nips nips2011 nips2011-256 knowledge-graph by maker-knowledge-mining

256 nips-2011-Solving Decision Problems with Limited Information


Source: pdf

Author: Denis D. Maua, Cassio Campos

Abstract: We present a new algorithm for exactly solving decision-making problems represented as an influence diagram. We do not require the usual assumptions of no forgetting and regularity, which allows us to solve problems with limited information. The algorithm, which implements a sophisticated variable elimination procedure, is empirically shown to outperform a state-of-the-art algorithm in randomly generated problems of up to 150 variables and 1064 strategies. 1

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 The algorithm, which implements a sophisticated variable elimination procedure, is empirically shown to outperform a state-of-the-art algorithm in randomly generated problems of up to 150 variables and 1064 strategies. [sent-7, score-0.256]

2 Influence diagrams [6] are representational devices for utility-based decision making under uncertainty. [sent-11, score-0.28]

3 Many popular decision-making frameworks such as finite-horizon POMDPs can be casted as influence diagrams [7]. [sent-12, score-0.186]

4 Traditionally, influence diagrams target problems involving a single, non-forgetful decision maker; this makes them unfitted to represent decision-making with limited information. [sent-13, score-0.28]

5 Limited memory influence diagrams (LIMIDs) generalize influence diagrams to allow for (explicit representation of) bounded memory policies and simultaneous decisions [1, 2]. [sent-14, score-0.5]

6 More precisely, LIMIDs relax the regularity and no forgetting assumptions of influence diagrams, namely, that there is a complete temporal ordering over the decisions, and that observations and decisions are permanently remembered. [sent-15, score-0.172]

7 Under certain graph-structural conditions (which no forgetting and regularity imply), Lauritzen and Nilsson [2] show that LIMIDs can be solved by dynamic programming with complexity exponential in the treewidth of the graph. [sent-18, score-0.19]

8 In this paper, we formally describe LIMIDs (Section 2) and show that policies can be partially ordered, and that the ordering can be extended monotonically, allowing for the generalized variable elimination procedure in Section 3. [sent-21, score-0.347]

9 In fact, our algorithm is orders of magnitude faster than the CR algorithm on randomly generated diagrams containing up to 150 variables. [sent-23, score-0.186]

10 Decision and chance variables have domains different from the empty domain, whereas value variables are always associated to the empty domain. [sent-29, score-0.287]

11 If x and y are sets of variables such that y ⊆ x ⊆ U, and x is an element of the domain Ωx , we write x↓y to denote the projection of x onto the smaller domain Ωy , that is, x↓y ∈ Ωy contains only the components of x that are compatible with the variables in y. [sent-34, score-0.174]

12 Arcs entering chance and value nodes denote stochastic and functional dependency, respectively; arcs entering decision nodes describe information awareness or relevance at the time the decision is made. [sent-51, score-0.406]

13 If X is a node in L, we denote by paX the set of parents of X, that is, the set of nodes of L from which there is an arc pointing to X. [sent-52, score-0.173]

14 , nodes to which there is an arc from X), and faX paX ∪ {X} denote pa its family. [sent-55, score-0.198]

15 Each chance variable C in C has an associated function pC C specifying the probability Pr(C = x↓C |paC = x↓paC ) of C assuming value x↓C ∈ ΩC given that the parents take on values x↓paC ∈ ΩpaC for all x ∈ ΩfaC . [sent-56, score-0.267]

16 We assume that the probabilities associated to any chance node respect the Markov condition, that is, that any variable X ∈ C is stochastically independent from its non-descendant non-parents given its parents. [sent-57, score-0.209]

17 Each value variable V ∈ V is associated to a bounded real-valued utility function uV over ΩpaV , which quantifies the (additive) contribution of the states of its parents to the overall utility. [sent-58, score-0.303]

18 Thus, the overall utility of a joint state x ∈ ΩC∪D is given by the sum of utility functions V ∈V uV (x↓paV ). [sent-59, score-0.256]

19 For any decision variable D ∈ D, a policy δD specifies an action for each possible state configuration of its parents, that is, δD : ΩpaD → ΩD . [sent-60, score-0.22]

20 The robot has 9 time steps to first reach a position sA of the grid, for which it receives 10 points, and then a position sB , for which it is rewarded with 20 points. [sent-64, score-0.174]

21 The S D function pSt−1 t associated to St specifies the probabilities Pr(St = st |St−1 = st−1 , Dt = dt ) of t transitioning to state St = st from a state St−1 = st−1 when the robot executes action Dt = dt . [sent-82, score-0.341]

22 The function pStt is associated to Ot and quantifies the likelihood of estimating position Ot = ot O when in position St = st . [sent-83, score-0.233]

23 The utility function uRt associated to Rt equals 10 if st = sA and At = n and Bt = n, 20 if st = sB and At = y and Bt = n, and zero otherwise. [sent-88, score-0.358]

24 Given a policy δD , let pD D denote a function such that pa for each x ∈ ΩfaD it equals one if x↓D = δD (x↓paD ) and zero otherwise. [sent-91, score-0.232]

25 There is a one-to-one correspondence between pa pa functions pD D and policies δD ∈ ∆D , and specifying a policy δD is equivalent to specifying pD D . [sent-93, score-0.397]

26 A strategy s induces a joint probability mass function over the variables in C ∪ D by pa pa ps pC C pD D , (1) C∈C D∈D and has an associated expected utility given by Es [L] uV (x↓paV ) = ps (x) x∈ΩC∪D V ∈V ps C∪D uV . [sent-95, score-0.727]

27 Given a LIMID L of treewidth ω, we can evaluate the expected utility of any strategy s in time and space at most exponential in ω. [sent-97, score-0.291]

28 In fact, computing the MEU is NP-hard even in bounded treewidth diagrams [8]. [sent-109, score-0.316]

29 Each valuation is associated to a subset of the variables in U, called its scope. [sent-114, score-0.361]

30 More concretely, we define a valuation φ with scope x as a pair (p, u) of bounded nonnegative real-valued functions p and u over the domain Ωx ; we refer to p and u as the probability and utility part, respectively, of φ. [sent-115, score-0.604]

31 Often, we write φx to make explicit the scope x of a valuation φ. [sent-116, score-0.404]

32 For any x ⊆ U, we denote the set of all possible valuations with scope x by Φx . [sent-117, score-0.565]

33 The set of all possible valuations is given by Φ x⊆U Φx . [sent-118, score-0.455]

34 If φ = (p, u) and ψ = (q, v) are valuations with scopes x and y, respectively, its combination φ ⊗ ψ is the valuation (pq, pv + qu) with scope x ∪ y. [sent-121, score-0.952]

35 If φ = (p, u) is a valuation with scope x, and y is a set of variables such that y ⊆ x, the marginal φ↓y is the valuation ( x\y p, x\y u) with scope y. [sent-123, score-0.846]

36 The following result shows that our framework respects the necessary conditions for computing efficiently with valuations (in the sense of keeping the scope of valuations minimal during the variable elimination procedure). [sent-125, score-1.238]

37 The system (Φ, U, ⊗, ↓) satisfies the following three axioms of a (weak) labeled valuation algebra [11, 12]. [sent-127, score-0.348]

38 To show (A3), consider any two valuations (p, u) and (q, v) with scopes x and y, respectively, and a set z such that x ⊆ z ⊆ x ∪ y. [sent-133, score-0.482]

39 y The framework of valuations allows us to compute the expected utility of a given strategy efficiently: Proposition 3. [sent-140, score-0.639]

40 Given a LIMID L and a strategy s = (δD )D∈D , let pa pa (pC C , 0) ⊗ φs C∈C (pD D , 0) ⊗ D∈D (1, uV ) , (3) V ∈V pa where, for each D, pD D is the function in PD associated with policy δD . [sent-141, score-0.484]

41 By definition of s pa combination, we have that φs = (ps , ps V ∈V uV ), where ps = X∈C∪D pX X as in (1). [sent-145, score-0.28]

42 Given any strategy s, we can use a variable elimination procedure to efficiently compute φ↓∅ and s hence its expected utility in time polynomial in the largest domain of a variable but exponential in the width of the elimination ordering. [sent-149, score-0.702]

43 1 Figure 2 shows a variable elimination procedure used to compute the expected utility of a strategy of the simple LIMID on the left-hand side. [sent-150, score-0.402]

44 For any two valuations φ = (p, u) and ψ = (q, v) in Φ, if φ and ψ have equal scope, p ≤ q and u ≤ v, then φ ≤ ψ holds. [sent-156, score-0.455]

45 The system (Φ, U, ⊗, ↓, ≤) satisfies the following two additional axioms of an ordered valuation algebra [13]. [sent-159, score-0.378]

46 Consider two valuations (px , ux ) and (qx , vx ) with scope x such that (px , ux ) ≤ (qx , vx ), and two valuations (py , uy ) and (qy , vy ) with scope y satisfying (py , uy ) ≤ (qy , vy ). [sent-164, score-1.83]

47 By definition of ≤, we have that px ≤ qx , ux ≤ vx , py ≤ qy and uy ≤ vy . [sent-165, score-0.767]

48 Since all functions are nonnegative, it follows that px py ≤ qx qy , px uy ≤ qx vy and py ux ≤ qy vx . [sent-166, score-1.184]

49 Hence, (px , ux ) ⊗ (py , uy ) = (px py , px uy + py ux ) ≤ (qx qy , qx vy + qy vx ) = (qx , vx ) ⊗ (qy , vy ). [sent-167, score-1.29]

50 It follows from monotonicity of ≤ with respect to addition of real numbers that (px , ux )↓y = ( x\y px , x\y ux ) ≤ ( x\y qx , x\y vx ) = (qx , vx )↓y . [sent-170, score-0.636]

51 To illustrate this, consider the variable elimination scheme in Figure 2 for two different strategies s s s s and s , and let ψ1 , ψ2 , ψ3 be the valuations produced in the propagation step for strategy s and s s s s s s s ψ1 , ψ2 , ψ3 the valuations for s . [sent-172, score-1.281]

52 As a consequence, we can abort variable elimination for s after the second iteration. [sent-174, score-0.218]

53 We can also exploit the redundancy between valuations produced during variable elimination for neighbor strategies. [sent-175, score-0.673]

54 If Ψx is a set of valuations with scope x and Ψy is a set of valuations with scope y the operation Ψx ⊗ Ψy {φx ⊗ φy : φx ∈ Ψx , φy ∈ Ψy } returns the set of combinations of a valuation in Ψx and a valuation in Ψy . [sent-178, score-1.741]

55 For X ∈ x, the operation Ψ−X {φ−X : φx ∈ Ψx } eliminates variable X x x from all valuations in Ψx . [sent-179, score-0.541]

56 Given a finite set of valuations Ψ ⊆ Φ, we say that a valuation φ ∈ Ψ is maximal if for all ψ ∈ Ψ such that φ ≤ ψ it holds that ψ ≤ φ. [sent-180, score-0.785]

57 The operator prune returns the set prune(Ψ) of maximal valuations of Ψ (by pruning non-maximal valuations). [sent-181, score-0.833]

58 1 The width of an elimination ordering is the maximum cardinality of the scope of a valuation produced during variable elimination minus one. [sent-182, score-0.909]

59 Consider a LIMID L and an elimination ordering X1 < · · · < Xn over the variables in C ∪ D. [sent-184, score-0.262]

60 The elimination ordering can be selected using the standard methods for Bayesian networks [9]. [sent-185, score-0.224]

61 Note that unlike standard algorithms for variable elimination in influence diagrams we allow any elimination ordering. [sent-186, score-0.559]

62 The algorithm is initialized by generating one set of valuations for each variable X in U as follows. [sent-187, score-0.518]

63 For each chance variable X ∈ C, add a singleton ΨX pa {(pX X , 0)} to V0 ; 2. [sent-190, score-0.265]

64 For each decision variable X ∈ D, add a set of valuations ΨX to V0 ; 3. [sent-191, score-0.612]

65 For each value variable X ∈ V, add a singleton ΨX pa pa {(pX X , 0) : pX X ∈ PX } {(1, uX )} to V0 . [sent-192, score-0.287]

66 Once V0 has been initialized with a set of valuations for each variable in the diagram, we recursively eliminate a variable Xi in C ∪ D in the given ordering and remove any non-maximal valuation: Propagation: For i = 1, . [sent-193, score-0.65]

67 Let Bi be the set of all valuations in Vi−1 whose scope contains Xi ; 2. [sent-197, score-0.565]

68 Finally, the algorithm outputs the utility part of the single maximal valuation in the set Termination: Return the real number u such that (p, u) ∈ prune( Ψ∈Vn Ψ∈Vn Ψ: Ψ). [sent-200, score-0.458]

69 u is a real number because the valuations in Ψ∈Vn Ψ have empty scope and thus both their probability and utility parts are identified with real numbers. [sent-201, score-0.739]

70 The following result is a straightforward extension of [14, Lemma 1(iv)] that is needed to guarantee the correctness of discarding non-maximal valuations in the propagation step. [sent-202, score-0.535]

71 If Ψx and Ψy are two sets of ordered valuations and z ⊆ x then (i) prune(Ψx ⊗prune(Ψy )) = prune(Ψx ⊗Ψy ) and (ii) prune(prune(Ψx )↓z ) = prune(Ψ↓z ). [sent-205, score-0.485]

72 x The result shows that, like marginalization, the prune operation distributes over any factorization of X∈U ΨX . [sent-206, score-0.365]

73 The following lemma shows that at any iteration i of the propagation step the combination of all sets in the current pool of sets Vi produces the set of maximal valuations of the initial factorization. [sent-207, score-0.606]

74 The basis is easily obtained by applying Lemmas 2 and 5 and the axioms of valuation algebra to prune([ Ψ∈V0 Ψ]−X1 ) in order to obtain prune( Ψ∈V1 Ψ). [sent-215, score-0.348]

75 By eliminating Xi+1 from both sides and then applying the prune operation we get to prune([prune([ Ψ∈V0 Ψ]−X1 ···Xi )]−Xi+1 ) = prune([prune( Ψ∈Vi Ψ)]−Xi+1 ). [sent-217, score-0.365]

76 According to Proposition 3, each element φ−X1 ···Xn in Ψ−X1 ···Xn is a valuation whose probability part is one and utility part equals s L Es [L]. [sent-221, score-0.479]

77 Thus, the maximal expected utility MEU[L] is the utility part of the single valuation in prune(Ψ−X1 ···Xn ). [sent-222, score-0.586]

78 It is not difficult to see that after the initialization step, the set V0 contains sets L 6 Ψ of valuations such that Ψ∈V0 Ψ = ΨL . [sent-223, score-0.486]

79 Hence, Lemma 6 states that after the last iteration, MPU produces a set Vn of sets of valuations such that prune( Ψ∈Vn Ψ) = prune(Ψ−X1 ···Xn ) = L MEU[L]. [sent-224, score-0.455]

80 The algorithm returns the utility part of a valuation (p, u) in prune( Ψ∈Vn Ψ), which, by Lemma 6 for i = n, equals prune([ Ψ∈V0 Ψ]↓∅ ). [sent-229, score-0.479]

81 By definition of V0 , any valuation φ in ( Ψ∈V0 Ψ) factorizes as in (3). [sent-230, score-0.294]

82 Also, there is exactly one valuation φ ∈ ( Ψ∈V0 Ψ) for each strategy in ∆. [sent-231, score-0.35]

83 Moreover, since functions with empty scope correspond to numbers, the relation ≤ specifies a total ordering over the valuations in ( Ψ∈V0 Ψ)↓∅ , which implies a single maximal element. [sent-233, score-0.716]

84 The time complexity of the algorithm is given by the cost of creating the sets of valuations in the initialization step plus the overall cost of the combination and marginalization operations performed during the propagation step. [sent-236, score-0.633]

85 For any decision variable D, let ρD |ΩD ||ΩpaD | denote the number of policies in ∆D (which coincides with the number of functions in PD ). [sent-238, score-0.217]

86 There is exactly one valuation in ΨD in V0 for every policy in ∆D . [sent-239, score-0.357]

87 Then the initialization loop for decision variables takes O(|D|ρ) time, which is exponential in the input (the sets of policies are not considered as an input of the problem). [sent-241, score-0.223]

88 As with any variable elimination procedure, the running time of propagating (sets of) valuations is exponential in the width of the given ordering, which is in the best case given by the treewidth of the diagram. [sent-243, score-0.813]

89 Consider the case of an ordering with bounded width ω and a bounded number of states per variable κ. [sent-244, score-0.211]

90 In the worst case, ν is equal to ρ|D| ≤ O(κ|D|κ ), that is, all sets associated to decision variables have been combined without discarding any valuation. [sent-251, score-0.192]

91 Hence, the worst-case complexity of the propagation step is exponential in the input, even if the ordering width and the number of states per variable are bounded. [sent-252, score-0.214]

92 Each LIMID is parameterized by the number of decision nodes d, the number of chance nodes c, the maximum cardinality of the domain of a chance variable ωC , and the maximum cardinality of the domain of a decision variable ωD . [sent-256, score-0.736]

93 Then we repeatedly add an arc from a decision node with no children to a value node with no parents (so that each decision node has at least one value node as children). [sent-262, score-0.4]

94 Finally, we repeatedly add an arc that neither makes the domain of a variable greater than the given bounds nor makes the treewidth more than 10, until no arcs can be added without exceeding the bounds. [sent-264, score-0.307]

95 2 Note that this generates diagrams where decision and chance variables have at most log2 ωD − 1 and log2 ωC − 1 parents, respectively. [sent-265, score-0.408]

96 We instead use a greedy heuristic that resulted in diagrams whose treewidth ranged from 5 to 10. [sent-269, score-0.293]

97 Within the limit of 12 hours, MPU was able to compute diagrams containing up to 1064 strategies, whereas CR solved diagrams with at most 1025 strategies. [sent-281, score-0.397]

98 5 Conclusion LIMIDs are highly expressive models for utility-based decision making that subsume influence diagrams and finite-horizon (partially observable) Markov decision processes. [sent-284, score-0.374]

99 This can be done by arbitrarily discarding valuations during the propagation step so as to bound the size of propagated sets. [sent-290, score-0.535]

