permitted set is permitted, and every superset of a forbidden set is forbidden. The present model can be seen as a general method for storing permitted sets in a recurrent network. This method introduces a neuron for each permitted set, relying on a unary or “grandmother cell” representation. In contrast, Xie et al.[9] used lateral inhibition in a single layer of neurons to store permitted sets. By introducing extra neurons, the present model achieves superior storage capacity, much as unary models of associative memory [1] surpass distributed models [5]. A.1 Unconditional winner-take-all in the W layer The synapses between two W-neurons have strengths 0 −α −α 0 The eigenvalues of this matrix are ±α. Therefore two W-neurons constitute a forbidden set if α > 1. By the nesting property, it follows more than two W-neurons is also a forbidden set, and that the W layer has the unconditional winner-take-all property. A.2 Part-whole combinations as permitted sets Theorem 1. Suppose that β < 1. If γ 2 < β + (1 − β)/k then any combination of k ≥ 1 parts consistent with a whole corresponds to a permitted set. Proof. Consider k parts belonging to a whole. They are represented by one W-neuron and k P-neurons, with synaptic connections given by the (k + 1) × (k + 1) matrix M= −β(11T − I) γ1 , γ1T 0 (5) where 1 is the k- dimensional vector whose elements are all equal to one. Two eigenvectors of M are of the form (1T c), and have the same eigenvalues as the 2 × 2 matrix −β(k − 1) γk γ 0 This matrix has eigenvalues less than one when γ 2 < β + (1 − β)/k and β(k − 1) + 2 > 0. The other k − 1 eigenvectors are of the form (dT , 0), where dT 1 = 0. These have eigenvalues β. Therefore all eigenvalues of W are less than one if the condition of the theorem is satisfied. A.3 Constraints on combining parts Here, we derive conditions under which the network can enforce the constraint that steady state activity be confined to parts that constitute a whole. Theorem 2. Suppose that β > 0 and σ 2 +β 2 +γ 2 +2σβγ > 1 If a W- neuron is active, then only P- neurons corresponding to parts contained in the relevant whole can be active at a stable steady state. Proof. Consider P- neurons Pi , Pj , and W- neuron Wa . Supa a pose that ξi = 1 but ξj = 0. As shown in Figure 5, the matrix of connections is given by: W = 0 −β γ −β 0 −σ γ −σ 0 (6) Wa γ Pi -σ -β Pj Figure 5: A set of one W- neuron and two P- neurons is forbidden if one part belongs to the whole and the other does not. This set is permitted if all eigenvalues of W − I have negative real parts. The characteristic equation of I − W is λ3 + b1 λ2 + b2 λ + b3 = 0, where b1 = 3, b2 = 3 − σ 2 − β 2 − γ 2 and b3 = 1−2σβγ−σ 2 −β 2 −γ 2 . According to the Routh- Hurwitz theorem, all the eigenvalues have negative real parts if and only if b1 > 0, b3 > 0 and b1 b2 > b3 . Clearly, the first condition is always satisfied. The second condition is more restrictive than the third. It is satisfied only when σ 2 + β 2 + γ 2 + 2σβγ < 1. Hence, one of the eigenvalues has a positive real part when this condition is broken, i.e., when σ 2 +β 2 +γ 2 +2σβγ > 1. By the nesting property, any larger set of P- neurons inconsistent with the W- neuron is also forbidden. A.4 Completion of wholes √ Theorem 3. If γ > β and a single W- neuron a is active at a steady state, then Pi > 0 a for all i such that ξi = 1. Proof. Suppose that the detected whole has k parts. At the steady state Pi = a ξi Bi − (β − γ 2 )Ptot 1−β + where Ptot = Pi = i 1 1 − β + (β − γ 2 )k k a B i ξi i=1 (7) A.5 Preventing runaway If feedback loops cause the network activity to diverge, then the preceding analyses are not relevant. Here we give a sufficient condition guaranteeing that runaway instability does not happen. It is not a necessary condition. Interestingly, the condition implies the condition of Theorem 1. Theorem 4. Suppose that P and W obey the dynamics of Eqs. (1) and (2), and define the objective function E 1−α 2 = − 2 Wa a α + 2 Wa a 1−β + 2 a Pi Wa ξi + σ Bi Pi − γ i 2 ia Pi2 i 2 β + 2 Pi i a (1 − ξi )Pi Wa . (8) ia Then E is a Lyapunov like function that, given β > γ 2 − dynamics to a stable steady state. 