nips nips2009 nips2009-142 knowledge-graph by maker-knowledge-mining

142 nips-2009-Locality-sensitive binary codes from shift-invariant kernels


Source: pdf

Author: Maxim Raginsky, Svetlana Lazebnik

Abstract: This paper addresses the problem of designing binary codes for high-dimensional data such that vectors that are similar in the original space map to similar binary strings. We introduce a simple distribution-free encoding scheme based on random projections, such that the expected Hamming distance between the binary codes of two vectors is related to the value of a shift-invariant kernel (e.g., a Gaussian kernel) between the vectors. We present a full theoretical analysis of the convergence properties of the proposed scheme, and report favorable experimental performance as compared to a recent state-of-the-art method, spectral hashing.

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 edu Abstract This paper addresses the problem of designing binary codes for high-dimensional data such that vectors that are similar in the original space map to similar binary strings. [sent-5, score-0.318]

2 We introduce a simple distribution-free encoding scheme based on random projections, such that the expected Hamming distance between the binary codes of two vectors is related to the value of a shift-invariant kernel (e. [sent-6, score-0.567]

3 We present a full theoretical analysis of the convergence properties of the proposed scheme, and report favorable experimental performance as compared to a recent state-of-the-art method, spectral hashing. [sent-9, score-0.205]

4 1 Introduction Recently, there has been a lot of interest in the problem of designing compact binary codes for reducing storage requirements and accelerating search and retrieval in large collections of highdimensional vector data [11, 13, 15]. [sent-10, score-0.295]

5 A desirable property of such coding schemes is that they should map similar data points to similar binary strings, i. [sent-11, score-0.177]

6 Hamming distances can be computed very efficiently in hardware, resulting in very fast retrieval of strings similar to a given query, even for brute-force search in a database consisting of millions of data points [11, 13]. [sent-14, score-0.333]

7 Moreover, if code strings can be effectively used as hash keys, then similarity searches can be carried out in sublinear time. [sent-15, score-0.404]

8 [11, 13], the notion of similarity between data points comes from supervisory information, e. [sent-18, score-0.09]

9 The binary encoder is then trained to reproduce this “semantic” similarity measure. [sent-21, score-0.129]

10 In this paper, we are more interested in unsupervised schemes, where the similarity is given by Euclidean distance or by a kernel defined on the original feature space. [sent-22, score-0.257]

11 [15] have recently proposed a spectral hashing approach motivated by the idea that a good encoding scheme should minimize the sum of Hamming distances between pairs of code strings weighted by the value of a Gaussian kernel between the corresponding feature vectors. [sent-24, score-1.272]

12 With appropriate heuristic simplifications, this objective can be shown to yield a very efficient encoding rule, where each bit of the code is given by the sign of a sine function applied to a one-dimensional projection of the feature vector. [sent-25, score-0.333]

13 Spectral hashing shows promising experimental results, but its behavior is not easy to characterize theoretically. [sent-26, score-0.515]

14 In particular, it is not clear whether the Hamming distance between spectral hashing code strings converges to any function of the Euclidean distance or the kernel value between the original vectors as the number of bits in the code increases. [sent-27, score-1.686]

15 In this paper, we propose a coding method that is similar to spectral hashing computationally, but is derived from completely different considerations, is amenable to full theoretical analysis, and shows better practical behavior as a function of code size. [sent-28, score-0.908]

16 We start with a low-dimensional mapping of the original data that is guaranteed to preserve the value of a shift-invariant kernel (specifically, the random Fourier features of Rahimi and Recht [8]), and convert this mapping to a binary one with similar guarantees. [sent-29, score-0.269]

17 In particular, we show that the normalized Hamming distance (i. [sent-30, score-0.153]

18 , Ham- ming distance divided by the number of bits in the code) between any two embedded points sharply concentrates around a well-defined continuous function of the kernel value. [sent-32, score-0.581]

19 Using entropy bounds from the theory of empirical processes, we also prove a stronger result of this type that holds for any compact domain of RD , provided the number of bits is proportional to the intrinsic dimension of the domain. [sent-34, score-0.27]

20 Our scheme is completely distribution-free with respect to the data: its structure depends only on the underlying kernel. [sent-35, score-0.138]

21 In this, it is similar to locality sensitive hashing (LSH) [1], which is a family of methods for deriving low-dimensional discrete representations of the data for sublinear near-neighbor search. [sent-36, score-0.52]

22 To the best of our knowledge, our scheme is among the first random projection methods for constructing a similarity-preserving embedding into a binary cube. [sent-38, score-0.261]

23 In addition to presenting a thorough theoretical analysis, we have evaluated our approach on both synthetic and real data (images from the LabelMe database [10] represented by high-dimensional GIST descriptors [7]) and compared its performance to that of spectral hashing. [sent-39, score-0.351]

24 2 Binary codes for shift-invariant kernels Consider a Mercer kernel K(·, ·) on RD that satisfies the following for all points x, y ∈ RD : (K1) It is translation-invariant (or shift-invariant), i. [sent-41, score-0.283]

25 The Gaussian kernel K(x, y) = exp(−γ x − y 2 /2) or the Laplacian kernel K(x, y) = exp(−γ x − y 1 ) are two well-known examples. [sent-48, score-0.248]

26 We would like to construct an embedding F n of RD into the binary cube {0, 1}n such that for any pair x, y the normalized Hamming distance 1 △ 1 dH (F n (x), F n (y)) = n n n i=1 1{Fi (x)=Fi (y)} between F n (x) = (F1 (x), . [sent-49, score-0.396]

27 In other words, we would like to map D-dimensional real vectors into n-bit binary strings in a locality-sensitive manner, where the notion of locality is induced by the kernel K. [sent-56, score-0.393]

28 Recently, Rahimi and Recht [8] gave a scheme that takes a Mercer kernel satisfying (K1) and (K2) and produces a random mapping Φn : RD → Rn , such that, with high probability, the inner product of any two transformed points approximates the kernel: Φn (x)·Φn (y) ≈ K(x−y) for all x, y. [sent-59, score-0.314]

29 Their scheme exploits Bochner’s theorem [9], a fundamental result in harmonic analysis which says that any such K is a Fourier transform of a uniquely defined probability measure PK on RD . [sent-60, score-0.151]

30 For example, for the Gaussian kernel K(s) = e−γ s /2 , we take ω ∼ Normal(0, γID×D ). [sent-62, score-0.124]

31 n From random Fourier features to random binary codes. [sent-75, score-0.093]

32 We will compose the RFFs with random binary quantizers. [sent-76, score-0.093]

33 Now, given a kernel K, we define a random map Ft,ω,b : RD → {0, 1} through 1 [1 + Qt (cos(ω · x + b))] , (2) 2 where t ∼ Unif[−1, 1], ω ∼ PK , and b ∼ Unif[0, 2π] are independent of one another. [sent-80, score-0.124]

34 In addition, we have the following upper and lower bounds solely in terms of the kernel value K(x−y): Lemma 2. [sent-91, score-0.124]