100 Ordered valuation algebras: a generic framework for approximating inference. [sent-373, score-0.294]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('valuations', 0.455), ('prune', 0.342), ('valuation', 0.294), ('limid', 0.259), ('mpu', 0.228), ('limids', 0.198), ('diagrams', 0.186), ('elimination', 0.155), ('px', 0.129), ('utility', 0.128), ('cr', 0.122), ('qx', 0.115), ('pa', 0.112), ('scope', 0.11), ('treewidth', 0.107), ('meu', 0.107), ('uv', 0.107), ('ux', 0.104), ('qy', 0.098), ('pd', 0.097), ('decision', 0.094), ('vx', 0.092), ('uy', 0.091), ('chance', 0.09), ('robot', 0.088), ('ps', 0.084), ('py', 0.075), ('st', 0.072), ('ordering', 0.069), ('policy', 0.063), ('variable', 0.063), ('vy', 0.063), ('uence', 0.061), ('parents', 0.06), ('policies', 0.06), ('vi', 0.059), ('equals', 0.057), ('strategy', 0.056), ('axioms', 0.054), ('pad', 0.054), ('vn', 0.052), ('sb', 0.05), ('domain', 0.049), ('propagation', 0.049), ('strategies', 0.048), ('empty', 0.046), ('ot', 0.046), ('decisions', 0.045), ('sa', 0.044), ('arc', 0.044), ('arcs', 0.044), ('position', 0.043), ('nodes', 0.042), ('marginalization', 0.041), ('diagram', 0.041), ('bi', 0.04), ('dt', 0.04), ('campos', 0.04), ('pav', 0.04), ('bt', 0.04), ('variables', 0.038), ('maximal', 0.036), ('lemma', 0.035), ('pv', 0.035), ('pac', 0.034), ('xn', 0.034), ('forgetting', 0.033), ('utilities', 0.033), ('width', 0.033), ('maker', 0.031), ('discarding', 0.031), ('qu', 0.031), ('combination', 0.031), ('initialization', 0.031), ('algebras', 0.03), ('commutativity', 0.03), ('distributivity', 0.03), ('idsia', 0.03), ('manno', 0.03), ('maximality', 0.03), ('pax', 0.03), ('ordered', 0.03), ('cardinality', 0.03), ('es', 0.03), ('pq', 0.029), ('associated', 0.029), ('proposition', 0.028), ('node', 0.027), ('cassio', 0.027), ('scopes', 0.027), ('operations', 0.026), ('specifying', 0.025), ('xi', 0.025), ('solved', 0.025), ('nothing', 0.025), ('regularity', 0.025), ('operation', 0.023), ('bounded', 0.023), ('denis', 0.023), ('lauritzen', 0.023)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.99999976 256 nips-2011-Solving Decision Problems with Limited Information

Author: Denis D. Maua, Cassio Campos

Abstract: We present a new algorithm for exactly solving decision-making problems represented as an influence diagram. We do not require the usual assumptions of no forgetting and regularity, which allows us to solve problems with limited information. The algorithm, which implements a sophisticated variable elimination procedure, is empirically shown to outperform a state-of-the-art algorithm in randomly generated problems of up to 150 variables and 1064 strategies. 1

2 0.10684506 106 nips-2011-Generalizing from Several Related Classification Tasks to a New Unlabeled Sample

Author: Gilles Blanchard, Gyemin Lee, Clayton Scott

Abstract: We consider the problem of assigning class labels to an unlabeled test data set, given several labeled training data sets drawn from similar distributions. This problem arises in several applications where data distributions fluctuate because of biological, technical, or other sources of variation. We develop a distributionfree, kernel-based approach to the problem. This approach involves identifying an appropriate reproducing kernel Hilbert space and optimizing a regularized empirical risk over the space. We present generalization error analysis, describe universal kernels, and establish universal consistency of the proposed methodology. Experimental results on flow cytometry data are presented. 1

3 0.086558446 162 nips-2011-Lower Bounds for Passive and Active Learning

Author: Maxim Raginsky, Alexander Rakhlin

Abstract: We develop unified information-theoretic machinery for deriving lower bounds for passive and active learning schemes. Our bounds involve the so-called Alexander’s capacity function. The supremum of this function has been recently rediscovered by Hanneke in the context of active learning under the name of “disagreement coefficient.” For passive learning, our lower bounds match the upper bounds of Gin´ and Koltchinskii up to constants and generalize analogous results of Mase sart and N´ d´ lec. For active learning, we provide first known lower bounds based e e on the capacity function rather than the disagreement coefficient. 1

4 0.083720684 90 nips-2011-Evaluating the inverse decision-making approach to preference learning

Author: Alan Jern, Christopher G. Lucas, Charles Kemp

Abstract: Psychologists have recently begun to develop computational accounts of how people infer others’ preferences from their behavior. The inverse decision-making approach proposes that people infer preferences by inverting a generative model of decision-making. Existing data sets, however, do not provide sufficient resolution to thoroughly evaluate this approach. We introduce a new preference learning task that provides a benchmark for evaluating computational accounts and use it to compare the inverse decision-making approach to a feature-based approach, which relies on a discriminative combination of decision features. Our data support the inverse decision-making approach to preference learning. A basic principle of decision-making is that knowing people’s preferences allows us to predict how they will behave: if you know your friend likes comedies and hates horror films, you can probably guess which of these options she will choose when she goes to the theater. Often, however, we do not know what other people like and we can only infer their preferences from their behavior. If you know that a different friend saw a comedy today, does that mean that he likes comedies in general? The conclusion you draw will likely depend on what else was playing and what movie choices he has made in the past. A goal for social cognition research is to develop a computational account of people’s ability to infer others’ preferences. One computational approach is based on inverse decision-making. This approach begins with a model of how someone’s preferences lead to a decision. Then, this model is inverted to determine the most likely preferences that motivated an observed decision. An alternative approach might simply learn a functional mapping between features of an observed decision and the preferences that motivated it. For instance, in your friend’s decision to see a comedy, perhaps the more movie options he turned down, the more likely it is that he has a true preference for comedies. The difference between the inverse decision-making approach and the feature-based approach maps onto the standard dichotomy between generative and discriminative models. Economists have developed an instance of the inverse decision-making approach known as the multinomial logit model [1] that has been widely used to infer consumer’s preferences from their choices. This model has recently been explored as a psychological model [2, 3, 4], but there are few behavioral data sets for evaluating it as a model of how people learn others’ preferences. Additionally, the data sets that do exist tend to be drawn from the developmental literature, which focuses on simple tasks that collect only one or two judgments from children [5, 6, 7]. The limitations of these data sets make it difficult to evaluate the multinomial logit model with respect to alternative accounts of preference learning like the feature-based approach. In this paper, we use data from a new experimental task that elicits a detailed set of preference judgments from a single participant in order to evaluate the predictions of several preference learning models from both the inverse decision-making and feature-based classes. Our task requires each participant to sort a large number of observed decisions on the basis of how strongly they indicate 1 (a) (b) (c) d c c (d) b b a a x d x d c b a x 1. Number of chosen effects (−/+) 2. Number of forgone effects (+/+) 3. Number of forgone options (+/+) 4. Number of forgone options containing x (−/−) 5. Max/min number of effects in a forgone option (+/−) 6. Is x in every option? (−/−) 7. Chose only option with x? (+/+) 8. Is x the only difference between options? (+/+) 9. Do all options have same number of effects? (+/+) 10. Chose option with max/min number of effects? (−/−) Figure 1: (a)–(c) Examples of the decisions used in the experiments. Each column represents one option and the boxes represent different effects. The chosen option is indicated by the black rectangle. (d) Features used by the weighted feature and ranked feature models. Features 5 and 10 involved maxima in Experiment 1, which focused on all positive effects, and minima in Experiment 2, which focused on all negative effects. The signs in parentheses indicate the direction of the feature that suggests a stronger preference in Experiment 1 / Experiment 2. a preference for a chosen item. Because the number of decisions is large and these decisions vary on multiple dimensions, predicting how people will order them offers a challenging benchmark on which to compare computational models of preference learning. Data sets from these sorts of detailed tasks have proved fruitful in other domains. For example, data reported by Shepard, Hovland, and Jenkins [8]; Osherson, Smith, Wilkie, L´ pez, and Shafir [9]; and Wasserman, Elek, Chatlosh, o and Baker [10] have motivated much subsequent research on category learning, inductive reasoning, and causal reasoning, respectively. We first describe our preference learning task in detail. We then present several inverse decisionmaking and feature-based models of preference learning and compare these models’ predictions to people’s judgments in two experiments. The data are well predicted by models that follow the inverse decision-making approach, suggesting that this computational approach may help explain how people learn others’ preferences. 1 Multi-attribute decisions and revealed preferences We designed a task that can be used to elicit a large number of preference judgments from a single participant. The task involves a set of observed multi-attribute decisions, some examples of which are represented visually in Figure 1. Each decision is among a set of options and each option produces a set of effects. Figure 1 shows several decisions involving a total of five effects distributed among up to five options. The differently colored boxes represent different effects and the chosen option is marked by a black rectangle. For example, 1a shows a choice between an option with four effects and an option with a single effect; here, the decision maker chose the second option. In our task, people are asked to rank a large number of these decisions by how strongly they suggest that the decision maker had a preference for a particular effect (e.g., effect x in Figure 1). By imposing some minimal constraints, the space of unique multi-attribute decisions is finite and we can obtain rankings for every decision in the space. For example, Figure 2c shows a complete list of 47 unique decisions involving up to five effects, subject to several constraints described later. Three of these decisions are shown in Figure 1. If all the effects are positive—pieces of candy, for example—the first decision (1a) suggests a strong preference for candy x, because the decision maker turned down four pieces in favor of one. The second decision (1b), however, offers much weaker evidence because nearly everyone would choose four pieces of candy over one, even without a specific preference for x. The third decision (1c) provides evidence that is strong but perhaps not quite as strong as the first decision. When all effects are negative—like electric shocks at different body locations—decision makers may still find some effects more tolerable than others, but different inferences are sometimes supported. For example, for negative effects, 1a provides weak evidence that x is relatively tolerable because nearly everyone would choose one shock over four. 2 A computational account of preference learning We now describe a simple computational model for learning a person’s preferences after observing that person make a decision like the ones in Figure 1. We assume that there are n available options 2 {o1 , . . . , on }, each of which produces one or more effects from the set {f1 , f2 , ..., fm }. For simplicity, we assume that effects are binary. Let ui denote the utility the decision maker assigns to effect fi . We begin by specifying a model of decision-making that makes the standard assumptions that decision makers tend to choose things with greater utility and that utilities are additive. That is, if fj is a binary vector indicating the effects produced by option oj and u is a vector of utilities assigned to each of the m effects, then the total utility associated with option oj can be expressed as Uj = fj T u. We complete the specification of the model by applying the Luce choice rule [11], a common psychological model of choice behavior, as the function that chooses among the options: p(c = oj |u, f ) = exp(Uj ) = exp(Uk ) n k=1 exp(fj T u) n T k=1 exp(fk u) (1) where c denotes the choice made. This model can predict the choice someone will make among a specified set of options, given the utilities that person assigns to the effects in each option. To obtain estimates of someone’s utilities, we invert this model by applying Bayes’ rule: p(u|c, F) = p(c|u, F)p(u) p(c|F) (2) where F = {f1 , . . . , fn } specifies the available options and their corresponding effects. This is the multinomial logit model [1], a standard econometric model. In order to apply Equation 2 we must specify a prior p(u) on the utilities. We adopt a standard approach that places independent Gaussian priors on the utilities: ui ∼ N (µ, σ 2 ). For decisions where effects are positive—like candies—we set µ = 2σ, which corresponds to a prior distribution that places approximately 2% of the probability mass below zero. Similarly, for negative effects—like electric shocks—we set µ = −2σ. 2.1 Ordering a set of observed decisions Equation 2 specifies a posterior probability distribution over utilities for a single observed decision but does not provide a way to compare the inferences drawn from multiple decisions for the purposes of ordering them. Suppose we are interested in a decision maker’s preference for effect x and we wish to order a set of decisions by how strongly they support this preference. Two criteria for ordering the decisions are as follows: Absolute utility Relative utility p(c|ux , F)p(ux ) p(c|F) p(c|∀j ux ≥ uj , F)p(∀j ux ≥ uj ) p(∀j ux ≥ uj |c, F) = p(c|F) E(ux |c, F) = Eux The absolute utility model orders decisions by the mean posterior utility for effect x. This criterion is perhaps the most natural way to assess how much a decision indicates a preference for x, but it requires an inference about the utility of x in isolation, and research suggests that people often think about the utility of an effect only in relation to other salient possibilities [12]. The relative utility model applies this idea to preference learning by ordering decisions based on how strongly they suggest that x has a greater utility than all other effects. The decisions in Figures 1b and 1c are cases where the two models lead to different predictions. If the effects are all negative (e.g., electric shocks), the absolute utility model predicts that 1b provides stronger evidence for a tolerance for x because the decision maker chose to receive four shocks instead of just one. The relative utility model predicts that 1c provides stronger evidence because 1b offers no way to determine the relative tolerance of the four chosen effects with respect to one another. Like all generative models, the absolute and relative models incorporate three qualitatively different components: the likelihood term p(c|u, F), the prior p(u), and the reciprocal of the marginal likelihood 1/p(c|F). We assume that the total number of effects is fixed in advance and, as a result, the prior term will be the same for all decisions that we consider. The two other components, however, will vary across decisions. The inverse decision-making approach predicts that both components should influence preference judgments, and we will test this prediction by comparing our 3 two inverse decision-making models to two alternatives that rely only one of these components as an ordering criterion: p(c|∀j ux ≥ uj , F) 1/p(c|F) Representativeness Surprise The representativeness model captures how likely the observed decision would be if the utility for x were high, and previous research has shown that people sometimes rely on a representativeness computation of this kind [13]. The surprise model captures how unexpected the observed decision is overall; surprising decisions may be best explained in terms of a strong preference for x, but unsurprising decisions provide little information about x in particular. 2.2 Feature-based models We also consider a class of feature-based models that use surface features to order decisions. The ten features that we consider are shown in Figure 1d, where x is the effect of interest. As an example, the first feature specifies the number of effects chosen; because x is always among the chosen effects, decisions where few or no other effects belong to the chosen option suggest the strongest preference for x (when all effects are positive). This and the second feature were previously identified by Newtson [14]; we included the eight additional features shown in Figure 1d in an attempt to include all possible features that seemed both simple and relevant. We consider two methods for combining this set of features to order a set of decisions by how strongly they suggest a preference for x. The first model is a standard linear regression model, which we refer to as the weighted feature model. The model learns a weight for each feature, and the rank of a given decision is determined by a weighted sum of its features. The second model is a ranked feature model that sorts the observed decisions with respect to a strict ranking of the features. The top-ranked feature corresponds to the primary sort key, the second-ranked feature to the secondary sort key, and so on. For example, suppose that the top-ranked feature is the number of chosen effects and the second-ranked feature is the number of forgone options. Sorting the three decisions in Figure 1 according to this criterion produces the following ordering: 1a,1c,1b. This notion of sorting items on the basis of ranked features has been applied before to decision-making [15, 16] and other domains of psychology [17], but we are not aware of any previous applications to preference learning. Although our inverse decision-making and feature-based models represent two very different approaches, both may turn out to be valuable. An inverse decision-making approach may be the appropriate account of preference learning at Marr’s [18] computational level, and a feature-based approach may capture the psychological processes by which the computational-level account is implemented. Our goal, therefore, is not necessarily to accept one of these approaches and dismiss the other. Instead, we entertain three distinct possibilities. First, both approaches may account well for the data, which would support the idea that they are valid accounts operating at different levels of analysis. Second, the inverse decision-making approach may offer a better account, suggesting that process-level accounts other than the feature-based approach should be explored. Finally, the feature-based approach may offer a better account, suggesting that inverse decision-making does not constitute an appropriate computational-level account of preference learning. 3 Experiment 1: Positive effects Our first experiment focuses on decisions involving only positive effects. The full set of 47 decisions we used is shown in Figure 2c. This set includes every possible unique decision with up to five different effects, subject to the following constraints: (1) one of the effects (effect x) must always appear in the chosen option, (2) there are no repeated options, (3) each effect may appear in an option at most once, (4) only effects in the chosen option may be repeated in other options, and (5) when effects appear in multiple options, the number of effects is held constant across options. The first constraint is necessary for the sorting task, the second two constraints create a finite space of decisions, and the final two constraints limit attention to what we deemed the most interesting cases. Method 43 Carnegie Mellon undergraduates participated for course credit. Each participant was given a set of cards, with one decision printed on each card. The decisions were represented visually 4 (a) (c) Decisions 42 40 45 Mean human rankings 38 30 23 20 22 17 13 12 11 10 9 8 7 6 19 18 31 34 28 21 26 36 35 33 37 27 29 32 25 24 16 15 14 5 4 3 2 1 Absolute utility model rankings (b) Mean human rankings (Experiment 1) 47 43 44 46 45 38 37 36 34 35 30 32 33 31 29 28 24 26 27 25 21 19 22 20 18 16 17 12 13 7 6 11 5 9 4 10 8 1 2 3 42 40 41 39 47 46 44 41 43 39 23 15 14 Mean human rankings (Experiment 2) 1. dcbax 2. cbax 3. bax 4. ax 5. x 6. dcax | bcax 7. dx | cx | bx | ax 8. cax | bax 9. bdx | bcx | bax 10. dcx | bax 11. bx | ax 12. bdx | cax | bax 13. cx | bx | ax 14. d | cbax 15. c | bax 16. b | ax 17. d | c | bax 18. dc | bax 19. c | b | ax 20. dc | bx | ax 21. bdc | bax 22. ad | cx | bx | ax 23. d | c | b | ax 24. bad | bcx | bax 25. ac | bx | ax 26. cb | ax 27. cbad | cbax 28. dc | b | ax 29. ad | ac | bx | ax 30. ab | ax 31. bad | bax 32. dc | ab | ax 33. dcb | ax 34. a | x 35. bad | bac | bax 36. ac | ab | ax 37. ad | ac | ab | ax 38. b | a | x 39. ba | x 40. c | b | a | x 41. cb | a | x 42. d | c | b | a | x 43. cba | x 44. dc | ba | x 45. dc | b | a | x 46. dcb | a | x 47. dcba | x Figure 2: (a) Comparison between the absolute utility model rankings and the mean human rankings for Experiment 1. Each point represents one decision, numbered with respect to the list in panel c. (b) Comparison between the mean human rankings in Experiments 1 and 2. In both scatter plots, the solid diagonal lines indicate a perfect correspondence between the two sets of rankings. (c) The complete set of decisions, ordered by the mean human rankings from Experiment 1. Options are separated by vertical bars and the chosen option is always at the far right. Participants were always asked about a preference for effect x. as in Figure 1 but without the letter labels. Participants were told that the effects were different types of candy and each option was a bag containing one or more pieces of candy. They were asked to sort the cards by how strongly each decision suggested that the decision maker liked a particular target candy, labeled x in Figure 2c. They sorted the cards freely on a table but reported their final rankings by writing them on a sheet of paper, from weakest to strongest evidence. They were instructed to order the cards as completely as possible, but were told that they could assign the same ranking to a set of cards if they believed those cards provided equal evidence. 3.1 Results Two participants were excluded as outliers based on the criterion that their rankings for at least five decisions were at least three standard deviations from the mean rankings. We performed a hierarchical clustering analysis of the remaining 41 participants’ rankings using rank correlation as a similarity metric. Participants’ rankings were highly correlated: cutting the resulting dendrogram at 0.2 resulted in one cluster that included 33 participants and the second largest cluster included 5 Surprise MAE = 17.8 MAE = 7.0 MAE = 4.3 MAE = 17.3 MAE = 9.5 Human rankings Experiment 2 Negative effects Representativeness MAE = 2.3 MAE = 6.7 Experiment 1 Positive effects Relative utility MAE = 2.3 Human rankings Absolute utility Model rankings Model rankings Model rankings Model rankings Figure 3: Comparison between human rankings in both experiments and predicted rankings from four models. The solid diagonal lines indicate a perfect correspondence between human and model rankings. only 3 participants. Thus, we grouped all participants together and analyzed their mean rankings. The 0.2 threshold was chosen because it produced the most informative clustering in Experiment 2. Inverse decision-making models We implemented the inverse decision-making models using importance sampling with 5 million samples drawn from the prior distribution p(u). Because all the effects were positive, we used a prior on utilities that placed nearly all probability mass above zero (µ = 4, σ = 2). The mean human rankings are compared with the absolute utility model rankings in Figure 2a, and the mean human rankings are listed in order in 2c. Fractional rankings were used for both the human data and the model predictions. The human rankings in the figure are the means of participants’ fractional rankings. The first row of Figure 3 contains similar plots that allow comparison of the four models we considered. In these plots, the solid diagonal lines indicate a perfect correspondence between model and human rankings. Thus, the largest deviations from this line represent the largest deviations in the data from the model’s predictions. Figure 3 shows that the absolute and relative utility models make virtually identical predictions and both models provide a strong account of the human rankings as measured by mean absolute error (MAE = 2.3 in both cases). Moreover, both models correctly predict the highest ranked decision and the set of lowest ranked decisions. The only clear discrepancy between the model predictions and the data is the cluster of points at the lower left, labeled as Decisions 6–13 in Figure 2a. These are all cases in which effect x appears in all options and therefore these decisions provide no information about a decision maker’s preference for x. Consequently, the models assign the same ranking to this group as to the group of decisions in which there is only a single option (Decisions 1–5). Although people appeared to treat these groups somewhat differently, the models still correctly predict that the entire group of decisions 1–13 is ranked lower than all other decisions. The surprise and representativeness models do not perform nearly as well (MAE = 7.0 and 17.8, respectively). Although the surprise model captures some of the general trends in the human rankings, it makes several major errors. For example, consider Decision 7: dx|cx|bx|ax. This decision provides no information about a preference for x because it appears in every option. The decision is surprising, however, because a decision maker choosing at random from these options would make the observed choice only 1/4 of the time. The representativeness model performs even worse, primarily because it does not take into account alternative explanations for why an option was chosen, such as the fact that no other options were available (e.g., Decision 1 in Figure 2c). The failure of these models to adequately account for the data suggests that both the likelihood p(c|u, F) and marginal likelihood p(c|F) are important components of the absolute and relative utility models. Feature-based models We compared the performance of the absolute and relative utility models to our two feature-based models: the weighted feature and ranked feature models. For each participant, 6 (b) Ranked feature 10 10 5 Figure 4: Results of the feature-based model analysis from Experiment 1 for (a) the weighted feature models and (b) the ranked feature models. The histograms show the minimum number of features needed to match the accuracy (measured by MAE) of the absolute utility model for each participant. 15 5 1 2 3 4 5 6 >6 15 1 2 3 4 5 6 7 8 9 10 >10 Number of participants (a) Weighted feature Number of features needed we considered every subset of features1 in Figure 1d in order to determine the minimum number of features needed by the two models to achieve the same level of accuracy as the absolute utility model, as measured by mean absolute error. The results of these analyses are shown in Figure 4. For the majority of participants, at least four features were needed by both models to match the accuracy of the absolute utility model. For the weighted feature model, 14 participants could not be fit as well as the absolute utility model even when all ten features were considered. These results indicate that a feature-based account of people’s inferences in our task must be supplied with a relatively large number of features. By contrast, the inverse decision-making approach provides a relatively parsimonious account of the data. 4 Experiment 2: Negative effects Experiment 2 focused on a setting in which all effects are negative, motivated by the fact that the inverse decision-making models predict several major differences in orderings when effects are negative rather than positive. For instance, the absolute utility model’s relative rankings of the decisions in Figures 1a and 1b are reversed when all effects are negative rather than positive. Method 42 Carnegie Mellon undergraduates participated for course credit. The experimental design was identical to Experiment 1 except that participants were told that the effects were electric shocks at different body locations. They were asked to sort the cards on the basis of how strongly each decision suggested that the decision maker finds shocks at the target location relatively tolerable. The model predictions were derived in the same way as for Experiment 1, but with a prior distribution on utilities that placed nearly all probability mass below zero (µ = −4, σ = 2) to reflect the fact that effects were all negative. 4.1 Results Three participants were excluded as outliers by the same criterion applied in Experiment 1. The resulting mean rankings are compared with the corresponding rankings from Experiment 1 in Figure 2b. The figure shows that responses based on positive and negative effects were substantially different in a number of cases. Figure 3 shows how the mean rankings compare to the predictions of the four models we considered. Although the relative utility model is fairly accurate, no model achieves the same level of accuracy as the absolute and relative utility models in Experiment 1. In addition, the relative utility model provides a poor account of the responses of many individual participants. To better understand responses at the individual level, we repeated the hierarchical clustering analysis described in Experiment 1, which revealed that 29 participants could be grouped into one of four clusters, with the remaining participants each in their own clusters. We analyzed these four clusters independently, excluding the 10 participants that could not be naturally grouped. We compared the mean rankings of each cluster to the absolute and relative utility models, as well as all one- and two-feature weighted feature and ranked feature models. Figure 5 shows that the mean rankings of participants in Cluster 1 (N = 8) were best fit by the absolute utility model, the mean rankings of participants in Cluster 2 (N = 12) were best fit by the relative utility model, and the mean rankings of participants in Clusters 3 (N = 3) and 4 (N = 6) were better fit by feature-based models than by either the absolute or relative utility models. 1 A maximum of six features was considered for the ranked feature model because considering more features was computationally intractable. 7 Cluster 4 N =6 MAE = 4.9 MAE = 14.0 MAE = 7.9 MAE = 5.3 MAE = 2.6 MAE = 13.0 MAE = 6.2 Human rankings Relative utility Cluster 3 N =3 MAE = 2.6 Absolute utility Cluster 2 N = 12 Human rankings Cluster 1 N =8 Factors: 1,3 Factors: 1,8 MAE = 2.3 MAE = 5.2 Model rankings Best−fitting weighted feature Factors: 6,7 MAE = 4.0 Model rankings Model rankings Model rankings Human rankings Factors: 3,8 MAE = 4.8 Figure 5: Comparison between human rankings for four clusters of participants identified in Experiment 2 and predicted rankings from three models. Each point in the plots corresponds to one decision and the solid diagonal lines indicate a perfect correspondence between human and model rankings. The third row shows the predictions of the best-fitting two-factor weighted feature model for each cluster. The two factors listed refer to Figure 1d. To examine how well the models accounted for individuals’ rankings within each cluster, we compared the predictions of the inverse decision-making models to the best-fitting two-factor featurebased model for each participant. In Cluster 1, 7 out of 8 participants were best fit by the absolute utility model; in Cluster 2, 8 out of 12 participants were best fit by the relative utility model; in Clusters 3 and 4, all participants were better fit by feature-based models. No single feature-based model provided the best fit for more than two participants, suggesting that participants not fit well by the inverse decision-making models were not using a single alternative strategy. Applying the feature-based model analysis from Experiment 1 to the current results revealed that the weighted feature model required an average of 6.0 features to match the performance of the absolute utility model for participants in Cluster 1, and an average of 3.9 features to match the performance of the relative utility model for participants in Cluster 2. Thus, although a single model did not fit all participants well in the current experiment, many participants were fit well by one of the two inverse decision-making models, suggesting that this general approach is useful for explaining how people reason about negative effects as well as positive effects. 5 Conclusion In two experiments, we found that an inverse decision-making approach offered a good computational account of how people make judgments about others’ preferences. Although this approach is conceptually simple, our analyses indicated that it captures the influence of a fairly large number of relevant decision features. Indeed, the feature-based models that we considered as potential process models of preference learning could only match the performance of the inverse decision-making approach when supplied with a relatively large number of features. We feel that this result rules out the feature-based approach as psychologically implausible, meaning that alternative process-level accounts will need to be explored. One possibility is sampling, which has been proposed as a psychological mechanism for approximating probabilistic inferences [19, 20]. However, even if process models that use large numbers of features are considered plausible, the inverse decision-making approach provides a valuable computational-level account that helps to explain which decision features are informative. Acknowledgments This work was supported in part by the Pittsburgh Life Sciences Greenhouse Opportunity Fund and by NSF grant CDI-0835797. 8 References [1] D. McFadden. Conditional logit analysis of qualitative choice behavior. In P. Zarembka, editor, Frontiers in Econometrics. Amademic Press, New York, 1973. [2] C. G. Lucas, T. L. Griffiths, F. Xu, and C. Fawcett. A rational model of preference learning and choice prediction by children. In Proceedings of Neural Information Processing Systems 21, 2009. [3] L. Bergen, O. R. Evans, and J. B. Tenenbaum. Learning structured preferences. In Proceedings of the 32nd Annual Conference of the Cognitive Science Society, 2010. [4] A. Jern and C. Kemp. Decision factors that support preference learning. In Proceedings of the 33rd Annual Conference of the Cognitive Science Society, 2011. [5] T. Kushnir, F. Xu, and H. M. Wellman. Young children use statistical sampling to infer the preferences of other people. Psychological Science, 21(8):1134–1140, 2010. [6] L. Ma and F. Xu. Young children’s use of statistical sampling evidence to infer the subjectivity of preferences. Cognition, in press. [7] M. J. Doherty. Theory of Mind: How Children Understand Others’ Thoughts and Feelings. Psychology Press, New York, 2009. [8] R. N. Shepard, C. I. Hovland, and H. M. Jenkins. Learning and memorization of classifications. Psychological Monographs, 75, Whole No. 517, 1961. [9] D. N. Osherson, E. E. Smith, O. Wilkie, A. L´ pez, and E. Shafir. Category-based induction. Psychological o Review, 97(2):185–200, 1990. [10] E. A. Wasserman, S. M. Elek, D. L. Chatlosh, and A. G. Baker. Rating causal relations: Role of probability in judgments of response-outcome contingency. Journal of Experimental Psychology: Learning, Memory, and Cognition, 19(1):174–188, 1993. [11] R. D. Luce. Individual choice behavior. John Wiley, 1959. [12] D. Ariely, G. Loewenstein, and D. Prelec. Tom Sawyer and the construction of value. Journal of Economic Behavior & Organization, 60:1–10, 2006. [13] D. Kahneman and A. Tversky. Subjective probability: A judgment of representativeness. Cognitive Psychology, 3(3):430–454, 1972. [14] D. Newtson. Dispositional inference from effects of actions: Effects chosen and effects forgone. Journal of Experimental Social Psychology, 10:489–496, 1974. [15] P. C. Fishburn. Lexicographic orders, utilities and decision rules: A survey. Management Science, 20(11):1442–1471, 1974. [16] G. Gigerenzer and P. M. Todd. Fast and frugal heuristics: The adaptive toolbox. Oxford University Press, New York, 1999. [17] A. Prince and P. Smolensky. Optimality Theory: Constraint Interaction in Generative Grammar. WileyBlackwell, 2004. [18] D. Marr. Vision. W. H. Freeman, San Francisco, 1982. [19] A. N. Sanborn, T. L. Griffiths, and D. J. Navarro. Rational approximations to rational models: Alternative algorithms for category learning. Psychological Review, 117:1144–1167, 2010. [20] L. Shi and T. L. Griffiths. Neural implementation of Bayesian inference by importance sampling. In Proceedings of Neural Information Processing Systems 22, 2009. 9

