nips nips2010 nips2010-288 knowledge-graph by maker-knowledge-mining

288 nips-2010-Worst-case bounds on the quality of max-product fixed-points


Source: pdf

Author: Meritxell Vinyals, Jes\'us Cerquides, Alessandro Farinelli, Juan A. Rodríguez-aguilar

Abstract: We study worst-case bounds on the quality of any fixed point assignment of the max-product algorithm for Markov Random Fields (MRF). We start providing a bound independent of the MRF structure and parameters. Afterwards, we show how this bound can be improved for MRFs with specific structures such as bipartite graphs or grids. Our results provide interesting insight into the behavior of max-product. For example, we prove that max-product provides very good results (at least 90% optimal) on MRFs with large variable-disjoint cycles1 . 1

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 Worst-case bounds on the quality of max-product fixed-points ´ Jesus Cerquides Artificial Intelligence Research Institute (IIIA) Spanish Scientific Research Council (CSIC) Campus UAB, Bellaterra, Spain cerquide@iiia. [sent-1, score-0.232]

2 es Abstract We study worst-case bounds on the quality of any fixed point assignment of the max-product algorithm for Markov Random Fields (MRF). [sent-9, score-0.492]

3 Afterwards, we show how this bound can be improved for MRFs with specific structures such as bipartite graphs or grids. [sent-11, score-0.37]

4 Many of these practical problems can be formulated as finding the maximum a posteriori (MAP) assignment, namely the most likely joint variable assignment in an MRF. [sent-15, score-0.335]

5 For weighted b-matching problems with a bipartite structure Huang and Jebara [15] establish that max-product algorithm always converges to the optimal. [sent-22, score-0.214]

6 Despite these guarantees provided in these particular cases, for arbitrary MRFs little is known on the quality of the max-product fixed-point assignments. [sent-23, score-0.229]

7 This bound 1 MRFs in which all cycles are variable-disjoint, namely that they do not share any edge and in which each cycle contains at least 20 variables. [sent-26, score-0.519]

8 1 is calculated after running the max-sum algorithm and depends on the particular MRF (structure and parameters) and therefore provide no guarantees on the quality of max-product assignments on arbitrary MRFs with cycles. [sent-27, score-0.337]

9 In this paper we provide quality guarantees for max-product fixed-points in general settings that can be calculated prior to the execution of the algorithm. [sent-28, score-0.183]

10 To this end, we define worst-case bounds on the quality of any max-product fixed-point for any MRF, independently of its structure and parameters. [sent-29, score-0.265]

11 For example, we prove that in 2-D grids max-product fixed points assignments have at least 33% of the quality of the optimum; and that for MRFs with large variable-disjoint cycles1 they have at least 90% of the quality of the optimum. [sent-31, score-0.449]

12 These results shed some light on the relationship between the quality of max-product assignments and the structure of MRFs. [sent-32, score-0.268]

13 Our results build upon two main components: (i) the characterization of any fixed-point max-product assignment as a neighbourhood maximum in a specific region of the MRF [17]; and (ii) the worstcase bounds on the quality of a neighbourhood maximum obtained in the K-optimality framework [18, 19]. [sent-33, score-0.892]

14 We combine these two results by: (i) generalising the worst-case bounds in [18, 19] to consider any arbitrary region; and (ii) assessing worst-case bounds for the specific region presented in [17] (for which any fixed-point max-product assignment is known to be maximal). [sent-34, score-0.63]

15 In this work we assume that: (i) there is a unique MAP assignment (as assumed in [17]); and (ii) all potentials θs and θst are non-negative. [sent-42, score-0.341]

16 The max-product algorithm is an iterative, local, message-passing algorithm for finding the MAP assignment in a discrete MRF as specified by equation 2. [sent-43, score-0.329]

17 The standard update rules for max-sum algorithm are:   mij (xj ) = αij + max θi (xi ) + θij (xi , xj ) + xi mki (xi ) k∈N (i)\j bi (xi ) = θi (xi ) + mki (xi ) k∈N (i) where αij is a normalization constant and N (i) is the set of indices for variables that are connected to xi . [sent-45, score-0.256]

18 At each following iteration, each variable xi aggregates all incoming messages and computes the belief bi (xi ), which is then used to obtain the maxsum assignment xM S . [sent-48, score-0.389]

19 i 2 x0 x1 x3 x2 (a) x0 x1 x2 x3 (b) x0 x1 x2 x3 (c) x0 x1 x2 x0 x3 x2 (d) x1 x3 (e) Figure 1: (a) 4-complete graph and (b)-(e) sets of variables covered by the SLT-region. [sent-50, score-0.322]

20 In particular, they find the conditions for a fixed-point max-sum assignment xM S to be neighbourhood maximum, namely greater than all other assignments in a specific large region around xM S . [sent-60, score-0.626]

21 Notice that characterising an assignment as neighbourhood maximum is weaker than a global maximum, but stronger than a local maximum. [sent-61, score-0.403]

22 introduce the notion of Single Loops and Trees (SLT) region to characterise the assignments in such region. [sent-63, score-0.261]

23 Hence, we say that an assignment xSLT is SLT-optimal if it is greater than any other assignment in its SLT region. [sent-66, score-0.52]

24 Figures 1(b)-(e) illustrate examples of assignments in the SLTregion in the complete graph of figure 1(a), here boldfaced nodes stand for variables that vary the assignment with respect to xSLT . [sent-68, score-0.567]

25 introduced worst-case bounds on the quality of a neighbourhood maximum in a region characterized by its size. [sent-70, score-0.489]

26 In this section we generalize these bounds to use them for any neighbourhood maximum in a region characterized by arbitrary criteria. [sent-73, score-0.408]

27 1 C-optimal bounds Hereafter we propose a general notion of region optimality, the so-called C-optimality, and describe how to calculate bounds for a C-optimal assignment, namely an assignment that is neighbourhood maximum in a region characterized by an arbitrary C criteria. [sent-76, score-0.925]

28 A region C ⊂ P(V ) is a set composed by subsets of V . [sent-81, score-0.163]

29 We say that A ⊆ V is covered by C if there is a C α ∈ C such that C α completely covers A. [sent-82, score-0.242]

30 Given two assignments xA and xB , we define D(xA , xB ) as the set containing the variables whose values in xA and xB differ. [sent-83, score-0.19]

31 An assignment is C-optimal if it cannot be improved by changing the values in any group of variables covered by C. [sent-84, score-0.499]

32 That is, an assignment xA is C-optimal if for every assignment xB s. [sent-85, score-0.549]

33 D(xA , xB ) is covered by C we have that θ(xA ) ≥ θ(xB ). [sent-87, score-0.157]

34 For every C α ∈ C, consider an assignment xα such that xα = xC if xi ∈ C α and xα = x∗ if xi ∈ C α . [sent-97, score-0.363]

35 (4) C α ∈C Notice that although θ(xα ) is defined as the sum of unary potentials and pairwise potentials values we can always get rid of unary potentials by combining them into pairwise potentials without changing the structure of the MRF. [sent-99, score-0.521]

36 We classify each edge S ∈ E into one of three disjoint groups, depending on whether C α covers S completely (T (C α )), partially (P (C α )), or not at all (N (C α )), so that θ(xα ) = S∈T (C α ) θS (xα ) + α α S∈P (C α ) θS (x ) + S∈N (C α ) θS (x ). [sent-101, score-0.156]

37 We can remove the partially covered potentials at the cost of obtaining a looser bound. [sent-102, score-0.238]

38 Now, by definition of xα , for every variable xi in a potential completely covered by C α we have that xα = x∗ , and for every variable xi in a potential not covered at all by C α we have that xα = xC . [sent-104, score-0.487]

39 2 Size-optimal bounds as a specific case of C-optimal bounds Now we present the main result in [18] as a specific case of C-optimality. [sent-113, score-0.21]

40 An assignment is k-sizeoptimal if it can not be improved by changing the value of any group of size k or fewer variables. [sent-114, score-0.26]

41 For any MRF and for any k-optimal assignment xk : θ(xk ) ≥ (k − 1) θ(x∗ ) (2|V | − k − 1) (6) Proof. [sent-116, score-0.26]

42 The k number of elements in C that completely cover S is cc(S, C) = |V |−2 (take the two variables in S k−2 plus k − 2 variables out of the remaining |V | − 2). [sent-119, score-0.292]

43 The number of elements in C that do not cover |−2 S at all is nc(S, C) = |V k (take k variables out of the remaining |V | − 2 variables). [sent-120, score-0.169]

44 d=2 d=4 d=8 d=128 d=1024 50 10 20 30 40 50 Minimum number of variables in each cycle (b) Bounds on MRFs with variable-disjoint cycles when varying the number of cycles and their size. [sent-123, score-0.563]

45 Figure 2: Percent optimal bounds for max-sum fixed point assignments in specific MRF structures. [sent-124, score-0.213]

46 4 Quality guarantees on max-sum fixed-point assignments In this section we define quality guarantees for max-sum fixed-point assignments in MRFs with arbitary and specific structures. [sent-125, score-0.455]

47 Our quality guarantees prove that the value of any max-sum fixedpoint assignments can not be less than a fraction of the optimum. [sent-126, score-0.291]

48 The main idea is that by virtue of the characterization of any max-sum fixed point assignment as SLT-optimal, we can select any region C composed of a combination of single cycles and trees of our graph and use it for computing its corresponding C-optimal bound by means of proposition 1. [sent-127, score-0.836]

49 We start by proving that bounds for a given graph apply to its subgraphs. [sent-128, score-0.188]

50 Then, we find that the bound for the complete graph applies to any MRF independently of its structure and parameters. [sent-129, score-0.247]

51 1 C-optimal bounds based on the SLT region In this section we show that C-optimal bounds based on SLT-optimality for a given graph can be applied to any of its subgraphs. [sent-132, score-0.407]

52 Then the bound of equation 3 for G holds for any SLT-optimal assignment in G . [sent-136, score-0.426]

53 We can compose a region C containing the same elements as C but removing those variables which are not contained in V . [sent-138, score-0.229]

54 Observe that the bound obtained by applying equation 3 to C is greater or equal than the bound obtained for C. [sent-140, score-0.263]

55 A direct conclusion of proposition 3 is that any bound based on the SLT-region of a complete graph of n variables can be directly applied to any subgraph of n or fewer variables regardless of its structure. [sent-142, score-0.501]

56 In what follows we assess the bound for a complete graph. [sent-143, score-0.162]

57 For any max-sum fixed point assignment xM S , 1 · θ(x∗ ). [sent-146, score-0.26]

58 Let C be a region containing every possible combination of three variables in V . [sent-148, score-0.225]

59 For any MRF, any max-sum fixed point assignment xM S satisfies equation 7. [sent-152, score-0.329]

60 Since any graph can be seen as a subgraph of the complete graph with the same number of variables, the corollary is straightforward given propositions 3 and 4. [sent-153, score-0.258]

61 2 SLT-bounds for specific MRF structures and independent of the MRF parameters In this section we show that for MRFs with specific structures, it is possible to provide bounds much tighter than the structure-independent bound provided by corollary 5. [sent-159, score-0.281]

62 These structures include, but are not limited to, bipartite graphs, 2-D grids, and variable-disjoint cycle graphs. [sent-160, score-0.373]

63 1 Bipartite graphs In this section we define the C-optimal bound of equation 3 for any max-sum fixed point assignment in an n-m bipartite MRF. [sent-163, score-0.652]

64 An n-m bipartite MRF is a graph whose vertices can be divided into two disjoint sets, one with n variables and another one with m variables, such that the n variables in the first set are connected to the m variables in the second set. [sent-164, score-0.51]

65 For any MRF with n-m bipartite structure where m ≥ n, and for any max-sum fixed point assignment xM S we have that: 1 m≥n+3 θ(xM S ) ≥ b(n, m) · θ(x∗ ) b(n, m) = n 2 (8) m 1 n when m < n + 3, and so equation 8 holds (details can be found in the additional material). [sent-167, score-0.543]

66 Figures 3(b)-(j) show the elements in the region C B composed of sets of four variables, two from each side. [sent-170, score-0.196]

67 For example, the only graph that does not include neither x0 nor x3 is the graph of figure 3(j). [sent-175, score-0.166]

68 Figure 2(a) plots the bound of equation 8 for bipartite graphs when varying the number of variables. [sent-177, score-0.392]

69 Note that although, also in this case, the value of the bound rapidly decreases with the number of variables, it is two times the values of the structure-independent bound (see equation 7). [sent-178, score-0.294]

70 2 Two-dimensional (2-D) grids In this section we define the C-optimal bound of equation 3 for any max-sum fixed point assignment in a two-dimensional grid MRF. [sent-181, score-0.513]

71 For any MRF with an n grid structure where n is an even number, for any max-sum fixed point assignment xM S we have that n θ(xM S ) ≥ · θ(x∗ ) (9) 3n − 4 Proof. [sent-186, score-0.293]

72 Each edge is completely covered by n elements and hence 2 2 cc∗ = n . [sent-192, score-0.302]

73 Finally2 , for each edge S, there are nc∗ = ( n − 1)( n − 2) elements of C that do not cover 2 2 2 S at all. [sent-193, score-0.158]

74 Figures 4 (b)-(e) show the vertex-induced subgraphs for each set of vertices in the region C formed by the combination of any pairs of rows in {(1, 3), (2, 4)} and pair of columns in {(1, 3), (2, 4)}. [sent-197, score-0.188]

75 For example, the edge that links the two first variables in the first row, namely x0 and x1 , is included in the subgraphs of figures (a) and (b). [sent-200, score-0.234]

76 Figure 2(a) plots the bound for 2-D grids when varying the number of variables. [sent-203, score-0.184]

77 Note that when compared with the bound for complete and bipartite structures, the bound for 2-D grids decreases smoothly and tends to stabilize as the number of variables increases. [sent-204, score-0.609]

78 In fact, observe that by equation 9, the bound for 2-D grids is never less that 1/3 independently of the grid size. [sent-205, score-0.253]

79 3 MRFs that are a union of variable-disjoint cycles In this section we assess a bound for MRFs composed of a set of variable-disjoint cycles, namely of cycles that do not share any variable. [sent-208, score-0.551]

80 A common pattern shared by the bounds assessed so far is that they decrease as the number of variables of an MRF grows. [sent-209, score-0.187]

81 Consider the MRF composed of two variable-disjoint cycles of size 4 depicted in figure 5(a). [sent-212, score-0.217]

82 Just by counting we observe that each edge is completely covered 6 times, so cc∗ = 6. [sent-216, score-0.269]

83 The following result generalizes the previous example to MRFs containing d variable-disjoint cycles of size larger or equal to l. [sent-219, score-0.168]

84 The proof generalizes the region explained in example 3 to any variable-disjoint cycle MRF by defining a region that includes an element for every possible edge removal from every cycle but one. [sent-223, score-0.647]

85 Equation 10 shows that the bound: (i) decreases with the number of cycles; and (ii) increases as the maximum number of variables in each cycle grows. [sent-225, score-0.295]

86 Figure 2(b) illustrates the relationship between the bound, the number of cycles (d), and the maximum size of the cycles (l). [sent-226, score-0.373]

87 The first thing we observe is that the size of the cycles has more impact on the bound than the number of cycles. [sent-227, score-0.265]

88 In fact, observe that by equation 10, the bound for a variable-disjoint cycle graph with a maximum cycle size of l is at least (l−2) , independently of the number of cycles. [sent-228, score-0.576]

89 Thus, if the minimum size of l a cycle is 20, the quality for a fixed point is guaranteed to be at least 90%. [sent-229, score-0.272]

90 Hence, quality guarantees for max-sum fixed points are good whenever: (i) the cycles in the MRF do not share any variables; and (ii) the smallest cycle in the MRF is large. [sent-230, score-0.496]

91 3 SLT-bounds for arbitrary MRF structures and independent of the MRF parameters In this section we discuss how to assess tight SLT-bounds for any arbitrary MRF structure. [sent-233, score-0.17]

92 Similarly to [18, 20], we can use linear fractional programming (LFP) to compute the structure specific SLT bounds in any MRF with arbitrary structure. [sent-234, score-0.184]

93 Let C be a region for all subsets in the SLT region of the graphical model G = V, E of an MRF. [sent-235, score-0.228]

94 For each S ∈ E, the LFP contains two LFP variables that represents the value of the edge S for the SLT-optimum, xM S , and for the MAP assignment, x∗ . [sent-236, score-0.153]