35 1 shows a comparison of the kernel approximation properties of the RFFs [8] with our scheme for the Gaussian kernel. [sent-104, score-0.234]

36 (a) (b) (c) Figure 1: (a) Approximating the Gaussian kernel by random features (green) and random signs (red). [sent-105, score-0.159]

37 (b) Relationship of normalized Hamming distance between random signs to functions of the kernel. [sent-106, score-0.188]

38 The scatter plots in (a) and (b) are obtained from a synthetic set of 500 uniformly distributed 2D points with n = 5000. [sent-107, score-0.247]

39 (c) Bounds for normalized Hamming distance in Lemmas 2. [sent-108, score-0.153]

40 Now we concatenate several mappings of the form Ft,ω,b to construct an embedding of X into the binary cube {0, 1}n . [sent-112, score-0.243]

41 , n As we will show next, this construction ensures that, for any two points x and y, the fraction of the bits where the binary strings F n (x) and F n (y) disagree sharply concentrates around hK (x − y), provided n is large enough. [sent-125, score-0.582]

42 , K(x − y) ∼ 1, most of the bits of F n (x) and F n (y) will agree, whereas for any two points x and y that are “dissimilar,” i. [sent-128, score-0.233]

43 We first prove a Johnson–Lindenstrauss type result which says that, for any finite subset of RD , the normalized Hamming distance respects the similarities between points. [sent-132, score-0.194]

44 , xN } ⊂ RD , F n is such that 1 dH (F n (xj ), F n (xk )) ≤ hK (xj − xk ) + δ (4) n 1 h1 (K(xj − xk )) − δ ≤ dH (F n (xj ), F n (xk )) ≤ h2 (K(xj − xk )) + δ (5) n 2 for all j, k with probability ≥ 1 − N 2 e−2nδ . [sent-139, score-0.195]

45 Thus, any N -point subset of RD can be embedded, with high probability, into the binary cube of dimension O(log N ) in a similarity-preserving way. [sent-141, score-0.232]

46 5 The Assouad dimension of X ⊂ RD , denoted by dX , is the smallest integer k, such that, for any ball B ⊂ RD , the set B ∩ X can be covered by 2k balls of half the radius of B. [sent-146, score-0.096]

47 The Assouad dimension is a widely used measure of the intrinsic dimension [2, 6, 3]. [sent-147, score-0.138]

48 6 Suppose that the kernel K is such that LK = Eω∼PK ω 2 < +∞. [sent-151, score-0.124]

49 We will bound these covering numbers by the covering numbers of X with respect to the Euclidean norm on RD . [sent-172, score-0.126]

50 It can be shown that, for any four points x, x′ , y, y ′ ∈ X , 1Ax,y − 1Ax′ ,y′ 2 L2 (µ) = 2 1{θ∈Ax,y } − 1{θ∈Ax′ ,y′ } dµ(θ) ≤ µ(Bx △Bx′ ) + µ(By △By′ ), △ where △ denotes symmetric difference of sets, and Bx = {(t, ω, b) : Qt (cos(ω · x + b)) = +1} (details omitted for lack of space). [sent-173, score-0.084]

51 By definition of the Assouad dimension, N (δ, X , · ) ≤ (2 diam X /δ)dX , 2dX so N (ǫ, A, · L2 (µ) ) ≤ 2LK diam X . [sent-180, score-0.274]

52 We can now estimate the integral in (9) by ǫ2 R(A) ≤ C1 LK dX diam X , n (10) diam X for some constant C1 > 0. [sent-181, score-0.274]

53 √ 2 For example, with the Gaussian kernel K(s) = e−γ s /2 on RD , we have LK = Dγ. [sent-184, score-0.124]

54 The kernel bandwidth γ is often chosen as γ ∝ 1/[D(diam X )2 ] (see, e. [sent-185, score-0.172]

55 8]); with this setting, the number of bits needed to guarantee the bound (6) is n = Ω((dX /δ 2 ) log(1/ǫ)). [sent-189, score-0.179]

56 It is possible, in principle, to construct a dimension-reducing embedding of X into a binary cube, provided the number of bits in the embedding is larger than the intrinsic dimension of X . [sent-190, score-0.479]

57 Our method Spectral hashing (a) (b) (c) (d) (e) (f) Figure 2: Synthetic results. [sent-191, score-0.459]

58 First row: scatter plots of normalized Hamming distance vs. [sent-192, score-0.288]

59 Euclidean distance for our method (a) and spectral hashing (b) with code size 32 bits. [sent-193, score-0.95]

60 Second row: scatter plots for our method (c) and spectral hashing (d) with code size 512 bits. [sent-195, score-0.988]

61 Third row: recall-precision plots for our method (e) and spectral hashing (f) for code sizes from 8 to 512 bits (best viewed in color). [sent-196, score-1.118]

62 3 Empirical Evaluation In this section, we present the results of our scheme with a Gaussian kernel, and compare our performance to spectral hashing [15]. [sent-197, score-0.774]

63 1 Spectral hashing is a recently introduced, state-of-the-art approach that has been reported to obtain better results than several other well-known methods, including LSH [1] and restricted Boltzmann machines [11]. [sent-198, score-0.459]

64 Unlike our method, spectral hashing chooses code parameters in a deterministic, data-dependent way, motivated by results on convergence of 1 We use the code made available by the authors of [15] at http://www. [sent-199, score-1.042]

65 Our method Spectral hashing Figure 3: Recall-precision curves for the LabelMe database for our method (left) and for spectral hashing (right). [sent-204, score-1.185]

66 Though spectral hashing is derived from completely different considerations than our method, its encoding scheme is similar to ours in terms of basic computation. [sent-207, score-0.84]

67 The structural similarity between spectral hashing and our method makes comparison between them appropriate. [sent-209, score-0.7]

68 To demonstrate the basic behavior of our method, we first report results for two-dimensional synthetic data using a protocol similar to [15] (we have also conducted tests on higher-dimensional synthetic data, with very similar results). [sent-210, score-0.143]

69 To distinguish true positives from false positives for evaluating retrieval performance, we select a “nominal” neighborhood radius so that each query point on average has 50 neighbors in the database. [sent-213, score-0.4]

70 Next, we rescale the data so that this radius is 1, and set the bandwidth of the kernel to γ = 1. [sent-214, score-0.221]

71 2 (a,c) shows scatter plots of normalized Hamming distance vs. [sent-216, score-0.288]

72 Euclidean distance for each query point paired with each database point for 32-bit and 512-bit codes. [sent-217, score-0.263]

73 As more bits are added to our code, the variance of the scatter plots decreases, and the points cluster tighter around the theoretically expected curve (Eq. [sent-218, score-0.368]

74 The scatter plots for spectral hashing are shown in Fig. [sent-221, score-0.799]

75 As the number of bits in the spectral hashing code is increased, normalized Hamming distance does not appear to converge to any clear function of the Euclidean distance. [sent-223, score-1.185]