1−γ 2 N −1 , ensures convergence of the Proof. (sketch) Differentiation of E with respect to time shows that that E is nonincreasing in the nonnegative orthant and constant only at steady states of the network dynamics. We must also show that E is radially unbounded, which is true if the quadratic part of E is copositive definite. Note that the last term of E is lower-bounded by zero and the previous term is upper bounded by γ ia Pi Wa . We assume α > 1. Thus, we can use Cauchy’s 2 2 inequality, i Pi2 ≥ ( i Pi ) /N , and the fact that a Wa ≤ ( a Wa )2 for Wa ≥ 0, to derive E≥ 1 2 Wa )2 + ( If β > γ 2 − unbounded. a 1−γ 2 N −1 , 1 − β + βN ( N Pi )2 − 2γ( i Wa a Pi ) i − Bi Pi . (9) i the quadratic form in the inequality is positive definite and E is radially References [1] E. B. Baum, J. Moody, and F. Wilczek. Internal representations for associative memory. Biol. Cybern., 59:217–228, 1988. [2] K. Fukushima. Neocognitron: a self organizing neural network model for a mechanism of pattern recognition unaffected by shift in position. Biol Cybern, 36(4):193–202, 1980. [3] R.H. Hahnloser, R. Sarpeshkar, M.A. Mahowald, R.J. Douglas, and H.S. Seung. Digital selection and analogue amplification coexist in a cortex-inspired silicon circuit. Nature, 405(6789):947– 51, Jun 22 2000. [4] R.H. Hahnloser, H.S. Seung, and J.-J. Slotine. Permitted and forbidden sets in symmetric threshold-linear networks. Neural Computation, 15:621–638, 2003. [5] J.J. Hopfield. Neural networks and physical systems with emergent collective computational abilities. Proc Natl Acad Sci U S A, 79(8):2554–8, Apr 1982. [6] Y. LeCun, B. Boser, J. S. Denker, D. Henderson, R. E. Howard, W. Hubbard, and L. D. Jackel. Backpropagation applied to handwritten zip code recognition. Neural Comput., 1:541–551, 1989. [7] J. L. McClelland and D. E. Rumelhart. An interactive activation model of context effects in letter perception: Part i. an account of basic findings. Psychological Review, 88(5):375–407, Sep 1981. [8] M Riesenhuber and T Poggio. Hierarchical models of object recognition in cortex. Nat Neurosci, 2(11):1019–25, Nov 1999. [9] X. Xie, R.H. Hahnloser, and H. S. Seung. Selectively grouping neurons in recurrent networks of lateral inhibition. Neural Computation, 14:2627–2646, 2002.
2 0.66246223 6 nips-2005-A Connectionist Model for Constructive Modal Reasoning
Author: Artur Garcez, Luis C. Lamb, Dov M. Gabbay
Abstract: We present a new connectionist model for constructive, intuitionistic modal reasoning. We use ensembles of neural networks to represent intuitionistic modal theories, and show that for each intuitionistic modal program there exists a corresponding neural network ensemble that computes the program. This provides a massively parallel model for intuitionistic modal reasoning, and sets the scene for integrated reasoning, knowledge representation, and learning of intuitionistic theories in neural networks, since the networks in the ensemble can be trained by examples using standard neural learning algorithms. 1
3 0.55188245 157 nips-2005-Principles of real-time computing with feedback applied to cortical microcircuit models
Author: Wolfgang Maass, Prashant Joshi, Eduardo D. Sontag
Abstract: The network topology of neurons in the brain exhibits an abundance of feedback connections, but the computational function of these feedback connections is largely unknown. We present a computational theory that characterizes the gain in computational power achieved through feedback in dynamical systems with fading memory. It implies that many such systems acquire through feedback universal computational capabilities for analog computing with a non-fading memory. In particular, we show that feedback enables such systems to process time-varying input streams in diverse ways according to rules that are implemented through internal states of the dynamical system. In contrast to previous attractor-based computational models for neural networks, these flexible internal states are high-dimensional attractors of the circuit dynamics, that still allow the circuit state to absorb new information from online input streams. In this way one arrives at novel models for working memory, integration of evidence, and reward expectation in cortical circuits. We show that they are applicable to circuits of conductance-based Hodgkin-Huxley (HH) neurons with high levels of noise that reflect experimental data on invivo conditions. 1
4 0.53329784 61 nips-2005-Dynamical Synapses Give Rise to a Power-Law Distribution of Neuronal Avalanches
Author: Anna Levina, Michael Herrmann
Abstract: There is experimental evidence that cortical neurons show avalanche activity with the intensity of firing events being distributed as a power-law. We present a biologically plausible extension of a neural network which exhibits a power-law avalanche distribution for a wide range of connectivity parameters. 1
5 0.49450538 181 nips-2005-Spiking Inputs to a Winner-take-all Network
Author: Matthias Oster, Shih-Chii Liu