95 5 Conclusions We provided worst-case bounds on the quality of any max-product fixed point. [sent-244, score-0.232]

96 With this aim, we have introduced C-optimality, which has proven a valuable tool to bound the quality of max-product fixed points. [sent-245, score-0.224]

97 Concretely, we have proven that independently of an MRF structure, max-product has a quality guarantee that decreases with the number of variables of the MRF. [sent-246, score-0.24]

98 Furthermore, our results allow to identify new classes of MRF structures, besides acyclic and single-cycle, for which we can provide theoretical guarantees on the quality of max-product assignments. [sent-247, score-0.234]

99 As an example, we defined significant bounds for 2-D grids and MRFs with variable-disjoint cycles. [sent-248, score-0.192]

100 Comparison of graph cuts with belief propagation for stereo, using identical mrf parameters. [sent-254, score-0.603]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('mrf', 0.433), ('assignment', 0.26), ('mrfs', 0.254), ('xc', 0.25), ('cc', 0.203), ('nc', 0.203), ('bipartite', 0.181), ('xm', 0.178), ('xs', 0.173), ('cycles', 0.168), ('covered', 0.157), ('cycle', 0.145), ('lfp', 0.136), ('quality', 0.127), ('slt', 0.116), ('region', 0.114), ('assignments', 0.108), ('neighbourhood', 0.106), ('bounds', 0.105), ('bound', 0.097), ('grids', 0.087), ('graph', 0.083), ('variables', 0.082), ('potentials', 0.081), ('xb', 0.079), ('edge', 0.071), ('xa', 0.071), ('equation', 0.069), ('spain', 0.067), ('bellaterra', 0.066), ('csic', 0.066), ('iiia', 0.066), ('spanish', 0.066), ('uab', 0.066), ('proposition', 0.065), ('subgraph', 0.058), ('mins', 0.058), ('guarantees', 0.056), ('cover', 0.054), ('campus', 0.053), ('st', 0.053), ('acyclic', 0.051), ('unary', 0.05), ('messages', 0.05), ('composed', 0.049), ('map', 0.048), ('structures', 0.047), ('arbitrary', 0.046), ('graphs', 0.045), ('xt', 0.045), ('propagation', 0.045), ('covers', 0.044), ('brendan', 0.044), ('farinelli', 0.044), ('kiekintveld', 0.044), ('koetter', 0.044), ('mceliece', 0.044), ('meritxell', 0.044), ('milind', 0.044), ('srinivas', 0.044), ('vinyals', 0.044), ('xslt', 0.044), ('subgraphs', 0.043), ('loopy', 0.043), ('belief', 0.042), ('concretely', 0.042), ('completely', 0.041), ('yair', 0.041), ('characterisation', 0.039), ('characterise', 0.039), ('pearce', 0.039), ('aji', 0.039), ('namely', 0.038), ('maximum', 0.037), ('xi', 0.037), ('sujay', 0.036), ('ralf', 0.036), ('mki', 0.036), ('alan', 0.035), ('council', 0.035), ('percent', 0.035), ('complete', 0.034), ('fields', 0.034), ('structure', 0.033), ('sanghavi', 0.033), ('aamas', 0.033), ('elements', 0.033), ('ii', 0.033), ('pairwise', 0.032), ('tighter', 0.032), ('afterwards', 0.032), ('assess', 0.031), ('decreases', 0.031), ('rows', 0.031), ('gures', 0.031), ('frey', 0.03), ('optimality', 0.029), ('every', 0.029), ('coordination', 0.029), ('mij', 0.028)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.99999976 288 nips-2010-Worst-case bounds on the quality of max-product fixed-points

Author: Meritxell Vinyals, Jes\'us Cerquides, Alessandro Farinelli, Juan A. Rodríguez-aguilar

Abstract: We study worst-case bounds on the quality of any fixed point assignment of the max-product algorithm for Markov Random Fields (MRF). We start providing a bound independent of the MRF structure and parameters. Afterwards, we show how this bound can be improved for MRFs with specific structures such as bipartite graphs or grids. Our results provide interesting insight into the behavior of max-product. For example, we prove that max-product provides very good results (at least 90% optimal) on MRFs with large variable-disjoint cycles1 . 1

2 0.16814907 169 nips-2010-More data means less inference: A pseudo-max approach to structured learning