76 Because the derivation of spectral hashing in [15] includes several heuristic steps, the behavior of the resulting scheme appears to be difficult to analyze, and shows some undesirable effects as the code size increases. [sent-224, score-0.99]

77 Figure 2 (e,f) compares recall-precision curves for both methods using a range of code sizes. [sent-225, score-0.189]

78 Since the normalized Hamming distance for our method converges to a monotonic function of the Euclidean distance, its performance keeps improving as a function of code size. [sent-226, score-0.342]

79 On the other hand, spectral hashing starts out with promising performance for very short codes (up to 32 bits), but then deteriorates for higher numbers of bits. [sent-227, score-0.827]

80 Next, we present retrieval results for 14,871 images taken from the LabelMe database [10]. [sent-228, score-0.167]

81 The images are represented by 320-dimensional GIST descriptors [7], which have proven to be effective at capturing perceptual similarity between scenes. [sent-229, score-0.097]

82 ” As with the synthetic experiments, a nominal threshold of the average distance to the 50th nearest neighbor is used to determine whether a database point returned for a given query is considered a true positive. [sent-231, score-0.41]

83 Figure 3 shows precisionrecall curves for code sizes ranging from 16 bits to 1024 bits. [sent-232, score-0.408]

84 As in the synthetic experiments, spectral hashing appears to have an advantage over our method for extremely small code sizes, up to about 32 bits. [sent-233, score-0.911]

85 For larger code sizes, our method begins to dominate. [sent-235, score-0.219]

86 For example, with a 128-bit code (which is equivalent to just two double-precision floating point numbers), our scheme achieves 0. [sent-236, score-0.299]

87 8 precision Euclidean neighbors 32 bit code 512 bit code Precision: 0. [sent-237, score-0.63]

88 96 Figure 4: Examples of retrieval for two query images on the LabelMe database. [sent-241, score-0.209]

89 The left column shows top 48 neighbors for each query according to Euclidean distance (the query image is in the top left of the collage). [sent-242, score-0.389]

90 right) column shows nearest neighbors according to normalized Hamming distance with a 32-bit (resp. [sent-244, score-0.27]

91 The precision of retrieval is evaluated as the proportion of top Hamming neighbors that are also Euclidean neighbors within the “nominal” radius. [sent-246, score-0.286]

92 Moreover, the performance of spectral hashing actually begins to decrease for code sizes above 256 bits. [sent-252, score-0.923]

93 Finally, Figure 4 shows retrieval results for our method on a couple of representative query images. [sent-253, score-0.174]

94 In addition to being completely distribution-free and exhibiting more desirable behavior as a function of code size, our scheme has one more practical advantage. [sent-254, score-0.354]

95 Unlike spectral hashing, we retain the kernel bandwidth γ as a “free parameter,” which gives us flexibility in terms of adapting to target neighborhood size, or setting a target Hamming distance for neighbors at a given Euclidean distance. [sent-255, score-0.587]

96 This can be especially useful for making sure that a significant fraction of neighbors for each query are mapped to strings whose Hamming distance from the query is no greater than 2. [sent-256, score-0.536]

97 This is a necesary condition for being able to use binary codes for hashing as opposed to brute-force search (although, as demonstrated in [11, 13], even brute-force search with binary codes can already be quite fast). [sent-257, score-0.855]

98 To ensure high recall within a low Hamming radius, we can progressively increase the kernel bandwidth γ as the code size increases, thus counteracting the increase in unnormalized Hamming distance that inevitably accompanies larger code sizes. [sent-258, score-0.647]

99 At present, our promising initial results, combined with our comprehensive theoretical analysis, convincingly demonstrate the potential usefulness of our scheme for large-scale indexing and search applications. [sent-261, score-0.139]

100 Near-optimal hashing algorithms for approximate nearest neighbor in high dimensions. [sent-267, score-0.492]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('hashing', 0.459), ('hamming', 0.391), ('cos', 0.205), ('spectral', 0.205), ('code', 0.189), ('bits', 0.179), ('strings', 0.147), ('rd', 0.145), ('diam', 0.137), ('kernel', 0.124), ('hk', 0.116), ('bx', 0.115), ('unif', 0.114), ('qt', 0.11), ('scheme', 0.11), ('lsh', 0.105), ('codes', 0.105), ('query', 0.104), ('dh', 0.102), ('dx', 0.099), ('ft', 0.099), ('distance', 0.097), ('euclidean', 0.097), ('binary', 0.093), ('cube', 0.092), ('scatter', 0.089), ('labelme', 0.089), ('fourier', 0.085), ('neighbors', 0.084), ('lk', 0.079), ('lindenstrauss', 0.078), ('pk', 0.074), ('retrieval', 0.07), ('xk', 0.065), ('sgn', 0.063), ('rahimi', 0.063), ('database', 0.062), ('bit', 0.06), ('assouad', 0.059), ('embedding', 0.058), ('johnson', 0.058), ('synthetic', 0.058), ('normalized', 0.056), ('nominal', 0.056), ('embedded', 0.055), ('points', 0.054), ('chapel', 0.052), ('rff', 0.052), ('rffs', 0.052), ('radius', 0.049), ('precision', 0.048), ('bandwidth', 0.048), ('dimension', 0.047), ('plots', 0.046), ('sine', 0.046), ('intrinsic', 0.044), ('fn', 0.043), ('hill', 0.042), ('says', 0.041), ('torralba', 0.041), ('xj', 0.041), ('sizes', 0.04), ('lazebnik', 0.039), ('encoding', 0.038), ('concentrates', 0.037), ('recht', 0.037), ('disagree', 0.037), ('lemma', 0.036), ('similarity', 0.036), ('mx', 0.035), ('signs', 0.035), ('sharply', 0.035), ('images', 0.035), ('gist', 0.034), ('mercer', 0.034), ('covering', 0.034), ('fi', 0.034), ('nearest', 0.033), ('positives', 0.032), ('sublinear', 0.032), ('dasgupta', 0.032), ('fix', 0.032), ('laplacian', 0.031), ('schemes', 0.03), ('omitted', 0.03), ('dissimilar', 0.03), ('begins', 0.03), ('behaves', 0.029), ('locality', 0.029), ('promising', 0.029), ('neighborhood', 0.029), ('numbers', 0.029), ('bn', 0.028), ('completely', 0.028), ('behavior', 0.027), ('pt', 0.027), ('designing', 0.027), ('mapping', 0.026), ('descriptors', 0.026), ('sin', 0.026)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.99999982 142 nips-2009-Locality-sensitive binary codes from shift-invariant kernels

Author: Maxim Raginsky, Svetlana Lazebnik

Abstract: This paper addresses the problem of designing binary codes for high-dimensional data such that vectors that are similar in the original space map to similar binary strings. We introduce a simple distribution-free encoding scheme based on random projections, such that the expected Hamming distance between the binary codes of two vectors is related to the value of a shift-invariant kernel (e.g., a Gaussian kernel) between the vectors. We present a full theoretical analysis of the convergence properties of the proposed scheme, and report favorable experimental performance as compared to a recent state-of-the-art method, spectral hashing.