Abstract: Recurrent networks that perform a winner-take-all computation have been studied extensively. Although some of these studies include spiking networks, they consider only analog input rates. We present results of this winner-take-all computation on a network of integrate-and-fire neurons which receives spike trains as inputs. We show how we can configure the connectivity in the network so that the winner is selected after a pre-determined number of input spikes. We discuss spiking inputs with both regular frequencies and Poisson-distributed rates. The robustness of the computation was tested by implementing the winner-take-all network on an analog VLSI array of 64 integrate-and-fire neurons which have an innate variance in their operating parameters. 1
7 0.44915405 124 nips-2005-Measuring Shared Information and Coordinated Activity in Neuronal Networks
8 0.39708999 40 nips-2005-CMOL CrossNets: Possible Neuromorphic Nanoelectronic Circuits
9 0.39540637 108 nips-2005-Layered Dynamic Textures
10 0.38073471 64 nips-2005-Efficient estimation of hidden state dynamics from spike trains
11 0.36906493 62 nips-2005-Efficient Estimation of OOMs
12 0.36699495 73 nips-2005-Fast biped walking with a reflexive controller and real-time policy searching
13 0.3668831 118 nips-2005-Learning in Silicon: Timing is Everything
14 0.35577345 196 nips-2005-Two view learning: SVM-2K, Theory and Practice
15 0.33794561 171 nips-2005-Searching for Character Models
16 0.33155504 106 nips-2005-Large-scale biophysical parameter estimation in single neurons via constrained linear regression
17 0.29995415 99 nips-2005-Integrate-and-Fire models with adaptation are good enough
18 0.29935116 134 nips-2005-Neural mechanisms of contrast dependent receptive field size in V1
19 0.29459611 143 nips-2005-Off-Road Obstacle Avoidance through End-to-End Learning
20 0.27380842 203 nips-2005-Visual Encoding with Jittering Eyes
topicId topicWeight
[(3, 0.026), (10, 0.017), (27, 0.014), (31, 0.567), (34, 0.048), (39, 0.012), (41, 0.011), (55, 0.015), (57, 0.028), (60, 0.012), (69, 0.048), (73, 0.017), (88, 0.058), (91, 0.039)]
simIndex simValue paperId paperTitle
same-paper 1 0.96502304 164 nips-2005-Representing Part-Whole Relationships in Recurrent Neural Networks
Author: Viren Jain, Valentin Zhigulin, H. S. Seung
Abstract: There is little consensus about the computational function of top-down synaptic connections in the visual system. Here we explore the hypothesis that top-down connections, like bottom-up connections, reflect partwhole relationships. We analyze a recurrent network with bidirectional synaptic interactions between a layer of neurons representing parts and a layer of neurons representing wholes. Within each layer, there is lateral inhibition. When the network detects a whole, it can rigorously enforce part-whole relationships by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the theory of permitted and forbidden sets [3, 4]. The network behaviors are illustrated by recreating Rumelhart and McClelland’s “interactive activation” model [7]. In neural network models of visual object recognition [2, 6, 8], patterns of synaptic connectivity often reflect part-whole relationships between the features that are represented by neurons. For example, the connections of Figure 1 reflect the fact that feature B both contains simpler features A1, A2, and A3, and is contained in more complex features C1, C2, and C3. Such connectivity allows neurons to follow the rule that existence of the part is evidence for existence of the whole. By combining synaptic input from multiple sources of evidence for a feature, a neuron can “decide” whether that feature is present. 1 The synapses shown in Figure 1 are purely bottom-up, directed from simple to complex features. However, there are also top-down connections in the visual system, and there is little consensus about their function. One possibility is that top-down connections also reflect part-whole relationships. They allow feature detectors to make decisions using the rule that existence of the whole is evidence for existence of its parts. In this paper, we analyze the dynamics of a recurrent network in which part-whole relationships are stored as bidirectional synaptic interactions, rather than the unidirectional interactions of Figure 1. The network has a number of interesting computational capabilities. When the network detects a whole, it can rigorously enforce part-whole relationships 1 Synaptic connectivity may reflect other relationships besides part-whole. For example, invariances can be implemented by connecting detectors of several instances of the same feature to the same target, which is consequently an invariant detector of the feature. C1 C2 C3 B A1 A2 A3 Figure 1: The synaptic connections (arrows) of neuron B represent part-whole relationships. Feature B both contains simpler features and is contained in more complex features. The synaptic interactions are drawn one-way, as in most models of visual