Author: David Sontag, Ofer Meshi, Amir Globerson, Tommi S. Jaakkola

Abstract: The problem of learning to predict structured labels is of key importance in many applications. However, for general graph structure both learning and inference are intractable. Here we show that it is possible to circumvent this difficulty when the distribution of training examples is rich enough, via a method similar in spirit to pseudo-likelihood. We show that our new method achieves consistency, and illustrate empirically that it indeed approaches the performance of exact methods when sufficiently large training sets are used. Many prediction problems in machine learning applications are structured prediction tasks. For example, in protein folding we are given a protein sequence and the goal is to predict the protein’s native structure [14]. In parsing for natural language processing, we are given a sentence and the goal is to predict the most likely parse tree [2]. In these and many other applications, we can formalize the structured prediction problem as taking an input x (e.g., primary sequence, sentence) and predicting ˆ y (e.g., structure, parse) according to y = arg maxy∈Y θ · φ(x, y ), where φ(x, y) is a function that ˆ maps any input and a candidate assignment to a feature vector, Y denotes the space of all possible assignments to the vector y, and θ is a weight vector to be learned. This paper addresses the problem of learning structured prediction models from data. In particular, given a set of labeled examples {(xm , y m )}M , our goal is to find a vector θ such that for each m=1 example m, y m = arg maxy∈Y θ · φ(xm , y), i.e. one which separates the training data. For many structured prediction models, maximization over Y is computationally intractable. This makes it difficult to apply previous algorithms for learning structured prediction models, such as structured perceptron [2], stochastic subgradient [10], and cutting-plane algorithms [5], which require making a prediction at every iteration (equivalent to repeatedly solving an integer linear program). Given training data, we can consider the space of parameters Θ that separate the data. This space can be defined by the intersection of a large number of linear inequalities. A recent approach to getting around the hardness of prediction is to use linear programming (LP) relaxations to approximate the maximization over Y [4, 6, 9]. However, separation with respect to a relaxation places stronger constraints on the parameters. The target solution, an integral vertex in the LP, must now distinguish itself also from possible fractional vertexes that arise due to the relaxation. The relaxations can therefore be understood as optimizing over an inner bound of Θ. This set may be empty even if the training data is separable with exact inference [6]. Another obstacle to using LP relaxations for learning is that solving the LPs can be very slow. In this paper we ask whether it is possible to learn while avoiding inference altogether. We propose a new learning algorithm, inspired by pseudo-likelihood [1], that optimizes over an outer bound of Θ. Learning involves optimizing over only a small number of constraints per data point, and thus can be performed quickly, even for complex structured prediction models. We show that, if the data available for learning is “nice”, this algorithm is consistent, i.e. it will find some θ ∈ Θ. This is an example of how having the right data can circumvent the hardness of learning for structured prediction. 1 We also investigate the limitations of the proposed method. We show that the problem of even deciding whether a given data set is separable is NP-hard, and thus learning in a strict sense is no easier than prediction. Thus, we should not expect for our algorithm, or any other polynomial time algorithm, to always succeed at learning from an arbitrary finite data set. To our knowledge, this is the first result characterizing the hardness of exact learning for structured prediction. Finally, we show empirically that our algorithm allows us to successfully learn the parameters for both multi-label prediction and protein side-chain placement. The performance of the algorithm is improved as more data becomes available, as our theoretical results anticipate. 1 Pseudo-Max method We consider the general structured prediction problem. The input space is denoted by X and the set of all possible assignments by Y. Each y ∈ Y corresponds to n variables y1 , . . . , yn , each with k possible states. The classifier uses a (given) function φ(x, y) : X , Y → Rd and (learned) weights θ ∈ Rd , and is defined as y(x; θ) = arg maxy∈Y f (ˆ ; x, θ) where f is the discriminant function y ˆ f (y; x, θ) = θ · φ(x, y). Our analysis will focus on functions φ whose scope is limited to small sets of the yi variables, but for now we keep the discussion general. Given a set of labeled examples {(xm , y m )}M , the goal of the typical learning problem is to find m=1 weights θ that correctly classify the training examples. Consider first the separable case. Define the set of separating weight vectors, Θ = θ | ∀m, y ∈ Y, f (y m ; xm , θ) ≥ f (y; xm , θ)+e(y, y m ) . e is a loss function (e.g., zero-one or Hamming) such that e(y m , y m ) = 0 and e(y, y m ) > 0 for y = y m , which serves to rule out the trivial solution θ = 0.1 The space Θ is defined by exponentially many constraints per example, one for each competing assignment. In this work we consider a much simpler set of constraints where, for each example, we only consider the competing assignments obtained by modifying a single label yi , while fixing the other labels to their value at y m . The pseudo-max set, which is an outer bound on Θ, is given by Here ym −i m Θps = θ | ∀m, i, yi , f (y m ; xm , θ) ≥ f (y m , yi ; xm , θ) + e(yi , yi ) . −i denotes the label y m (1) without the assignment to yi . When the data is not separable, Θ will be the empty set. Instead, we may choose to minimize the hinge loss, (θ) = m maxy f (y; xm , θ) − f (y m ; xm , θ) + e(y, y m ) , which can be shown to be an upper bound on the training error [13]. When the data is separable, minθ (θ) = 0. Note that regularization may be added to this objective. The corresponding pseudo-max objective replaces the maximization over all of y with maximization over a single variable yi while fixing the other labels to their value at y m :2,3 M ps (θ) n = m=1 i=1 m max f (y m , yi ; xm , θ) − f (y m ; xm , θ) + e(yi , yi ) . −i yi Analogous to before, we have minθ ps (θ) (2) = 0 if and only if θ ∈ Θps . The objective in Eq. 2 is similar in spirit to pseudo-likelihood objectives used for maximum likelihood estimation of parameters of Markov random fields (MRFs) [1]. The pseudo-likelihood estimate is provably consistent when the data generating distribution is a MRF of the same structure as used in the pseudo-likelihood objective. However, our setting is different since we only get to view the maximizing assignment of the MRF rather than samples from it. Thus, a particular x will always be paired with the same y rather than samples y drawn from the conditional distribution p(y|x; θ). The pseudo-max constraints in Eq. 1 are also related to cutting plane approaches to inference [4, 5]. In the latter, the learning problem is solved by repeatedly looking for assignments that violate the separability constraint (or its hinge version). Our constraints can be viewed as using a very small 1 An alternative formulation, which we use in the next section, is to break the symmetry by having part of the input not be multiplied by any weight. This will also rule out the trivial solution θ = 0. P 2 It is possible to use maxi instead of i , and some of our consistency results will still hold. 3 The pseudo-max approach is markedly different from a learning method which predicts each label yi independently, since the objective considers all i simultaneously (both at learning and test time). 2 x2 0.2 J ∗ + x1 = 0 y = (0, 1) y = (1, 1) g(J12) x2 = 0 x1 J ∗ + x1 + x2 = 0 y = (0, 0) c1=0 c1=1 c1= 1 0.15 0.1 J + x2 = 0 ∗ 0.05 y = (1, 0) x1 = 0 0 1 0.5 0 J 0.5 1 Figure 1: Illustrations for a model with two variables. Left: Partitioning of X induced by configurations y(x) for some J ∗ > 0. Blue lines carve out the exact regions. Red lines denote the pseudo-max constraints that hold with equality. Pseudo-max does not obtain the diagonal constraint coming from comparing configurations y = (1, 1) and (0, 0), since these differ by more than one coordinate. Right: One strictly-convex component of the ps (J ) function (see Eq. 9). The function is shown for different values of c1 , the mean of the x1 variable. subset of assignments for the set of candidate constraint violators. We also note that when exact maximization over the discriminant function f (y; x, θ) is hard, the standard cutting plane algorithm cannot be employed since it is infeasible to find a violated constraint. For the pseudo-max objective, finding a constraint violation is simple and linear in the number of variables.4 It is easy to see (as will be elaborated on next) that the pseudo-max method does not in general yield a consistent estimate of θ, even in the separable case. However, as we show, consistency can be shown to be achieved under particular assumptions on the data generating distribution p(x). 2 Consistency of the Pseudo-Max method In this section we show that if the feature generating distribution p(x) satisfies particular assumptions, then the pseudo-max approach yields a consistent estimate. In other words, if the training data is of the form {(xm , y(xm ; θ ∗ ))}M for some true parameter vector θ ∗ , then as M → ∞ the m=1 minimum of the pseudo-max objective will converge to θ ∗ (up to equivalence transformations). The section is organized as follows. First, we provide intuition for the consistency results by considering a model with only two variables. Then, in Sec. 2.1, we show that any parameter θ ∗ can be identified to within arbitrary accuracy by choosing a particular training set (i.e., choice of xm ). This in itself proves consistency, as long as there is a non-zero probability of sampling this set. In Sec. 2.2 we give a more direct proof of consistency by using strict convexity arguments. For ease of presentation, we shall work with a simplified instance of the structured learning setting. We focus on binary variables, yi ∈ {0, 1}, and consider discriminant functions corresponding to Ising models, a special case of pairwise MRFs (J denotes the vector of “interaction” parameters): f (y; x, J ) = ij∈E Jij yi yj + i yi xi (3) The singleton potential for variable yi is yi xi and is not dependent on the model parameters. We could have instead used Ji yi xi , which would be more standard. However, this would make the parameter vector J invariant to scaling, complicating the identifiability analysis. In the consistency analysis we will assume that the data is generated using a true parameter vector J ∗ . We will show that as the data size goes to infinity, minimization of ps (J ) yields J ∗ . We begin with an illustrative analysis of the pseudo-max constraints for a model with only two variables, i.e. f (y; x, J) = Jy1 y2 + y1 x1 + y2 x2 . The purpose of the analysis is to demonstrate general principles for when pseudo-max constraints may succeed or fail. Assume that training samples are generated via y(x) = argmaxy f (y; x, J ∗ ). We can partition the input space X into four regions, ˆ ˆ {x ∈ X : y(x) = y } for each of the four configurations y , shown in Fig. 1 (left). The blue lines outline the exact decision boundaries of f (y; x, J ∗ ), with the lines being given by the constraints 4 The methods differ substantially in the non-separable setting where we minimize ps (θ), using a slack variable for every node and example, rather than just one slack variable per example as in (θ). 3 in Θ that hold with equality. The red lines denote the pseudo-max constraints in Θps that hold with equality. For x such that y(x) = (1, 0) or (0, 1), the pseudo-max and exact constraints are identical. We can identify J ∗ by obtaining samples x = (x1 , x2 ) that explore both sides of one of the decision boundaries that depends on J ∗ . The pseudo-max constraints will fail to identify J ∗ if the samples do not sufficiently explore the transitions between y = (0, 1) and y = (1, 1) or between y = (1, 0) and y = (1, 1). This can happen, for example, when the input samples are dependent, giving only rise to the configurations y = (0, 0) and y = (1, 1). For points labeled (1, 1) around the decision line J ∗ + x1 + x2 = 0, pseudo-max can only tell that they respect J ∗ + x1 ≥ 0 and J ∗ + x2 ≥ 0 (dashed red lines), or x1 ≤ 0 and x2 ≤ 0 for points labeled (0, 0). Only constraints that depend on the parameter are effective for learning. For pseudo-max to be able to identify J ∗ , the input samples must be continuous, densely populating the two parameter dependent decision lines that pseudo-max can use. The two point sets in the figure illustrate good and bad input distributions for pseudo-max. The diagonal set would work well with the exact constraints but badly with pseudo-max, and the difference can be arbitrarily large. However, the input distribution on the right, populating the J ∗ + x2 = 0 decision line, would permit pseudo-max to identify J ∗ . 2.1 Identifiability of True Parameters In this section, we show that it is possible to approximately identify the true model parameters, up to model equivalence, using the pseudo-max constraints and a carefully chosen linear number of data points. Consider the learning problem for structured prediction defined on a fixed graph G = (V, E) where the parameters to be learned are pairwise potential functions θij (yi , yj ) for ij ∈ E and single node fields θi (yi ) for i ∈ V . We consider discriminant functions of the form f (y; x, θ) = ij∈E θij (yi , yj ) + i θi (yi ) + i xi (yi ), (4) where the input space X = R|V |k specifies the single node potentials. Without loss of generality, we remove the additional degrees of freedom in θ by restricting it to be in a canonical form: θ ∈ Θcan if for all edges θij (yi , yj ) = 0 whenever yi = 0 or yj = 0, and if for all nodes, θi (yi ) = 0 when yi = 0. As a result, assuming the training set comes from a model in this class, and the input fields xi (yi ) exercise the discriminant function appropriately, we can hope to identify θ ∗ ∈ Θcan . Indeed, we show that, for some data sets, the pseudo-max constraints are sufficient to identify θ ∗ . Let Θps ({y m , xm }) be the set of parameters that satisfy the pseudo-max classification constraints m Θps ({y m , xm }) = θ | ∀m, i, yi = yi , f (y m ; xm , θ) ≥ f (y m , yi ; xm , θ) . −i (5) m e(yi , yi ), For simplicity we omit the margin losses since the input fields xi (yi ) already suffice to rule out the trivial solution θ = 0. Proposition 2.1. For any θ ∗ ∈ Θcan , there is a set of 2|V |(k − 1) + 2|E|(k − 1)2 examples, {xm , y(xm ; θ ∗ )}, such that any pseudo-max consistent θ ∈ Θps ({y m , xm }) ∩ Θcan is arbitrarily close to θ ∗ . The proof is given in the supplementary material. To illustrate the key ideas, we consider the simpler binary discriminant function discussed in Eq. 3. Note that the binary model is already in the canonical form since Jij yi yj = 0 whenever yi = 0 or yj = 0. For any ij ∈ E, we show how to choose two input examples x1 and x2 such that any J consistent with the pseudo-max constraints for these ∗ ∗ two examples will have Jij ∈ [Jij − , Jij + ]. Repeating this for all of the edge parameters then gives the complete set of examples. The input examples we need for this will depend on J ∗ . For the first example, we set the input fields for all neighbors of i (except j) in such a way that ∗ we force the corresponding labels to be zero. More formally, we set x1 < −|N (k)| maxl |Jkl | for k 1 k ∈ N (i)\j, resulting in yk = 0, where y 1 = y(x1 ). In contrast, we set x1 to a large value, e.g. j ∗ 1 ∗ x1 > |N (j)| maxl |Jjl |, so that yj = 1. Finally, for node i, we set x1 = −Jij + so as to obtain a j i 1 slight preference for yi = 1. All other input fields can be set arbitrarily. As a result, the pseudo-max constraints pertaining to node i are f (y 1 ; x1 , J ) ≥ f (y 1 , yi ; x1 , J ) for yi = 0, 1. By taking into −i 1 account the label assignments for yi and its neighbors, and by removing terms that are the same on both sides of the equation, we get Jij + x1 + x1 ≥ Jij yi + yi x1 + x1 , which, for yi = 0, implies i j i j ∗ that Jij + x1 ≥ 0 or Jij − Jij + ≥ 0. The second example x2 differs only in terms of the input i ∗ 2 ∗ field for i. In particular, we set x2 = −Jij − so that yi = 0. This gives Jij ≤ Jij + , as desired. i 4 2.2 Consistency via Strict Convexity In this section we prove the consistency of the pseudo-max approach by showing that it corresponds to minimizing a strictly convex function. Our proof only requires that p(x) be non-zero for all x ∈ Rn (a simple example being a multi-variate Gaussian) and that J ∗ is finite. We use a discriminant function as in Eq. 3. Now, assume the input points xm are distributed according to p(x) and that y m are obtained via y m = arg maxy f (y; xm , J ∗ ). We can write the ps (J ) objective for finite data, and its limit when M → ∞, compactly as: 1 m m = max (yi − yi ) xm + Jki yk ps (J ) i M m i yi k∈N (i) p(x) max (yi − yi (x)) xi + → yi i Jki yk (x) dx (6) k∈N (i) ∗ where yi (x) is the label of i for input x when using parameters J . Starting from the above, consider the terms separately for each i. We partition the integral over x ∈ Rn into exclusive regions according to the predicted labels of the neighbors of i (given x). Define Sij = {x : yj (x) = 1 and yk (x) = 0 for k ∈ N (i)\j}. Eq. 6 can then be written as ps (J ) = gi ({Jik }k∈N (i) ) + ˆ i gik (Jik ) , (7) k∈N (i) where gik (Jik ) = x∈Sik p(x) maxyi [(yi −yi (x))(xi +Jik )]dx and gi ({Jik }k∈N (i) ) contains all of ˆ the remaining terms, i.e. where either zero or more than one neighbor is set to one. The function gi ˆ is convex in J since it is a sum of integrals over convex functions. We proceed to show that gik (Jik ) is strictly convex for all choices of i and k ∈ N (i). This will show that ps (J ) is strictly convex since it is a sum over functions strictly convex in each one of the variables in J . For all values xi ∈ (−∞, ∞) there is some x in Sij . This is because for any finite xi and finite J ∗ , the other xj ’s can be chosen so as to give the y configuration corresponding to Sij . Now, since p(x) has full support, we have P (Sij ) > 0 and p(x) > 0 for any x in Sij . As a result, this also holds for the marginal pi (xi |Sij ) over xi within Sij . After some algebra, we obtain: gij (Jij ) = P (Sij ) ∞ p(x)yi (x)(xi + Jij )dx pi (xi |Sij ) max [0, xi + Jij ] dxi − −∞ x∈Sij The integral over the yi (x)(xi + Jij ) expression just adds a linear term to gij (Jij ). The relevant remaining term is (for brevity we drop P (Sij ), a strictly positive constant, and the ij index): h(J) = ∞ pi (xi |Sij ) max [0, xi + J] dxi = −∞ ∞ ˆ pi (xi |Sij )h(xi , J)dxi (8) −∞ ˆ ˆ where we define h(xi , J) = max [0, xi + J]. Note that h(J) is convex since h(xi , J) is convex in J for all xi . We want to show that h(J) is strictly convex. Consider J < J and α ∈ (0, 1) and define ˆ ˆ the interval I = [−J, −αJ − (1 − α)J ]. For xi ∈ I it holds that: αh(xi , J) + (1 − α)h(xi , J ) > ˆ i , αJ + (1 − α)J ) (since the first term is strictly positive and the rest are zero). For all other x, h(x ˆ this inequality holds but is not necessarily strict (since h is always convex in J). We thus have after integrating over x that αh(J) + (1 − α)h(J ) > h(αJ + (1 − α)J ), implying h is strictly convex, as required. Note that we used the fact that p(x) has full support when integrating over I. The function ps (J ) is thus a sum of strictly convex functions in all its variables (namely g(Jik )) plus other convex functions of J , hence strictly convex. We can now proceed to show consistency. By strict convexity, the pseudo-max objective is minimized at a unique point J . Since we know that ps (J ∗ ) = 0 and zero is a lower bound on the value of ps (J ), it follows that J ∗ is the unique minimizer. Thus we have that as M → ∞, the minimizer of the pseudo-max objective is the true parameter vector, and thus we have consistency. As an example, consider the case of two variables y1 , y2 , with x1 and x2 distributed according to ∗ N (c1 , 1), N (0, 1) respectively. Furthermore assume J12 = 0. Then simple direct calculation yields: 2 2 2 c1 + J12 −c1 1 1 √ (9) e−x /2 dx − √ e−c1 /2 + √ e−(J12 +c1 ) /2 2π 2π 2π −J12 −c1 which is indeed a strictly convex function that is minimized at J = 0 (see Fig. 1 for an illustration). g(J12 ) = 5 3 Hardness of Structured Learning Most structured prediction learning algorithms use some form of inference as a subroutine. However, the corresponding prediction task is generally NP-hard. For example, maximizing the discriminant function defined in Eq. 3 is equivalent to solving Max-Cut, which is known to be NP-hard. This raises the question of whether it is possible to bypass prediction during learning. Although prediction may be intractable for arbitrary MRFs, what does this say about the difficulty of learning with a polynomial number of data points? In this section, we show that the problem of deciding whether there exists a parameter vector that separates the training data is NP-hard. Put in the context of the positive results in this paper, these hardness results show that, although in some cases the pseudo-max constraints yield a consistent estimate, we cannot hope for a certificate of optimality. Put differently, although the pseudo-max constraints in the separable case always give an outer bound on Θ (and may even be a single point), Θ could be the empty set – and we would never know the difference. Theorem 3.1. Given labeled examples {(xm , y m )}M for a fixed but arbitrary graph G, it is m=1 NP-hard to decide whether there exists parameters θ such that ∀m, y m = arg maxy f (y; xm , θ). Proof. Any parameters θ have an equivalent parameterization in canonical form (see section Sec. 2.1, also supplementary). Thus, the examples will be separable if and only if they are separable by some θ ∈ Θcan . We reduce from unweighted Max-Cut. The Max-Cut problem is to decide, given an undirected graph G, whether there exists a cut of at least K edges. Let G be the same graph as G, with k = 3 states per variable. We construct a small set of examples where a parameter vector will exist that separates the data if and only if there is no cut of K or more edges in G. Let θ be parameters in canonical form equivalent to θij (yi , yj ) = 1 if (yi , yj ) ∈ {(1, 2), (2, 1)}, 0 if yi = yj , and −n2 if (yi , yj ) ∈ {(1, 3), (2, 3), (3, 1), (3, 2)}. We first construct 4n + 8|E| examples, using the technique described in Sec. 2.1 (also supplementary material), which when restricted to the space Θcan , constrain the parameters to equal θ. We then use one more example (xm , y m ) where y m = 3 (every node is in state 3) and, for all i, xm (3) = K−1 and xm (1) = xm (2) = 0. The first i i i n two states encode the original Max-Cut instance, while the third state is used to construct a labeling y m that has value equal to K − 1, and is otherwise not used. Let K ∗ be the value of the maximum cut in G. If in any assignment to the last example there is a variable taking the state 3 and another variable taking the state 1 or 2, then the assignment’s value will be at most K ∗ − n2 , which is less than zero. By construction, the 3 assignment has value K − 1. Thus, the optimal assignment must either be 3 with value K − 1, or some combination of states 1 and 2, which has value at most K ∗ . If K ∗ > K − 1 then 3 is not optimal and the examples are not separable. If K ∗ ≤ K − 1, the examples are separable. This result illustrates the potential difficulty of learning in worst-case graphs. Nonetheless, many problems have a more restricted dependence on the input. For example, in computer vision, edge potentials may depend only on the difference in color between two adjacent pixels. Our results do not preclude positive results of learnability in such restricted settings. By establishing hardness of learning, we also close the open problem of relating hardness of inference and learning in structured prediction. If inference problems can be solved in polynomial time, then so can learning (using, e.g., structured perceptron). Thus, when learning is hard, inference must be hard as well. 4 Experiments To evaluate our learning algorithm, we test its performance on both synthetic and real-world datasets. We show that, as the number of training samples grows, the accuracy of the pseudo-max method improves and its speed-up gain over competing algorithms increases. Our learning algorithm corresponds to solving the following, where we add L2 regularization and use a scaled 0-1 loss, m m e(yi , yi ) = 1{yi = yi }/nm (nm is the number of labels in example m): min θ C m nm M nm m=1 i=1 m max f (y m , yi ; xm , θ) − f (y m ; xm , θ) + e(yi , yi ) + θ −i yi 2 . (10) We will compare the pseudo-max method with learning using structural SVMs, both with exact inference and LP relaxations [see, e.g., 4]. We use exact inference for prediction at test time. 6 (a) Synthetic (b) Reuters 0.4 exact LP−relaxation pseudo−max 0.15 Test error Test error 0.2 0.1 0.05 0 1 10 2 10 0.2 0.1 0 1 10 3 10 Train size exact LP−relaxation pseudo−max 0.3 2 10 3 10 4 10 Train size Figure 2: Test error as a function of train size for various algorithms. Subfigure (a) shows results for a synthetic setting, while (b) shows performance on the Reuters data. In the synthetic setting we use the discriminant function f (y; x, θ) = ij∈E θij (yi , yj ) + xi θi (yi ), which is similar to Eq. 4. We take a fully connected graph over n = 10 binary labels. i For a weight vector θ ∗ (sampled once, uniformly in the range [−1, 1], and used for all train/test sets) we generate train and test instances by sampling xm uniformly in the range [−5, 5] and then computing the optimal labels y m = arg maxy∈Y f (y; xm , θ ∗ ). We generate train sets of increasing size (M = {10, 50, 100, 500, 1000, 5000}), run the learning algorithms, and measure the test error for the learned weights (with 1000 test samples). For each train size we average the test error over 10 repeats of sampling and training. Fig. 2(a) shows a comparison of the test error for the three learning algorithms. For small numbers of training examples, the test error of pseudo-max is larger than that of the other algorithms. However, as the train size grows, the error converges to that of exact learning, as our consistency results predict. We also test the performance of our algorithm on a multi-label document classification task from the Reuters dataset [7]. The data consists of M = 23149 training samples, and we use a reduction of the dataset to the 5 most frequent labels. The 5 label variables form a fully connected pairwise graph structure (see [4] for a similar setting). We use random subsamples of increasing size from the train set to learn the parameters, and then measure the test error using 20000 additional samples. For each sample size and learning algorithm, we optimize the trade-off parameter C using 30% of the training data as a hold-out set. Fig. 2(b) shows that for the large data regime the performance of pseudo-max learning gets close to that of the other methods. However, unlike the synthetic setting there is still a small gap, even after seeing the entire train set. This could be because the full dataset is not yet large enough to be in the consistent regime (note that exact learning has not flattened either), or because the consistency conditions are not fully satisfied: the data might be non-separable or the support of the input distribution p(x) may be partial. We next apply our method to the problem of learning the energy function for protein side-chain placement, mirroring the learning setup of [14], where the authors train a conditional random field (CRF) using tree-reweighted belief propagation to maximize a lower bound on the likelihood.5 The prediction problem for side-chain placement corresponds to finding the most likely assignment in a pairwise MRF, and fits naturally into our learning framework. There are only 8 parameters to be learned, corresponding to a reweighting of known energy terms. The dataset consists of 275 proteins, where each MRF has several hundred variables (one per residue of the protein) and each variable has on average 20 states. For prediction we use CPLEX’s ILP solver. Fig. 3 shows a comparison of the pseudo-max method and a cutting-plane algorithm which uses an LP relaxation, solved with CPLEX, for finding violated constraints.6 We generate training sets of increasing size (M = {10, 50, 100, 274}), and measure the test error for the learned weights on the remaining examples.7 For M = 10, 50, 100 we average the test error over 3 random train/test splits, whereas for M = 274 we do 1-fold cross validation. We use C = 1 for both algorithms. 5 The authors’ data and results are available from: http://cyanover.fhcrc.org/recomb-2007/ We significantly optimized the cutting-plane algorithm, e.g. including a large number of initial cuttingplanes and restricting the weight vector to be positive (which we know to hold at optimality). 7 Specifically, for each protein we compute the fraction of correctly predicted χ1 and χ2 angles for all residues (except when trivial, e.g. just 1 state). Then, we compute the median of this value across all proteins. 6 7 Time to train (minutes) Test error (χ1 and χ2) 0.27 0.265 pseudo−max LP−relaxation Soft rep 0.26 0.255 0.25 0 50 100 150 200 Train size 250 250 200 pseudo−max LP−relaxation 150 100 50 0 0 50 100 150 200 Train size 250 Figure 3: Training time (for one train/test split) and test error as a function of train size for both the pseudomax method and a cutting-plane algorithm which uses a LP relaxation for inference, applied to the problem of learning the energy function for protein side-chain placement. The pseudo-max method obtains better accuracy than both the LP relaxation and HCRF (given roughly five times more data) for a fraction of the training time. The original weights (“Soft rep” [3]) used for this energy function have 26.7% error across all 275 proteins. The best previously reported parameters, learned in [14] using a Hidden CRF, obtain 25.6% error (their training set included 55 of these 275 proteins, so this is an optimistic estimate). To get a sense of the difficulty of this learning task, we also tried a random positive weight vector, uniformly sampled from the range [0, 1], obtaining an error of 34.9% (results would be much worse if we allowed the weights to be negative). Training using pseudo-max with 50 examples, we learn parameters in under a minute that give better accuracy than the HCRF. The speed-up of training with pseudo-max (using CPLEX’s QP solver) versus cutting-plane is striking. For example, for M = 10, pseudo-max takes only 3 seconds, a 1000-fold speedup. Unfortunately the cutting-plane algorithm took a prohibitive amount of time to be able to run on the larger training sets. Since the data used in learning for protein side-chain placement is both highly non-separable and relatively little, these positive results illustrate the potential wide-spread applicability of the pseudo-max method. 5 Discussion The key idea of our method is to find parameters that prefer the true assignment y m over assignments that differ from it in only one variable, in contrast to all other assignments. Perhaps surprisingly, this weak requirement is sufficient to achieve consistency given a rich enough input distribution. One extension of our approach is to add constraints for assignments that differ from y m in more than one variable. This would tighten the outer bound on Θ and possibly result in improved performance, but would also increase computational complexity. We could also add such competing assignments via a cutting-plane scheme so that optimization is performed only over a subset of these constraints. Our work raises a number of important open problems: It would be interesting to derive generalization bounds to understand the convergence rate of our method, as well as understanding the effect of the distribution p(x) on these rates. The distribution p(x) needs to have two key properties. On the one hand, it needs to explore the space Y in the sense that a sufficient number of labels need to be obtained as the correct label for the true parameters (this is indeed used in our consistency proofs). On the other hand, p(x) needs to be sufficiently sensitive close to the decision boundaries so that the true parameters can be inferred. We expect that generalization analysis will depend on these two properties of p(x). Note that [11] studied active learning schemes for structured data and may be relevant in the current context. How should one apply this learning algorithm to non-separable data sets? We suggested one approach, based on using a hinge loss for each of the pseudo constraints. One question in this context is, how resilient is this learning algorithm to label noise? Recent work has analyzed the sensitivity of pseudo-likelihood methods to model mis-specification [8], and it would be interesting to perform a similar analysis here. Also, is it possible to give any guarantees for the empirical and expected risks (with respect to exact inference) obtained by outer bound learning versus exact learning? Finally, our algorithm demonstrates a phenomenon where more data can make computation easier. Such a scenario was recently analyzed in the context of supervised learning [12], and it would be interesting to combine the approaches. Acknowledgments: We thank Chen Yanover for his assistance with the protein data. This work was supported by BSF grant 2008303 and a Google Research Grant. D.S. was supported by a Google PhD Fellowship. 8 References [1] J. Besag. The analysis of non-lattice data. The Statistician, 24:179–195, 1975. [2] M. Collins. Discriminative training methods for hidden Markov models: Theory and experiments with perceptron algorithms. In EMNLP, 2002. [3] G. Dantas, C. Corrent, S. L. Reichow, J. J. Havranek, Z. M. Eletr, N. G. Isern, B. Kuhlman, G. Varani, E. A. Merritt, and D. Baker. High-resolution structural and thermodynamic analysis of extreme stabilization of human procarboxypeptidase by computational protein design. Journal of Molecular Biology, 366(4):1209 – 1221, 2007. [4] T. Finley and T. Joachims. Training structural SVMs when exact inference is intractable. In Proceedings of the 25th International Conference on Machine Learning 25, pages 304–311. ACM, 2008. [5] T. Joachims, T. Finley, and C.-N. Yu. Cutting-plane training of structural SVMs. Machine Learning, 77(1):27–59, 2009. [6] A. Kulesza and F. Pereira. Structured learning with approximate inference. In Advances in Neural Information Processing Systems 20, pages 785–792. 2008. [7] D. Lewis, , Y. Yang, T. Rose, and F. Li. RCV1: a new benchmark collection for text categorization research. JMLR, 5:361–397, 2004. [8] P. Liang and M. I. Jordan. An asymptotic analysis of generative, discriminative, and pseudolikelihood estimators. In Proceedings of the 25th international conference on Machine learning, pages 584–591, New York, NY, USA, 2008. ACM Press. [9] A. F. T. Martins, N. A. Smith, and E. P. Xing. Polyhedral outer approximations with application to natural language parsing. In ICML 26, pages 713–720, 2009. [10] N. Ratliff, J. A. D. Bagnell, and M. Zinkevich. (Online) subgradient methods for structured prediction. In AISTATS, 2007. [11] D. Roth and K. Small. Margin-based active learning for structured output spaces. In Proc. of the European Conference on Machine Learning (ECML). Springer, September 2006. [12] S. Shalev-Shwartz and N. Srebro. SVM optimization: inverse dependence on training set size. In Proceedings of the 25th international conference on Machine learning, pages 928–935. ACM, 2008. [13] B. Taskar, C. Guestrin, and D. Koller. Max margin Markov networks. In Advances in Neural Information Processing Systems 16, pages 25–32. 2004. [14] C. Yanover, O. Schueler-Furman, and Y. Weiss. Minimizing and learning energy functions for side-chain prediction. Journal of Computational Biology, 15(7):899–911, 2008. 9