2 0.43046418 135 nips-2009-Learning to Hash with Binary Reconstructive Embeddings

Author: Brian Kulis, Trevor Darrell

Abstract: Fast retrieval methods are increasingly critical for many large-scale analysis tasks, and there have been several recent methods that attempt to learn hash functions for fast and accurate nearest neighbor searches. In this paper, we develop an algorithm for learning hash functions based on explicitly minimizing the reconstruction error between the original distances and the Hamming distances of the corresponding binary embeddings. We develop a scalable coordinate-descent algorithm for our proposed hashing objective that is able to efficiently learn hash functions in a variety of settings. Unlike existing methods such as semantic hashing and spectral hashing, our method is easily kernelized and does not require restrictive assumptions about the underlying distribution of the data. We present results over several domains to demonstrate that our method outperforms existing state-of-the-art techniques. 1

3 0.10261431 220 nips-2009-Slow Learners are Fast

Author: Martin Zinkevich, John Langford, Alex J. Smola

Abstract: Online learning algorithms have impressive convergence properties when it comes to risk minimization and convex games on very large problems. However, they are inherently sequential in their design which prevents them from taking advantage of modern multi-core architectures. In this paper we prove that online learning with delayed updates converges well, thereby facilitating parallel online learning. 1

4 0.099583738 139 nips-2009-Linear-time Algorithms for Pairwise Statistical Problems

Author: Parikshit Ram, Dongryeol Lee, William March, Alexander G. Gray

Abstract: Several key computational bottlenecks in machine learning involve pairwise distance computations, including all-nearest-neighbors (ďŹ nding the nearest neighbor(s) for each point, e.g. in manifold learning) and kernel summations (e.g. in kernel density estimation or kernel machines). We consider the general, bichromatic case for these problems, in addition to the scientiďŹ c problem of N-body simulation. In this paper we show for the ďŹ rst time O(đ?‘ ) worst case runtimes for practical algorithms for these problems based on the cover tree data structure [1]. 1

5 0.098861314 229 nips-2009-Statistical Analysis of Semi-Supervised Learning: The Limit of Infinite Unlabelled Data