object recognition. Existence of the part is regarded as evidence for existence of the whole. This paper makes the interactions bidirectional, allowing the existence of the whole to be evidence for the existence of its parts. by ignoring parts that do not belong. The network can complete the whole by filling in missing parts. The network can refuse to recognize a whole, if the activated parts do not conform to a stored part-whole relationship. Parameter regimes in which these behaviors happen are identified using the recently developed theory of permitted and forbidden sets [3, 4]. Our model is closely related to the interactive activation model of word recognition, which was proposed by McClelland and Rumelhart to explain the word superiority effect studied by visual psychologists [7]. Here our concern is not to model a psychological effect, but to characterize mathematically how computations involving part-whole relationships can be carried out by a recurrent network. 1 Network model Suppose that we are given a set of part-whole relationships specified by a ξi = 1, if part i is contained in whole a 0, otherwise We assume that every whole contains at least one part, and every part is contained in at least one whole. The stimulus drives a layer of neurons that detect parts. These neurons also interact with a layer of neurons that detect wholes. We will refer to part-detectors as “P-neurons” and whole-detectors as “W-neurons.” The part-whole relationships are directly stored in the synaptic connections between P and a W neurons. If ξi = 1, the ith neuron in the P layer and the ath neuron in the W layer have a an excitatory interaction of strength γ. If ξi = 0, the neurons have an inhibitory interaction of strength σ. Furthermore, the P-neurons inhibit each other with strength β, and the Wneurons inhibit each other with strength α. All of these interactions are symmetric, and all activation functions are the rectification nonlinearity [z]+ = max{z, 0}. Then the dynamics of the network takes the form ˙ Wa + Wa a Pi ξ i − σ = γ i + a (1 − ξi )Pi − α i Wb , + ˙ Pi + Pi (1) b=a a Wa ξ i − σ = γ a a (1 − ξi )Wa − β a Pj + B i . j=i (2) where Bi is the input to the P layer from the stimulus. Figure 2 shows an example of a network with two wholes. Each whole contains two parts. One of the parts is contained in both wholes. -α Wa excitation γ -σ inhibition P1 B1 -β } W layer Wb -σ P2 -β B2 P3 } P layer B3 Figure 2: Model in example configuration: ξ = {(1, 1, 0), (0, 1, 1)}. When a stimulus is presented, it activates some of the P-neurons, which activate some of the W-neurons. The network eventually converges to a stable steady state. We will assume that α > 1. In the Appendix, we prove that this leads to unconditional winner-take-all behavior in the W layer. In other words, no more than one W-neuron can be active at a stable steady state. If a single W-neuron is active, then a whole has been detected. Potentially there are also many P-neurons active, indicating detection of parts. This representation may have different properties, depending on the choice of parameters β, γ, and σ. As discussed below, these include rigorous enforcement of part-whole relationships, completion of wholes by “filling in” missing parts, and non-recognition of parts that do not conform to a whole. 2 Enforcement of part-whole relationships Suppose that a single W-neuron is active at a stable steady state, so that a whole has been detected. Part-whole relationships are said to be enforced if the network always ignores parts that are not contained in the detected whole, despite potentially strong bottom-up evidence for them. It can be shown that enforcement follows from the inequality σ 2 + β 2 + γ 2 + 2σβγ > 1. (3) which guarantees that neuron i in the P layer is inactive, if neuron a in the W layer is a active and ξi = 0. When part-whole relations are enforced, prior knowledge about legal combinations of parts strictly constrains what may be perceived. This result is proven in the Appendix, and only an intuitive explanation is given here. Enforcement is easiest to understand when there is interlayer inhibition (σ > 0). In this case, the active W-neuron directly inhibits the forbidden P-neurons. The case of σ = 0 is more subtle. Then enforcement is mediated by lateral inhibition in the P layer. Excitatory feedback from the W-neuron has the effect of counteracting the lateral inhibition between the P-neurons that belong to the whole. As a result, these P-neurons become strongly activated enough to inhibit the rest of the P layer. 