5 0.079153448 20 nips-2011-Active Learning Ranking from Pairwise Preferences with Almost Optimal Query Complexity

Author: Nir Ailon

Abstract: Given a set V of n elements we wish to linearly order them using pairwise preference labels which may be non-transitive (due to irrationality or arbitrary noise). The goal is to linearly order the elements while disagreeing with as few pairwise preference labels as possible. Our performance is measured by two parameters: The number of disagreements (loss) and the query complexity (number of pairwise preference labels). Our algorithm adaptively queries at most O(n poly(log n, ε−1 )) preference labels for a regret of ε times the optimal loss. This is strictly better, and often significantly better than what non-adaptive sampling could achieve. Our main result helps settle an open problem posed by learning-to-rank (from pairwise information) theoreticians and practitioners: What is a provably correct way to sample preference labels? 1

6 0.078528523 207 nips-2011-Optimal learning rates for least squares SVMs using Gaussian kernels

7 0.066005044 16 nips-2011-A reinterpretation of the policy oscillation phenomenon in approximate policy iteration

8 0.062799394 211 nips-2011-Penalty Decomposition Methods for Rank Minimization

9 0.062689364 142 nips-2011-Large-Scale Sparse Principal Component Analysis with Application to Text Data

10 0.061573513 174 nips-2011-Monte Carlo Value Iteration with Macro-Actions

11 0.057042744 215 nips-2011-Policy Gradient Coagent Networks

12 0.056608856 88 nips-2011-Environmental statistics and the trade-off between model-based and TD learning in humans

13 0.05606943 205 nips-2011-Online Submodular Set Cover, Ranking, and Repeated Active Learning

14 0.054343149 212 nips-2011-Periodic Finite State Controllers for Efficient POMDP and DEC-POMDP Planning

15 0.05397116 226 nips-2011-Projection onto A Nonnegative Max-Heap

16 0.053388529 233 nips-2011-Rapid Deformable Object Detection using Dual-Tree Branch-and-Bound

17 0.052075647 300 nips-2011-Variance Reduction in Monte-Carlo Tree Search

18 0.051048405 79 nips-2011-Efficient Offline Communication Policies for Factored Multiagent POMDPs

19 0.05096304 237 nips-2011-Reinforcement Learning using Kernel-Based Stochastic Factorization

20 0.050562706 18 nips-2011-Action-Gap Phenomenon in Reinforcement Learning


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, 0.127), (1, -0.07), (2, 0.0), (3, 0.033), (4, -0.063), (5, 0.001), (6, -0.05), (7, -0.054), (8, -0.028), (9, -0.044), (10, -0.027), (11, -0.007), (12, 0.013), (13, -0.01), (14, 0.032), (15, -0.053), (16, 0.086), (17, 0.044), (18, 0.02), (19, 0.03), (20, 0.063), (21, -0.088), (22, 0.067), (23, -0.05), (24, -0.057), (25, 0.011), (26, -0.05), (27, -0.008), (28, -0.158), (29, 0.008), (30, 0.113), (31, -0.013), (32, -0.021), (33, -0.028), (34, 0.015), (35, 0.111), (36, 0.071), (37, -0.051), (38, -0.17), (39, -0.017), (40, -0.031), (41, -0.051), (42, -0.05), (43, -0.003), (44, -0.028), (45, -0.066), (46, -0.047), (47, -0.014), (48, -0.045), (49, 0.109)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.92159492 256 nips-2011-Solving Decision Problems with Limited Information

Author: Denis D. Maua, Cassio Campos

Abstract: We present a new algorithm for exactly solving decision-making problems represented as an influence diagram. We do not require the usual assumptions of no forgetting and regularity, which allows us to solve problems with limited information. The algorithm, which implements a sophisticated variable elimination procedure, is empirically shown to outperform a state-of-the-art algorithm in randomly generated problems of up to 150 variables and 1064 strategies. 1

2 0.65404093 106 nips-2011-Generalizing from Several Related Classification Tasks to a New Unlabeled Sample

Author: Gilles Blanchard, Gyemin Lee, Clayton Scott

Abstract: We consider the problem of assigning class labels to an unlabeled test data set, given several labeled training data sets drawn from similar distributions. This problem arises in several applications where data distributions fluctuate because of biological, technical, or other sources of variation. We develop a distributionfree, kernel-based approach to the problem. This approach involves identifying an appropriate reproducing kernel Hilbert space and optimizing a regularized empirical risk over the space. We present generalization error analysis, describe universal kernels, and establish universal consistency of the proposed methodology. Experimental results on flow cytometry data are presented. 1

3 0.53562307 207 nips-2011-Optimal learning rates for least squares SVMs using Gaussian kernels

Author: Mona Eberts, Ingo Steinwart

Abstract: We prove a new oracle inequality for support vector machines with Gaussian RBF kernels solving the regularized least squares regression problem. To this end, we apply the modulus of smoothness. With the help of the new oracle inequality we then derive learning rates that can also be achieved by a simple data-dependent parameter selection method. Finally, it turns out that our learning rates are asymptotically optimal for regression functions satisfying certain standard smoothness conditions. 1

4 0.52762771 201 nips-2011-On the Completeness of First-Order Knowledge Compilation for Lifted Probabilistic Inference