Author: Boaz Nadler, Nathan Srebro, Xueyuan Zhou

Abstract: We study the behavior of the popular Laplacian Regularization method for SemiSupervised Learning at the regime of a fixed number of labeled points but a large number of unlabeled points. We show that in Rd , d 2, the method is actually not well-posed, and as the number of unlabeled points increases the solution degenerates to a noninformative function. We also contrast the method with the Laplacian Eigenvector method, and discuss the “smoothness” assumptions associated with this alternate method. 1 Introduction and Setup In this paper we consider the limit behavior of two popular semi-supervised learning (SSL) methods based on the graph Laplacian: the regularization approach [15] and the spectral approach [3]. We consider the limit when the number of labeled points is fixed and the number of unlabeled points goes to infinity. This is a natural limit for SSL as the basic SSL scenario is one in which unlabeled data is virtually infinite. We can also think of this limit as “perfect” SSL, having full knowledge of the marginal density p(x). The premise of SSL is that the marginal density p(x) is informative about the unknown mapping y(x) we are trying to learn, e.g. since y(x) is expected to be “smooth” in some sense relative to p(x). Studying the infinite-unlabeled-data limit, where p(x) is fully known, allows us to formulate and understand the underlying smoothness assumptions of a particular SSL method, and judge whether it is well-posed and sensible. Understanding the infinite-unlabeled-data limit is also a necessary first step to studying the convergence of the finite-labeled-data estimator. We consider the following setup: Let p(x) be an unknown smooth density on a compact domain Ω ⊂ Rd with a smooth boundary. Let y : Ω → Y be the unknown function we wish to estimate. In case of regression Y = R whereas in binary classification Y = {−1, 1}. The standard (transductive) semisupervised learning problem is formulated as follows: Given l labeled points, (x1 , y1 ), . . . , (xl , yl ), with yi = y(xi ), and u unlabeled points xl+1 , . . . , xl+u , with all points xi sampled i.i.d. from p(x), the goal is to construct an estimate of y(xl+i ) for any unlabeled point xl+i , utilizing both the labeled and the unlabeled points. We denote the total number of points by n = l + u. We are interested in the regime where l is fixed and u → ∞. 1 2 SSL with Graph Laplacian Regularization We first consider the following graph-based approach formulated by Zhu et. al. [15]: y (x) = arg min In (y) ˆ subject to y(xi ) = yi , i = 1, . . . , l y where 1 n2 In (y) = Wi,j (y(xi ) − y(xj ))2 (1) (2) i,j is a Laplacian regularization term enforcing “smoothness” with respect to the n×n similarity matrix W . This formulation has several natural interpretations in terms of, e.g. random walks and electrical circuits [15]. These interpretations, however, refer to a fixed graph, over a finite set of points with given similarities. In contrast, our focus here is on the more typical scenario where the points xi ∈ Rd are a random sample from a density p(x), and W is constructed based on this sample. We would like to understand the behavior of the method in terms of the density p(x), particularly in the limit where the number of unlabeled points grows. Under what assumptions on the target labeling y(x) and on the density p(x) is the method (1) sensible? The answer, of course, depends on how the matrix W is constructed. We consider the common situation where the similarities are obtained by applying some decay filter to the distances: xi −xj σ Wi,j = G (3) where G : R+ → R+ is some function with an adequately fast decay. Popular choices are the 2 Gaussian filter G(z) = e−z /2 or the ǫ-neighborhood graph obtained by the step filter G(z) = 1z<1 . For simplicity, we focus here on the formulation (1) where the solution is required to satisfy the constraints at the labeled points exactly. In practice, the hard labeling constraints are often replaced with a softer loss-based data term, which is balanced against the smoothness term In (y), e.g. [14, 6]. Our analysis and conclusions apply to such variants as well. Limit of the Laplacian Regularization Term As the number of unlabeled examples grows the regularization term (2) converges to its expectation, where the summation is replaced by integration w.r.t. the density p(x): lim In (y) = I (σ) (y) = n→∞ G Ω Ω x−x′ σ (y(x) − y(x′ ))2 p(x)p(x′ )dxdx′ . (4) In the above limit, the bandwidth σ is held fixed. Typically, one would also drive the bandwidth σ to zero as n → ∞. There are two reasons for this choice. First, from a practical perspective, this makes the similarity matrix W sparse so it can be stored and processed. Second, from a theoretical perspective, this leads to a clear and well defined limit of the smoothness regularization term In (y), at least when σ → 0 slowly enough1 , namely when σ = ω( d log n/n). If σ → 0 as n → ∞, and as long as nσ d / log n → ∞, then after appropriate normalization, the regularizer converges to a density weighted gradient penalty term [7, 8]: d lim d+2 In (y) n→∞ Cσ (σ) d (y) d+2 I σ→0 Cσ = lim ∇y(x) 2 p(x)2 dx = J(y) = (5) Ω where C = Rd z 2 G( z )dz, and assuming 0 < C < ∞ (which is the case for both the Gaussian and the step filters). This energy functional J(f ) therefore encodes the notion of “smoothness” with respect to p(x) that is the basis of the SSL formulation (1) with the graph constructions specified by (3). To understand the behavior and appropriateness of (1) we must understand this functional and the associated limit problem: y (x) = arg min J(y) ˆ subject to y(xi ) = yi , i = 1, . . . , l (6) y p When σ = o( d 1/n) then all non-diagonal weights Wi,j vanish (points no longer have any “close by” p neighbors). We are not aware of an analysis covering the regime where σ decays roughly as d 1/n, but would be surprised if a qualitatively different meaningful limit is reached. 1 2 3 Graph Laplacian Regularization in R1 We begin by considering the solution of (6) for one dimensional data, i.e. d = 1 and x ∈ R. We first consider the situation where the support of p(x) is a continuous interval Ω = [a, b] ⊂ R (a and/or b may be infinite). Without loss of generality, we assume the labeled data is sorted in increasing order a x1 < x2 < · · · < xl b. Applying the theory of variational calculus, the solution y (x) ˆ satisfies inside each interval (xi , xi+1 ) the Euler-Lagrange equation d dy p2 (x) = 0. dx dx Performing two integrations and enforcing the constraints at the labeled points yields y(x) = yi + x 1/p2 (t)dt xi (yi+1 xi+1 1/p2 (t)dt xi − yi ) for xi x xi+1 (7) with y(x) = x1 for a x x1 and y(x) = xl for xl x b. If the support of p(x) is a union of disjoint intervals, the above analysis and the form of the solution applies in each interval separately. The solution (7) seems reasonable and desirable from the point of view of the “smoothness” assumptions: when p(x) is uniform, the solution interpolates linearly between labeled data points, whereas across low-density regions, where p(x) is close to zero, y(x) can change abruptly. Furthermore, the regularizer J(y) can be interpreted as a Reproducing Kernel Hilbert Space (RKHS) squared semi-norm, giving us additional insight into this choice of regularizer: b 1 Theorem 1. Let p(x) be a smooth density on Ω = [a, b] ⊂ R such that Ap = 4 a 1/p2 (t)dt < ∞. 2 Then, J(f ) can be written as a squared semi-norm J(f ) = f Kp induced by the kernel x′ ′ Kp (x, x ) = Ap − 1 2 x with a null-space of all constant functions. That is, f the RKHS induced by Kp . 1 p2 (t) dt Kp . (8) is the norm of the projection of f onto If p(x) is supported on several disjoint intervals, Ω = ∪i [ai , bi ], then J(f ) can be written as a squared semi-norm induced by the kernel 1 bi dt 4 ai p2 (t) ′ Kp (x, x ) = − 1 2 x′ dt x p2 (t) if x, x′ ∈ [ai , bi ] (9) if x ∈ [ai , bi ], x′ ∈ [aj , bj ], i = j 0 with a null-space spanned by indicator functions 1[ai ,bi ] (x) on the connected components of Ω. Proof. For any f (x) = i αi Kp (x, xi ) in the RKHS induced by Kp : df dx J(f ) = 2 p2 (x)dx = αi αj Jij (10) i,j where Jij = d d Kp (x, xi ) Kp (x, xj )p2 (x)dx dx dx When xi and xj are in different connected components of Ω, the gradients of Kp (·, xi ) and Kp (·, xj ) are never non-zero together and Jij = 0 = Kp (xi , xj ). When they are in the same connected component [a, b], and assuming w.l.o.g. a xi xj b: Jij = = xi 1 4 1 4 a b a 1 dt + p2 (t) 1 1 dt − p2 (t) 2 xj xi xj xi −1 dt + p2 (t) xj 1 dt p2 (t) 1 dt = Kp (xi , xj ). p2 (t) Substituting Jij = Kp (xi , xj ) into (10) yields J(f ) = 3 b αi αj Kp (xi , xj ) = f (11) Kp . Combining Theorem 1 with the Representer Theorem [13] establishes that the solution of (6) (or of any variant where the hard constraints are replaced by a data term) is of the form: l y(x) = αj Kp (x, xj ) + βi 1[ai ,bi ] (x), j=1 i where i ranges over the connected components [ai , bi ] of Ω, and we have: l J(y) = αi αj Kp (xi , xj ). (12) i,j=1 Viewing the regularizer as y 2 p suggests understanding (6), and so also its empirical approximaK tion (1), by interpreting Kp (x, x′ ) as a density-based “similarity measure” between x and x′ . This similarity measure indeed seems sensible: for a uniform density it is simply linearly decreasing as a function of the distance. When the density is non-uniform, two points are relatively similar only if they are connected by a region in which 1/p2 (x) is low, i.e. the density is high, but are much less “similar”, i.e. related to each other, when connected by a low-density region. Furthermore, there is no dependence between points in disjoint components separated by zero density regions. 