3 Completion of wholes by filling in missing parts If a W-neuron is active, it excites the P-neurons that belong to the whole. As a result, even if one of these P-neurons receives no bottom-up input (Bi = 0), it is still active. We call this phenomenon “completion,” and it is guaranteed to happen when (4) γ> β The network may thus “imagine” parts that are consistent with the recognized whole, but are not actually present in the stimulus. As with enforcement, this condition depends on top-down connections. √ In the special case γ = β, the interlayer excitation between a W-neuron and its P-neurons exactly cancels out the lateral inhibition between the P-neurons at a steady state. So the recurrent connections effectively vanish, letting the activity of the P-neurons be determined by their feedforward inputs. When the interlayer excitation is stronger than this, the inequality (4) holds, and completion occurs. 4 Non-recognition of a whole If there is no interlayer inhibition (σ = 0), then a single W-neuron is always active, assuming that there is some activity in the P layer. To see this, suppose for the sake of contradiction that all the W-neurons are inactive. Then they receive no inhibition to counteract the excitation from the P layer. This means some of them must be active, which contradicts our assumption. This means that the network always recognizes a whole, even if the stimulus is very different from any part-whole combination that is stored in the network. However, if interlayer inhibition is sufficiently strong (large σ), the network may refuse to recognize a whole. Neurons in the P layer are activated, but there is no activity in the W layer. Formal conditions on σ can be derived, but are not given here because of space limitations. In case of non-recognition, constraints on the P-layer are not enforced. It is possible for the network to detect a configuration of parts that is not consistent with any stored whole. 5 Example: Interactive Activation model To illustrate the computational capabilities of our network, we use it to recreate the interactive activation (IA) model of McClelland and Rumelhart. Figure 3 shows numerical simulations of a network containing three layers of neurons representing strokes, letters, and words, respectively. There are 16 possible strokes in each of four letter positions. For each stroke, there are two neurons, one signaling the presence of the stroke and the other signaling its absence. Letter neurons represent each letter of the alphabet in each of four positions. Word neurons represent each of 1200 common four letter words. The letter and word layers correspond to the P and W layers that were introduced previously. There are bidirectional interactions between the letter and word layers, and lateral inhibition within the layers. The letter neurons also receive input from the stroke neurons, but this interaction is unidirectional. Our network differs in two ways from the original IA model. First, all interactions involving letter and word neurons are symmetric. In the original model, the interactions between the letter and word layers were asymmetric. In particular, inhibitory connections only ran from letter neurons to word neurons, and not vice versa. Second, the only nonlinearity in our model is rectification. These two aspects allow us to apply the full machinery of the theory of permitted and forbidden sets. Figure 3 shows the result of presenting the stimulus “MO M” for four different settings of parameters. In each of the four cases, the word layer of the network converges to the same result, detecting the word “MOON”, which is the closest stored word to the stimulus. However, the activity in the letter layer is different in the four cases. input: P layer reconstruction W layer P layer reconstruction W layer completion noncompletion enforcement non-enforcement Figure 3: Simulation of 4 different parameter regimes in a letter- word recognition network. Within each panel, the middle column presents a feature- layer reconstruction based on the letter activity shown in the left column. W layer activity is shown in the right column. The top row shows the network state after 10 iterations of the dynamics. The bottom row shows the steady state. In the left column, the parameters obey the inequality (3), so that part- whole relationships are enforced. The activity of the letter layer is visualized by activating the strokes corresponding to each active letter neuron. The activated letters are part of the word “MOON”. In the top left, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom left, completion does not occur. In the simulations of the right column, parameters are such that part- whole relationships are not enforced. Consequently, the word layer is much more active. Bottom- up input provides evidence for several other letters, which is not suppressed. In