3 0.15718842 164 nips-2010-MAP Estimation for Graphical Models by Likelihood Maximization

Author: Akshat Kumar, Shlomo Zilberstein

Abstract: Computing a maximum a posteriori (MAP) assignment in graphical models is a crucial inference problem for many practical applications. Several provably convergent approaches have been successfully developed using linear programming (LP) relaxation of the MAP problem. We present an alternative approach, which transforms the MAP problem into that of inference in a mixture of simple Bayes nets. We then derive the Expectation Maximization (EM) algorithm for this mixture that also monotonically increases a lower bound on the MAP assignment until convergence. The update equations for the EM algorithm are remarkably simple, both conceptually and computationally, and can be implemented using a graph-based message passing paradigm similar to max-product computation. Experiments on the real-world protein design dataset show that EM’s convergence rate is significantly higher than the previous LP relaxation based approach MPLP. EM also achieves a solution quality within 95% of optimal for most instances. 1

4 0.14307968 165 nips-2010-MAP estimation in Binary MRFs via Bipartite Multi-cuts

Author: Sashank J. Reddi, Sunita Sarawagi, Sundar Vishwanathan

Abstract: We propose a new LP relaxation for obtaining the MAP assignment of a binary MRF with pairwise potentials. Our relaxation is derived from reducing the MAP assignment problem to an instance of a recently proposed Bipartite Multi-cut problem where the LP relaxation is guaranteed to provide an O(log k) approximation where k is the number of vertices adjacent to non-submodular edges in the MRF. We then propose a combinatorial algorithm to efficiently solve the LP and also provide a lower bound by concurrently solving its dual to within an approximation. The algorithm is up to an order of magnitude faster and provides better MAP scores and bounds than the state of the art message passing algorithm of [1] that tightens the local marginal polytope with third-order marginal constraints. 1