4 Graph Laplacian Regularization in Higher Dimensions The analysis of the previous section seems promising, at it shows that in one dimension, the SSL method (1) is well posed and converges to a sensible limit. Regretfully, in higher dimensions this is not the case anymore. In the following theorem we show that the infimum of the limit problem (6) is zero and can be obtained by a sequence of functions which are certainly not a sensible extrapolation of the labeled points. Theorem 2. Let p(x) be a smooth density over Rd , d 2, bounded from above by some constant pmax , and let (x1 , y1 ), . . . , (xl , yl ) be any (non-repeating) set of labeled examples. There exist continuous functions yǫ (x), for any ǫ > 0, all satisfying the constraints yǫ (xj ) = yj , j = 1, . . . , l, such ǫ→0 ǫ→0 that J(yǫ ) −→ 0 but yǫ (x) −→ 0 for all x = xj , j = 1, . . . , l. Proof. We present a detailed proof for the case of l = 2 labeled points. The generalization of the proof to more labeled points is straightforward. Furthermore, without loss of generality, we assume the first labeled point is at x0 = 0 with y(x0 ) = 0 and the second labeled point is at x1 with x1 = 1 and y(x1 ) = 1. In addition, we assume that the ball B1 (0) of radius one centered around the origin is contained in Ω = {x ∈ Rd | p(x) > 0}. We first consider the case d > 2. Here, for any ǫ > 0, consider the function x ǫ yǫ (x) = min ,1 which indeed satisfies the two constraints yǫ (xi ) = yi , i = 0, 1. Then, J(yǫ ) = Bǫ (0) p2 (x) dx ǫ2 pmax ǫ2 dx = p2 Vd ǫd−2 max (13) Bǫ (0) where Vd is the volume of a unit ball in Rd . Hence, the sequence of functions yǫ (x) satisfy the constraints, but for d > 2, inf ǫ J(yǫ ) = 0. For d = 2, a more extreme example is necessary: consider the functions 2 x yǫ (x) = log +ǫ ǫ log 1+ǫ ǫ for x 1 and yǫ (x) = 1 for x > 1. These functions satisfy the two constraints yǫ (xi ) = yi , i = 0, 1 and: J(yǫ ) = 4 h “ ”i 1+ǫ 2 log ǫ 4πp2 max h “ ”i 1+ǫ 2 log ǫ x B1 (0) log ( x 1+ǫ ǫ 2 2 +ǫ)2 p2 (x)dx 4p2 h “ max ”i2 1+ǫ log ǫ 4πp2 max ǫ→0 = −→ 0. log 1+ǫ ǫ 4 1 0 r2 (r 2 +ǫ)2 2πrdr The implication of Theorem 2 is that regardless of the values at the labeled points, as u → ∞, the solution of (1) is not well posed. Asymptotically, the solution has the form of an almost everywhere constant function, with highly localized spikes near the labeled points, and so no learning is performed. In particular, an interpretation in terms of a density-based kernel Kp , as in the onedimensional case, is not possible. Our analysis also carries over to a formulation where a loss-based data term replaces the hard label constraints, as in l 1 y = arg min ˆ (y(xj ) − yj )2 + γIn (y) y(x) l j=1 In the limit of infinite unlabeled data, functions of the form yǫ (x) above have a zero data penalty term (since they exactly match the labels) and also drive the regularization term J(y) to zero. Hence, it is possible to drive the entire objective functional (the data term plus the regularization term) to zero with functions that do not generalize at all to unlabeled points. 4.1 Numerical Example We illustrate the phenomenon detailed by Theorem 2 with a simple example. Consider a density p(x) in R2 , which is a mixture of two unit variance spherical Gaussians, one per class, centered at the origin and at (4, 0). We sample a total of n = 3000 points, and label two points from each of the two components (four total). We then construct a similarity matrix using a Gaussian filter with σ = 0.4. Figure 1 depicts the predictor y (x) obtained from (1). In fact, two different predictors are shown, ˆ obtained by different numerical methods for solving (1). Both methods are based on the observation that the solution y (x) of (1) satisfies: ˆ n y (xi ) = ˆ n Wij y (xj ) / ˆ j=1 Wij on all unlabeled points i = l + 1, . . . , l + u. (14) j=1 Combined with the constraints of (1), we obtain a system of linear equations that can be solved by Gaussian elimination (here invoked through MATLAB’s backslash operator). This is the method used in the top panels of Figure 1. Alternatively, (14) can be viewed as an update equation for y (xi ), ˆ which can be solved via the power method, or label propagation [2, 6]: start with zero labels on the unlabeled points and iterate (14), while keeping the known labels on x1 , . . . , xl . This is the method used in the bottom panels of Figure 1. As predicted, y (x) is almost constant for almost all unlabeled points. Although all values are very ˆ close to zero, thresholding at the “right” threshold does actually produce sensible results in terms of the true -1/+1 labels. However, beyond being inappropriate for regression, a very flat predictor is still problematic even from a classification perspective. First, it is not possible to obtain a meaningful confidence measure for particular labels. Second, especially if the size of each class is not known apriori, setting the threshold between the positive and negative classes is problematic. In our example, setting the threshold to zero yields a generalization error of 45%. The differences between the two numerical methods for solving (1) also point out to another problem with the ill-posedness of the limit problem: the solution is numerically very un-stable. A more quantitative evaluation, that also validates that the effect in Figure 1 is not a result of choosing a “wrong” bandwidth σ, is given in Figure 2. We again simulated data from a mixture of two Gaussians, one Gaussian per class, this time in 20 dimensions, with one labeled point per class, and an increasing number of unlabeled points. In Figure 2 we plot the squared error, and the classification error of the resulting predictor y (x). We plot the classification error both when a threshold ˆ of zero is used (i.e. the class is determined by sign(ˆ(x))) and with the ideal threshold minimizing y the test error. For each unlabeled sample size, we choose the bandwidth σ yielding the best test performance (this is a “cheating” approach which provides a lower bound on the error of the best method for selecting the bandwidth). As the number of unlabeled examples increases the squared error approaches 1, indicating a flat predictor. Using a threshold of zero leads to an increase in the classification error, possibly due to numerical instability. Interestingly, although the predictors become very flat, the classification error using the ideal threshold actually improves slightly. Note that 5 DIRECT INVERSION SQUARED ERROR SIGN ERROR: 45% OPTIMAL BANDWIDTH 1 0.9 1 5 0 4 2 0.85 y(x) > 0 y(x) < 0 6 0.95 10 0 0 −1 10 0 200 400 600 800 0−1 ERROR (THRESHOLD=0) 0.32 −5 10 0 5 −10 0 −10 −5 −5 0 5 10 10 1 0 0 200 400 600 800 OPTIMAL BANDWIDTH 0.5 0 0 200 400 600 800 0−1 ERROR (IDEAL THRESHOLD) 0.19 5 200 400 600 800 OPTIMAL BANDWIDTH 1 0.28 SIGN ERR: 17.1 0.3 0.26 POWER METHOD 0 1.5 8 0 0.18 −1 10 6 0.17 4 −5 10 0 5 −10 0 −5 −10 −5 0 5 10 Figure 1: Left plots: Minimizer of Eq. (1). Right plots: the resulting classification according to sign(y). The four labeled points are shown by green squares. Top: minimization via Gaussian elimination (MATLAB backslash). Bottom: minimization via label propagation with 1000 iterations - the solution has not yet converged, despite small residuals of the order of 2 · 10−4 . 0.16 0 200 400 600 800 2 0 200 400 600 800 Figure 2: Squared error (top), classification error with a threshold of zero (center) and minimal classification error using ideal threhold (bottom), of the minimizer of (1) as a function of number of unlabeled points. For each error measure and sample size, the bandwidth minimizing the test error was used, and is plotted. ideal classification performance is achieved with a significantly larger bandwidth than the bandwidth minimizing the squared loss, i.e. when the predictor is even flatter. 4.2 Probabilistic Interpretation, Exit and Hitting Times As mentioned above, the Laplacian regularization method (1) has a probabilistic interpretation in terms of a random walk on the weighted graph. Let x(t) denote a random walk on the graph with transition matrix M = D−1 W where D is a diagonal matrix with Dii = j Wij . Then, for the binary classification case with yi = ±1 we have [15]: y (xi ) = 2 Pr x(t) hits a point labeled +1 before hitting a point labeled -1 x(0) = xi − 1 ˆ We present an interpretation of our analysis in terms of the limiting properties of this random walk. Consider, for simplicity, the case where the two classes are separated by a low density region. Then, the random walk has two intrinsic quantities of interest. The first is the mean exit time from one cluster to the other, and the other is the mean hitting time to the labeled points in that cluster. As the number of unlabeled points increases and σ → 0, the random walk converges to a diffusion process [12]. While the mean exit time then converges to a finite value corresponding to its diffusion analogue, the hitting time to a labeled point increases to infinity (as these become absorbing boundaries of measure zero). With more and more unlabeled data the random walk will fully mix, forgetting where it started, before it hits any label. Thus, the probability of hitting +1 before −1 will become uniform across the entire graph, independent of the starting location xi , yielding a flat predictor. 5 Keeping σ Finite At this point, a reader may ask whether the problems found in higher dimensions are due to taking the limit σ → 0. One possible objection is that there is an intrinsic characteristic scale for the data σ0 where (with high probability) all points at a distance xi − xj < σ0 have the same label. If this is the case, then it may not necessarily make sense to take values of σ < σ0 in constructing W . However, keeping σ finite while taking the number of unlabeled points to infinity does not resolve the problem. On the contrary, even the one-dimensional case becomes ill-posed in this case. To see this, consider a function y(x) which is zero everywhere except at the labeled points, where y(xj ) = yj . With a finite number of labeled points of measure zero, I (σ) (y) = 0 in any dimension 6 50 points 500 points 3500 points 1 1 0.5 0.5 0.5 0 0 0 −0.5 y 1 −0.5 −0.5 −1 −2 0 2 4 6 −1 −2 0 2 4 6 −1 −2 0 2 4 6 x Figure 3: Minimizer of (1) for a 1-d problem with a fixed σ = 0.4, two labeled points and an increasing number of unlabeled points. and for any fixed σ > 0. While this limiting function is discontinuous, it is also possible to construct ǫ→0 a sequence of continuous functions yǫ that all satisfy the constraints and for which I (σ) (yǫ ) −→ 0. This behavior is illustrated in Figure 3. We generated data from a mixture of two 1-D Gaussians centered at the origin and at x = 4, with one Gaussian labeled −1 and the other +1. We used two labeled points at the centers of the Gaussians and an increasing number of randomly drawn unlabeled points. As predicted, with a fixed σ, although the solution is reasonable when the number of unlabeled points is small, it becomes flatter, with sharp spikes on the labeled points, as u → ∞. 