Author: Guy Broeck

Abstract: Probabilistic logics are receiving a lot of attention today because of their expressive power for knowledge representation and learning. However, this expressivity is detrimental to the tractability of inference, when done at the propositional level. To solve this problem, various lifted inference algorithms have been proposed that reason at the first-order level, about groups of objects as a whole. Despite the existence of various lifted inference approaches, there are currently no completeness results about these algorithms. The key contribution of this paper is that we introduce a formal definition of lifted inference that allows us to reason about the completeness of lifted inference algorithms relative to a particular class of probabilistic models. We then show how to obtain a completeness result using a first-order knowledge compilation approach for theories of formulae containing up to two logical variables. 1 Introduction and related work Probabilistic logic models build on first-order logic to capture relational structure and on graphical models to represent and reason about uncertainty [1, 2]. Due to their expressivity, these models can concisely represent large problems with many interacting random variables. While the semantics of these logics is often defined through grounding the models [3], performing inference at the propositional level is – as for first-order logic – inefficient. This has motivated the quest for lifted inference methods that exploit the structure of probabilistic logic models for efficient inference, by reasoning about groups of objects as a whole and avoiding repeated computations. The first approaches to exact lifted inference have upgraded the variable elimination algorithm to the first-order level [4, 5, 6]. More recent work is based on methods from logical inference [7, 8, 9, 10], such as knowledge compilation. While these approaches often yield dramatic improvements in runtime over propositional inference methods on specific problems, it is still largely unclear for which classes of models these lifted inference operators will be useful and for which ones they will eventually have to resort to propositional inference. One notable exception in this regard is lifted belief propagation [11], which performs exact lifted inference on any model whose factor graph representation is a tree. A first contribution of this paper is that we introduce a notion of domain lifted inference, which formally defines what lifting means, and which can be used to characterize the classes of probabilistic models to which lifted inference applies. Domain lifted inference essentially requires that probabilistic inference runs in polynomial time in the domain size of the logical variables appearing in the model. As a second contribution we show that the class of models expressed as 2-WFOMC formulae (weighted first-order model counting with up to 2 logical variables per formula) can be domain lifted using an extended first-order knowledge compilation approach [10]. The resulting approach allows for lifted inference even in the presence of (anti-) symmetric or total relations in a theory. These are extremely common and useful concepts that cannot be lifted by any of the existing first-order knowledge compilation inference rules. 1 2 Background We will use standard concepts of function-free first-order logic (FOL). An atom p(t1 , . . . , tn ) consists of a predicate p/n of arity n followed by n arguments, which are either constants or logical variables. An atom is ground if it does not contain any variables. A literal is an atom a or its negation ¬a. A clause is a disjunction l1 ∨ ... ∨ lk of literals. If k = 1, it is a unit clause. An expression is an atom, literal or clause. The pred(a) function maps an atom to its predicate and the vars(e) function maps an expression to its logical variables. A theory in conjunctive normal form (CNF) is a conjunction of clauses. We often represent theories by their set of clauses and clauses by their set of literals. Furthermore, we will assume that all logical variables are universally quantified. In addition, we associate a set of constraints with each clause or atom, either of the form X = t, where X is a logical variable and t is a constant or variable, or of the form X ∈ D, where D is a domain, or the negation of these constraints. These define a finite domain for each logical variable. Abusing notation, we will use constraints of the form X = t to denote a substitution of X by t. The function atom(e) maps an expression e to its atoms, now associating the constraints on e with each atom individually. To add the constraint c to an expression e, we use the notation e ∧ c. Two atoms unify if there is a substitution which makes them identical and if the conjunction of the constraints on both atoms with the substitution is satisfiable. Two expressions e1 and e2 are independent, written e1 ⊥ e2 , if no atom a1 ∈ atom(e1 ) unifies with an atom a2 ∈ atom(e2 ). ⊥ We adopt the Weighted First-Order Model Counting (WFOMC) [10] formalism to represent probabilistic logic models, building on the notion of a Herbrand interpretation. Herbrand interpretations are subsets of the Herbrand base HB (T ), which consists of all ground atoms that can be constructed with the available predicates and constant symbols in T . The atoms in a Herbrand interpretation are assumed to be true. All other atoms in HB (T ) are assumed to be false. An interpretation I satisfies a theory T , written as I |= T , if it satisfies all the clauses c ∈ T . The WFOMC problem is defined on a weighted logic theory T , which is a logic theory augmented with a positive weight function w and a negative weight function w, which assign a weight to each predicate. The WFOMC problem involves computing wmc(T, w, w) = w(pred(a)) I|=T a∈I 3 3.1 w(pred(a)). (1) a∈HB(T )\I First-order knowledge compilation for lifted probabilistic inference Lifted probabilistic inference A first-order probabilistic model defines a probability distribution P over the set of Herbrand interpretations H. Probabilistic inference in these models is concerned with computing the posterior probability P(q|e) of query q given evidence e, where q and e are logical expressions in general: P(q|e) = h∈H,h|=q∧e P(h) h∈H,h|=e P(h) (2) We propose one notion of lifted inference for first-order probabilistic models, defined in terms of the computational complexity of inference w.r.t. the domains of logical variables. It is clear that other notions of lifted inference are conceivable, especially in the case of approximate inference. Definition 1 (Domain Lifted Probabilistic Inference). A probabilistic inference procedure is domain lifted for a model m, query q and evidence e iff the inference procedure runs in polynomial time in |D1 |, . . . , |Dk | with Di the domain of the logical variable vi ∈ vars(m, q, e). Domain lifted inference does not prohibit the algorithm to be exponential in the size of the vocabulary, that is, the number of predicates, arguments and constants, of the probabilistic model, query and evidence. In fact, the definition allows inference to be exponential in the number of constants which occur in arguments of atoms in the theory, query or evidence, as long as it is polynomial in the cardinality of the logical variable domains. This definition of lifted inference stresses the ability to efficiently deal with the domains of the logical variables that arise, regardless of their size, and formalizes what seems to be generally accepted in the lifted inference literature. 2 A class of probabilistic models is a set of probabilistic models expressed in a particular formalism. As examples, consider Markov logic networks (MLN) [12] or parfactors [4], or the weighted FOL theories for WFOMC that we introduced above, when the weights are normalized. Definition 2 (Completeness). Restricting queries to atoms and evidence to a conjunction of literals, a procedure that is domain lifted for all probabilistic models m in a class of models M and for all queries q and evidence e, is called complete for M . 3.2 First-order knowledge compilation First-order knowledge compilation is an approach to lifted probabilistic inference consisting of the following three steps (see Van den Broeck et al. [10] for details): 1. Convert the probabilistic logical model to a weighted CNF. Converting MLNs or parfactors requires adding new atoms to the theory that represent the (truth) value of each factor or formula. set-disjunction 2 friends(X, Y ) ∧ smokes(X) ⇒ smokes(Y ) Smokers ⊆ People decomposable conjunction unit clause leaf (a) MLN Model ∧ smokes(X), X ∈ Smokers smokes(Y ) ∨ ¬ smokes(X) ∨¬ friends(X, Y ) ∨ ¬ f(X, Y ) friends(X, Y ) ∨ f(X, Y ) smokes(X) ∨ f(X, Y ) ¬ smokes(Y ) ∨ f(X, Y ). ∧ f(X, Y ), Y ∈ Smokers ∧ ¬ smokes(Y ), Y ∈ Smokers / ∧ f(X, Y ), X ∈ Smokers, Y ∈ Smokers / / x ∈ Smokers (b) CNF Theory deterministic disjunction Predicate friends smokes f w 1 1 e2 w 1 1 1 y ∈ Smokers / ∨ set-conjunction ∧ f(x, y) (c) Weight Functions ¬ friends(x, y) ∧ friends(x, y) ¬ f(x, y) (d) First-Order d-DNNF Circuit Figure 1: Friends-smokers example (taken from [10]) Example 1. The MLN in Figure 1a assigns a weight to a formula in FOL. Figure 1b represents the same model as a weighted CNF, introducing a new atom f(X, Y ) to encode the truth value of the MLN formula. The probabilistic information is captured by the weight functions in Figure 1c. 2. Compile the logical theory into a First-Order d-DNNF (FO d-DNNF) circuit. Figure 1d shows an example of such a circuit. Leaves represent unit clauses. Inner nodes represent the disjunction or conjunction of their children l and r, but with the constraint that disjunctions must be deterministic (l ∧ r is unsatisfiable) and conjunctions must be decomposable (l ⊥ r). ⊥ 3. Perform WFOMC inference to compute posterior probabilities. In a FO d-DNNF circuit, WFOMC is polynomial in the size of the circuit and the cardinality of the domains. To compile the CNF theory into a FO d-DNNF circuit, Van den Broeck et al. [10] propose a set of compilation rules, which we will refer to as CR 1 . We will now briefly describe these rules. Unit Propagation introduces a decomposable conjunction when the theory contains a unit clause. Independence creates a decomposable conjunction when the theory contains independent subtheories. Shannon decomposition applies when the theory contains ground atoms and introduces a deterministic disjunction between two modified theories: one where the ground atom is true, and one where it is false. Shattering splits clauses in the theory until all pairs of atoms represent either a disjoint or identical set of ground atoms. Example 2. In Figure 2a, the first two clauses are made independent from the friends(X, X) clause and split off in a decomposable conjunction by unit propagation. The unit clause becomes a leaf of the FO d-DNNF circuit, while the other operand requires further compilation. 3 dislikes(X, Y ) ∨ friends(X, Y ) fun(X) ∨ ¬ friends(X, Y ) friends(X, Y ) ∨ dislikes(X, Y ) ¬ friends(X, Y ) ∨ likes(X, Y ) friends(X, X) fun(X) ∨ ¬ friends(X, Y ) fun(X) ∨ ¬ friends(Y, X) ∧ FunPeople ⊆ People x ∈ People friends(X, X) friends(X, Y ) ∨ dislikes(X, Y ), X = Y ¬ friends(X, Y ) ∨ likes(X, Y ), X = Y likes(X, X) dislikes(x, Y ) ∨ friends(x, Y ) fun(x) ∨ ¬ friends(x, Y ) fun(X), X ∈ FunPeople ¬ fun(X), X ∈ FunPeople / fun(X) ∨ ¬ friends(X, Y ) fun(X) ∨ ¬ friends(Y, X) (a) Unit propagation of friends(X, X) (b) Independent partial grounding (c) Atom counting of fun(X) Figure 2: Examples of compilation rules. Circles are FO d-DNNF inner nodes. White rectangles show theories before and after applying the rule. All variable domains are People. (taken from [10]) Independent Partial Grounding creates a decomposable conjunction over a set of child circuits, which are identical up to the value of a grounding constant. Since they are structurally identical, only one child circuit is actually compiled. Atom Counting applies when the theory contains an atom with a single logical variable X ∈ D. It explicitly represents the domain D ⊆ D of X for which the atom is true. It compiles the theory into a deterministic disjunction between all possible such domains. Again, these child circuits are identical up to the value of D and only one is compiled. Example 3. The theory in Figure 2b is compiled into a decomposable set-conjunction of theories that are independent and identical up to the value of the x constant. The theory in Figure 2c contains an atom with one logical variable: fun(X). Atom counting compiles it into a deterministic setdisjunction over theories that differ in FunPeople, which is the domain of X for which fun(X) is true. Subsequent steps of unit propagation remove the fun(X) atoms from the theory entirely. 3.3 Completeness We will now characterize those theories where the CR 1 compilation rules cannot be used, and where the inference procedure has to resort to grounding out the theory to propositional logic. For these, first-order knowledge compilation using CR 1 is not yet domain lifted. When a logical theory contains symmetric, anti-symmetric or total relations, such as friends(X, Y ) ⇒ friends(Y, X), parent(X, Y ) ⇒ ¬ parent(Y, X), X = Y, ≤ (X, Y) ∨ ≤ (Y, X), or more general formulas, such as enemies(X, Y ) ⇒ ¬ friend(X, Y ) ∧ ¬ friend(Y, X), (3) (4) (5) (6) none of the CR 1 rules apply. Intuitively, the underlying problem is the presence of either: • Two unifying (not independent) atoms in the same clause which contain the same logical variable in different positions of the argument list. Examples include (the CNF of) Formulas 3, 4 and 5, where the X and Y variable are bound by unifying two atoms from the same clause. • Two logical variables that bind when unifying one pair of atoms but appear in different positions of the argument list of two other unifying atoms. Examples include Formula 6, which in CNF is ¬ friend(X, Y ) ∨ ¬ enemies(X, Y ) ¬ friend(Y, X) ∨ ¬ enemies(X, Y ) Here, unifying the enemies(X, Y ) atoms binds the X variables from both clauses, which appear in different positions of the argument lists of the unifying atoms friend(X, Y ) and friend(Y, X). Both of these properties preclude the use of CR 1 rules. Also in the context of other model classes, such as MLNs, probabilistic versions of the above formulas cannot be processed by CR 1 rules. 4 Even though first-order knowledge compilation with CR 1 rules does not have a clear completeness result, we can show some properties of theories to which none of the compilation rules apply. First, we need to distinguish between the arity of an atom and its dimension. A predicate with arity two might have atoms with dimension one, when one of the arguments is ground or both are identical. Definition 3 (Dimension of an Expression). The dimension of an expression e is the number of logical variables it contains: dim(e) = | vars(e)|. Lemma 1 (CR 1 Postconditions). The CR 1 rules remove all atoms from the theory T which have zero or one logical variable arguments, such that afterwards ∀a ∈ atom(T ) : dim(a) > 1. When no CR 1 rule applies, the theory is shattered and contains no independent subtheories. Proof. Ground atoms are removed by the Shannon decomposition operator followed by unit propagation. Atoms with a single logical variable (including unary relations) are removed by the atom counting operator followed by unit propagation. If T contains independent subtheories, the independence operator can be applied. Shattering is always applied when T is not yet shattered. 4 Extending first-order knowledge compilation In this section we introduce a new operator which does apply to the theories from Section 3.3. 4.1 Logical variable properties To formally define the operator we propose, and prove its correctness, we first introduce some mathematical concepts related to the logical variables in a theory (partly after Jha et al. [8]). Definition 4 (Binding Variables). Two logical variables X, Y are directly binding b(X, Y ) if they are bound by unifying a pair of atoms in the theory. The binding relationship b+ (X, Y ) is the transitive closure of the directly binding relation b(X, Y ). Example 4. In the theory ¬ p(W, X) ∨ ¬ q(X) r(Y ) ∨ ¬ q(Y ) ¬ r(Z) ∨ s(Z) the variable pairs (X, Y ) and (Y, Z) are directly binding. The variables X, Y and Z are binding. Variable W does not bind to any other variable. Note that the binding relationship b+ (X, Y ) is an equivalence relation that defines two equivalence classes: {X, Y, Z} and {W }. Lemma 2 (Binding Domains). After shattering, binding logical variables have identical domains. Proof. During shattering (see Section 3.2), when two atoms unify, binding two variables with partially overlapping domains, the atoms’ clauses are split up into clauses where the domain of the variables is identical, and clauses where the domains are disjoint and the atoms no longer unify. Definition 5 (Root Binding Class). A root variable is a variable that appears in all the atoms in its clause. A root binding class is an equivalence class of binding variables where all variables are root. Example 5. In the theory of Example 4, {X, Y, Z} is a root binding class and {W } is not. 4.2 Domain recursion We will now introduce the new domain recursion operator, starting with its preconditions. Definition 6. A theory allows for domain recursion when (i) the theory is shattered, (ii) the theory contains no independent subtheories and (iii) there exists a root binding class. From now on, we will denote with C the set of clauses of the theory at hand and with B a root binding class guaranteed to exist if C allows for domain recursion. Lemma 2 states that all variables in B have identical domains. We will denote the domain of these variables with D. The intuition behind the domain recursion operator is that it modifies D by making one element explicit: D = D ∪ {xD } with xD ∈ D . This explicit domain element is introduced by the S PLIT D / function, which splits clauses w.r.t. the new subdomain D and element xD . 5 Definition 7 (S PLIT D). For a clause c and given set of variables Vc ⊆ vars(c) with domain D, let S PLIT D(c, Vc ) = c, if Vc = ∅ S PLIT D(c1 , Vc \ {V }) ∪ S PLIT D(c2 , Vc \ {V }), if Vc = ∅ (7) where c1 = c ∧ (V = xD ) and c2 = c ∧ (V = xD ) ∧ (V ∈ D ) for some V ∈ Vc . For a set of clauses C and set of variables V with domain D: S PLIT D(C, V) = c∈C S PLIT D(c, V ∩ vars(c)). The domain recursion operator creates three sets of clauses: S PLIT D(C, B) = Cx ∪ Cv ∪ Cr , with Cx = {c ∧ (V = xD )|c ∈ C}, (8) V ∈B∩vars(c) Cv = {c ∧ (V = xD ) ∧ (V ∈ D )|c ∈ C}, (9) V ∈B∩vars(c) Cr = S PLIT D(C, B) \ Cx \ Cv . (10) Proposition 3. The conjunction of the domain recursion sets is equivalent to the original theory: c∈Cr c . c∈Cv c ∧ c∈C c ≡ c∈S PLIT D(C,B) c and therefore c∈C c ≡ c∈Cx c ∧ We will now show that these sets are independent and that their conjunction is decomposable. Theorem 4. The theories Cx , Cv and Cr are independent: Cx ⊥ Cv , Cx ⊥ Cr and Cv ⊥ Cr . ⊥ ⊥ ⊥ The proof of Theorem 4 relies on the following Lemma. Lemma 5. If the theory allows for domain recursion, all clauses and atoms contain the same number of variables from B: ∃n, ∀c ∈ C, ∀a ∈ atom(C) : | vars(c) ∩ B | = | vars(a) ∩ B | = n. c Proof. Denote with Cn the clauses in C that contain n logical variables from B and with Cn its compliment in C. If C is nonempty, there is a n > 0 for which Cn is nonempty. Then every atom in Cn contains exactly n variables from B (Definition 5). Since the theory contains no independent c c subtheories, there must be an atom a in Cn which unifies with an atom ac in Cn , or Cn is empty. After shattering, all unifications bind one variable from a to a single variable from ac . Because a contains exactly n variables from B, ac must also contain exactly n (Definition 4), and because B is c a root binding class, the clause of ac also contains exactly n, which contradicts the definition of Cn . c Therefore, Cn is empty, and because the variables in B are root, they also appear in all atoms. Proof of Theorem 4. From Lemma 5, all atoms in C contain the same number of variables from B. In Cx , these variables are all constrained to be equal to xD , while in Cv and Cr at least one variable is constrained to be different from xD . An attempt to unify an atom from Cx with an atom from Cv or Cr therefore creates an unsatisfiable set of constraints. Similarly, atoms from Cv and Cr cannot be unified. Finally, we extend the FO d-DNNF language proposed in Van den Broeck et al. [10] with a new node, the recursive decomposable conjunction ∧ r , and define the domain recursion compilation rule. Definition 8 ( ∧ r ). The FO d-DNNF node ∧ r (nx , nr , D, D , V) represents a decomposable conjunction between the d-DNNF nodes nx , nr and a d-DNNF node isomorphic to the ∧ r node itself. In particular, the isomorphic operand is identical to the node itself, except for the size of the domain of the variables in V, which becomes one smaller, going from D to D in the isomorphic operand. We have shown that the conjunction between sets Cx , Cv and Cr is decomposable (Theorem 4) and logically equivalent to the original theory (Proposition 3). Furthermore, Cv is identical to C, up to the constraints on the domain of the variables in B. This leads us to the following definition of domain recursion. Definition 9 (Domain Recursion). The domain recursion compilation rule compiles C into ∧ r (nx , nr , D, D , B), where nx , nr are the compiled circuits for Cx , Cr . The third set Cv is represented by the recursion on D, according to Definition 8. 6 nv Cr ∧r ¬ friends(x, X) ∨ friends(X, x), X = x ¬ friends(X, x) ∨ friends(x, X), X = x P erson ← P erson \ {x} nr nx ∨ Cx ¬ friends(x, x) ∨ friends(x, x) x ∈P erson x =x ¬ friends(x, x) friends(x, x) ∨ ∧ ¬ friends(x, x ) ¬ friends(x , x) ∧ friends(x, x ) friends(x , x) Figure 3: Circuit for the symmetric relation in Equation 3, rooted in a recursive conjunction. Example 6. Figure 3 shows the FO d-DNNF circuit for Equation 3. The theory is split up into three independent theories: Cr and Cx , shown in the Figure 3, and Cv = {¬ friends(X, Y ) ∨ friends(Y, X), X = x, Y = x}. The conjunction of these theories is equivalent to Equation 3. Theory Cv is identical to Equation 3, up to the inequality constraints on X and Y . Theorem 6. Given a function size, which maps domains to their size, the weighted first-order model count of a ∧ r (nx , nr , D, D , V) node is size(D) wmc( ∧ r (nx , nr , D, D , V), size) = wmc(nx , size)size(D) wmc(nr , size ∪{D → s}), s=0 (11) where size ∪{D → s} adds to the size function that the subdomain D has cardinality s. Proof. If C allows for domain recursion, due to Theorem 4, the weighted model count is wmc(C, size) = 1, if size(D) = 0 wmc(Cx ) · wmc(Cv , size ) · wmc(Cr , size ) if size(D) > 0 (12) where size = size ∪{D → size(D) − 1}. Theorem 7. The Independent Partial Grounding compilation rule is a special case of the domain recursion rule, where ∀c ∈ C : | vars(c) ∩ B | = 1 (and therefore Cr = ∅). 4.3 Completeness In this section, we introduce a class of models for which first-order knowledge compilation with domain recursion is complete. Definition 10 (k-WFOMC). The class of k-WFOMC consist of WFOMC theories with clauses that have up to k logical variables. A first completeness result is for 2-WFOMC, using the set of knowledge compilation rules CR 2 , which are the rules in CR 1 extended with domain recursion. Theorem 8 (Completeness for 2-WFOMC). First-order knowledge compilation using the CR 2 compilation rules is a complete domain lifted probabilistic inference algorithm for 2-WFOMC. Proof. From Lemma 1, after applying the CR 1 rules, the theory contains only atoms with dimension larger than or equal to two. From Definition 10, each clause has dimension smaller than or equal to two. Therefore, each logical variable in the theory is a root variable and according to Definition 5, every equivalence class of binding variables is a root binding class. Because of Lemma 1, the theory allows for domain recursion, which requires further compilation of two theories: Cx and Cr into nx and nr . Both have dimension smaller than 2 and can be lifted by CR 1 compilation rules. The properties of 2-WFOMC are a sufficient but not necessary condition for first-order knowledge compilation to be domain lifted. We can obtain a similar result for MLNs or parfactors by reducing them to a WFOMC problem. If an MLN contains only formulae with up to k logical variables, then its WFOMC representation will be in k-WFOMC. 7 This result for 2-WFOMC is not trivial. Van den Broeck et al. [10] showed in their experiments that counting first-order variable elimination (C-FOVE) [6] fails to lift the “Friends Smoker Drinker” problem, which is in 2-WFOMC. We will show in the next section that the CR 1 rules fail to lift the theory in Figure 4a, which is in 2-WFOMC. Note that there are also useful theories that are not in 2-WFOMC, such as those containing the transitive relation friends(X, Y ) ∧ friends(Y, Z) ⇒ friends(X, Z). 5 Empirical evaluation To complement the theoretical results of the previous section, we extended the WFOMC implementation1 with the domain recursion rule. We performed experiments with the theory in Figure 4a, which is a version of the friends and smokers model [11] extended with the symmetric relation of Equation 3. We evaluate the performance querying P(smokes(bob)) with increasing domain size, comparing our approach to the existing WFOMC implementation and its propositional counterpart, which first grounds the theory and then compiles it with the c2d compiler [13] to a propositional d-DNNF circuit. We did not compare to C-FOVE [6] because it cannot perform lifted inference on this model. 2 smokes(X) ∧ friends(X, Y ) ⇒ smokes(Y ) friends(X, Y ) ⇒ friends(Y, X). Runtime [s] Propositional inference quickly becomes intractable when there are more than 20 people. The lifted inference algorithms scale much better. The CR 1 rules can exploit some regularities in the model. For example, they eliminate all the smokes(X) atoms from the theory. They do, however, resort to grounding at a later stage of the compilation process. With the domain recursion rule, there is no need for grounding. This advantage is clear in the experiments, our approach having an almost constant inference time in this range of domains sizes. Note that the runtimes for c2d include compilation and evaluation of the circuit, whereas the WFOMC runtimes only represent evaluation of the FO d-DNNF. After all, propositional compilation depends on the domain size but first-order compilation does not. First-order compilation takes a constant two seconds for both rule sets. 10000 1000 100 10 1 0.1 0.01 c2d WFOMC - CR1 WFOMC - CR2 10 20 30 40 50 60 Number of People 70 80 (b) Evaluation Runtime (a) MLN Model Figure 4: Symmetric friends and smokers experiment, comparing propositional knowledge compilation (c2d) to WFOMC using compilation rules CR 1 and CR 2 (which includes domain recursion). 6 Conclusions We proposed a definition of complete domain lifted probabilistic inference w.r.t. classes of probabilistic logic models. This definition considers algorithms to be lifted if they are polynomial in the size of logical variable domains. Existing first-order knowledge compilation turns out not to admit an intuitive completeness result. Therefore, we generalized the existing Independent Partial Grounding compilation rule to the domain recursion rule. With this one extra rule, we showed that first-order knowledge compilation is complete for a significant class of probabilistic logic models, where the WFOMC representation has up to two logical variables per clause. Acknowledgments The author would like to thank Luc De Raedt, Jesse Davis and the anonymous reviewers for valuable feedback. This work was supported by the Research Foundation-Flanders (FWO-Vlaanderen). 1 http://dtai.cs.kuleuven.be/wfomc/ 8 References [1] Lise Getoor and Ben Taskar, editors. An Introduction to Statistical Relational Learning. MIT Press, 2007. [2] Luc De Raedt, Paolo Frasconi, Kristian Kersting, and Stephen Muggleton, editors. Probabilistic inductive logic programming: theory and applications. Springer-Verlag, Berlin, Heidelberg, 2008. [3] Daan Fierens, Guy Van den Broeck, Ingo Thon, Bernd Gutmann, and Luc De Raedt. Inference in probabilistic logic programs using weighted CNF’s. In Proceedings of UAI, pages 256–265, 2011. [4] David Poole. First-order probabilistic inference. In Proceedings of IJCAI, pages 985–991, 2003. [5] Rodrigo de Salvo Braz, Eyal Amir, and Dan Roth. Lifted first-order probabilistic inference. In Proceedings of IJCAI, pages 1319–1325, 2005. [6] Brian Milch, Luke S. Zettlemoyer, Kristian Kersting, Michael Haimes, and Leslie Pack Kaelbling. Lifted Probabilistic Inference with Counting Formulas. In Proceedings of AAAI, pages 1062–1068, 2008. [7] Vibhav Gogate and Pedro Domingos. Exploiting Logical Structure in Lifted Probabilistic Inference. In Proceedings of StarAI, 2010. [8] Abhay Jha, Vibhav Gogate, Alexandra Meliou, and Dan Suciu. Lifted Inference Seen from the Other Side: The Tractable Features. In Proceedings of NIPS, 2010. [9] Vibhav Gogate and Pedro Domingos. Probabilistic theorem proving. In Proceedings of UAI, pages 256–265, 2011. [10] Guy Van den Broeck, Nima Taghipour, Wannes Meert, Jesse Davis, and Luc De Raedt. Lifted Probabilistic Inference by First-Order Knowledge Compilation. In Proceedings of IJCAI, pages 2178–2185, 2011. [11] Parag Singla and Pedro Domingos. Lifted first-order belief propagation. In Proceedings of AAAI, pages 1094–1099, 2008. [12] Matthew Richardson and Pedro Domingos. Markov logic networks. Machine Learning, 62(1): 107–136, 2006. [13] Adnan Darwiche. New advances in compiling CNF to decomposable negation normal form. In Proceedings of ECAI, pages 328–332, 2004. 9