5 0.11666221 103 nips-2010-Generating more realistic images using gated MRF's

Author: Marc'aurelio Ranzato, Volodymyr Mnih, Geoffrey E. Hinton

Abstract: Probabilistic models of natural images are usually evaluated by measuring performance on rather indirect tasks, such as denoising and inpainting. A more direct way to evaluate a generative model is to draw samples from it and to check whether statistical properties of the samples match the statistics of natural images. This method is seldom used with high-resolution images, because current models produce samples that are very different from natural images, as assessed by even simple visual inspection. We investigate the reasons for this failure and we show that by augmenting existing models so that there are two sets of latent variables, one set modelling pixel intensities and the other set modelling image-specific pixel covariances, we are able to generate high-resolution images that look much more realistic than before. The overall model can be interpreted as a gated MRF where both pair-wise dependencies and mean intensities of pixels are modulated by the states of latent variables. Finally, we confirm that if we disallow weight-sharing between receptive fields that overlap each other, the gated MRF learns more efficient internal representations, as demonstrated in several recognition tasks. 1 Introduction and Prior Work The study of the statistical properties of natural images has a long history and has influenced many fields, from image processing to computational neuroscience [1]. In this work we focus on probabilistic models of natural images. These models are useful for extracting representations [2, 3, 4] that can be used for discriminative tasks and they can also provide adaptive priors [5, 6, 7] that can be used in applications like denoising and inpainting. Our main focus, however, will be on improving the quality of the generative model, rather than exploring its possible applications. Markov Random Fields (MRF’s) provide a very general framework for modelling natural images. In an MRF, an image is assigned a probability which is a normalized product of potential functions, with each function typically being defined over a subset of the observed variables. In this work we consider a very versatile class of MRF’s in which potential functions are defined over both pixels and latent variables, thus allowing the states of the latent variables to modulate or gate the effective interactions between the pixels. This type of MRF, that we dub gated MRF, was proposed as an image model by Geman and Geman [8]. Welling et al. [9] showed how an MRF in this family1 could be learned for small image patches and their work was extended to high-resolution images by Roth and Black [6] who also demonstrated its success in some practical applications [7]. Besides their practical use, these models were specifically designed to match the statistical properties of natural images, and therefore, it seems natural to evaluate them in those terms. Indeed, several authors [10, 7] have proposed that these models should be evaluated by generating images and 1 Product of Student’s t models (without pooling) may not appear to have latent variables but each potential can be viewed as an infinite mixture of zero-mean Gaussians where the inverse variance of the Gaussian is the latent variable. 1 checking whether the samples match the statistical properties observed in natural images. It is, therefore, very troublesome that none of the existing models can generate good samples, especially for high-resolution images (see for instance fig. 2 in [7] which is one of the best models of highresolution images reported in the literature so far). In fact, as our experiments demonstrate the generated samples from these models are more similar to random images than to natural images! When MRF’s with gated interactions are applied to small image patches, they actually seem to work moderately well, as demonstrated by several authors [11, 12, 13]. The generated patches have some coherent and elongated structure and, like natural image patches, they are predominantly very smooth with sudden outbreaks of strong structure. This is unsurprising because these models have a built-in assumption that images are very smooth with occasional strong violations of smoothness [8, 14, 15]. However, the extension of these patch-based models to high-resolution images by replicating filters across the image has proven to be difficult. The receptive fields that are learned no longer resemble Gabor wavelets but look random [6, 16] and the generated images lack any of the long range structure that is so typical of natural images [7]. The success of these methods in applications such as denoising is a poor measure of the quality of the generative model that has been learned: Setting the parameters to random values works almost as well for eliminating independent Gaussian noise [17], because this can be done quite well by just using a penalty for high-frequency variation. In this work, we show that the generative quality of these models can be drastically improved by jointly modelling both pixel mean intensities and pixel covariances. This can be achieved by using two sets of latent variables, one that gates pair-wise interactions between pixels and another one that sets the mean intensities of pixels, as we already proposed in some earlier work [4]. Here, we show that this modelling choice is crucial to make the gated MRF work well on high-resolution images. Finally, we show that the most widely used method of sharing weights in MRF’s for high-resolution images is overly constrained. Earlier work considered homogeneous MRF’s in which each potential is replicated at all image locations. This has the subtle effect of making learning very difficult because of strong correlations at nearby sites. Following Gregor and LeCun [18] and also Tang and Eliasmith [19], we keep the number of parameters under control by using local potentials, but unlike Roth and Black [6] we only share weights between potentials that do not overlap. 2 Augmenting Gated MRF’s with Mean Hidden Units A Product of Student’s t (PoT) model [15] is a gated MRF defined on small image patches that can be viewed as modelling image-specific, pair-wise relationships between pixel values by using the states of its latent variables. It is very good at representing the fact that two-pixel have very similar intensities and no good at all at modelling what these intensities are. Failure to model the mean also leads to impoverished modelling of the covariances when the input images have nonzero mean intensity. The covariance RBM (cRBM) [20] is another model that shares the same limitation since it only differs from PoT in the distribution of its latent variables: The posterior over the latent variables is a product of Bernoulli distributions instead of Gamma distributions as in PoT. We explain the fundamental limitation of these models by using a simple toy example: Modelling two-pixel images using a cRBM with only one binary hidden unit, see fig. 1. This cRBM assumes that the conditional distribution over the input is a zero-mean Gaussian with a covariance that is determined by the state of the latent variable. Since the latent variable is binary, the cRBM can be viewed as a mixture of two zero-mean full covariance Gaussians. The latent variable uses the pairwise relationship between pixels to decide which of the two covariance matrices should be used to model each image. When the input data is pre-proessed by making each image have zero mean intensity (the empirical histogram is shown in the first row and first column), most images lie near the origin because most of the times nearby pixels are strongly correlated. Less frequently we encounter edge images that exhibit strong anti-correlation between the pixels, as shown by the long tails along the anti-diagonal line. A cRBM could model this data by using two Gaussians (first row and second column): one that is spherical and tight at the origin for smooth images and another one that has a covariance elongated along the anti-diagonal for structured images. If, however, the whole set of images is normalized by subtracting from every pixel the mean value of all pixels over all images (second row and first column), the cRBM fails at modelling structured images (second row and second column). It can fit a Gaussian to the smooth images by discovering 2 Figure 1: In the first row, each image is zero mean. In the second row, the whole set of data points is centered but each image can have non-zero mean. The first column shows 8x8 images picked at random from natural images. The images in the second column are generated by a model that does not account for mean intensity. The images in the third column are generated by a model that has both “mean” and “covariance” hidden units. The contours in the first column show the negative log of the empirical distribution of (tiny) natural two-pixel images (x-axis being the first pixel and the y-axis the second pixel). The plots in the other columns are toy examples showing how each model could represent the empirical distribution using a mixture of Gaussians with components that have one of two possible covariances (corresponding to the state of a binary “covariance” latent variable). Models that can change the means of the Gaussians (mPoT and mcRBM) can represent better structured images (edge images lie along the anti-diagonal and are fitted by the Gaussians shown in red) while the other models (PoT and cRBM) fail, overall when each image can have non-zero mean. the direction of strong correlation along the main diagonal, but it is very likely to fail to discover the direction of anti-correlation, which is crucial to represent discontinuities, because structured images with different mean intensity appear to be evenly spread over the whole input space. If the model has another set of latent variables that can change the means of the Gaussian distributions in the mixture (as explained more formally below and yielding the mPoT and mcRBM models), then the model can represent both changes of mean intensity and the correlational structure of pixels (see last column). The mean latent variables effectively subtract off the relevant mean from each data-point, letting the covariance latent variable capture the covariance structure of the data. As before, the covariance latent variable needs only to select between two covariance matrices. In fact, experiments on real 8x8 image patches confirm these conjectures. Fig. 1 shows samples drawn from PoT and mPoT. mPoT (and similarly mcRBM [4]) is not only better at modelling zero mean images but it can also represent images that have non zero mean intensity well. We now describe mPoT, referring the reader to [4] for a detailed description of mcRBM. In PoT [9] the energy function is: E PoT (x, hc ) = i 1 [hc (1 + (Ci T x)2 ) + (1 − γ) log hc ] i i 2 (1) where x is a vectorized image patch, hc is a vector of Gamma “covariance” latent variables, C is a filter bank matrix and γ is a scalar parameter. The joint probability over input pixels and latent variables is proportional to exp(−E PoT (x, hc )). Therefore, the conditional distribution over the input pixels is a zero-mean Gaussian with covariance equal to: Σc = (Cdiag(hc )C T )−1 . (2) In order to make the mean of the conditional distribution non-zero, we define mPoT as the normalized product of the above zero-mean Gaussian that models the covariance and a spherical covariance Gaussian that models the mean. The overall energy function becomes: E mPoT (x, hc , hm ) = E PoT (x, hc ) + E m (x, hm ) 3 (3) Figure 2: Illustration of different choices of weight-sharing scheme for a RBM. Links converging to one latent variable are filters. Filters with the same color share the same parameters. Kinds of weight-sharing scheme: A) Global, B) Local, C) TConv and D) Conv. E) TConv applied to an image. Cells correspond to neighborhoods to which filters are applied. Cells with the same color share the same parameters. F) 256 filters learned by a Gaussian RBM with TConv weight-sharing scheme on high-resolution natural images. Each filter has size 16x16 pixels and it is applied every 16 pixels in both the horizontal and vertical directions. Filters in position (i, j) and (1, 1) are applied to neighborhoods that are (i, j) pixels away form each other. Best viewed in color. where hm is another set of latent variables that are assumed to be Bernoulli distributed (but other distributions could be used). The new energy term is: E m (x, hm ) = 1 T x x− 2 hm Wj T x j (4) j yielding the following conditional distribution over the input pixels: p(x|hc , hm ) = N (Σ(W hm ), Σ), Σ = (Σc + I)−1 (5) with Σc defined in eq. 2. As desired, the conditional distribution has non-zero mean2 . Patch-based models like PoT have been extended to high-resolution images by using spatially localized filters [6]. While we can subtract off the mean intensity from independent image patches to successfully train PoT, we cannot do that on a high-resolution image because overlapping patches might have different mean. Unfortunately, replicating potentials over the image ignoring variations of mean intensity has been the leading strategy to date [6]3 . This is the major reason why generation of high-resolution images is so poor. Sec. 4 shows that generation can be drastically improved by explicitly accounting for variations of mean intensity, as performed by mPoT and mcRBM. 3 Weight-Sharing Schemes By integrating out the latent variables, we can write the density function of any gated MRF as a normalized product of potential functions (for mPoT refer to eq. 6). In this section we investigate different ways of constraining the parameters of the potentials of a generic MRF. Global: The obvious way to extend a patch-based model like PoT to high-resolution images is to define potentials over the whole image; we call this scheme global. This is not practical because 1) the number of parameters grows about quadratically with the size of the image making training too slow, 2) we do not need to model interactions between very distant pairs of pixels since their dependence is negligible, and 3) we would not be able to use the model on images of different size. Conv: The most popular way to handle big images is to define potentials on small subsets of variables (e.g., neighborhoods of size 5x5 pixels) and to replicate these potentials across space while 2 The need to model the means was clearly recognized in [21] but they used conjunctive latent features that simultaneously represented a contribution to the “precision matrix” in a specific direction and the mean along that same direction. 3 The success of PoT-like models in Bayesian denoising is not surprising since the noisy image effectively replaces the reconstruction term from the mean hidden units (see eq. 5), providing a set of noisy mean intensities that are cleaned up by the patterns of correlation enforced by the covariance latent variables. 4 sharing their parameters at each image location [23, 24, 6]. This yields a convolutional weightsharing scheme, also called homogeneous field in the statistics literature. This choice is justified by the stationarity of natural images. This weight-sharing scheme is extremely concise in terms of number of parameters, but also rather inefficient in terms of latent representation. First, if there are N filters at each location and these filters are stepped by one pixel then the internal representation is about N times overcomplete. The internal representation has not only high computational cost, but it is also highly redundant. Since the input is mostly smooth and the parameters are the same across space, the latent variables are strongly correlated as well. This inefficiency turns out to be particularly harmful for a model like PoT causing the learned filters to become “random” looking (see fig 3-iii). A simple intuition follows from the equivalence between PoT and square ICA [15]. If the filter matrix C of eq. 1 is square and invertible, we can marginalize out the latent variables and write: p(y) = i S(yi ), where yi = Ci T x and S is a Student’s t distribution. In other words, there is an underlying assumption that filter outputs are independent. However, if the filters of matrix C are shifted and overlapping versions of each other, this clearly cannot be true. Training PoT with the Conv weight-sharing scheme forces the model to find filters that make filter outputs as independent as possible, which explains the very high-frequency patterns that are usually discovered [6]. Local: The Global and Conv weight-sharing schemes are at the two extremes of a spectrum of possibilities. For instance, we can define potentials on a small subset of input variables but, unlike Conv, each potential can have its own set of parameters, as shown in fig. 2-B. This is called local, or inhomogeneous field. Compared to Conv the number of parameters increases only slightly but the number of latent variables required and their redundancy is greatly reduced. In fact, the model learns different receptive fields at different locations as a better strategy for representing the input, overall when the number of potentials is limited (see also fig. 2-F). TConv: Local would not allow the model to be trained and tested on images of different resolution, and it might seem wasteful not to exploit the translation invariant property of images. We therefore advocate the use of a weight-sharing scheme that we call tiled-convolutional (TConv) shown in fig. 2-C and E [18]. Each filter tiles the image without overlaps with copies of itself (i.e. the stride equals the filter diameter). This reduces spatial redundancy of latent variables and allows the input images to have arbitrary size. At the same time, different filters do overlap with each other in order to avoid tiling artifacts. Fig. 2-F shows filters that were (jointly) learned by a Restricted Boltzmann Machine (RBM) [29] with Gaussian input variables using the TConv weight-sharing scheme. 4 Experiments We train gated MRF’s with and without mean hidden units using different weight-sharing schemes. The training procedure is very similar in all cases. We perform approximate maximum likelihood by using Fast Persistence Contrastive Divergence (FPCD) [25] and we draw samples by using Hybrid Monte Carlo (HMC) [26]. Since all latent variables can be exactly marginalized out we can use HMC on the free energy (negative logarithm of the marginal distribution over the input pixels). For mPoT this is: F mPoT (x) = − log(p(x))+const. = k,i 1 1 γ log(1+ (Cik T xk )2 )+ xT x− 2 2 T log(1+exp(Wjk xk )) (6) k,j where the index k runs over spatial locations and xk is the k-th image patch. FPCD keeps samples, called negative particles, that it uses to represent the model distribution. These particles are all updated after each weight update. For each mini-batch of data-points a) we compute the derivative of the free energy w.r.t. the training samples, b) we update the negative particles by running HMC for one HMC step consisting of 20 leapfrog steps. We start at the previous set of negative particles and use as parameters the sum of the regular parameters and a small perturbation vector, c) we compute the derivative of the free energy at the negative particles, and d) we update the regular parameters by using the difference of gradients between step a) and c) while the perturbation vector is updated using the gradient from c) only. The perturbation is also strongly decayed to zero and is subject to a larger learning rate. The aim is to encourage the negative particles to explore the space more quickly by slightly and temporarily raising the energy at their current position. Note that the use of FPCD as opposed to other estimation methods (like Persistent Contrastive Divergence [27]) turns out to be crucial to achieve good mixing of the sampler even after training. We train on mini-batches of 32 samples using gray-scale images of approximate size 160x160 pixels randomly cropped from the Berkeley segmentation dataset [28]. We perform 160,000 weight updates decreasing the learning by a factor of 4 by the end of training. The initial learning rate is set to 0.1 for the covariance 5 Figure 3: 160x160 samples drawn by A) mPoT-TConv, B) mHPoT-TConv, C) mcRBM-TConv and D) PoTTConv. On the side also i) a subset of 8x8 “covariance” filters learned by mPoT-TConv (the plot below shows how the whole set of filters tile a small patch; each bar correspond to a Gabor fit of a filter and colors identify filters applied at the same 8x8 location, each group is shifted by 2 pixels down the diagonal and a high-resolution image is tiled by replicating this pattern every 8 pixels horizontally and vertically), ii) a subset of 8x8 “mean” filters learned by the same mPoT-TConv, iii) filters learned by PoT-Conv and iv) by PoT-TConv. filters (matrix C of eq. 1), 0.01 for the mean parameters (matrix W of eq. 4), and 0.001 for the other parameters (γ of eq. 1). During training we condition on the borders and initialize the negative particles at zero in order to avoid artifacts at the border of the image. We learn 8x8 filters and pre-multiply the covariance filters by a whitening transform retaining 99% of the variance; we also normalize the norm of the covariance filters to prevent some of them from decaying to zero during training4 . Whenever we use the TConv weight-sharing scheme the model learns covariance filters that mostly resemble localized and oriented Gabor functions (see fig. 3-i and iv), while the Conv weight-sharing scheme learns structured but poorly localized high-frequency patterns (see fig. 3-iii) [6]. The TConv models re-use the same 8x8 filters every 8 pixels and apply a diagonal offset of 2 pixels between neighboring filters with different weights in order to reduce tiling artifacts. There are 4 sets of filters, each with 64 filters for a total of 256 covariance filters (see bottom plot of fig. 3). Similarly, we have 4 sets of mean filters, each with 32 filters. These filters have usually non-zero mean and exhibit on-center off-surround and off-center on-surround patterns, see fig. 3-ii. In order to draw samples from the learned models, we run HMC for a long time (10,000 iterations, each composed of 20 leap-frog steps). Some samples of size 160x160 pixels are reported in fig. 3 A)D). Without modelling the mean intensity, samples lack structure and do not seem much different from those that would be generated by a simple Gaussian model merely fitting the second order statistics (see fig. 3 in [1] and also fig. 2 in [7]). By contrast, structure, sharp boundaries and some simple texture emerge only from models that have mean latent variables, namely mcRBM, mPoT and mHPoT which differs from mPoT by having a second layer pooling matrix on the squared covariance filter outputs [11]. A more quantitative comparison is reported in table 1. We first compute marginal statistics of filter responses using the generated images, natural images from the test set, and random images. The statistics are the normalized histogram of individual filter responses to 24 Gabor filters (8 orientations and 3 scales). We then calculate the KL divergence between the histograms on random images and generated images and the KL divergence between the histograms on natural images and generated images. The table also reports the average difference of energies between random images and natural images. All results demonstrate that models that account for mean intensity generate images 4 The code used in the experiments can be found at the first author’s web-page. 6 MODEL F (R) − F (T ) (104 ) KL(R G) KL(T G) KL(R G) − KL(T PoT - Conv 2.9 0.3 0.6 PoT - TConv 2.8 0.4 1.0 -0.6 mPoT - TConv 5.2 1.0 0.2 0.8 mHPoT - TConv 4.9 1.7 0.8 0.9 mcRBM - TConv 3.5 1.5 1.0 G) -0.3 0.5 Table 1: Comparing MRF’s by measuring: difference of energy (negative log ratio of probabilities) between random images (R) and test natural images (T), the KL divergence between statistics of random images (R) and generated images (G), KL divergence between statistics of test natural images (T) and generated images (G), and difference of these two KL divergences. Statistics are computed using 24 Gabor filters. that are closer to natural images than to random images, whereas models that do not account for the mean (like the widely used PoT-Conv) produce samples that are actually closer to random images. 4.1 Discriminative Experiments on Weight-Sharing Schemes In future work, we intend to use the features discovered by the generative model for recognition. To understand how the different weight sharing schemes affect recognition performance we have done preliminary tests using the discriminative performance of a simpler model on simpler data. We consider one of the simplest and most versatile models, namely the RBM [29]. Since we also aim to test the Global weight-sharing scheme we are constrained to using fairly low resolution datasets such as the MNIST dataset of handwritten digits [30] and the CIFAR 10 dataset of generic object categories [22]. The MNIST dataset has soft binary images of size 28x28 pixels, while the CIFAR 10 dataset has color images of size 32x32 pixels. CIFAR 10 has 10 classes, 5000 training samples per class and 1000 test samples per class. MNIST also has 10 classes with, on average, 6000 training samples per class and 1000 test samples per class. The energy function of the RBM trained on the CIFAR 10 dataset, modelling input pixels with 3 (R,G,B) Gaussian variables [31], is exactly the one shown in eq. 4; while the RBM trained on MNIST uses logistic units for the pixels and the energy function is again the same as before but without any quadratic term. All models are trained in an unsupervised way to approximately maximize the likelihood in the training set using Contrastive Divergence [32]. They are then used to represent each input image with a feature vector (mean of the posterior over the latent variables) which is fed to a multinomial logistic classifier for discrimination. Models are compared in terms of: 1) recognition accuracy, 2) convergence time and 3) dimensionality of the representation. In general, assuming filters much smaller than the input image and assuming equal number of latent variables, Conv, TConv and Local models process each sample faster than Global by a factor approximately equal to the ratio between the area of the image and the area of the filters, which can be very large in practice. In the first set of experiments reported on the left of fig. 4 we study the internal representation in terms of discrimination and dimensionality using the MNIST dataset. For each choice of dimensionality all models are trained using the same number of operations. This is set to the amount necessary to complete one epoch over the training set using the Global model. This experiment shows that: 1) Local outperforms all other weight-sharing schemes for a wide range of dimensionalities, 2) TConv does not perform as well as Local probably because the translation invariant assumption is clearly violated for these relatively small, centered, images, 3) Conv performs well only when the internal representation is very high dimensional (10 times overcomplete) otherwise it severely underfits, 4) Global performs well when the representation is compact but its performance degrades rapidly as this increases because it needs more than the allotted training time. The right hand side of fig. 4 shows how the recognition performance evolves as we increase the number of operations (or training time) using models that produce a twice overcomplete internal representation. With only very few filters Conv still underfits and it does not improve its performance by training for longer, but Global does improve and eventually it reaches the performance of Local. If we look at the crossing of the error rate at 2% we can see that Local is about 4 times faster than Global. To summarize, Local provides more compact representations than Conv, is much faster than Global while achieving 7 6 2.4 error rate % 5 error rate % 2.6 Global Local TConv Conv 4 3 2 1 0 2.2 Global Local 2 Conv 1.8 1000 2000 3000 4000 5000 dimensionality 6000 7000 1.6 0 8000 2 4 6 8 # flops (relative to # flops per epoch of Global model) 10 Figure 4: Experiments on MNIST using RBM’s with different weight-sharing schemes. Left: Error rate as a function of the dimensionality of the latent representation. Right: Error rate as a function of the number of operations (normalized to those needed to perform one epoch in the Global model); all models have a twice overcomplete latent representation. similar performance in discrimination. Also, Local can easily scale to larger images while Global cannot. Similar experiments are performed using the CIFAR 10 dataset [22] of natural images. Using the same protocol introduced in earlier work by Krizhevsky [22], the RBM’s are trained in an unsupervised way on a subset of the 80 million tiny images dataset [33] and then “fine-tuned” on the CIFAR 10 dataset by supervised back-propagation of the error through the linear classifier and feature extractor. All models produce an approximately 10,000 dimensional internal representation to make a fair comparison. Models using local filters learn 16x16 filters that are stepped every pixel. Again, we do not experiment with the TConv weight-sharing scheme because the image is not large enough to allow enough replicas. Similarly to fig. 3-iii the Conv weight-sharing scheme was very difficult to train and did not produce Gabor-like features. Indeed, careful injection of sparsity and long training time seem necessary [31] for these RBM’s. By contrast, both Local and Global produce Gabor-like filters similar to those shown in fig. 2 F). The model trained with Conv weight-sharing scheme yields an accuracy equal to 56.6%, while Local and Global yield much better performance, 63.6% and 64.8% [22], respectively. Although Local and Global have similar performance, training with the Local weight-sharing scheme took under an hour while using the Global weight-sharing scheme required more than a day. 5 Conclusions and Future Work This work is motivated by the poor generative quality of currently popular MRF models of natural images. These models generate images that are actually more similar to white noise than to natural images. Our contribution is to recognize that current models can benefit from 1) the addition of a simple model of the mean intensities and from 2) the use of a less constrained weight-sharing scheme. By augmenting these models with an extra set of latent variables that model mean intensity we can generate samples that look much more realistic: they are characterized by smooth regions, sharp boundaries and some simple high frequency texture. We validate our approach by comparing the statistics of filter outputs on natural images and generated images. In the future, we plan to integrate these MRF’s into deeper hierarchical models and to use their internal representation to perform object recognition in high-resolution images. The hope is to further improve generation by capturing longer range dependencies and to exploit this to better cope with missing values and ambiguous sensory inputs. References [1] E.P. Simoncelli. Statistical modeling of photographic images. Handbook of Image and Video Processing, pages 431–441, 2005. 8 [2] A. Hyvarinen, J. Karhunen, and E. Oja. Independent Component Analysis. John Wiley & Sons, 2001. [3] G.E. Hinton and R. R Salakhutdinov. Reducing the dimensionality of data with neural networks. Science, 313(5786):504–507, 2006. [4] M. Ranzato and G.E. Hinton. Modeling pixel means and covariances using factorized third-order boltzmann machines. In CVPR, 2010. [5] M.J. Wainwright and E.P. Simoncelli. Scale mixtures of gaussians and the statistics of natural images. In NIPS, 2000. [6] S. Roth and M.J. Black. Fields of experts: A framework for learning image priors. In CVPR, 2005. [7] U. Schmidt, Q. Gao, and S. Roth. A generative perspective on mrfs in low-level vision. In CVPR, 2010. [8] S. Geman and D. Geman. Stochastic relaxation, gibbs distributions, and the bayesian restoration of images. PAMI, 6:721–741, 1984. [9] M. Welling, G.E. Hinton, and S. Osindero. Learning sparse topographic representations with products of student-t distributions. In NIPS, 2003. [10] S.C. Zhu and D. Mumford. Prior learning and gibbs reaction diffusion. PAMI, pages 1236–1250, 1997. [11] S. Osindero, M. Welling, and G. E. Hinton. Topographic product models applied to natural scene statistics. Neural Comp., 18:344–381, 2006. [12] S. Osindero and G. E. Hinton. Modeling image patches with a directed hierarchy of markov random fields. In NIPS, 2008. [13] Y. Karklin and M.S. Lewicki. Emergence of complex cell properties by learning to generalize in natural scenes. Nature, 457:83–86, 2009. [14] B. A. Olshausen and D. J. Field. Sparse coding with an overcomplete basis set: a strategy employed by v1? Vision Research, 37:3311–3325, 1997. [15] Y. W. Teh, M. Welling, S. Osindero, and G. E. Hinton. Energy-based models for sparse overcomplete representations. JMLR, 4:1235–1260, 2003. [16] Y. Weiss and W.T. Freeman. What makes a good model of natural images? In CVPR, 2007. [17] S. Roth and M. J. Black. Fields of experts. Int. Journal of Computer Vision, 82:205–229, 2009. [18] K. Gregor and Y. LeCun. Emergence of complex-like cells in a temporal product network with local receptive fields. arXiv:1006.0448, 2010. [19] C. Tang and C. Eliasmith. Deep networks for robust visual recognition. In ICML, 2010. [20] M. Ranzato, A. Krizhevsky, and G.E. Hinton. Factored 3-way restricted boltzmann machines for modeling natural images. In AISTATS, 2010. [21] N. Heess, C.K.I. Williams, and G.E. Hinton. Learning generative texture models with extended fields-ofexperts. In BMCV, 2009. [22] A. Krizhevsky. Learning multiple layers of features from tiny images, 2009. MSc Thesis, Dept. of Comp. Science, Univ. of Toronto. [23] A. Waibel, T. Hanazawa, G. Hinton, K. Shikano, and K. Lang. Phoneme recognition using time-delay neural networks. IEEE Acoustics Speech and Signal Proc., 37:328–339, 1989. [24] Y. LeCun, L. Bottou, Y. Bengio, and P. Haffner. Gradient-based learning applied to document recognition. Proceedings of the IEEE, 86(11):2278–2324, 1998. [25] T. Tieleman and G.E. Hinton. Using fast weights to improve persistent contrastive divergence. In ICML, 2009. [26] R.M. Neal. Bayesian learning for neural networks. Springer-Verlag, 1996. [27] T. Tieleman. Training restricted boltzmann machines using approximations to the likelihood gradient. In ICML, 2008. [28] http://www.cs.berkeley.edu/projects/vision/grouping/segbench/. [29] M. Welling, M. Rosen-Zvi, and G.E. Hinton. Exponential family harmoniums with an application to information retrieval. In NIPS, 2005. [30] http://yann.lecun.com/exdb/mnist/. [31] H. Lee, R. Grosse, R. Ranganath, and A. Y. Ng. Convolutional deep belief networks for scalable unsupervised learning of hierarchical representations. In Proc. ICML, 2009. [32] G.E. Hinton. Training products of experts by minimizing contrastive divergence. Neural Computation, 14:1771–1800, 2002. [33] A. Torralba, R. Fergus, and W.T. Freeman. 80 million tiny images: a large dataset for non-parametric object and scene recognition. PAMI, 30:1958–1970, 2008. 9