6 Fourier-Eigenvector Based Methods Before we conclude, we discuss a different approach for SSL, also based on the Graph Laplacian, suggested by Belkin and Niyogi [3]. Instead of using the Laplacian as a regularizer, constraining candidate predictors y(x) non-parametrically to those with small In (y) values, here the predictors are constrained to the low-dimensional space spanned by the first few eigenvectors of the Laplacian: The similarity matrix W is computed as before, and the Graph Laplacian matrix L = D − W is considered (recall D is a diagonal matrix with Dii = j Wij ). Only predictors p j=1 aj ej y (x) = ˆ (15) spanned by the first p eigenvectors e1 , . . . , ep of L (with smallest eigenvalues) are considered. The coefficients aj are chosen by minimizing a loss function on the labeled data, e.g. the squared loss: (ˆ1 , . . . , ap ) = arg min a ˆ l j=1 (yj − y (xj ))2 . ˆ (16) Unlike the Laplacian Regularization method (1), the Laplacian Eigenvector method (15)–(16) is well posed in the limit u → ∞. This follows directly from the convergence of the eigenvectors of the graph Laplacian to the eigenfunctions of the corresponding Laplace-Beltrami operator [10, 4]. Eigenvector based methods were shown empirically to provide competitive generalization performance on a variety of simulated and real world problems. Belkin and Niyogi [3] motivate the approach by arguing that ‘the eigenfunctions of the Laplace-Beltrami operator provide a natural basis for functions on the manifold and the desired classification function can be expressed in such a basis’. In our view, the success of the method is actually not due to data lying on a low-dimensional manifold, but rather due to the low density separation assumption, which states that different class labels form high-density clusters separated by low density regions. Indeed, under this assumption and with sufficient separation between the clusters, the eigenfunctions of the graph Laplace-Beltrami operator are approximately piecewise constant in each of the clusters, as in spectral clustering [12, 11], providing a basis for a labeling that is constant within clusters but variable across clusters. In other settings, such as data uniformly distributed on a manifold but without any significant cluster structure, the success of eigenvector based methods critically depends on how well can the unknown classification function be approximated by a truncated expansion with relatively few eigenvectors. We illustrate this issue with the following three-dimensional example: Let p(x) denote the uniform density in the box [0, 1] × [0, 0.8] × [0, 0.6], where the box lengths are different to prevent eigenvalue multiplicity. Consider learning three different functions, y1 (x) = 1x1 >0.5 , y2 (x) = 1x1 >x2 /0.8 and y3 (x) = 1x2 /0.8>x3 /0.6 . Even though all three functions are relatively simple, all having a linear separating boundary between the classes on the manifold, as shown in the experiment described in Figure 4, the Eigenvector based method (15)–(16) gives markedly different generalization performances on the three targets. This happens both when the number of eigenvectors p is set to p = l/5 as suggested by Belkin and Niyogi, as well as for the optimal (oracle) value of p selected on the test set (i.e. a “cheating” choice representing an upper bound on the generalization error of this method). 7 Prediction Error (%) p = #labeled points/5 40 optimal p 20 labeled points 40 Approx. Error 50 20 20 0 20 20 40 60 # labeled points 0 10 20 40 60 # labeled points 0 0 5 10 15 # eigenvectors 0 0 5 10 15 # eigenvectors Figure 4: Left three panels: Generalization Performance of the Eigenvector Method (15)–(16) for the three different functions described in the text. All panels use n = 3000 points. Prediction counts the number of sign agreements with the true labels. Rightmost panel: best fit when many (all 3000) points are used, representing the best we can hope for with a few leading eigenvectors. The reason for this behavior is that y2 (x) and even more so y3 (x) cannot be as easily approximated by the very few leading eigenfunctions—even though they seem “simple” and “smooth”, they are significantly more complicated than y1 (x) in terms of measure of simplicity implied by the Eigenvector Method. Since the density is uniform, the graph Laplacian converges to the standard Laplacian and its eigenfunctions have the form ψi,j,k (x) = cos(iπx1 ) cos(jπx2 /0.8) cos(kπx3 /0.6), making it hard to represent simple decision boundaries which are not axis-aligned. 7 Discussion Our results show that a popular SSL method, the Laplacian Regularization method (1), is not wellbehaved in the limit of infinite unlabeled data, despite its empirical success in various SSL tasks. The empirical success might be due to two reasons. First, it is possible that with a large enough number of labeled points relative to the number of unlabeled points, the method is well behaved. This regime, where the number of both labeled and unlabeled points grow while l/u is fixed, has recently been analyzed by Wasserman and Lafferty [9]. However, we do not find this regime particularly satisfying as we would expect that having more unlabeled data available should improve performance, rather than require more labeled points or make the problem ill-posed. It also places the user in a delicate situation of choosing the “just right” number of unlabeled points without any theoretical guidance. Second, in our experiments we noticed that although the predictor y (x) becomes extremely flat, in ˆ binary tasks, it is still typically possible to find a threshold leading to a good classification performance. We do not know of any theoretical explanation for such behavior, nor how to characterize it. Obtaining such an explanation would be very interesting, and in a sense crucial to the theoretical foundation of the Laplacian Regularization method. On a very practical level, such a theoretical understanding might allow us to correct the method so as to avoid the numerical instability associated with flat predictors, and perhaps also make it appropriate for regression. The reason that the Laplacian regularizer (1) is ill-posed in the limit is that the first order gradient is not a sufficient penalty in high dimensions. This fact is well known in spline theory, where the Sobolev Embedding Theorem [1] indicates one must control at least d+1 derivatives in Rd . In the 2 context of Laplacian regularization, this can be done using the iterated Laplacian: replacing the d+1 graph Laplacian matrix L = D − W , where D is the diagonal degree matrix, with L 2 (matrix to d+1 the 2 power). In the infinite unlabeled data limit, this corresponds to regularizing all order- d+1 2 (mixed) partial derivatives. In the typical case of a low-dimensional manifold in a high dimensional ambient space, the order of iteration should correspond to the intrinsic, rather then ambient, dimensionality, which poses a practical problem of estimating this usually unknown dimensionality. We are not aware of much practical work using the iterated Laplacian, nor a good understanding of its appropriateness for SSL. A different approach leading to a well-posed solution is to include also an ambient regularization term [5]. However, the properties of the solution and in particular its relation to various assumptions about the “smoothness” of y(x) relative to p(x) remain unclear. Acknowledgments The authors would like to thank the anonymous referees for valuable suggestions. The research of BN was supported by the Israel Science Foundation (grant 432/06). 8 References [1] R.A. Adams, Sobolev Spaces, Academic Press (New York), 1975. [2] A. Azran, The rendevous algorithm: multiclass semi-supervised learning with Markov Random Walks, ICML, 2007. [3] M. Belkin, P. Niyogi, Using manifold structure for partially labelled classification, NIPS, vol. 15, 2003. [4] M. Belkin and P. Niyogi, Convergence of Laplacian Eigenmaps, NIPS 19, 2007. [5] M. Belkin, P. Niyogi and S. Sindhwani, Manifold Regularization: A Geometric Framework for Learning from Labeled and Unlabeled Examples, JMLR, 7:2399-2434, 2006. [6] Y. Bengio, O. Delalleau, N. Le Roux, label propagation and quadratic criterion, in Semi-Supervised Learning, Chapelle, Scholkopf and Zien, editors, MIT Press, 2006. [7] O. Bosquet, O. Chapelle, M. Hein, Measure Based Regularization, NIPS, vol. 16, 2004. [8] M. Hein, Uniform convergence of adaptive graph-based regularization, COLT, 2006. [9] J. Lafferty, L. Wasserman, Statistical Analysis of Semi-Supervised Regression, NIPS, vol. 20, 2008. [10] U. von Luxburg, M. Belkin and O. Bousquet, Consistency of spectral clustering, Annals of Statistics, vol. 36(2), 2008. [11] M. Meila, J. Shi. A random walks view of spectral segmentation, AI and Statistics, 2001. [12] B. Nadler, S. Lafon, I.G. Kevrekidis, R.R. Coifman, Diffusion maps, spectral clustering and eigenfunctions of Fokker-Planck operators, NIPS, vol. 18, 2006. [13] B. Sch¨ lkopf, A. Smola, Learning with Kernels, MIT Press, 2002. o [14] D. Zhou, O. Bousquet, T. Navin Lal, J. Weston, B. Sch¨ lkopf, Learning with local and global consistency, o NIPS, vol. 16, 2004. [15] X. Zhu, Z. Ghahramani, J. Lafferty, Semi-Supervised Learning using Gaussian fields and harmonic functions, ICML, 2003. 9