5 0.45592797 237 nips-2011-Reinforcement Learning using Kernel-Based Stochastic Factorization

Author: Andre S. Barreto, Doina Precup, Joelle Pineau

Abstract: Kernel-based reinforcement-learning (KBRL) is a method for learning a decision policy from a set of sample transitions which stands out for its strong theoretical guarantees. However, the size of the approximator grows with the number of transitions, which makes the approach impractical for large problems. In this paper we introduce a novel algorithm to improve the scalability of KBRL. We resort to a special decomposition of a transition matrix, called stochastic factorization, to fix the size of the approximator while at the same time incorporating all the information contained in the data. The resulting algorithm, kernel-based stochastic factorization (KBSF), is much faster but still converges to a unique solution. We derive a theoretical upper bound for the distance between the value functions computed by KBRL and KBSF. The effectiveness of our method is illustrated with computational experiments on four reinforcement-learning problems, including a difficult task in which the goal is to learn a neurostimulation policy to suppress the occurrence of seizures in epileptic rat brains. We empirically demonstrate that the proposed approach is able to compress the information contained in KBRL’s model. Also, on the tasks studied, KBSF outperforms two of the most prominent reinforcement-learning algorithms, namely least-squares policy iteration and fitted Q-iteration. 1

6 0.45142785 300 nips-2011-Variance Reduction in Monte-Carlo Tree Search

7 0.44640318 59 nips-2011-Composite Multiclass Losses

8 0.44183078 174 nips-2011-Monte Carlo Value Iteration with Macro-Actions

9 0.4255119 50 nips-2011-Budgeted Optimization with Concurrent Stochastic-Duration Experiments

10 0.38794112 212 nips-2011-Periodic Finite State Controllers for Efficient POMDP and DEC-POMDP Planning

11 0.37195531 122 nips-2011-How Do Humans Teach: On Curriculum Learning and Teaching Dimension

12 0.36892897 36 nips-2011-Analysis and Improvement of Policy Gradient Estimation

13 0.36449724 20 nips-2011-Active Learning Ranking from Pairwise Preferences with Almost Optimal Query Complexity

14 0.36432803 79 nips-2011-Efficient Offline Communication Policies for Factored Multiagent POMDPs

15 0.35666704 215 nips-2011-Policy Gradient Coagent Networks

16 0.35015929 16 nips-2011-A reinterpretation of the policy oscillation phenomenon in approximate policy iteration

17 0.34783494 10 nips-2011-A Non-Parametric Approach to Dynamic Programming

18 0.333352 152 nips-2011-Learning in Hilbert vs. Banach Spaces: A Measure Embedding Viewpoint

19 0.32636276 52 nips-2011-Clustering via Dirichlet Process Mixture Models for Portable Skill Discovery

20 0.31825405 3 nips-2011-A Collaborative Mechanism for Crowdsourcing Prediction Problems


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(0, 0.078), (4, 0.028), (20, 0.02), (26, 0.055), (31, 0.084), (33, 0.019), (43, 0.052), (45, 0.072), (53, 0.319), (57, 0.034), (74, 0.047), (83, 0.041), (99, 0.036)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.75816905 256 nips-2011-Solving Decision Problems with Limited Information

Author: Denis D. Maua, Cassio Campos

Abstract: We present a new algorithm for exactly solving decision-making problems represented as an influence diagram. We do not require the usual assumptions of no forgetting and regularity, which allows us to solve problems with limited information. The algorithm, which implements a sophisticated variable elimination procedure, is empirically shown to outperform a state-of-the-art algorithm in randomly generated problems of up to 150 variables and 1064 strategies. 1

2 0.72914761 120 nips-2011-History distribution matching method for predicting effectiveness of HIV combination therapies

Author: Jasmina Bogojeska

Abstract: This paper presents an approach that predicts the effectiveness of HIV combination therapies by simultaneously addressing several problems affecting the available HIV clinical data sets: the different treatment backgrounds of the samples, the uneven representation of the levels of therapy experience, the missing treatment history information, the uneven therapy representation and the unbalanced therapy outcome representation. The computational validation on clinical data shows that, compared to the most commonly used approach that does not account for the issues mentioned above, our model has significantly higher predictive power. This is especially true for samples stemming from patients with longer treatment history and samples associated with rare therapies. Furthermore, our approach is at least as powerful for the remaining samples. 1

3 0.70196223 79 nips-2011-Efficient Offline Communication Policies for Factored Multiagent POMDPs

Author: João V. Messias, Matthijs Spaan, Pedro U. Lima

Abstract: Factored Decentralized Partially Observable Markov Decision Processes (DecPOMDPs) form a powerful framework for multiagent planning under uncertainty, but optimal solutions require a rigid history-based policy representation. In this paper we allow inter-agent communication which turns the problem in a centralized Multiagent POMDP (MPOMDP). We map belief distributions over state factors to an agent’s local actions by exploiting structure in the joint MPOMDP policy. The key point is that when sparse dependencies between the agents’ decisions exist, often the belief over its local state factors is sufficient for an agent to unequivocally identify the optimal action, and communication can be avoided. We formalize these notions by casting the problem into convex optimization form, and present experimental results illustrating the savings in communication that we can obtain.