6 0.10120932 270 nips-2010-Tight Sample Complexity of Large-Margin Learning

7 0.099774122 101 nips-2010-Gaussian sampling by local perturbations

8 0.095860586 77 nips-2010-Epitome driven 3-D Diffusion Tensor image segmentation: on extracting specific structures

9 0.091985308 49 nips-2010-Computing Marginal Distributions over Continuous Markov Networks for Statistical Relational Learning

10 0.080795079 63 nips-2010-Distributed Dual Averaging In Networks

11 0.074944489 232 nips-2010-Sample Complexity of Testing the Manifold Hypothesis

12 0.074687265 1 nips-2010-(RF)^2 -- Random Forest Random Field

13 0.074065246 239 nips-2010-Sidestepping Intractable Inference with Structured Ensemble Cascades

14 0.070323467 259 nips-2010-Subgraph Detection Using Eigenvector L1 Norms

15 0.069508284 283 nips-2010-Variational Inference over Combinatorial Spaces

16 0.068959273 254 nips-2010-Stability Approach to Regularization Selection (StARS) for High Dimensional Graphical Models

17 0.068952687 144 nips-2010-Learning Efficient Markov Networks

18 0.066231124 21 nips-2010-Accounting for network effects in neuronal responses using L1 regularized point process models

19 0.063008651 83 nips-2010-Evidence-Specific Structures for Rich Tractable CRFs

20 0.062577561 180 nips-2010-Near-Optimal Bayesian Active Learning with Noisy Observations


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, 0.176), (1, 0.038), (2, 0.042), (3, 0.06), (4, -0.066), (5, -0.057), (6, -0.047), (7, -0.014), (8, 0.112), (9, 0.046), (10, -0.16), (11, -0.095), (12, 0.023), (13, 0.037), (14, -0.102), (15, -0.022), (16, -0.067), (17, -0.153), (18, 0.062), (19, 0.006), (20, 0.065), (21, -0.006), (22, -0.064), (23, 0.101), (24, 0.024), (25, -0.079), (26, -0.145), (27, 0.034), (28, -0.081), (29, -0.023), (30, -0.03), (31, -0.073), (32, -0.041), (33, 0.043), (34, 0.116), (35, -0.145), (36, -0.033), (37, 0.027), (38, 0.004), (39, -0.093), (40, 0.037), (41, 0.001), (42, -0.017), (43, 0.006), (44, -0.117), (45, 0.061), (46, -0.036), (47, -0.025), (48, -0.024), (49, 0.082)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.9579407 288 nips-2010-Worst-case bounds on the quality of max-product fixed-points

Author: Meritxell Vinyals, Jes\'us Cerquides, Alessandro Farinelli, Juan A. Rodríguez-aguilar

Abstract: We study worst-case bounds on the quality of any fixed point assignment of the max-product algorithm for Markov Random Fields (MRF). We start providing a bound independent of the MRF structure and parameters. Afterwards, we show how this bound can be improved for MRFs with specific structures such as bipartite graphs or grids. Our results provide interesting insight into the behavior of max-product. For example, we prove that max-product provides very good results (at least 90% optimal) on MRFs with large variable-disjoint cycles1 . 1

2 0.81341618 165 nips-2010-MAP estimation in Binary MRFs via Bipartite Multi-cuts

Author: Sashank J. Reddi, Sunita Sarawagi, Sundar Vishwanathan

Abstract: We propose a new LP relaxation for obtaining the MAP assignment of a binary MRF with pairwise potentials. Our relaxation is derived from reducing the MAP assignment problem to an instance of a recently proposed Bipartite Multi-cut problem where the LP relaxation is guaranteed to provide an O(log k) approximation where k is the number of vertices adjacent to non-submodular edges in the MRF. We then propose a combinatorial algorithm to efficiently solve the LP and also provide a lower bound by concurrently solving its dual to within an approximation. The algorithm is up to an order of magnitude faster and provides better MAP scores and bounds than the state of the art message passing algorithm of [1] that tightens the local marginal polytope with third-order marginal constraints. 1

3 0.7806223 164 nips-2010-MAP Estimation for Graphical Models by Likelihood Maximization

Author: Akshat Kumar, Shlomo Zilberstein

Abstract: Computing a maximum a posteriori (MAP) assignment in graphical models is a crucial inference problem for many practical applications. Several provably convergent approaches have been successfully developed using linear programming (LP) relaxation of the MAP problem. We present an alternative approach, which transforms the MAP problem into that of inference in a mixture of simple Bayes nets. We then derive the Expectation Maximization (EM) algorithm for this mixture that also monotonically increases a lower bound on the MAP assignment until convergence. The update equations for the EM algorithm are remarkably simple, both conceptually and computationally, and can be implemented using a graph-based message passing paradigm similar to max-product computation. Experiments on the real-world protein design dataset show that EM’s convergence rate is significantly higher than the previous LP relaxation based approach MPLP. EM also achieves a solution quality within 95% of optimal for most instances. 1

4 0.54594678 49 nips-2010-Computing Marginal Distributions over Continuous Markov Networks for Statistical Relational Learning

Author: Matthias Broecheler, Lise Getoor

Abstract: Continuous Markov random fields are a general formalism to model joint probability distributions over events with continuous outcomes. We prove that marginal computation for constrained continuous MRFs is #P-hard in general and present a polynomial-time approximation scheme under mild assumptions on the structure of the random field. Moreover, we introduce a sampling algorithm to compute marginal distributions and develop novel techniques to increase its efficiency. Continuous MRFs are a general purpose probabilistic modeling tool and we demonstrate how they can be applied to statistical relational learning. On the problem of collective classification, we evaluate our algorithm and show that the standard deviation of marginals serves as a useful measure of confidence. 1

5 0.53823543 169 nips-2010-More data means less inference: A pseudo-max approach to structured learning