the top right, the inequality (4) is satisfied, so that the missing “O” in the stimulus is filled in. In the bottom right, the “O” neuron is not activated in the third position, so there is no completion. However, some letter neurons for the third position are activated, due to the input from neurons that indicate the absence of strokes. input: non-recognition event multi-stability Figure 4: Simulation of a non- recognition event and example of multistability. Figure 4 shows simulations for large σ, deep in the enforcement regime where non- recognition is a possibility. From one initial condition, the network converges to a state in which no W neurons are active, a non- recognition. From another initial condition, the network detects the word “NORM”. Deep in the enforcement regime, the top- down feedback can be so strong that the network has multiple stable states, many of which bear little resemblance to the stimulus at all. This is a problematic aspect of this network. It can be prevented by setting parameters at the edge of the enforcement regime. 6 Discussion We have analyzed a recurrent network that performs computations involving part- whole relationships. The network can fill in missing parts and suppress parts that do not belong. These two computations are distinct and can be dissociated from each other, as shown in Figure 3. While these two computations can also be performed by associative memory models, they are not typically dissociable in these models. For example, in the Hopfield model pattern completion and noise suppression are both the result of recall of one of a finite number of stereotyped activity patterns. We believe that our model is more appropriate for perceptual systems, because its behavior is piecewise linear, due its reliance on rectification nonlinearity. Therefore, analog aspects of computation are able to coexist with the part-whole relationships. Furthermore, in our model the stimulus is encoded in maintained synaptic input to the network, rather than as an initial condition of the dynamics. A Appendix: Permitted and forbidden sets Our mathematical results depend on the theory of permitted and forbidden sets [3, 4], which is summarized briefly here. The theory is applicable to neural networks with rectification nonlinearity, of the form xi + xi = [bi + j Wij xj ]+ . Neuron i is said to be active when ˙ xi > 0. For a network of N neurons, there are 2N possible sets of active neurons. For each active set, consider the submatrix of Wij corresponding to the synapses between active neurons. If all eigenvalues of this submatrix have real parts less than or equal to unity, then the active set is said to be permitted. Otherwise the active set is said to be forbidden. A set is permitted if and only if there exists an input vector b such that those neurons are active at a stable steady state. Permitted sets can be regarded as memories stored in the synaptic connections Wij . If Wij is a symmetric matrix, the nesting property holds: every subset of a permitted set is permitted, and every superset of a forbidden set is forbidden. The present model can be seen as a general method for storing permitted sets in a recurrent network. This method introduces a neuron for each permitted set, relying on a unary or “grandmother cell” representation. In contrast, Xie et al.[9] used lateral inhibition in a single layer of neurons to store permitted sets. By introducing extra neurons, the present model achieves superior storage capacity, much as unary models of associative memory [1] surpass distributed models [5]. A.1 Unconditional winner-take-all in the W layer The synapses between two W-neurons have strengths 0 −α −α 0 The eigenvalues of this matrix are ±α. Therefore two W-neurons constitute a forbidden set if α > 1. By the nesting property, it follows more than two W-neurons is also a forbidden set, and that the W layer has the unconditional winner-take-all property. A.2 Part-whole combinations as permitted sets Theorem 1. Suppose that β < 1. If γ 2 < β + (1 − β)/k then any combination of k ≥ 1 parts consistent with a whole corresponds to a permitted set. Proof. Consider k parts belonging to a whole. They are represented by one W-neuron and k P-neurons, with synaptic connections given by the (k + 1) × (k + 1) matrix M= −β(11T − I) γ1 , γ1T 0 (5) where 1 is the k- dimensional vector whose elements are all equal to one. Two eigenvectors of M are of the form (1T c), and have the same eigenvalues as the 2 × 2 matrix −β(k − 1) γk γ 0 This matrix has eigenvalues less than one when γ 2 < β + (1 − β)/k and β(k − 1) + 2 > 0. The other k − 1 eigenvectors are of the form (dT , 0), where dT 1 = 0. These have eigenvalues β. Therefore all eigenvalues