4 0.63777369 5 nips-2011-A Denoising View of Matrix Completion

Author: Weiran Wang, Zhengdong Lu, Miguel Á. Carreira-Perpiñán

Abstract: In matrix completion, we are given a matrix where the values of only some of the entries are present, and we want to reconstruct the missing ones. Much work has focused on the assumption that the data matrix has low rank. We propose a more general assumption based on denoising, so that we expect that the value of a missing entry can be predicted from the values of neighboring points. We propose a nonparametric version of denoising based on local, iterated averaging with meanshift, possibly constrained to preserve local low-rank manifold structure. The few user parameters required (the denoising scale, number of neighbors and local dimensionality) and the number of iterations can be estimated by cross-validating the reconstruction error. Using our algorithms as a postprocessing step on an initial reconstruction (provided by e.g. a low-rank method), we show consistent improvements with synthetic, image and motion-capture data. Completing a matrix from a few given entries is a fundamental problem with many applications in machine learning, computer vision, network engineering, and data mining. Much interest in matrix completion has been caused by recent theoretical breakthroughs in compressed sensing [1, 2] as well as by the now celebrated Netflix challenge on practical prediction problems [3, 4]. Since completion of arbitrary matrices is not a well-posed problem, it is often assumed that the underlying matrix comes from a restricted class. Matrix completion models almost always assume a low-rank structure of the matrix, which is partially justified through factor models [4] and fast convex relaxation [2], and often works quite well when the observations are sparse and/or noisy. The low-rank structure of the matrix essentially asserts that all the column vectors (or the row vectors) live on a low-dimensional subspace. This assumption is arguably too restrictive for problems with richer structure, e.g. when each column of the matrix represents a snapshot of a seriously corrupted motion capture sequence (see section 3), for which a more flexible model, namely a curved manifold, is more appropriate. In this paper, we present a novel view of matrix completion based on manifold denoising, which conceptually generalizes the low-rank assumption to curved manifolds. Traditional manifold denoising is performed on fully observed data [5, 6], aiming to send the data corrupted by noise back to the correct surface (defined in some way). However, with a large proportion of missing entries, we may not have a good estimate of the manifold. Instead, we start with a poor estimate and improve it iteratively. Therefore the “noise” may be due not just to intrinsic noise, but mostly to inaccurately estimated missing entries. We show that our algorithm can be motivated from an objective purely based on denoising, and prove its convergence under some conditions. We then consider a more general case with a nonlinear low-dimensional manifold and use a stopping criterion that works successfully in practice. Our model reduces to a low-rank model when we require the manifold to be flat, showing a relation with a recent thread of matrix completion models based on alternating projection [7]. In our experiments, we show that our denoising-based matrix completion model can make better use of the latent manifold structure on both artificial and real-world data sets, and yields superior recovery of the missing entries. The paper is organized as follows: section 1 reviews nonparametric denoising methods based on mean-shift updates, section 2 extends this to matrix completion by using denoising with constraints, section 3 gives experimental results, and section 4 discusses related work. 1 1 Denoising with (manifold) blurring mean-shift algorithms (GBMS/MBMS) In Gaussian blurring mean-shift (GBMS), denoising is performed in a nonparametric way by local averaging: each data point moves to the average of its neighbors (to a certain scale), and the process is repeated. We follow the derivation in [8]. Consider a dataset {xn }N ⊂ RD and define a n=1 Gaussian kernel density estimate p(x) = 1 N N Gσ (x, xn ) (1) n=1 1 with bandwidth σ > 0 and kernel Gσ (x, xn ) ∝ exp − 2 ( x − xn /σ)2 (other kernels may be used, such as the Epanechnikov kernel, which results in sparse affinities). The (non-blurring) mean-shift algorithm rearranges the stationary point equation ∇p(x) = 0 into the iterative scheme x(τ +1) = f (x(τ ) ) with N x (τ +1) = f (x (τ ) p(n|x )= (τ ) )xn p(n|x (τ ) n=1 )= exp − 1 (x(τ ) − xn )/σ 2 N n′ =1 2 exp − 1 (x(τ ) − xn′ )/σ 2 2 . (2) This converges to a mode of p from almost every initial x ∈ RD , and can be seen as taking selfadapting step sizes along the gradient (since the mean shift f (x) − x is parallel to ∇p(x)). This iterative scheme was originally proposed by [9] and it or variations of it have found widespread application in clustering [8, 10–12] and denoising of 3D point sets (surface fairing; [13, 14]) and manifolds in general [5, 6]. The blurring mean-shift algorithm applies one step of the previous scheme, initialized from every point, in parallel for all points. That is, given the dataset X = {x1 , . . . , xN }, for each xn ∈ X ˜ we obtain a new point xn = f (xn ) by applying one step of the mean-shift algorithm, and then we ˜ replace X with the new dataset X, which is a blurred (shrunk) version of X. By iterating this process we obtain a sequence of datasets X(0) , X(1) , . . . (and a corresponding sequence of kernel density estimates p(0) (x), p(1) (x), . . .) where X(0) is the original dataset and X(τ ) is obtained by blurring X(τ −1) with one mean-shift step. We can see this process as maximizing the following objective function [10] by taking parallel steps of the form (2) for each point: N p(xn ) = E(X) = n=1 1 N N N 1 e− 2 Gσ (xn , xm ) ∝ xn −xm σ 2 . (3) n,m=1 n,m=1 This process eventually converges to a dataset X(∞) where all points are coincident: a completely denoised dataset where all structure has been erased. As shown by [8], this process can be stopped early to return clusters (= locally denoised subsets of points); the number of clusters obtained is controlled by the bandwidth σ. However, here we are interested in the denoising behavior of GBMS. ˜ The GBMS step can be formulated in a matrix form reminiscent of spectral clustering [8] as X = X P where X = (x1 , . . . , xN ) is a D×N matrix of data points; W is the N ×N matrix of Gaussian N affinities wnm = Gσ (xn , xm ); D = diag ( n=1 wnm ) is the degree matrix; and P = WD−1 is N an N × N stochastic matrix: pnm = p(n|xm ) ∈ (0, 1) and n=1 pnm = 1. P (or rather its transpose) is the stochastic matrix of the random walk in a graph [15], which in GBMS represents the posterior probabilities of each point under the kernel density estimate (1). P is similar to the 1 1 matrix N = D− 2 WD− 2 derived from the normalized graph Laplacian commonly used in spectral clustering, e.g. in the normalized cut [16]. Since, by the Perron-Frobenius theorem [17, ch. 8], all left eigenvalues of P(X) have magnitude less than 1 except for one that equals 1 and is associated with ˜ an eigenvector of constant entries, iterating X = X P(X) converges to the stationary distribution of each P(X), where all points coincide. ˜ From this point of view, the product X = X P(X) can be seen as filtering the dataset X with a datadependent low-pass filter P(X), which makes clear the denoising behavior. This also suggests using ˜ other filters [12] X = X φ(P(X)) as long as φ(1) = 1 and |φ(r)| < 1 for r ∈ [0, 1), such as explicit schemes φ(P) = (1 − η)I + ηP for η ∈ (0, 2], power schemes φ(P) = Pn for n = 1, 2, 3 . . . or implicit schemes φ(P) = ((1 + η)I − ηP)−1 for η > 0. One important problem with GBMS is that it denoises equally in all directions. When the data lies on a low-dimensional manifold, denoising orthogonally to it removes out-of-manifold noise, but 2 denoising tangentially to it perturbs intrinsic degrees of freedom of the data and causes shrinkage of the entire manifold (most strongly near its boundary). To prevent this, the manifold blurring meanshift algorithm (MBMS) [5] first computes a predictor averaging step with GBMS, and then for each point xn a corrector projective step removes the step direction that lies in the local tangent space of xn (obtained from local PCA run on its k nearest neighbors). In practice, both GBMS and MBMS must be stopped early to prevent excessive denoising and manifold distortions. 2 Blurring mean-shift denoising algorithms for matrix completion We consider the natural extension of GBMS to the matrix completion case by adding the constraints given by the present values. We use the subindex notation XM and XP to indicate selection of the missing or present values of the matrix XD×N , where P ⊂ U , M = U \ P and U = {(d, n): d = 1, . . . , D, n = 1, . . . , N }. The indices P and values XP of the present matrix entries are the data of the problem. Then we have the following constrained optimization problem: N Gσ (xn , xm ) max E(X) = X s.t. XP = XP . (4) n,m=1 This is similar to low-rank formulations for matrix completion that have the same constraints but use as objective function the reconstruction error with a low-rank assumption, e.g. X − ABX 2 with AD×L , BL×D and L < D. We initialize XM to the output of some other method for matrix completion, such as singular value projection (SVP; [7]). For simple constraints such as ours, gradient projection algorithms are attractive. The gradient of E wrt X is a matrix of D × N whose nth column is: ∇xn E(X) = 2 σ2 N 1 e− 2 xn −xm σ 2 N (xm − xn ) ∝ m=1 2 p(m|xn )xm p(xn ) −xn + σ2 m=1 (5) and its projection on the constraint space is given by zeroing its entries having indices in P; call ΠP this projection operator. Then, we have the following step of length α ≥ 0 along the projected gradient: (τ +1) X(τ +1) = X(τ ) + αΠP (∇X E(X(τ ) )) ⇐⇒ XM (τ ) = XM + α ΠP (∇X E(X(τ ) )) M (6) which updates only the missing entries XM . Since our search direction is ascent and makes an angle with the gradient that is bounded away from π/2, and E is lower bounded, continuously differentiable and has bounded Hessian (thus a Lipschitz continuous gradient) in RN L , by carrying out a line search that satisfies the Wolfe conditions, we are guaranteed convergence to a local stationary point, typically a maximizer [18, th. 3.2]. However, as reasoned later, we do not perform a line search at all, instead we fix the step size to the GBMS self-adapting step size, which results in a simple and faster algorithm consisting of carrying out a GBMS step on X (i.e., X(τ +1) = X(τ ) P(X(τ ) )) and then refilling XP to the present values. While we describe the algorithm in this way for ease of explanation, in practice we do not actually compute the GBMS step for all xdn values, but only for the missing ones, which is all we need. Thus, our algorithm carries out GBMS denoising steps within the missing-data subspace. We can derive this result in a different way by starting from N the unconstrained optimization problem maxXP E(X) = n,m=1 Gσ (xn , xm ) (equivalent to (4)), computing its gradient wrt XP , equating it to zero and rearranging (in the same way the mean-shift algorithm is derived) to obtain a fixed-point iteration identical to our update above. Fig. 1 shows the pseudocode for our denoising-based matrix completion algorithms (using three nonparametric denoising algorithms: GBMS, MBMS and LTP). Convergence and stopping criterion As noted above, we have guaranteed convergence by simply satisfying standard line search conditions, but a line search is costly. At present we do not have (τ +1) a proof that the GBMS step size satisfies such conditions, or indeed that the new iterate XM increases or leaves unchanged the objective, although we have never encountered a counterexample. In fact, it turns out that none of the work about GBMS that we know about proves that either: [10] proves that ∅(X(τ +1) ) ≤ ∅(X(τ ) ) for 0 < ρ < 1, where ∅(·) is the set diameter, while [8, 12] 3 notes that P(X) has a single eigenvalue of value 1 and all others of magnitued less than 1. While this shows that all points converge to the same location, which indeed is the global maximum of (3), it does not necessarily follow that each step decreases E. GBMS (k, σ) with full or k-nn graph: given XD×N , M repeat for n = 1, . . . , N Nn ← {1, . . . , N } (full graph) or k nearest neighbors of xn (k-nn graph) Gσ (xn ,xm ) mean-shift xm ∂xn ← −xn + m∈Nn step m′ ∈Nn Gσ (xn ,xm′ ) end XM ← XM + (∂X)M move points’ missing entries until validation error increases return X However, the question of convergence as τ → ∞ has no practical interest in a denoising setting, because achieving a total denoising almost never yields a good matrix completion. What we want is to achieve just enough denoising and stop the algorithm, as was the case with GBMS clustering, and as is the case in algorithms for image denoising. We propose to determine the optimal number of iterations, as well as the bandwidth σ and any other parameters, by cross-validation. Specifically, we select a held-out set by picking a random subset of the present entries and considering them as missing; this allows us to evaluate an error between our completion for them and the ground truth. We stop iterating when this error increases. MBMS (L, k, σ) with full or k-nn graph: given XD×N , M repeat for n = 1, . . . , N Nn ← {1, . . . , N } (full graph) or k nearest neighbors of xn (k-nn graph) Gσ (xn ,xm ) mean-shift xm ∂xn ← −xn + m∈Nn step m′ ∈Nn Gσ (xn ,xm′ ) Xn ← k nearest neighbors of xn (µn , Un ) ← PCA(Xn , L) estimate L-dim tangent space at xn subtract parallel motion ∂xn ← (I − Un UT )∂xn n end XM ← XM + (∂X)M move points’ missing entries until validation error increases return X This argument justifies an algorithmic, as opposed to an opLTP (L, k) with k-nn graph: given XD×N , M timization, view of denoisingrepeat based matrix completion: apfor n = 1, . . . , N ply a denoising step, refill the Xn ← k nearest neighbors of xn present values, iterate until the (µn , Un ) ← PCA(Xn , L) estimate L-dim tangent space at xn validation error increases. This project point onto tangent space allows very general definitions ∂xn ← (I − Un UT )(µn − xn ) n end of denoising, and indeed a lowXM ← XM + (∂X)M move points’ missing entries rank projection is a form of deuntil validation error increases noising where points are not alreturn X lowed outside the linear manifold. Our formulation using Figure 1: Our denoising matrix completion algorithms, based on the objective function (4) is still Manifold Blurring Mean Shift (MBMS) and its particular cases useful in that it connects our Local Tangent Projection (LTP, k-nn graph, σ = ∞) and Gauss- denoising assumption with the ian Blurring Mean Shift (GBMS, L = 0); see [5] for details. Nn more usual low-rank assumption contains all N points (full graph) or only xn ’s nearest neighbors that has been used in much ma(k-nn graph). The index M selects the components of its input trix completion work, and juscorresponding to missing values. Parameters: denoising scale σ, tifies the refilling step as renumber of neighbors k, local dimensionality L. sulting from the present-data constraints under a gradientprojection optimization. MBMS denoising for matrix completion Following our algorithmic-based approach to denois˜ ing, we could consider generalized GBMS steps of the form X = X φ(P(X)). For clustering, Carreira-Perpi˜ an [12] found an overrelaxed explicit step φ(P) = (1 − η)I + ηP with η ≈ 1.25 to n´ achieve similar clusterings but faster. Here, we focus instead on the MBMS variant of GBMS that allows only for orthogonal, not tangential, point motions (defined wrt their local tangent space as estimated by local PCA), with the goal of preserving low-dimensional manifold structure. MBMS has 3 user parameters: the bandwidth σ (for denoising), and the latent dimensionality L and the 4 number of neighbors k (for the local tangent space and the neighborhood graph). A special case of MBMS called local tangent projection (LTP) results by using a neighborhood graph and setting σ = ∞ (so only two user parameters are needed: L and k). LTP can be seen as doing a low-rank matrix completion locally. LTP was found in [5] to have nearly as good performance as the best σ in several problems. MBMS also includes as particular cases GBMS (L = 0), PCA (k = N , σ = ∞), and no denoising (σ = 0 or L = D). Note that if we apply MBMS to a dataset that lies on a linear manifold of dimensionality d using L ≥ d then no denoising occurs whatsoever because the GBMS updates lie on the d-dimensional manifold and are removed by the corrector step. In practice, even if the data are assumed noiseless, the reconstruction from a low-rank method will lie close to but not exactly on the d-dimensional manifold. However, this suggests using largish ranks for the low-rank method used to reconstruct X and lower L values in the subsequent MBMS run. In summary, this yields a matrix completion algorithm where we apply an MBMS step, refill the present values, and iterate until the validation error increases. Again, in an actual implementation we compute the MBMS step only for the missing entries of X. The shrinking problem of GBMS is less pronounced in our matrix completion setting, because we constrain some values not to change. Still, in agreement with [5], we find MBMS to be generally superior to GBMS. Computational cost With a full graph, the cost per iteration of GBMS and MBMS is O(N 2 D) and O(N 2 D + N (D + k) min(D, k)2 ), respectively. In practice with high-dimensional data, best denoising results are obtained using a neighborhood graph [5], so that the sums over points in eqs. (3) or (4) extend only to the neighbors. With a k-nearest-neighbor graph and if we do not update the neighbors at each iteration (which affects the result little), the respective cost per iteration is O(N kD) and O(N kD + N (D + k) min(D, k)2 ), thus linear in N . The graph is constructed on the initial X we use, consisting of the present values and an imputation for the missing ones achieved with a standard matrix completion method, and has a one-off cost of O(N 2 D). The cost when we have a fraction µ = |M| ∈ [0, 1] of missing data is simply the above times µ. Hence the run time ND of our mean-shift-based matrix completion algorithms is faster the more present data we have, and thus faster than the usual GBMS or MBMS case, where all data are effectively missing. 