Author: David Sontag, Ofer Meshi, Amir Globerson, Tommi S. Jaakkola

Abstract: The problem of learning to predict structured labels is of key importance in many applications. However, for general graph structure both learning and inference are intractable. Here we show that it is possible to circumvent this difficulty when the distribution of training examples is rich enough, via a method similar in spirit to pseudo-likelihood. We show that our new method achieves consistency, and illustrate empirically that it indeed approaches the performance of exact methods when sufficiently large training sets are used. Many prediction problems in machine learning applications are structured prediction tasks. For example, in protein folding we are given a protein sequence and the goal is to predict the protein’s native structure [14]. In parsing for natural language processing, we are given a sentence and the goal is to predict the most likely parse tree [2]. In these and many other applications, we can formalize the structured prediction problem as taking an input x (e.g., primary sequence, sentence) and predicting ˆ y (e.g., structure, parse) according to y = arg maxy∈Y θ · φ(x, y ), where φ(x, y) is a function that ˆ maps any input and a candidate assignment to a feature vector, Y denotes the space of all possible assignments to the vector y, and θ is a weight vector to be learned. This paper addresses the problem of learning structured prediction models from data. In particular, given a set of labeled examples {(xm , y m )}M , our goal is to find a vector θ such that for each m=1 example m, y m = arg maxy∈Y θ · φ(xm , y), i.e. one which separates the training data. For many structured prediction models, maximization over Y is computationally intractable. This makes it difficult to apply previous algorithms for learning structured prediction models, such as structured perceptron [2], stochastic subgradient [10], and cutting-plane algorithms [5], which require making a prediction at every iteration (equivalent to repeatedly solving an integer linear program). Given training data, we can consider the space of parameters Θ that separate the data. This space can be defined by the intersection of a large number of linear inequalities. A recent approach to getting around the hardness of prediction is to use linear programming (LP) relaxations to approximate the maximization over Y [4, 6, 9]. However, separation with respect to a relaxation places stronger constraints on the parameters. The target solution, an integral vertex in the LP, must now distinguish itself also from possible fractional vertexes that arise due to the relaxation. The relaxations can therefore be understood as optimizing over an inner bound of Θ. This set may be empty even if the training data is separable with exact inference [6]. Another obstacle to using LP relaxations for learning is that solving the LPs can be very slow. In this paper we ask whether it is possible to learn while avoiding inference altogether. We propose a new learning algorithm, inspired by pseudo-likelihood [1], that optimizes over an outer bound of Θ. Learning involves optimizing over only a small number of constraints per data point, and thus can be performed quickly, even for complex structured prediction models. We show that, if the data available for learning is “nice”, this algorithm is consistent, i.e. it will find some θ ∈ Θ. This is an example of how having the right data can circumvent the hardness of learning for structured prediction. 1 We also investigate the limitations of the proposed method. We show that the problem of even deciding whether a given data set is separable is NP-hard, and thus learning in a strict sense is no easier than prediction. Thus, we should not expect for our algorithm, or any other polynomial time algorithm, to always succeed at learning from an arbitrary finite data set. To our knowledge, this is the first result characterizing the hardness of exact learning for structured prediction. Finally, we show empirically that our algorithm allows us to successfully learn the parameters for both multi-label prediction and protein side-chain placement. The performance of the algorithm is improved as more data becomes available, as our theoretical results anticipate. 1 Pseudo-Max method We consider the general structured prediction problem. The input space is denoted by X and the set of all possible assignments by Y. Each y ∈ Y corresponds to n variables y1 , . . . , yn , each with k possible states. The classifier uses a (given) function φ(x, y) : X , Y → Rd and (learned) weights θ ∈ Rd , and is defined as y(x; θ) = arg maxy∈Y f (ˆ ; x, θ) where f is the discriminant function y ˆ f (y; x, θ) = θ · φ(x, y). Our analysis will focus on functions φ whose scope is limited to small sets of the yi variables, but for now we keep the discussion general. Given a set of labeled examples {(xm , y m )}M , the goal of the typical learning problem is to find m=1 weights θ that correctly classify the training examples. Consider first the separable case. Define the set of separating weight vectors, Θ = θ | ∀m, y ∈ Y, f (y m ; xm , θ) ≥ f (y; xm , θ)+e(y, y m ) . e is a loss function (e.g., zero-one or Hamming) such that e(y m , y m ) = 0 and e(y, y m ) > 0 for y = y m , which serves to rule out the trivial solution θ = 0.1 The space Θ is defined by exponentially many constraints per example, one for each competing assignment. In this work we consider a much simpler set of constraints where, for each example, we only consider the competing assignments obtained by modifying a single label yi , while fixing the other labels to their value at y m . The pseudo-max set, which is an outer bound on Θ, is given by Here ym −i m Θps = θ | ∀m, i, yi , f (y m ; xm , θ) ≥ f (y m , yi ; xm , θ) + e(yi , yi ) . −i denotes the label y m (1) without the assignment to yi . When the data is not separable, Θ will be the empty set. Instead, we may choose to minimize the hinge loss, (θ) = m maxy f (y; xm , θ) − f (y m ; xm , θ) + e(y, y m ) , which can be shown to be an upper bound on the training error [13]. When the data is separable, minθ (θ) = 0. Note that regularization may be added to this objective. The corresponding pseudo-max objective replaces the maximization over all of y with maximization over a single variable yi while fixing the other labels to their value at y m :2,3 M ps (θ) n = m=1 i=1 m max f (y m , yi ; xm , θ) − f (y m ; xm , θ) + e(yi , yi ) . −i yi Analogous to before, we have minθ ps (θ) (2) = 0 if and only if θ ∈ Θps . The objective in Eq. 2 is similar in spirit to pseudo-likelihood objectives used for maximum likelihood estimation of parameters of Markov random fields (MRFs) [1]. The pseudo-likelihood estimate is provably consistent when the data generating distribution is a MRF of the same structure as used in the pseudo-likelihood objective. However, our setting is different since we only get to view the maximizing assignment of the MRF rather than samples from it. Thus, a particular x will always be paired with the same y rather than samples y drawn from the conditional distribution p(y|x; θ). The pseudo-max constraints in Eq. 1 are also related to cutting plane approaches to inference [4, 5]. In the latter, the learning problem is solved by repeatedly looking for assignments that violate the separability constraint (or its hinge version). Our constraints can be viewed as using a very small 1 An alternative formulation, which we use in the next section, is to break the symmetry by having part of the input not be multiplied by any weight. This will also rule out the trivial solution θ = 0. P 2 It is possible to use maxi instead of i , and some of our consistency results will still hold. 3 The pseudo-max approach is markedly different from a learning method which predicts each label yi independently, since the objective considers all i simultaneously (both at learning and test time). 2 x2 0.2 J ∗ + x1 = 0 y = (0, 1) y = (1, 1) g(J12) x2 = 0 x1 J ∗ + x1 + x2 = 0 y = (0, 0) c1=0 c1=1 c1= 1 0.15 0.1 J + x2 = 0 ∗ 0.05 y = (1, 0) x1 = 0 0 1 0.5 0 J 0.5 1 Figure 1: Illustrations for a model with two variables. Left: Partitioning of X induced by configurations y(x) for some J ∗ > 0. Blue lines carve out the exact regions. Red lines denote the pseudo-max constraints that hold with equality. Pseudo-max does not obtain the diagonal constraint coming from comparing configurations y = (1, 1) and (0, 0), since these differ by more than one coordinate. Right: One strictly-convex component of the ps (J ) function (see Eq. 9). The function is shown for different values of c1 , the mean of the x1 variable. subset of assignments for the set of candidate constraint violators. We also note that when exact maximization over the discriminant function f (y; x, θ) is hard, the standard cutting plane algorithm cannot be employed since it is infeasible to find a violated constraint. For the pseudo-max objective, finding a constraint violation is simple and linear in the number of variables.4 It is easy to see (as will be elaborated on next) that the pseudo-max method does not in general yield a consistent estimate of θ, even in the separable case. However, as we show, consistency can be shown to be achieved under particular assumptions on the data generating distribution p(x). 2 Consistency of the Pseudo-Max method In this section we show that if the feature generating distribution p(x) satisfies particular assumptions, then the pseudo-max approach yields a consistent estimate. In other words, if the training data is of the form {(xm , y(xm ; θ ∗ ))}M for some true parameter vector θ ∗ , then as M → ∞ the m=1 minimum of the pseudo-max objective will converge to θ ∗ (up to equivalence transformations). The section is organized as follows. First, we provide intuition for the consistency results by considering a model with only two variables. Then, in Sec. 2.1, we show that any parameter θ ∗ can be identified to within arbitrary accuracy by choosing a particular training set (i.e., choice of xm ). This in itself proves consistency, as long as there is a non-zero probability of sampling this set. In Sec. 2.2 we give a more direct proof of consistency by using strict convexity arguments. For ease of presentation, we shall work with a simplified instance of the structured learning setting. We focus on binary variables, yi ∈ {0, 1}, and consider discriminant functions corresponding to Ising models, a special case of pairwise MRFs (J denotes the vector of “interaction” parameters): f (y; x, J ) = ij∈E Jij yi yj + i yi xi (3) The singleton potential for variable yi is yi xi and is not dependent on the model parameters. We could have instead used Ji yi xi , which would be more standard. However, this would make the parameter vector J invariant to scaling, complicating the identifiability analysis. In the consistency analysis we will assume that the data is generated using a true parameter vector J ∗ . We will show that as the data size goes to infinity, minimization of ps (J ) yields J ∗ . We begin with an illustrative analysis of the pseudo-max constraints for a model with only two variables, i.e. f (y; x, J) = Jy1 y2 + y1 x1 + y2 x2 . The purpose of the analysis is to demonstrate general principles for when pseudo-max constraints may succeed or fail. Assume that training samples are generated via y(x) = argmaxy f (y; x, J ∗ ). We can partition the input space X into four regions, ˆ ˆ {x ∈ X : y(x) = y } for each of the four configurations y , shown in Fig. 1 (left). The blue lines outline the exact decision boundaries of f (y; x, J ∗ ), with the lines being given by the constraints 4 The methods differ substantially in the non-separable setting where we minimize ps (θ), using a slack variable for every node and example, rather than just one slack variable per example as in (θ). 3 in Θ that hold with equality. The red lines denote the pseudo-max constraints in Θps that hold with equality. For x such that y(x) = (1, 0) or (0, 1), the pseudo-max and exact constraints are identical. We can identify J ∗ by obtaining samples x = (x1 , x2 ) that explore both sides of one of the decision boundaries that depends on J ∗ . The pseudo-max constraints will fail to identify J ∗ if the samples do not sufficiently explore the transitions between y = (0, 1) and y = (1, 1) or between y = (1, 0) and y = (1, 1). This can happen, for example, when the input samples are dependent, giving only rise to the configurations y = (0, 0) and y = (1, 1). For points labeled (1, 1) around the decision line J ∗ + x1 + x2 = 0, pseudo-max can only tell that they respect J ∗ + x1 ≥ 0 and J ∗ + x2 ≥ 0 (dashed red lines), or x1 ≤ 0 and x2 ≤ 0 for points labeled (0, 0). Only constraints that depend on the parameter are effective for learning. For pseudo-max to be able to identify J ∗ , the input samples must be continuous, densely populating the two parameter dependent decision lines that pseudo-max can use. The two point sets in the figure illustrate good and bad input distributions for pseudo-max. The diagonal set would work well with the exact constraints but badly with pseudo-max, and the difference can be arbitrarily large. However, the input distribution on the right, populating the J ∗ + x2 = 0 decision line, would permit pseudo-max to identify J ∗ . 2.1 Identifiability of True Parameters In this section, we show that it is possible to approximately identify the true model parameters, up to model equivalence, using the pseudo-max constraints and a carefully chosen linear number of data points. Consider the learning problem for structured prediction defined on a fixed graph G = (V, E) where the parameters to be learned are pairwise potential functions θij (yi , yj ) for ij ∈ E and single node fields θi (yi ) for i ∈ V . We consider discriminant functions of the form f (y; x, θ) = ij∈E θij (yi , yj ) + i θi (yi ) + i xi (yi ), (4) where the input space X = R|V |k specifies the single node potentials. Without loss of generality, we remove the additional degrees of freedom in θ by restricting it to be in a canonical form: θ ∈ Θcan if for all edges θij (yi , yj ) = 0 whenever yi = 0 or yj = 0, and if for all nodes, θi (yi ) = 0 when yi = 0. As a result, assuming the training set comes from a model in this class, and the input fields xi (yi ) exercise the discriminant function appropriately, we can hope to identify θ ∗ ∈ Θcan . Indeed, we show that, for some data sets, the pseudo-max constraints are sufficient to identify θ ∗ . Let Θps ({y m , xm }) be the set of parameters that satisfy the pseudo-max classification constraints m Θps ({y m , xm }) = θ | ∀m, i, yi = yi , f (y m ; xm , θ) ≥ f (y m , yi ; xm , θ) . −i (5) m e(yi , yi ), For simplicity we omit the margin losses since the input fields xi (yi ) already suffice to rule out the trivial solution θ = 0. Proposition 2.1. For any θ ∗ ∈ Θcan , there is a set of 2|V |(k − 1) + 2|E|(k − 1)2 examples, {xm , y(xm ; θ ∗ )}, such that any pseudo-max consistent θ ∈ Θps ({y m , xm }) ∩ Θcan is arbitrarily close to θ ∗ . The proof is given in the supplementary material. To illustrate the key ideas, we consider the simpler binary discriminant function discussed in Eq. 3. Note that the binary model is already in the canonical form since Jij yi yj = 0 whenever yi = 0 or yj = 0. For any ij ∈ E, we show how to choose two input examples x1 and x2 such that any J consistent with the pseudo-max constraints for these ∗ ∗ two examples will have Jij ∈ [Jij − , Jij + ]. Repeating this for all of the edge parameters then gives the complete set of examples. The input examples we need for this will depend on J ∗ . For the first example, we set the input fields for all neighbors of i (except j) in such a way that ∗ we force the corresponding labels to be zero. More formally, we set x1 < −|N (k)| maxl |Jkl | for k 1 k ∈ N (i)\j, resulting in yk = 0, where y 1 = y(x1 ). In contrast, we set x1 to a large value, e.g. j ∗ 1 ∗ x1 > |N (j)| maxl |Jjl |, so that yj = 1. Finally, for node i, we set x1 = −Jij + so as to obtain a j i 1 slight preference for yi = 1. All other input fields can be set arbitrarily. As a result, the pseudo-max constraints pertaining to node i are f (y 1 ; x1 , J ) ≥ f (y 1 , yi ; x1 , J ) for yi = 0, 1. By taking into −i 1 account the label assignments for yi and its neighbors, and by removing terms that are the same on both sides of the equation, we get Jij + x1 + x1 ≥ Jij yi + yi x1 + x1 , which, for yi = 0, implies i j i j ∗ that Jij + x1 ≥ 0 or Jij − Jij + ≥ 0. The second example x2 differs only in terms of the input i ∗ 2 ∗ field for i. In particular, we set x2 = −Jij − so that yi = 0. This gives Jij ≤ Jij + , as desired. i 4 2.2 Consistency via Strict Convexity In this section we prove the consistency of the pseudo-max approach by showing that it corresponds to minimizing a strictly convex function. Our proof only requires that p(x) be non-zero for all x ∈ Rn (a simple example being a multi-variate Gaussian) and that J ∗ is finite. We use a discriminant function as in Eq. 3. Now, assume the input points xm are distributed according to p(x) and that y m are obtained via y m = arg maxy f (y; xm , J ∗ ). We can write the ps (J ) objective for finite data, and its limit when M → ∞, compactly as: 1 m m = max (yi − yi ) xm + Jki yk ps (J ) i M m i yi k∈N (i) p(x) max (yi − yi (x)) xi + → yi i Jki yk (x) dx (6) k∈N (i) ∗ where yi (x) is the label of i for input x when using parameters J . Starting from the above, consider the terms separately for each i. We partition the integral over x ∈ Rn into exclusive regions according to the predicted labels of the neighbors of i (given x). Define Sij = {x : yj (x) = 1 and yk (x) = 0 for k ∈ N (i)\j}. Eq. 6 can then be written as ps (J ) = gi ({Jik }k∈N (i) ) + ˆ i gik (Jik ) , (7) k∈N (i) where gik (Jik ) = x∈Sik p(x) maxyi [(yi −yi (x))(xi +Jik )]dx and gi ({Jik }k∈N (i) ) contains all of ˆ the remaining terms, i.e. where either zero or more than one neighbor is set to one. The function gi ˆ is convex in J since it is a sum of integrals over convex functions. We proceed to show that gik (Jik ) is strictly convex for all choices of i and k ∈ N (i). This will show that ps (J ) is strictly convex since it is a sum over functions strictly convex in each one of the variables in J . For all values xi ∈ (−∞, ∞) there is some x in Sij . This is because for any finite xi and finite J ∗ , the other xj ’s can be chosen so as to give the y configuration corresponding to Sij . Now, since p(x) has full support, we have P (Sij ) > 0 and p(x) > 0 for any x in Sij . As a result, this also holds for the marginal pi (xi |Sij ) over xi within Sij . After some algebra, we obtain: gij (Jij ) = P (Sij ) ∞ p(x)yi (x)(xi + Jij )dx pi (xi |Sij ) max [0, xi + Jij ] dxi − −∞ x∈Sij The integral over the yi (x)(xi + Jij ) expression just adds a linear term to gij (Jij ). The relevant remaining term is (for brevity we drop P (Sij ), a strictly positive constant, and the ij index): h(J) = ∞ pi (xi |Sij ) max [0, xi + J] dxi = −∞ ∞ ˆ pi (xi |Sij )h(xi , J)dxi (8) −∞ ˆ ˆ where we define h(xi , J) = max [0, xi + J]. Note that h(J) is convex since h(xi , J) is convex in J for all xi . We want to show that h(J) is strictly convex. Consider J < J and α ∈ (0, 1) and define ˆ ˆ the interval I = [−J, −αJ − (1 − α)J ]. For xi ∈ I it holds that: αh(xi , J) + (1 − α)h(xi , J ) > ˆ i , αJ + (1 − α)J ) (since the first term is strictly positive and the rest are zero). For all other x, h(x ˆ this inequality holds but is not necessarily strict (since h is always convex in J). We thus have after integrating over x that αh(J) + (1 − α)h(J ) > h(αJ + (1 − α)J ), implying h is strictly convex, as required. Note that we used the fact that p(x) has full support when integrating over I. The function ps (J ) is thus a sum of strictly convex functions in all its variables (namely g(Jik )) plus other convex functions of J , hence strictly convex. We can now proceed to show consistency. By strict convexity, the pseudo-max objective is minimized at a unique point J . Since we know that ps (J ∗ ) = 0 and zero is a lower bound on the value of ps (J ), it follows that J ∗ is the unique minimizer. Thus we have that as M → ∞, the minimizer of the pseudo-max objective is the true parameter vector, and thus we have consistency. As an example, consider the case of two variables y1 , y2 , with x1 and x2 distributed according to ∗ N (c1 , 1), N (0, 1) respectively. Furthermore assume J12 = 0. Then simple direct calculation yields: 2 2 2 c1 + J12 −c1 1 1 √ (9) e−x /2 dx − √ e−c1 /2 + √ e−(J12 +c1 ) /2 2π 2π 2π −J12 −c1 which is indeed a strictly convex function that is minimized at J = 0 (see Fig. 1 for an illustration). g(J12 ) = 5 3 Hardness of Structured Learning Most structured prediction learning algorithms use some form of inference as a subroutine. However, the corresponding prediction task is generally NP-hard. For example, maximizing the discriminant function defined in Eq. 3 is equivalent to solving Max-Cut, which is known to be NP-hard. This raises the question of whether it is possible to bypass prediction during learning. Although prediction may be intractable for arbitrary MRFs, what does this say about the difficulty of learning with a polynomial number of data points? In this section, we show that the problem of deciding whether there exists a parameter vector that separates the training data is NP-hard. Put in the context of the positive results in this paper, these hardness results show that, although in some cases the pseudo-max constraints yield a consistent estimate, we cannot hope for a certificate of optimality. Put differently, although the pseudo-max constraints in the separable case always give an outer bound on Θ (and may even be a single point), Θ could be the empty set – and we would never know the difference. Theorem 3.1. Given labeled examples {(xm , y m )}M for a fixed but arbitrary graph G, it is m=1 NP-hard to decide whether there exists parameters θ such that ∀m, y m = arg maxy f (y; xm , θ). Proof. Any parameters θ have an equivalent parameterization in canonical form (see section Sec. 2.1, also supplementary). Thus, the examples will be separable if and only if they are separable by some θ ∈ Θcan . We reduce from unweighted Max-Cut. The Max-Cut problem is to decide, given an undirected graph G, whether there exists a cut of at least K edges. Let G be the same graph as G, with k = 3 states per variable. We construct a small set of examples where a parameter vector will exist that separates the data if and only if there is no cut of K or more edges in G. Let θ be parameters in canonical form equivalent to θij (yi , yj ) = 1 if (yi , yj ) ∈ {(1, 2), (2, 1)}, 0 if yi = yj , and −n2 if (yi , yj ) ∈ {(1, 3), (2, 3), (3, 1), (3, 2)}. We first construct 4n + 8|E| examples, using the technique described in Sec. 2.1 (also supplementary material), which when restricted to the space Θcan , constrain the parameters to equal θ. We then use one more example (xm , y m ) where y m = 3 (every node is in state 3) and, for all i, xm (3) = K−1 and xm (1) = xm (2) = 0. The first i i i n two states encode the original Max-Cut instance, while the third state is used to construct a labeling y m that has value equal to K − 1, and is otherwise not used. Let K ∗ be the value of the maximum cut in G. If in any assignment to the last example there is a variable taking the state 3 and another variable taking the state 1 or 2, then the assignment’s value will be at most K ∗ − n2 , which is less than zero. By construction, the 3 assignment has value K − 1. Thus, the optimal assignment must either be 3 with value K − 1, or some combination of states 1 and 2, which has value at most K ∗ . If K ∗ > K − 1 then 3 is not optimal and the examples are not separable. If K ∗ ≤ K − 1, the examples are separable. This result illustrates the potential difficulty of learning in worst-case graphs. Nonetheless, many problems have a more restricted dependence on the input. For example, in computer vision, edge potentials may depend only on the difference in color between two adjacent pixels. Our results do not preclude positive results of learnability in such restricted settings. By establishing hardness of learning, we also close the open problem of relating hardness of inference and learning in structured prediction. If inference problems can be solved in polynomial time, then so can learning (using, e.g., structured perceptron). Thus, when learning is hard, inference must be hard as well. 4 Experiments To evaluate our learning algorithm, we test its performance on both synthetic and real-world datasets. We show that, as the number of training samples grows, the accuracy of the pseudo-max method improves and its speed-up gain over competing algorithms increases. Our learning algorithm corresponds to solving the following, where we add L2 regularization and use a scaled 0-1 loss, m m e(yi , yi ) = 1{yi = yi }/nm (nm is the number of labels in example m): min θ C m nm M nm m=1 i=1 m max f (y m , yi ; xm , θ) − f (y m ; xm , θ) + e(yi , yi ) + θ −i yi 2 . (10) We will compare the pseudo-max method with learning using structural SVMs, both with exact inference and LP relaxations [see, e.g., 4]. We use exact inference for prediction at test time. 6 (a) Synthetic (b) Reuters 0.4 exact LP−relaxation pseudo−max 0.15 Test error Test error 0.2 0.1 0.05 0 1 10 2 10 0.2 0.1 0 1 10 3 10 Train size exact LP−relaxation pseudo−max 0.3 2 10 3 10 4 10 Train size Figure 2: Test error as a function of train size for various algorithms. Subfigure (a) shows results for a synthetic setting, while (b) shows performance on the Reuters data. In the synthetic setting we use the discriminant function f (y; x, θ) = ij∈E θij (yi , yj ) + xi θi (yi ), which is similar to Eq. 4. We take a fully connected graph over n = 10 binary labels. i For a weight vector θ ∗ (sampled once, uniformly in the range [−1, 1], and used for all train/test sets) we generate train and test instances by sampling xm uniformly in the range [−5, 5] and then computing the optimal labels y m = arg maxy∈Y f (y; xm , θ ∗ ). We generate train sets of increasing size (M = {10, 50, 100, 500, 1000, 5000}), run the learning algorithms, and measure the test error for the learned weights (with 1000 test samples). For each train size we average the test error over 10 repeats of sampling and training. Fig. 2(a) shows a comparison of the test error for the three learning algorithms. For small numbers of training examples, the test error of pseudo-max is larger than that of the other algorithms. However, as the train size grows, the error converges to that of exact learning, as our consistency results predict. We also test the performance of our algorithm on a multi-label document classification task from the Reuters dataset [7]. The data consists of M = 23149 training samples, and we use a reduction of the dataset to the 5 most frequent labels. The 5 label variables form a fully connected pairwise graph structure (see [4] for a similar setting). We use random subsamples of increasing size from the train set to learn the parameters, and then measure the test error using 20000 additional samples. For each sample size and learning algorithm, we optimize the trade-off parameter C using 30% of the training data as a hold-out set. Fig. 2(b) shows that for the large data regime the performance of pseudo-max learning gets close to that of the other methods. However, unlike the synthetic setting there is still a small gap, even after seeing the entire train set. This could be because the full dataset is not yet large enough to be in the consistent regime (note that exact learning has not flattened either), or because the consistency conditions are not fully satisfied: the data might be non-separable or the support of the input distribution p(x) may be partial. We next apply our method to the problem of learning the energy function for protein side-chain placement, mirroring the learning setup of [14], where the authors train a conditional random field (CRF) using tree-reweighted belief propagation to maximize a lower bound on the likelihood.5 The prediction problem for side-chain placement corresponds to finding the most likely assignment in a pairwise MRF, and fits naturally into our learning framework. There are only 8 parameters to be learned, corresponding to a reweighting of known energy terms. The dataset consists of 275 proteins, where each MRF has several hundred variables (one per residue of the protein) and each variable has on average 20 states. For prediction we use CPLEX’s ILP solver. Fig. 3 shows a comparison of the pseudo-max method and a cutting-plane algorithm which uses an LP relaxation, solved with CPLEX, for finding violated constraints.6 We generate training sets of increasing size (M = {10, 50, 100, 274}), and measure the test error for the learned weights on the remaining examples.7 For M = 10, 50, 100 we average the test error over 3 random train/test splits, whereas for M = 274 we do 1-fold cross validation. We use C = 1 for both algorithms. 5 The authors’ data and results are available from: http://cyanover.fhcrc.org/recomb-2007/ We significantly optimized the cutting-plane algorithm, e.g. including a large number of initial cuttingplanes and restricting the weight vector to be positive (which we know to hold at optimality). 7 Specifically, for each protein we compute the fraction of correctly predicted χ1 and χ2 angles for all residues (except when trivial, e.g. just 1 state). Then, we compute the median of this value across all proteins. 6 7 Time to train (minutes) Test error (χ1 and χ2) 0.27 0.265 pseudo−max LP−relaxation Soft rep 0.26 0.255 0.25 0 50 100 150 200 Train size 250 250 200 pseudo−max LP−relaxation 150 100 50 0 0 50 100 150 200 Train size 250 Figure 3: Training time (for one train/test split) and test error as a function of train size for both the pseudomax method and a cutting-plane algorithm which uses a LP relaxation for inference, applied to the problem of learning the energy function for protein side-chain placement. The pseudo-max method obtains better accuracy than both the LP relaxation and HCRF (given roughly five times more data) for a fraction of the training time. The original weights (“Soft rep” [3]) used for this energy function have 26.7% error across all 275 proteins. The best previously reported parameters, learned in [14] using a Hidden CRF, obtain 25.6% error (their training set included 55 of these 275 proteins, so this is an optimistic estimate). To get a sense of the difficulty of this learning task, we also tried a random positive weight vector, uniformly sampled from the range [0, 1], obtaining an error of 34.9% (results would be much worse if we allowed the weights to be negative). Training using pseudo-max with 50 examples, we learn parameters in under a minute that give better accuracy than the HCRF. The speed-up of training with pseudo-max (using CPLEX’s QP solver) versus cutting-plane is striking. For example, for M = 10, pseudo-max takes only 3 seconds, a 1000-fold speedup. Unfortunately the cutting-plane algorithm took a prohibitive amount of time to be able to run on the larger training sets. Since the data used in learning for protein side-chain placement is both highly non-separable and relatively little, these positive results illustrate the potential wide-spread applicability of the pseudo-max method. 5 Discussion The key idea of our method is to find parameters that prefer the true assignment y m over assignments that differ from it in only one variable, in contrast to all other assignments. Perhaps surprisingly, this weak requirement is sufficient to achieve consistency given a rich enough input distribution. One extension of our approach is to add constraints for assignments that differ from y m in more than one variable. This would tighten the outer bound on Θ and possibly result in improved performance, but would also increase computational complexity. We could also add such competing assignments via a cutting-plane scheme so that optimization is performed only over a subset of these constraints. Our work raises a number of important open problems: It would be interesting to derive generalization bounds to understand the convergence rate of our method, as well as understanding the effect of the distribution p(x) on these rates. The distribution p(x) needs to have two key properties. On the one hand, it needs to explore the space Y in the sense that a sufficient number of labels need to be obtained as the correct label for the true parameters (this is indeed used in our consistency proofs). On the other hand, p(x) needs to be sufficiently sensitive close to the decision boundaries so that the true parameters can be inferred. We expect that generalization analysis will depend on these two properties of p(x). Note that [11] studied active learning schemes for structured data and may be relevant in the current context. How should one apply this learning algorithm to non-separable data sets? We suggested one approach, based on using a hinge loss for each of the pseudo constraints. One question in this context is, how resilient is this learning algorithm to label noise? Recent work has analyzed the sensitivity of pseudo-likelihood methods to model mis-specification [8], and it would be interesting to perform a similar analysis here. Also, is it possible to give any guarantees for the empirical and expected risks (with respect to exact inference) obtained by outer bound learning versus exact learning? Finally, our algorithm demonstrates a phenomenon where more data can make computation easier. Such a scenario was recently analyzed in the context of supervised learning [12], and it would be interesting to combine the approaches. Acknowledgments: We thank Chen Yanover for his assistance with the protein data. This work was supported by BSF grant 2008303 and a Google Research Grant. D.S. was supported by a Google PhD Fellowship. 8 References [1] J. Besag. The analysis of non-lattice data. The Statistician, 24:179–195, 1975. [2] M. Collins. Discriminative training methods for hidden Markov models: Theory and experiments with perceptron algorithms. In EMNLP, 2002. [3] G. Dantas, C. Corrent, S. L. Reichow, J. J. Havranek, Z. M. Eletr, N. G. Isern, B. Kuhlman, G. Varani, E. A. Merritt, and D. Baker. High-resolution structural and thermodynamic analysis of extreme stabilization of human procarboxypeptidase by computational protein design. Journal of Molecular Biology, 366(4):1209 – 1221, 2007. [4] T. Finley and T. Joachims. Training structural SVMs when exact inference is intractable. In Proceedings of the 25th International Conference on Machine Learning 25, pages 304–311. ACM, 2008. [5] T. Joachims, T. Finley, and C.-N. Yu. Cutting-plane training of structural SVMs. Machine Learning, 77(1):27–59, 2009. [6] A. Kulesza and F. Pereira. Structured learning with approximate inference. In Advances in Neural Information Processing Systems 20, pages 785–792. 2008. [7] D. Lewis, , Y. Yang, T. Rose, and F. Li. RCV1: a new benchmark collection for text categorization research. JMLR, 5:361–397, 2004. [8] P. Liang and M. I. Jordan. An asymptotic analysis of generative, discriminative, and pseudolikelihood estimators. In Proceedings of the 25th international conference on Machine learning, pages 584–591, New York, NY, USA, 2008. ACM Press. [9] A. F. T. Martins, N. A. Smith, and E. P. Xing. Polyhedral outer approximations with application to natural language parsing. In ICML 26, pages 713–720, 2009. [10] N. Ratliff, J. A. D. Bagnell, and M. Zinkevich. (Online) subgradient methods for structured prediction. In AISTATS, 2007. [11] D. Roth and K. Small. Margin-based active learning for structured output spaces. In Proc. of the European Conference on Machine Learning (ECML). Springer, September 2006. [12] S. Shalev-Shwartz and N. Srebro. SVM optimization: inverse dependence on training set size. In Proceedings of the 25th international conference on Machine learning, pages 928–935. ACM, 2008. [13] B. Taskar, C. Guestrin, and D. Koller. Max margin Markov networks. In Advances in Neural Information Processing Systems 16, pages 25–32. 2004. [14] C. Yanover, O. Schueler-Furman, and Y. Weiss. Minimizing and learning energy functions for side-chain prediction. Journal of Computational Biology, 15(7):899–911, 2008. 9