of W are less than one if the condition of the theorem is satisfied. A.3 Constraints on combining parts Here, we derive conditions under which the network can enforce the constraint that steady state activity be confined to parts that constitute a whole. Theorem 2. Suppose that β > 0 and σ 2 +β 2 +γ 2 +2σβγ > 1 If a W- neuron is active, then only P- neurons corresponding to parts contained in the relevant whole can be active at a stable steady state. Proof. Consider P- neurons Pi , Pj , and W- neuron Wa . Supa a pose that ξi = 1 but ξj = 0. As shown in Figure 5, the matrix of connections is given by: W = 0 −β γ −β 0 −σ γ −σ 0 (6) Wa γ Pi -σ -β Pj Figure 5: A set of one W- neuron and two P- neurons is forbidden if one part belongs to the whole and the other does not. This set is permitted if all eigenvalues of W − I have negative real parts. The characteristic equation of I − W is λ3 + b1 λ2 + b2 λ + b3 = 0, where b1 = 3, b2 = 3 − σ 2 − β 2 − γ 2 and b3 = 1−2σβγ−σ 2 −β 2 −γ 2 . According to the Routh- Hurwitz theorem, all the eigenvalues have negative real parts if and only if b1 > 0, b3 > 0 and b1 b2 > b3 . Clearly, the first condition is always satisfied. The second condition is more restrictive than the third. It is satisfied only when σ 2 + β 2 + γ 2 + 2σβγ < 1. Hence, one of the eigenvalues has a positive real part when this condition is broken, i.e., when σ 2 +β 2 +γ 2 +2σβγ > 1. By the nesting property, any larger set of P- neurons inconsistent with the W- neuron is also forbidden. A.4 Completion of wholes √ Theorem 3. If γ > β and a single W- neuron a is active at a steady state, then Pi > 0 a for all i such that ξi = 1. Proof. Suppose that the detected whole has k parts. At the steady state Pi = a ξi Bi − (β − γ 2 )Ptot 1−β + where Ptot = Pi = i 1 1 − β + (β − γ 2 )k k a B i ξi i=1 (7) A.5 Preventing runaway If feedback loops cause the network activity to diverge, then the preceding analyses are not relevant. Here we give a sufficient condition guaranteeing that runaway instability does not happen. It is not a necessary condition. Interestingly, the condition implies the condition of Theorem 1. Theorem 4. Suppose that P and W obey the dynamics of Eqs. (1) and (2), and define the objective function E 1−α 2 = − 2 Wa a α + 2 Wa a 1−β + 2 a Pi Wa ξi + σ Bi Pi − γ i 2 ia Pi2 i 2 β + 2 Pi i a (1 − ξi )Pi Wa . (8) ia Then E is a Lyapunov like function that, given β > γ 2 − dynamics to a stable steady state. 1−γ 2 N −1 , ensures convergence of the Proof. (sketch) Differentiation of E with respect to time shows that that E is nonincreasing in the nonnegative orthant and constant only at steady states of the network dynamics. We must also show that E is radially unbounded, which is true if the quadratic part of E is copositive definite. Note that the last term of E is lower-bounded by zero and the previous term is upper bounded by γ ia Pi Wa . We assume α > 1. Thus, we can use Cauchy’s 2 2 inequality, i Pi2 ≥ ( i Pi ) /N , and the fact that a Wa ≤ ( a Wa )2 for Wa ≥ 0, to derive E≥ 1 2 Wa )2 + ( If β > γ 2 − unbounded. a 1−γ 2 N −1 , 1 − β + βN ( N Pi )2 − 2γ( i Wa a Pi ) i − Bi Pi . (9) i the quadratic form in the inequality is positive definite and E is radially References [1] E. B. Baum, J. Moody, and F. Wilczek. Internal representations for associative memory. Biol. Cybern., 59:217–228, 1988. [2] K. Fukushima. Neocognitron: a self organizing neural network model for a mechanism of pattern recognition unaffected by shift in position. Biol Cybern, 36(4):193–202, 1980. [3] R.H. Hahnloser, R. Sarpeshkar, M.A. Mahowald, R.J. Douglas, and H.S. Seung. Digital selection and analogue amplification coexist in a cortex-inspired silicon circuit. Nature, 405(6789):947– 51, Jun 22 2000. [4] R.H. Hahnloser, H.S. Seung, and J.-J. Slotine. Permitted and forbidden sets in symmetric threshold-linear networks. Neural Computation, 15:621–638, 2003. [5] J.J. Hopfield. Neural networks and physical systems with emergent collective computational abilities. Proc Natl Acad Sci U S A, 79(8):2554–8, Apr 1982. [6] Y. LeCun, B. Boser, J. S. Denker, D. Henderson, R. E. Howard, W. Hubbard, and L. D. Jackel. Backpropagation applied to handwritten zip code recognition. Neural Comput., 1:541–551, 1989. [7] J. L. McClelland and D. E. Rumelhart. An interactive activation model of context effects in letter perception: Part i. an account of basic findings. Psychological Review, 88(5):375–407, Sep 1981. [8] M Riesenhuber and T Poggio. Hierarchical models of object recognition in cortex. Nat Neurosci, 2(11):1019–25, Nov 1999. [9] X. Xie, R.H. Hahnloser, and H. S. Seung. Selectively grouping neurons in recurrent networks of lateral inhibition. Neural Computation, 14:2627–2646, 2002.