3 Experimental results We compare with representative methods of several approaches: a low-rank matrix completion method, singular value projection (SVP [7], whose performance we found similar to that of alternating least squares, ALS [3, 4]); fitting a D-dimensional Gaussian model with EM and imputing the missing values of each xn as the conditional mean E {xn,Mn |xn,Pn } (we use the implementation of [19]); and the nonlinear method of [20] (nlPCA). We initialize GBMS and MBMS from some or all of these algorithms. For methods with user parameters, we set them by cross-validation in the following way: we randomly select 10% of the present entries and pretend they are missing as well, we run the algorithm on the remaining 90% of the present values, and we evaluate the reconstruction at the 10% entries we kept earlier. We repeat this over different parameters’ values and pick the one with lowest reconstruction error. We then run the algorithm with these parameters values on the entire present data and report the (test) error with the ground truth for the missing values. 100D Swissroll We created a 3D swissroll data set with 3 000 points and lifted it to 100D with a random orthonormal mapping, and added a little noise (spherical Gaussian with stdev 0.1). We selected uniformly at random 6.76% of the entries to be present. We use the Gaussian model and SVP (fixed rank = 3) as initialization for our algorithm. We typically find that these initial X are very noisy (fig. 3), with some reconstructed points lying between different branches of the manifold and causing a big reconstruction error. We fixed L = 2 (the known dimensionality) for MBMS and cross-validated the other parameters: σ and k for MBMS and GBMS (both using k-nn graph), and the number of iterations τ to be used. Table 1 gives the performance of MBMS and GBMS for testing, along with their optimal parameters. Fig. 3 shows the results of different methods at a few iterations. MBMS initialized from the Gaussian model gives the most remarkable denoising effect. To show that there is a wide range of σ and number of iterations τ that give good performance with GBMS and MBMS, we fix k = 50 and run the algorithm with varying σ values and plot the reconstruction error for missing entries over iterations in fig. 2. Both GBMS can achieve good 5 Methods Gaussian + GBMS (∞, 10, 0, 1) + MBMS (1, 20, 2, 25) SVP + GBMS (3, 50, 0, 1) + MBMS (3, 50, 2, 2) RSSE 168.1 165.8 157.2 156.8 151.4 151.8 mean 2.63 2.57 2.36 1.94 1.89 1.87 stdev 1.59 1.61 1.63 2.10 2.02 2.05 Methods nlPCA SVP + GBMS (400,140,0,1) + MBMS (500,140,9,5) Table 1: Swissroll data set: reconstruction errors obtained by different algorithms along with their optimal parameters (σ, k, L, no. iterations τ ). The three columns show the root sum of squared errors on missing entries, the mean, and the standard deviation of the pointwise reconstruction error, resp. SVP + GBMS error (RSSE) 180 170 SVP + MBMS Gaussian + GBMS 180 180 170 170 170 160 160 ∞ 160 150 0 1 2 3 4 5 6 7 8 910 12 14 16 18 20 iteration τ stdev 42.6 39.3 37.7 34.9 Gaussian + MBMS 180 8 10 15 25 mean 26.1 21.8 18.8 17.0 Table 2: MNIST-7 data set: errors of the different algorithms and their optimal parameters (σ, k, L, no. iterations τ ). The three columns show the root sum of squared errors on missing entries (×10−4 ), the mean, and the standard deviation of pixel errors, respectively. 160 0.3 0.5 1 2 3 5 RSSE 7.77 6.99 6.54 6.03 150 0 1 2 3 4 5 6 7 8 910 12 14 16 18 20 iteration τ 150 0 1 2 3 4 5 6 7 8 910 12 14 16 18 20 iteration τ 150 0 1 2 3 4 5 6 7 8 910 12 14 16 18 20 iteration τ Figure 2: Reconstruction error of GBMS/MBMS over iterations (each curve is a different σ value). denoising (and reconstruction), but MBMS is more robust, with good results occurring for a wide range of iterations, indicating it is able to preserve the manifold structure better. Mocap data We use the running-motion sequence 09 01 from the CMU mocap database with 148 samples (≈ 1.7 cycles) with 150 sensor readings (3D positions of 50 joints on a human body). The motion is intrinsically 1D, tracing a loop in 150D. We compare nlPCA, SVP, the Gaussian model, and MBMS initialized from the first three algorithms. For nlPCA, we do a grid search for the weight decay coefficient while fixing its structure to be 2 × 10 × 150 units, and use an early stopping criterion. For SVP, we do grid search on {1, 2, 3, 5, 7, 10} for the rank. For MBMS (L = 1) and GBMS (L = 0), we do grid search for σ and k. We report the reconstruction error as a function of the proportion of missing entries from 50% to 95%. For each missing-data proportion, we randomly select 5 different sets of present values and run all algorithms for them. Fig. 4 gives the mean errors of all algorithms. All methods perform well when missing-data proportion is small. nlPCA, being prone to local optima, is less stable than SVP and the Gaussian model, especially when the missing-data proportion is large. The Gaussian model gives the best and most stable initialization. At 95%, all methods fail to give an acceptable reconstruction, but up to 90% missing entries, MBMS and GBMS always beat the other algorithms. Fig. 4 shows selected reconstructions from all algorithms. MNIST digit ‘7’ The MNIST digit ‘7’ data set contains 6 265 greyscale (0–255) images of size 28 × 28. We create missing entries in a way reminiscent of run-length errors in transmission. We generate 16 to 26 rectangular boxes of an area approximately 25 pixels at random locations in each image and use them to black out pixels. In this way, we create a high dimensional data set (784 dimensions) with about 50% entries missing on average. Because of the loss of spatial correlations within the blocks, this missing data pattern is harder than random. The Gaussian model cannot handle such a big data set because it involves inverting large covariance matrices. nlPCA is also very slow and we cannot afford cross-validating its structure or the weight decay coefficient, so we picked a reasonable structure (10 × 30 × 784 units), used the default weight decay parameter in the code (10−3 ), and allowed up to 500 iterations. We only use SVP as initialization for our algorithm. Since the intrinsic dimension of MNIST is suspected to be not very high, 6 SVP τ =0 SVP + GBMS τ =1 SVP + MBMS τ =2 Gaussian τ =0 Gaussian + GBMS τ =1 Gaussian + MBMS τ = 25 20 20 20 20 20 20 15 15 15 15 15 15 10 10 10 10 10 10 5 5 5 5 5 5 0 0 0 0 0 0 −5 −5 −5 −5 −5 −5 −10 −10 −15 −15 −10 −5 0 5 10 15 20 −10 −15 −15 −10 −5 0 5 10 15 20 −15 −15 −10 −10 −5 0 5 10 15 20 −15 −15 −10 −10 −5 0 5 10 15 20 −15 −15 −10 −10 −5 0 5 10 15 20 −15 −15 −10 −5 0 5 10 15 20 Figure 3: Denoising effect of the different algorithms. For visualization, we project the 100D data to 3D with the projection matrix used for creating the data. Present values are refilled for all plots. 7000 6000 error 5000 4000 frame 2 (leg distance) frame 10 (foot pose) frame 147 (leg pose) nlPCA nlPCA + GBMS nlPCA + MBMS SVP SVP + GBMS SVP + MBMS Gaussian Gaussian + GBMS Gaussian + MBMS 3000 2000 1000 0 50 60 70 80 85 90 95 % of missing data Figure 4: Left: mean of errors (RSSE) of 5 runs obtained by different algorithms for varying percentage of missing values. Errorbars shown only for Gaussian + MBMS to avoid clutter. Right: sample reconstructions when 85% percent data is missing. Row 1: initialization. Row 2: init+GBMS. Row 3: init+MBMS. Color indicates different initialization: black, original data; red, nlPCA; blue, SVP; green, Gaussian. we used rank 10 for SVP and L = 9 for MBMS. We also use the same k = 140 as in [5]. So we only had to choose σ and the number of iterations via cross-validation. Table 2 shows the methods and their corresponding error. Fig. 5 shows some representative reconstructions from different algorithms, with present values refilled. The mean-shift averaging among closeby neighbors (a soft form of majority voting) helps to eliminate noise, unusual strokes and other artifacts created by SVP, which by their nature tend to occur in different image locations over the neighborhood of images. 4 Related work Matrix completion is widely studied in theoretical compressed sensing [1, 2] as well as practical recommender systems [3, 4]. Most matrix completion models rely on a low-rank assumption, and cannot fully exploit a more complex structure of the problem, such as curved manifolds. Related work is on multi-task learning in a broad sense, which extracts the common structure shared by multiple related objects and achieves simultaneous learning on them. This includes applications such as alignment of noise-corrupted images [21], recovery of images with occlusion [22], and even learning of multiple related regressors or classifiers [23]. Again, all these works are essentially based on a subspace assumption, and do not generalize to more complex situations. A line of work based on a nonlinear low-rank assumption (with a latent variable z of dimensionN 2 ality L < D) involves setting up a least-squares error function minf ,Z n=1 xn − f (zn ) = N,D 2 n,d=1 (xdn − fd (zn )) where one ignores the terms for which xdn is missing, and estimates the function f and the low-dimensional data projections Z by alternating optimization. Linear functions f have been used in the homogeneity analysis literature [24], where this approach is called “missing data deleted”. Nonlinear functions f have been used recently (neural nets [20]; Gaussian processes for collaborative filtering [25]). Better results are obtained if adding a projection term N 2 and optimizing over the missing data as well [26]. n=1 zn − F(xn ) 7 Orig Missing nlPCA SVP GBMS MBMS Orig Missing nlPCA SVP GBMS MBMS Figure 5: Selected reconstructions of MNIST block-occluded digits ‘7’ with different methods. Prior to our denoising-based work there have been efforts to extend the low-rank models to smooth manifolds, mostly in the context of compressed sensing. Baraniuk and Wakin [27] show that certain random measurements, e.g. random projection to a low-dimensional subspace, can preserve the metric of the manifold fairly well, if the intrinsic dimension and the curvature of the manifold are both small enough. However, these observations are not suitable for matrix completion and no algorithm is given for recovering the signal. Chen et al. [28] explicitly model a pre-determined manifold, and use this to regularize the signal when recovering the missing values. They estimate the manifold given complete data, while no complete data is assumed in our matrix completion setting. Another related work is [29], where the manifold modeled with Isomap is used in estimating the positions of satellite cameras in an iterative manner. Finally, our expectation that the value of a missing entry can be predicted from the values of neighboring points is similar to one category of collaborative filtering methods that essentially use similar users/items to predict missing values [3, 4]. 5 Conclusion We have proposed a new paradigm for matrix completion, denoising, which generalizes the commonly used assumption of low rank. Assuming low-rank implies a restrictive form of denoising where the data is forced to have zero variance away from a linear manifold. More general definitions of denoising can potentially handle data that lives in a low-dimensional manifold that is nonlinear, or whose dimensionality varies (e.g. a set of manifolds), or that does not have low rank at all, and naturally they handle noise in the data. Denoising works because of the fundamental fact that a missing value can be predicted by averaging nearby present values. Although we motivate our framework from a constrained optimization point of view (denoise subject to respecting the present data), we argue for an algorithmic view of denoising-based matrix completion: apply a denoising step, refill the present values, iterate until the validation error increases. In turn, this allows different forms of denoising, such as based on low-rank projection (earlier work) or local averaging with blurring mean-shift (this paper). Our nonparametric choice of mean-shift averaging further relaxes assumptions about the data and results in a simple algorithm with very few user parameters that afford user control (denoising scale, local dimensionality) but can be set automatically by cross-validation. Our algorithms are intended to be used as a postprocessing step over a user-provided initialization of the missing values, and we show they consistently improve upon existing algorithms. The MBMS-based algorithm bridges the gap between pure denoising (GBMS) and local low rank. Other definitions of denoising should be possible, for example using temporal as well as spatial neighborhoods, and even applicable to discrete data if we consider denoising as a majority voting among the neighbours of a vector (with suitable definitions of votes and neighborhood). Acknowledgments Work supported by NSF CAREER award IIS–0754089. 8 References [1] Emmanuel J. Cand` s and Benjamin Recht. Exact matrix completion via convex optimization. Foundations e of Computational Mathematics, 9(6):717–772, December 2009. [2] Emmanuel J. Cand` s and Terence Tao. The power of convex relaxation: Near-optimal matrix completion. e IEEE Trans. Information Theory, 56(5):2053–2080, April 2010. [3] Yehuda Koren. Factorization meets the neighborhood: A multifaceted collaborative filtering model. SIGKDD 2008, pages 426–434, Las Vegas, NV, August 24–27 2008. [4] Robert Bell and Yehuda Koren. Scalable collaborative filtering with jointly derived neighborhood interpolation weights. ICDM 2007, pages 43–52, October 28–31 2007. ´ [5] Weiran Wang and Miguel A. Carreira-Perpi˜ an. Manifold blurring mean shift algorithms for manifold n´ denoising. CVPR 2010, pages 1759–1766, San Francisco, CA, June 13–18 2010. [6] Matthias Hein and Markus Maier. Manifold denoising. NIPS 2006, 19:561–568. MIT Press, 2007. [7] Prateek Jain, Raghu Meka, and Inderjit S. Dhillon. Guaranteed rank minimization via singular value projection. NIPS 2010, 23:937–945. MIT Press, 2011. ´ [8] Miguel A. Carreira-Perpi˜ an. Fast nonparametric clustering with Gaussian blurring mean-shift. ICML n´ 2006, pages 153–160. Pittsburgh, PA, June 25–29 2006. [9] Keinosuke Fukunaga and Larry D. Hostetler. The estimation of the gradient of a density function, with application in pattern recognition. IEEE Trans. Information Theory, 21(1):32–40, January 1975. [10] Yizong Cheng. Mean shift, mode seeking, and clustering. IEEE Trans. PAMI, 17(8):790–799, 1995. [11] Dorin Comaniciu and Peter Meer. Mean shift: A robust approach toward feature space analysis. IEEE Trans. PAMI, 24(5):603–619, May 2002. ´ [12] Miguel A. Carreira-Perpi˜ an. Generalised blurring mean-shift algorithms for nonparametric clustering. n´ CVPR 2008, Anchorage, AK, June 23–28 2008. [13] Gabriel Taubin. A signal processing approach to fair surface design. SIGGRAPH 1995, pages 351–358. [14] Mathieu Desbrun, Mark Meyer, Peter Schr¨ der, and Alan H. Barr. Implicit fairing of irregular meshes o using diffusion and curvature flow. SIGGRAPH 1999, pages 317–324. [15] Fan R. K. Chung. Spectral Graph Theory. American Mathematical Society, Providence, RI, 1997. [16] Jianbo Shi and Jitendra Malik. Normalized cuts and image segmentation. IEEE Trans. PAMI, 22(8):888– 905, August 2000. [17] Roger A. Horn and Charles R. Johnson. Matrix Analysis. Cambridge University Press, 1986. [18] Jorge Nocedal and Stephen J. Wright. Numerical Optimization. Springer-Verlag, New York, second edition, 2006. [19] Tapio Schneider. Analysis of incomplete climate data: Estimation of mean values and covariance matrices and imputation of missing values. Journal of Climate, 14(5):853–871, March 2001. [20] Matthias Scholz, Fatma Kaplan, Charles L. Guy, Joachim Kopka, and Joachim Selbig. Non-linear PCA: A missing data approach. Bioinformatics, 21(20):3887–3895, October 15 2005. [21] Yigang Peng, Arvind Ganesh, John Wright, Wenli Xu, and Yi Ma. RASL: Robust alignment by sparse and low-rank decomposition for linearly correlated images. CVPR 2010, pages 763–770, 2010. [22] A. M. Buchanan and A. W. Fitzgibbon. Damped Newton algorithms for matrix factorization with missing data. CVPR 2005, pages 316–322, San Diego, CA, June 20–25 2005. [23] Andreas Argyriou, Theodoros Evgeniou, and Massimiliano Pontil. Multi-task feature learning. NIPS 2006, 19:41–48. MIT Press, 2007. [24] Albert Gifi. Nonlinear Multivariate Analysis. John Wiley & Sons, 1990. [25] Neil D. Lawrence and Raquel Urtasun. Non-linear matrix factorization with Gaussian processes. ICML 2009, Montreal, Canada, June 14–18 2009. ´ [26] Miguel A. Carreira-Perpi˜ an and Zhengdong Lu. Manifold learning and missing data recovery through n´ unsupervised regression. ICDM 2011, December 11–14 2011. [27] Richard G. Baraniuk and Michael B. Wakin. Random projections of smooth manifolds. Foundations of Computational Mathematics, 9(1):51–77, February 2009. [28] Minhua Chen, Jorge Silva, John Paisley, Chunping Wang, David Dunson, and Lawrence Carin. Compressive sensing on manifolds using a nonparametric mixture of factor analyzers: Algorithm and performance bounds. IEEE Trans. Signal Processing, 58(12):6140–6155, December 2010. [29] Michael B. Wakin. A manifold lifting algorithm for multi-view compressive imaging. In Proc. 27th Conference on Picture Coding Symposium (PCS’09), pages 381–384, 2009. 9