6 0.47936609 234 nips-2010-Segmentation as Maximum-Weight Independent Set

7 0.47721955 77 nips-2010-Epitome driven 3-D Diffusion Tensor image segmentation: on extracting specific structures

8 0.43464977 1 nips-2010-(RF)^2 -- Random Forest Random Field

9 0.43030229 84 nips-2010-Exact inference and learning for cumulative distribution functions on loopy graphs

10 0.42279333 101 nips-2010-Gaussian sampling by local perturbations

11 0.41841185 58 nips-2010-Decomposing Isotonic Regression for Efficiently Solving Large Problems

12 0.39242682 83 nips-2010-Evidence-Specific Structures for Rich Tractable CRFs

13 0.38856837 270 nips-2010-Tight Sample Complexity of Large-Margin Learning

14 0.38057372 90 nips-2010-Fast Large-scale Mixture Modeling with Component-specific Data Partitions

15 0.37526774 215 nips-2010-Probabilistic Deterministic Infinite Automata

16 0.3737947 126 nips-2010-Inference with Multivariate Heavy-Tails in Linear Models

17 0.36589199 148 nips-2010-Learning Networks of Stochastic Differential Equations

18 0.36380047 102 nips-2010-Generalized roof duality and bisubmodular functions

19 0.35262764 108 nips-2010-Graph-Valued Regression

20 0.33777922 239 nips-2010-Sidestepping Intractable Inference with Structured Ensemble Cascades


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(13, 0.042), (17, 0.01), (27, 0.049), (30, 0.099), (35, 0.014), (45, 0.179), (50, 0.059), (52, 0.035), (60, 0.068), (72, 0.308), (77, 0.042), (90, 0.025)]

similar papers list:

simIndex simValue paperId paperTitle

1 0.78141046 129 nips-2010-Inter-time segment information sharing for non-homogeneous dynamic Bayesian networks

Author: Dirk Husmeier, Frank Dondelinger, Sophie Lebre

Abstract: Conventional dynamic Bayesian networks (DBNs) are based on the homogeneous Markov assumption, which is too restrictive in many practical applications. Various approaches to relax the homogeneity assumption have recently been proposed, allowing the network structure to change with time. However, unless time series are very long, this flexibility leads to the risk of overfitting and inflated inference uncertainty. In the present paper we investigate three regularization schemes based on inter-segment information sharing, choosing different prior distributions and different coupling schemes between nodes. We apply our method to gene expression time series obtained during the Drosophila life cycle, and compare the predicted segmentation with other state-of-the-art techniques. We conclude our evaluation with an application to synthetic biology, where the objective is to predict a known in vivo regulatory network of five genes in yeast. 1

same-paper 2 0.76762736 288 nips-2010-Worst-case bounds on the quality of max-product fixed-points

Author: Meritxell Vinyals, Jes\'us Cerquides, Alessandro Farinelli, Juan A. Rodríguez-aguilar

Abstract: We study worst-case bounds on the quality of any fixed point assignment of the max-product algorithm for Markov Random Fields (MRF). We start providing a bound independent of the MRF structure and parameters. Afterwards, we show how this bound can be improved for MRFs with specific structures such as bipartite graphs or grids. Our results provide interesting insight into the behavior of max-product. For example, we prove that max-product provides very good results (at least 90% optimal) on MRFs with large variable-disjoint cycles1 . 1

3 0.72219342 213 nips-2010-Predictive Subspace Learning for Multi-view Data: a Large Margin Approach

Author: Ning Chen, Jun Zhu, Eric P. Xing

Abstract: Learning from multi-view data is important in many applications, such as image classification and annotation. In this paper, we present a large-margin learning framework to discover a predictive latent subspace representation shared by multiple views. Our approach is based on an undirected latent space Markov network that fulfills a weak conditional independence assumption that multi-view observations and response variables are independent given a set of latent variables. We provide efficient inference and parameter estimation methods for the latent subspace model. Finally, we demonstrate the advantages of large-margin learning on real video and web image data for discovering predictive latent representations and improving the performance on image classification, annotation and retrieval.

4 0.66560626 270 nips-2010-Tight Sample Complexity of Large-Margin Learning

Author: Sivan Sabato, Nathan Srebro, Naftali Tishby

Abstract: We obtain a tight distribution-specific characterization of the sample complexity of large-margin classification with L2 regularization: We introduce the γ-adapted-dimension, which is a simple function of the spectrum of a distribution’s covariance matrix, and show distribution-specific upper and lower bounds on the sample complexity, both governed by the γ-adapted-dimension of the source distribution. We conclude that this new quantity tightly characterizes the true sample complexity of large-margin classification. The bounds hold for a rich family of sub-Gaussian distributions. 1

5 0.59071183 260 nips-2010-Sufficient Conditions for Generating Group Level Sparsity in a Robust Minimax Framework

Author: Hongbo Zhou, Qiang Cheng

Abstract: Regularization technique has become a principled tool for statistics and machine learning research and practice. However, in most situations, these regularization terms are not well interpreted, especially on how they are related to the loss function and data. In this paper, we propose a robust minimax framework to interpret the relationship between data and regularization terms for a large class of loss functions. We show that various regularization terms are essentially corresponding to different distortions to the original data matrix. This minimax framework includes ridge regression, lasso, elastic net, fused lasso, group lasso, local coordinate coding, multiple kernel learning, etc., as special cases. Within this minimax framework, we further give mathematically exact definition for a novel representation called sparse grouping representation (SGR), and prove a set of sufficient conditions for generating such group level sparsity. Under these sufficient conditions, a large set of consistent regularization terms can be designed. This SGR is essentially different from group lasso in the way of using class or group information, and it outperforms group lasso when there appears group label noise. We also provide some generalization bounds in a classification setting. 1

6 0.59029102 193 nips-2010-Online Learning: Random Averages, Combinatorial Parameters, and Learnability

7 0.5889641 155 nips-2010-Learning the context of a category

8 0.58847225 88 nips-2010-Extensions of Generalized Binary Search to Group Identification and Exponential Costs

9 0.58846587 222 nips-2010-Random Walk Approach to Regret Minimization

10 0.58832014 75 nips-2010-Empirical Risk Minimization with Approximations of Probabilistic Grammars

11 0.58766866 1 nips-2010-(RF)^2 -- Random Forest Random Field

12 0.58740038 148 nips-2010-Learning Networks of Stochastic Differential Equations

13 0.58708739 280 nips-2010-Unsupervised Kernel Dimension Reduction

14 0.58692169 220 nips-2010-Random Projection Trees Revisited

15 0.58667743 87 nips-2010-Extended Bayesian Information Criteria for Gaussian Graphical Models

16 0.58649331 117 nips-2010-Identifying graph-structured activation patterns in networks

17 0.5862565 63 nips-2010-Distributed Dual Averaging In Networks

18 0.5842427 49 nips-2010-Computing Marginal Distributions over Continuous Markov Networks for Statistical Relational Learning

19 0.58395851 51 nips-2010-Construction of Dependent Dirichlet Processes based on Poisson Processes

20 0.58360076 92 nips-2010-Fast global convergence rates of gradient methods for high-dimensional statistical recovery