2 0.93049872 204 nips-2005-Walk-Sum Interpretation and Analysis of Gaussian Belief Propagation
Author: Dmitry Malioutov, Alan S. Willsky, Jason K. Johnson
Abstract: This paper presents a new framework based on walks in a graph for analysis and inference in Gaussian graphical models. The key idea is to decompose correlations between variables as a sum over all walks between those variables in the graph. The weight of each walk is given by a product of edgewise partial correlations. We provide a walk-sum interpretation of Gaussian belief propagation in trees and of the approximate method of loopy belief propagation in graphs with cycles. This perspective leads to a better understanding of Gaussian belief propagation and of its convergence in loopy graphs. 1
3 0.89922345 108 nips-2005-Layered Dynamic Textures
Author: Antoni B. Chan, Nuno Vasconcelos
Abstract: A dynamic texture is a video model that treats a video as a sample from a spatio-temporal stochastic process, specifically a linear dynamical system. One problem associated with the dynamic texture is that it cannot model video where there are multiple regions of distinct motion. In this work, we introduce the layered dynamic texture model, which addresses this problem. We also introduce a variant of the model, and present the EM algorithm for learning each of the models. Finally, we demonstrate the efficacy of the proposed model for the tasks of segmentation and synthesis of video.
4 0.88729209 65 nips-2005-Estimating the wrong Markov random field: Benefits in the computation-limited setting
Author: Martin J. Wainwright
Abstract: Consider the problem of joint parameter estimation and prediction in a Markov random field: i.e., the model parameters are estimated on the basis of an initial set of data, and then the fitted model is used to perform prediction (e.g., smoothing, denoising, interpolation) on a new noisy observation. Working in the computation-limited setting, we analyze a joint method in which the same convex variational relaxation is used to construct an M-estimator for fitting parameters, and to perform approximate marginalization for the prediction step. The key result of this paper is that in the computation-limited setting, using an inconsistent parameter estimator (i.e., an estimator that returns the “wrong” model even in the infinite data limit) is provably beneficial, since the resulting errors can partially compensate for errors made by using an approximate prediction technique. En route to this result, we analyze the asymptotic properties of M-estimators based on convex variational relaxations, and establish a Lipschitz stability property that holds for a broad class of variational methods. We show that joint estimation/prediction based on the reweighted sum-product algorithm substantially outperforms a commonly used heuristic based on ordinary sum-product. 1 Keywords: Markov random fields; variational method; message-passing algorithms; sum-product; belief propagation; parameter estimation; learning. 1
5 0.88106167 145 nips-2005-On Local Rewards and Scaling Distributed Reinforcement Learning
Author: Drew Bagnell, Andrew Y. Ng
Abstract: We consider the scaling of the number of examples necessary to achieve good performance in distributed, cooperative, multi-agent reinforcement learning, as a function of the the number of agents n. We prove a worstcase lower bound showing that algorithms that rely solely on a global reward signal to learn policies confront a fundamental limit: They require a number of real-world examples that scales roughly linearly in the number of agents. For settings of interest with a very large number of agents, this is impractical. We demonstrate, however, that there is a class of algorithms that, by taking advantage of local reward signals in large distributed Markov Decision Processes, are able to ensure good performance with a number of samples that scales as O(log n). This makes them applicable even in settings with a very large number of agents n. 1
6 0.61481857 78 nips-2005-From Weighted Classification to Policy Search
7 0.5774954 153 nips-2005-Policy-Gradient Methods for Planning
8 0.54705822 46 nips-2005-Consensus Propagation
9 0.53290457 142 nips-2005-Oblivious Equilibrium: A Mean Field Approximation for Large-Scale Dynamic Games
10 0.5240798 111 nips-2005-Learning Influence among Interacting Markov Chains
11 0.51335216 187 nips-2005-Temporal Abstraction in Temporal-difference Networks
12 0.50322604 154 nips-2005-Preconditioner Approximations for Probabilistic Graphical Models
13 0.49474859 124 nips-2005-Measuring Shared Information and Coordinated Activity in Neuronal Networks
14 0.4876163 36 nips-2005-Bayesian models of human action understanding
15 0.48512825 181 nips-2005-Spiking Inputs to a Winner-take-all Network
16 0.48443264 96 nips-2005-Inference with Minimal Communication: a Decision-Theoretic Variational Approach
17 0.47491607 139 nips-2005-Non-iterative Estimation with Perturbed Gaussian Markov Processes
18 0.46942332 72 nips-2005-Fast Online Policy Gradient Learning with SMD Gain Vector Adaptation
19 0.46061122 122 nips-2005-Logic and MRF Circuitry for Labeling Occluding and Thinline Visual Contours
20 0.46000838 90 nips-2005-Hot Coupling: A Particle Approach to Inference and Normalization on Pairwise Undirected Graphs