5 0.45429659 250 nips-2011-Shallow vs. Deep Sum-Product Networks

Author: Olivier Delalleau, Yoshua Bengio

Abstract: We investigate the representational power of sum-product networks (computation networks analogous to neural networks, but whose individual units compute either products or weighted sums), through a theoretical analysis that compares deep (multiple hidden layers) vs. shallow (one hidden layer) architectures. We prove there exist families of functions that can be represented much more efficiently with a deep network than with a shallow one, i.e. with substantially fewer hidden units. Such results were not available until now, and contribute to motivate recent research involving learning of deep sum-product networks, and more generally motivate research in Deep Learning. 1 Introduction and prior work Many learning algorithms are based on searching a family of functions so as to identify one member of said family which minimizes a training criterion. The choice of this family of functions and how members of that family are parameterized can be a crucial one. Although there is no universally optimal choice of parameterization or family of functions (or “architecture”), as demonstrated by the no-free-lunch results [37], it may be the case that some architectures are appropriate (or inappropriate) for a large class of learning tasks and data distributions, such as those related to Artificial Intelligence (AI) tasks [4]. Different families of functions have different characteristics that can be appropriate or not depending on the learning task of interest. One of the characteristics that has spurred much interest and research in recent years is depth of the architecture. In the case of a multi-layer neural network, depth corresponds to the number of (hidden and output) layers. A fixedkernel Support Vector Machine is considered to have depth 2 [4] and boosted decision trees to have depth 3 [7]. Here we use the word circuit or network to talk about a directed acyclic graph, where each node is associated with some output value which can be computed based on the values associated with its predecessor nodes. The arguments of the learned function are set at the input nodes of the circuit (which have no predecessor) and the outputs of the function are read off the output nodes of the circuit. Different families of functions correspond to different circuits and allowed choices of computations in each node. Learning can be performed by changing the computation associated with a node, or rewiring the circuit (possibly changing the number of nodes). The depth of the circuit is the length of the longest path in the graph from an input node to an output node. Deep Learning algorithms [3] are tailored to learning circuits with variable depth, typically greater than depth 2. They are based on the idea of multiple levels of representation, with the intuition that the raw input can be represented at different levels of abstraction, with more abstract features of the input or more abstract explanatory factors represented by deeper circuits. These algorithms are often based on unsupervised learning, opening the door to semi-supervised learning and efficient 1 use of large quantities of unlabeled data [3]. Analogies with the structure of the cerebral cortex (in particular the visual cortex) [31] and similarities between features learned with some Deep Learning algorithms and those hypothesized in the visual cortex [17] further motivate investigations into deep architectures. It has been suggested that deep architectures are more powerful in the sense of being able to more efficiently represent highly-varying functions [4, 3]. In this paper, we measure “efficiency” in terms of the number of computational units in the network. An efficient representation is important mainly because: (i) it uses less memory and is faster to compute, and (ii) given a fixed amount of training samples and computational power, better generalization is expected. The first successful algorithms for training deep architectures appeared in 2006, with efficient training procedures for Deep Belief Networks [14] and deep auto-encoders [13, 27, 6], both exploiting the general idea of greedy layer-wise pre-training [6]. Since then, these ideas have been investigated further and applied in many settings, demonstrating state-of-the-art learning performance in object recognition [16, 28, 18, 15] and segmentation [20], audio classification [19, 10], natural language processing [9, 36, 21, 32], collaborative filtering [30], modeling textures [24], modeling motion [34, 33], information retrieval [29, 26], and semi-supervised learning [36, 22]. Poon and Domingos [25] introduced deep sum-product networks as a method to compute partition functions of tractable graphical models. These networks are analogous to traditional artificial neural networks but with nodes that compute either products or weighted sums of their inputs. Analogously to neural networks, we define “hidden” nodes as those nodes that are neither input nodes nor output nodes. If the nodes are organized in layers, we define the “hidden” layers to be those that are neither the input layer nor the output layer. Poon and Domingos [25] report experiments with networks much deeper (30+ hidden layers) than those typically used until now, e.g. in Deep Belief Networks [14, 3], where the number of hidden layers is usually on the order of three to five. Whether such deep architectures have theoretical advantages compared to so-called “shallow” architectures (i.e. those with a single hidden layer) remains an open question. After all, in the case of a sum-product network, the output value can always be written as a sum of products of input variables (possibly raised to some power by allowing multiple connections from the same input), and consequently it is easily rewritten as a shallow network with a sum output unit and product hidden units. The argument supported by our theoretical analysis is that a deep architecture is able to compute some functions much more efficiently than a shallow one. Until recently, very few theoretical results supported the idea that deep architectures could present an advantage in terms of representing some functions more efficiently. Most related results originate from the analysis of boolean circuits (see e.g. [2] for a review). Well-known results include the proof that solving the n-bit parity task with a depth-2 circuit requires an exponential number of gates [1, 38], and more generally that there exist functions computable with a polynomial-size depthk circuit that would require exponential size when restricted to depth k − 1 [11]. Another recent result on boolean circuits by Braverman [8] offers proof of a longstanding conjecture, showing that bounded-depth boolean circuits are unable to distinguish some (non-uniform) input distributions from the uniform distribution (i.e. they are “fooled” by such input distributions). In particular, Braverman’s result suggests that shallow circuits can in general be fooled more easily than deep ones, i.e., that they would have more difficulty efficiently representing high-order dependencies (those involving many input variables). It is not obvious that circuit complexity results (that typically consider only boolean or at least discrete nodes) are directly applicable in the context of typical machine learning algorithms such as neural networks (that compute continuous representations of their input). Orponen [23] surveys theoretical results in computational complexity that are relevant to learning algorithms. For instance, H˚ stad and Goldmann [12] extended some results to the case of networks of linear threshold units a with positivity constraints on the weights. Bengio et al. [5, 7] investigate, respectively, complexity issues in networks of Gaussian radial basis functions and decision trees, showing intrinsic limitations of these architectures e.g. on tasks similar to the parity problem. Utgoff and Stracuzzi [35] informally discuss the advantages of depth in boolean circuit in the context of learning architectures. Bengio [3] suggests that some polynomials could be represented more efficiently by deep sumproduct networks, but without providing any formal statement or proofs. This work partly addresses this void by demonstrating families of circuits for which a deep architecture can be exponentially more efficient than a shallow one in the context of real-valued polynomials. Note that we do not address in this paper the problem of learning these parameters: even if an efficient deep representation exists for the function we seek to approximate, in general there is no 2 guarantee for standard optimization algorithms to easily converge to this representation. This paper focuses on the representational power of deep sum-product circuits compared to shallow ones, and studies it by considering particular families of target functions (to be represented by the learner). We first formally define sum-product networks. We consider two families of functions represented by deep sum-product networks (families F and G). For each family, we establish a lower bound on the minimal number of hidden units a depth-2 sum-product network would require to represent a function of this family, showing it is much less efficient than the deep representation. 2 Sum-product networks Definition 1. A sum-product network is a network composed of units that either compute the product of their inputs or a weighted sum of their inputs (where weights are strictly positive). Here, we restrict our definition of the generic term “sum-product network” to networks whose summation units have positive incoming weights1 , while others are called “negative-weight” networks. Definition 2. A “negative-weight“ sum-product network may contain summation units whose weights are non-positive (i.e. less than or equal to zero). Finally, we formally define what we mean by deep vs. shallow networks in the rest of the paper. Definition 3. A “shallow“ sum-product network contains a single hidden layer (i.e. a total of three layers when counting the input and output layers, and a depth equal to two). Definition 4. A “deep“ sum-product network contains more than one hidden layer (i.e. a total of at least four layers, and a depth at least three). The family F 3 3.1 Definition The first family of functions we study, denoted by F, is made of functions built from deep sumproduct networks that alternate layers of product and sum units with two inputs each (details are provided below). The basic idea we use here is that composing layers (i.e. using a deep architecture) is equivalent to using a factorized representation of the polynomial function computed by the network. Such a factorized representation can be exponentially more compact than its expansion as a sum of products (which can be associated to a shallow network with product units in its hidden layer and a sum unit as output). This is what we formally show in what follows. + ℓ2 = λ11ℓ1 + µ11ℓ1 = x1x2 + x3x4 = f (x1, x2, x3, x4) 2 1 1 λ11 = 1 µ11 = 1 × ℓ1 = x1x2 1 x1 x2 × ℓ1 = x3x4 2 x3 x4 Figure 1: Sum-product network computing the function f ∈ F such that i = λ11 = µ11 = 1. Let n = 4i , with i a positive integer value. Denote by ℓ0 the input layer containing scalar variables {x1 , . . . , xn }, such that ℓ0 = xj for 1 ≤ j ≤ n. Now define f ∈ F as any function computed by a j sum-product network (deep for i ≥ 2) composed of alternating product and sum layers: • ℓ2k+1 = ℓ2k · ℓ2k for 0 ≤ k ≤ i − 1 and 1 ≤ j ≤ 22(i−k)−1 2j−1 2j j • ℓ2k = λjk ℓ2k−1 + µjk ℓ2k−1 for 1 ≤ k ≤ i and 1 ≤ j ≤ 22(i−k) j 2j 2j−1 where the weights λjk and µjk of the summation units are strictly positive. The output of the network is given by f (x1 , . . . , xn ) = ℓ2i ∈ R, the unique unit in the last layer. 1 The corresponding (shallow) network for i = 1 and additive weights set to one is shown in Figure 1 1 This condition is required by some of the proofs presented here. 3 (this architecture is also the basic building block of bigger networks for i > 1). Note that both the input size n = 4i and the network’s depth 2i increase with parameter i. 3.2 Theoretical results The main result of this section is presented below in Corollary 1, providing a lower bound on the minimum number of hidden units required by a shallow sum-product network to represent a function f ∈ F. The high-level proof sketch consists in the following steps: (1) Count the number of unique products found in the polynomial representation of f (Lemma 1 and Proposition 1). (2) Show that the only possible architecture for a shallow sum-product network to compute f is to have a hidden layer made of product units, with a sum unit as output (Lemmas 2 to 5). (3) Conclude that the number of hidden units must be at least the number of unique products computed in step 3.2 (Lemma 6 and Corollary 1). Lemma 1. Any element ℓk can be written as a (positively) weighted sum of products of input varij ables, such that each input variable xt is used in exactly one unit of ℓk . Moreover, the number mk of products found in the sum computed by ℓk does not depend on j and obeys the following recurrence j rule for k ≥ 0: if k + 1 is odd, then mk+1 = m2 , otherwise mk+1 = 2mk . k Proof. We prove the lemma by induction on k. It is obviously true for k = 0 since ℓ0 = xj . j Assuming this is true for some k ≥ 0, we consider two cases: k+1 k • If k + 1 is odd, then ℓj = ℓk 2j−1 · ℓ2j . By the inductive hypothesis, it is the product of two (positively) weighted sums of products of input variables, and no input variable can k appear in both ℓk 2j−1 and ℓ2j , so the result is also a (positively) weighted sum of products k of input variables. Additionally, if the number of products in ℓk 2j−1 and ℓ2j is mk , then 2 mk+1 = mk , since all products involved in the multiplication of the two units are different (since they use disjoint subsets of input variables), and the sums have positive weights. Finally, by the induction assumption, an input variable appears in exactly one unit of ℓk . This unit is an input to a single unit of ℓk+1 , that will thus be the only unit of ℓk+1 where this input variable appears. k • If k + 1 is even, then ℓk+1 = λjk ℓk 2j−1 + µjk ℓ2j . Again, from the induction assumption, it j must be a (positively) weighted sum of products of input variables, but with mk+1 = 2mk such products. As in the previous case, an input variable will appear in the single unit of ℓk+1 that has as input the single unit of ℓk in which this variable must appear. 2i Proposition 1. The number of products in the sum computed in the output unit l1 of a network √ n−1 . computing a function in F is m2i = 2 Proof. We first prove by induction on k ≥ 1 that for odd k, mk = 22 k 22 1+1 2 2 k+1 2 −2 , and for even k, . This is obviously true for k = 1 since 2 = 2 = 1, and all units in ℓ1 are mk = 2 single products of the form xr xs . Assuming this is true for some k ≥ 1, then: −1 0 −2 • if k + 1 is odd, then from Lemma 1 and the induction assumption, we have: mk+1 = m2 = k 2 k 22 2 −1 k +1 = 22 2 • if k + 1 is even, then instead we have: mk+1 = 2mk = 2 · 22 k+1 2 −2 −2 = 22 = 22 (k+1)+1 2 (k+1) 2 −2 −1 which shows the desired result for k + 1, and thus concludes the induction proof. Applying this result with k = 2i (which is even) yields 2i m2i = 22 2 −1 √ =2 4 22i −1 √ =2 n−1 . 2i Lemma 2. The products computed in the output unit l1 can be split in two groups, one with products containing only variables x1 , . . . , x n and one containing only variables x n +1 , . . . , xn . 2 2 Proof. This is obvious since the last unit is a “sum“ unit that adds two terms whose inputs are these two groups of variables (see e.g. Fig. 1). 2i Lemma 3. The products computed in the output unit l1 involve more than one input variable. k Proof. It is straightforward to show by induction on k ≥ 1 that the products computed by lj all involve more than one input variable, thus it is true in particular for the output layer (k = 2i). Lemma 4. Any shallow sum-product network computing f ∈ F must have a “sum” unit as output. Proof. By contradiction, suppose the output unit of such a shallow sum-product network is multiplicative. This unit must have more than one input, because in the case that it has only one input, the output would be either a (weighted) sum of input variables (which would violate Lemma 3), or a single product of input variables (which would violate Proposition 1), depending on the type (sum or product) of the single input hidden unit. Thus the last unit must compute a product of two or more hidden units. It can be re-written as a product of two factors, where each factor corresponds to either one hidden unit, or a product of multiple hidden units (it does not matter here which specific factorization is chosen among all possible ones). Regardless of the type (sum or product) of the hidden units involved, those two factors can thus be written as weighted sums of products of variables xt (with positive weights, and input variables potentially raised to powers above one). From Lemma 1, both x1 and xn must be present in the final output, and thus they must appear in at least one of these two factors. Without loss of generality, assume x1 appears in the first factor. Variables x n +1 , . . . , xn then cannot be present in the second factor, since otherwise one product in the output 2 would contain both x1 and one of these variables (this product cannot cancel out since weights must be positive), violating Lemma 2. But with a similar reasoning, since as a result xn must appear in the first factor, variables x1 , . . . , x n cannot be present in the second factor either. Consequently, no 2 input variable can be present in the second factor, leading to the desired contradiction. Lemma 5. Any shallow sum-product network computing f ∈ F must have only multiplicative units in its hidden layer. Proof. By contradiction, suppose there exists a “sum“ unit in the hidden layer, written s = t∈S αt xt with S the set of input indices appearing in this sum, and αt > 0 for all t ∈ S. Since according to Lemma 4 the output unit must also be a sum (and have positive weights according to Definition 1), then the final output will also contain terms of the form βt xt for t ∈ S, with βt > 0. This violates Lemma 3, establishing the contradiction. Lemma 6. Any shallow negative-weight sum-product network (see Definition 2) computing f ∈ F √ must have at least 2 n−1 hidden units, if its output unit is a sum and its hidden units are products. Proof. Such a network computes a weighted sum of its hidden units, where each hidden unit is a γ product of input variables, i.e. its output can be written as Σj wj Πt xt jt with wj ∈ R and γjt ∈ {0, 1}. In order to compute a function in F, this shallow network thus needs a number of hidden units at least equal to the number of unique products in that function. From Proposition 1, this √ number is equal to 2 n−1 . √ Corollary 1. Any shallow sum-product network computing f ∈ F must have at least 2 units. n−1 hidden Proof. This is a direct corollary of Lemmas 4 (showing the output unit is a sum), 5 (showing that hidden units are products), and 6 (showing the desired result for any shallow network with this specific structure – regardless of the sign of weights). 5 3.3 Discussion Corollary 1 above shows that in order to compute some function in F with n inputs, the number of √ √ units in a shallow network has to be at least 2 n−1 , (i.e. grows exponentially in n). On another hand, the total number of units in the deep (for i > 1) network computing the same function, as described in Section 3.1, is equal to 1 + 2 + 4 + 8 + . . . + 22i−1 (since all units are binary), which is √ also equal to 22i − 1 = n − 1 (i.e. grows only quadratically in n). It shows that some deep sumproduct network with n inputs and depth O(log n) can represent with O(n) units what would √ require O(2 n ) units for a depth-2 network. Lemma 6 also shows a similar result regardless of the sign of the weights in the summation units of the depth-2 network, but assumes a specific architecture for this network (products in the hidden layer with a sum as output). 4 The family G In this section we present similar results with a different family of functions, denoted by G. Compared to F, one important difference of deep sum-product networks built to define functions in G is that they can vary their input size independently of their depth. Their analysis thus provides additional insight when comparing the representational efficiency of deep vs. shallow sum-product networks in the case of a fixed dataset. 4.1 Definition Networks in family G also alternate sum and product layers, but their units have as inputs all units from the previous layer except one. More formally, define the family G = ∪n≥2,i≥0 Gin of functions represented by sum-product networks, where the sub-family Gin is made of all sum-product networks with n input variables and 2i + 2 layers (including the input layer ℓ0 ), such that: 1. ℓ1 contains summation units; further layers alternate multiplicative and summation units. 2. Summation units have positive weights. 3. All layers are of size n, except the last layer ℓ2i+1 that contains a single sum unit that sums all units in the previous layer ℓ2i . k−1 4. In each layer ℓk for 1 ≤ k ≤ 2i, each unit ℓk takes as inputs {ℓm |m = j}. j An example of a network belonging to G1,3 (i.e. with three layers and three input variables) is shown in Figure 2. ℓ3 = x2 + x2 + x2 + 3(x1x2 + x1x3 + x2x3) = g(x1, x2, x3) 3 2 1 1 + ℓ2 = x2 + x1x2 × 1 1 +x1x3 + x2x3 ℓ1 = x2 + x3 1 × ℓ2 = . . . 2 × ℓ2 = x2 + x1x2 3 3 +x1x3 + x2x3 + + ℓ1 = x1 + x3 2 + ℓ1 = x1 + x2 3 x1 x2 x3 Figure 2: Sum-product network computing a function of G1,3 (summation units’ weights are all 1’s). 4.2 Theoretical results The main result is stated in Proposition 3 below, establishing a lower bound on the number of hidden units of a shallow sum-product network computing g ∈ G. The proof sketch is as follows: 1. We show that the polynomial expansion of g must contain a large set of products (Proposition 2 and Corollary 2). 2. We use both the number of products in that set as well as their degree to establish the desired lower bound (Proposition 3). 6 We will also need the following lemma, which states that when n − 1 items each belong to n − 1 sets among a total of n sets, then we can associate to each item one of the sets it belongs to without using the same set for different items. Lemma 7. Let S1 , . . . , Sn be n sets (n ≥ 2) containing elements of {P1 , . . . , Pn−1 }, such that for any q, r, |{r|Pq ∈ Sr }| ≥ n − 1 (i.e. each element Pq belongs to at least n − 1 sets). Then there exist r1 , . . . , rn−1 different indices such that Pq ∈ Srq for 1 ≤ q ≤ n − 1. Proof. Omitted due to lack of space (very easy to prove by construction). Proposition 2. For any 0 ≤ j ≤ i, and any product of variables P = Πn xαt such that αt ∈ N and t=1 t j 2j whose computed value, when expanded as a weighted t αt = (n − 1) , there exists a unit in ℓ sum of products, contains P among these products. Proof. We prove this proposition by induction on j. First, for j = 0, this is obvious since any P of this form must be made of a single input variable xt , that appears in ℓ0 = xt . t Suppose now the proposition is true for some j < i. Consider a product P = Πn xαt such that t=1 t αt ∈ N and t αt = (n − 1)j+1 . P can be factored in n − 1 sub-products of degree (n − 1)j , β i.e. written P = P1 . . . Pn−1 with Pq = Πn xt qt , βqt ∈ N and t βqt = (n − 1)j for all q. By t=1 the induction hypothesis, each Pq can be found in at least one unit ℓ2j . As a result, by property 4 kq (in the definition of family G), each Pq will also appear in the additive layer ℓ2j+1 , in at least n − 1 different units (the only sum unit that may not contain Pq is the one that does not have ℓ2j as input). kq By Lemma 7, we can thus find a set of units ℓ2j+1 such that for any 1 ≤ q ≤ n − 1, the product rq Pq appears in ℓ2j+1 , with indices rq being different from each other. Let 1 ≤ s ≤ n be such that rq 2(j+1) s = rq for all q. Then, from property 4 of family G, the multiplicative unit ℓs computes the n−1 2j+1 product Πq=1 ℓrq , and as a result, when expanded as a sum of products, it contains in particular P1 . . . Pn−1 = P . The proposition is thus true for j + 1, and by induction, is true for all j ≤ i. Corollary 2. The output gin of a sum-product network in Gin , when expanded as a sum of products, contains all products of variables of the form Πn xαt such that αt ∈ N and t αt = (n − 1)i . t=1 t Proof. Applying Proposition 2 with j = i, we obtain that all products of this form can be found in the multiplicative units of ℓ2i . Since the output unit ℓ2i+1 computes a sum of these multiplicative 1 units (weighted with positive weights), those products are also present in the output. Proposition 3. A shallow negative-weight sum-product network computing gin ∈ Gin must have at least (n − 1)i hidden units. Proof. First suppose the output unit of the shallow network is a sum. Then it may be able to compute gin , assuming we allow multiplicative units in the hidden layer in the hidden layer to use powers of their inputs in the product they compute (which we allow here for the proof to be more generic). However, it will require at least as many of these units as the number of unique products that can be found in the expansion of gin . In particular, from Corollary 2, it will require at least the number n of unique tuples of the form (α1 , . . . , αn ) such that αt ∈ N and t=1 αt = (n − 1)i . Denoting ni dni = (n − 1)i , this number is known to be equal to n+dni −1 , and it is easy to verify it is higher d than (or equal to) dni for any n ≥ 2 and i ≥ 0. Now suppose the output unit is multiplicative. Then there can be no multiplicative hidden unit, otherwise it would mean one could factor some input variable xt in the computed function output: this is not possible since by Corollary 2, for any variable xt there exist products in the output function that do not involve xt . So all hidden units must be additive, and since the computed function contains products of degree dni , there must be at least dni such hidden units. 7 4.3 Discussion Proposition 3 shows that in order to compute the same function as gin ∈ Gin , the number of units in the shallow network has to grow exponentially in i, i.e. in the network’s depth (while the deep network’s size grows linearly in i). The shallow network also needs to grow polynomially in the number of input variables n (with a degree equal to i), while the deep network grows only linearly in n. It means that some deep sum-product network with n inputs and depth O(i) can represent with O(ni) units what would require O((n − 1)i ) units for a depth-2 network. Note that in the similar results found for family F, the depth-2 network computing the same function as a function in F had to be constrained to either have a specific combination of sum and hidden units (in Lemma 6) or to have non-negative weights (in Corollary 1). On the contrary, the result presented here for family G holds without requiring any of these assumptions. 5 Conclusion We compared a deep sum-product network and a shallow sum-product network representing the same function, taken from two families of functions F and G. For both families, we have shown that the number of units in the shallow network has to grow exponentially, compared to a linear growth in the deep network, so as to represent the same functions. The deep version thus offers a much more compact representation of the same functions. This work focuses on two specific families of functions: finding more general parameterization of functions leading to similar results would be an interesting topic for future research. Another open question is whether it is possible to represent such functions only approximately (e.g. up to an error bound ǫ) with a much smaller shallow network. Results by Braverman [8] on boolean circuits suggest that similar results as those presented in this paper may still hold, but this topic has yet to be formally investigated in the context of sum-product networks. A related problem is also to look into functions defined only on discrete input variables: our proofs do not trivially extend to this situation because we cannot assume anymore that two polynomials yielding the same output values must have the same expansion coefficients (since the number of input combinations becomes finite). Acknowledgments The authors would like to thank Razvan Pascanu and David Warde-Farley for their help in improving this manuscript, as well as the anonymous reviewers for their careful reviews. This work was partially funded by NSERC, CIFAR, and the Canada Research Chairs. References [1] Ajtai, M. (1983). P1 1 -formulae on finite structures. Annals of Pure and Applied Logic, 24(1), 1–48. [2] Allender, E. (1996). Circuit complexity before the dawn of the new millennium. In 16th Annual Conference on Foundations of Software Technology and Theoretical Computer Science, pages 1–18. Lecture Notes in Computer Science 1180, Springer Verlag. [3] Bengio, Y. (2009). Learning deep architectures for AI. Foundations and Trends in Machine Learning, 2(1), 1–127. Also published as a book. Now Publishers, 2009. [4] Bengio, Y. and LeCun, Y. (2007). Scaling learning algorithms towards AI. In L. Bottou, O. Chapelle, D. DeCoste, and J. Weston, editors, Large Scale Kernel Machines. MIT Press. [5] Bengio, Y., Delalleau, O., and Le Roux, N. (2006). The curse of highly variable functions for local kernel machines. In NIPS’05, pages 107–114. MIT Press, Cambridge, MA. [6] Bengio, Y., Lamblin, P., Popovici, D., and Larochelle, H. (2007). Greedy layer-wise training of deep networks. In NIPS 19, pages 153–160. MIT Press. [7] Bengio, Y., Delalleau, O., and Simard, C. (2010). Decision trees do not generalize to new variations. Computational Intelligence, 26(4), 449–467. [8] Braverman, M. (2011). Poly-logarithmic independence fools bounded-depth boolean circuits. Communications of the ACM, 54(4), 108–115. [9] Collobert, R. and Weston, J. (2008). A unified architecture for natural language processing: Deep neural networks with multitask learning. In ICML 2008, pages 160–167. [10] Dahl, G. E., Ranzato, M., Mohamed, A., and Hinton, G. E. (2010). Phone recognition with the meancovariance restricted boltzmann machine. In Advances in Neural Information Processing Systems (NIPS). 8 [11] H˚ stad, J. (1986). Almost optimal lower bounds for small depth circuits. In Proceedings of the 18th a annual ACM Symposium on Theory of Computing, pages 6–20, Berkeley, California. ACM Press. [12] H˚ stad, J. and Goldmann, M. (1991). On the power of small-depth threshold circuits. Computational a Complexity, 1, 113–129. [13] Hinton, G. E. and Salakhutdinov, R. (2006). Reducing the dimensionality of data with neural networks. Science, 313(5786), 504–507. [14] Hinton, G. E., Osindero, S., and Teh, Y. (2006). A fast learning algorithm for deep belief nets. Neural Computation, 18, 1527–1554. [15] Kavukcuoglu, K., Sermanet, P., Boureau, Y.-L., Gregor, K., Mathieu, M., and LeCun, Y. (2010). Learning convolutional feature hierarchies for visual recognition. In NIPS’10. [16] Larochelle, H., Erhan, D., Courville, A., Bergstra, J., and Bengio, Y. (2007). An empirical evaluation of deep architectures on problems with many factors of variation. In ICML’07, pages 473–480. ACM. [17] Lee, H., Ekanadham, C., and Ng, A. (2008). Sparse deep belief net model for visual area V2. In NIPS’07, pages 873–880. MIT Press, Cambridge, MA. [18] Lee, H., Grosse, R., Ranganath, R., and Ng, A. Y. (2009a). Convolutional deep belief networks for scalable unsupervised learning of hierarchical representations. In ICML 2009. Montreal (Qc), Canada. [19] Lee, H., Pham, P., Largman, Y., and Ng, A. (2009b). Unsupervised feature learning for audio classification using convolutional deep belief networks. In NIPS’09, pages 1096–1104. [20] Levner, I. (2008). Data Driven Object Segmentation. Ph.D. thesis, Department of Computer Science, University of Alberta. [21] Mnih, A. and Hinton, G. E. (2009). A scalable hierarchical distributed language model. In NIPS’08, pages 1081–1088. [22] Mobahi, H., Collobert, R., and Weston, J. (2009). Deep learning from temporal coherence in video. In ICML’2009, pages 737–744. [23] Orponen, P. (1994). Computational complexity of neural networks: a survey. Nordic Journal of Computing, 1(1), 94–110. [24] Osindero, S. and Hinton, G. E. (2008). Modeling image patches with a directed hierarchy of markov random field. In NIPS’07, pages 1121–1128, Cambridge, MA. MIT Press. [25] Poon, H. and Domingos, P. (2011). Sum-product networks: A new deep architecture. In UAI’2011, Barcelona, Spain. [26] Ranzato, M. and Szummer, M. (2008). Semi-supervised learning of compact document representations with deep networks. In ICML. [27] Ranzato, M., Poultney, C., Chopra, S., and LeCun, Y. (2007). Efficient learning of sparse representations with an energy-based model. In NIPS’06, pages 1137–1144. MIT Press. [28] Ranzato, M., Boureau, Y.-L., and LeCun, Y. (2008). Sparse feature learning for deep belief networks. In NIPS’07, pages 1185–1192, Cambridge, MA. MIT Press. [29] Salakhutdinov, R. and Hinton, G. E. (2007). Semantic hashing. In Proceedings of the 2007 Workshop on Information Retrieval and applications of Graphical Models (SIGIR 2007), Amsterdam. Elsevier. [30] Salakhutdinov, R., Mnih, A., and Hinton, G. E. (2007). Restricted Boltzmann machines for collaborative filtering. In ICML 2007, pages 791–798, New York, NY, USA. [31] Serre, T., Kreiman, G., Kouh, M., Cadieu, C., Knoblich, U., and Poggio, T. (2007). A quantitative theory of immediate visual recognition. Progress in Brain Research, Computational Neuroscience: Theoretical Insights into Brain Function, 165, 33–56. [32] Socher, R., Lin, C., Ng, A. Y., and Manning, C. (2011). Learning continuous phrase representations and syntactic parsing with recursive neural networks. In ICML’2011. [33] Taylor, G. and Hinton, G. (2009). Factored conditional restricted Boltzmann machines for modeling motion style. In ICML 2009, pages 1025–1032. [34] Taylor, G., Hinton, G. E., and Roweis, S. (2007). Modeling human motion using binary latent variables. In NIPS’06, pages 1345–1352. MIT Press, Cambridge, MA. [35] Utgoff, P. E. and Stracuzzi, D. J. (2002). Many-layered learning. Neural Computation, 14, 2497–2539. [36] Weston, J., Ratle, F., and Collobert, R. (2008). Deep learning via semi-supervised embedding. In ICML 2008, pages 1168–1175, New York, NY, USA. [37] Wolpert, D. H. (1996). The lack of a priori distinction between learning algorithms. Neural Computation, 8(7), 1341–1390. [38] Yao, A. (1985). Separating the polynomial-time hierarchy by oracles. In Proceedings of the 26th Annual IEEE Symposium on Foundations of Computer Science, pages 1–10. 9

6 0.44754589 22 nips-2011-Active Ranking using Pairwise Comparisons

7 0.44485557 92 nips-2011-Expressive Power and Approximation Errors of Restricted Boltzmann Machines

8 0.44357559 12 nips-2011-A Two-Stage Weighting Framework for Multi-Source Domain Adaptation

9 0.44259921 297 nips-2011-Universal low-rank matrix recovery from Pauli measurements

10 0.44009408 57 nips-2011-Comparative Analysis of Viterbi Training and Maximum Likelihood Estimation for HMMs

11 0.43936649 204 nips-2011-Online Learning: Stochastic, Constrained, and Smoothed Adversaries

12 0.43851602 229 nips-2011-Query-Aware MCMC

13 0.43838674 197 nips-2011-On Tracking The Partition Function

14 0.43804362 186 nips-2011-Noise Thresholds for Spectral Clustering

15 0.43578863 84 nips-2011-EigenNet: A Bayesian hybrid of generative and conditional models for sparse learning

16 0.43560943 159 nips-2011-Learning with the weighted trace-norm under arbitrary sampling distributions

17 0.43551761 17 nips-2011-Accelerated Adaptive Markov Chain for Partition Function Computation

18 0.43534389 96 nips-2011-Fast and Balanced: Efficient Label Tree Learning for Large Scale Object Recognition

19 0.43516526 102 nips-2011-Generalised Coupled Tensor Factorisation

20 0.43314531 206 nips-2011-Optimal Reinforcement Learning for Gaussian Systems