6 0.088656314 116 nips-2009-Information-theoretic lower bounds on the oracle complexity of convex optimization

7 0.084305011 198 nips-2009-Rank-Approximate Nearest Neighbor Search: Retaining Meaning and Speed in High Dimensions

8 0.07963378 91 nips-2009-Fast, smooth and adaptive regression in metric spaces

9 0.078861073 169 nips-2009-Nonlinear Learning using Local Coordinate Coding

10 0.077926598 120 nips-2009-Kernels and learning curves for Gaussian process regression on random graphs

11 0.077471599 212 nips-2009-Semi-Supervised Learning in Gigantic Image Collections

12 0.075655505 118 nips-2009-Kernel Choice and Classifiability for RKHS Embeddings of Probability Distributions

13 0.07489647 190 nips-2009-Polynomial Semantic Indexing

14 0.074109927 260 nips-2009-Zero-shot Learning with Semantic Output Codes

15 0.074072063 137 nips-2009-Learning transport operators for image manifolds

16 0.073079243 213 nips-2009-Semi-supervised Learning using Sparse Eigenfunction Bases

17 0.071842797 143 nips-2009-Localizing Bugs in Program Executions with Graphical Models

18 0.07175716 74 nips-2009-Efficient Bregman Range Search

19 0.070884325 77 nips-2009-Efficient Match Kernel between Sets of Features for Visual Recognition

20 0.069517538 34 nips-2009-Anomaly Detection with Score functions based on Nearest Neighbor Graphs


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, -0.208), (1, 0.101), (2, -0.043), (3, 0.071), (4, -0.032), (5, 0.017), (6, -0.143), (7, 0.125), (8, -0.018), (9, 0.072), (10, -0.008), (11, 0.057), (12, 0.14), (13, 0.109), (14, -0.042), (15, -0.156), (16, 0.131), (17, 0.157), (18, -0.023), (19, -0.093), (20, -0.034), (21, 0.361), (22, 0.173), (23, 0.093), (24, 0.027), (25, -0.135), (26, -0.052), (27, 0.218), (28, 0.003), (29, 0.073), (30, -0.074), (31, -0.004), (32, 0.15), (33, 0.072), (34, 0.216), (35, 0.001), (36, -0.027), (37, -0.002), (38, -0.08), (39, -0.015), (40, 0.031), (41, -0.066), (42, -0.109), (43, 0.066), (44, -0.014), (45, 0.096), (46, 0.091), (47, 0.037), (48, -0.101), (49, -0.016)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.95901853 142 nips-2009-Locality-sensitive binary codes from shift-invariant kernels

Author: Maxim Raginsky, Svetlana Lazebnik

Abstract: This paper addresses the problem of designing binary codes for high-dimensional data such that vectors that are similar in the original space map to similar binary strings. We introduce a simple distribution-free encoding scheme based on random projections, such that the expected Hamming distance between the binary codes of two vectors is related to the value of a shift-invariant kernel (e.g., a Gaussian kernel) between the vectors. We present a full theoretical analysis of the convergence properties of the proposed scheme, and report favorable experimental performance as compared to a recent state-of-the-art method, spectral hashing.

2 0.9426111 135 nips-2009-Learning to Hash with Binary Reconstructive Embeddings

Author: Brian Kulis, Trevor Darrell

Abstract: Fast retrieval methods are increasingly critical for many large-scale analysis tasks, and there have been several recent methods that attempt to learn hash functions for fast and accurate nearest neighbor searches. In this paper, we develop an algorithm for learning hash functions based on explicitly minimizing the reconstruction error between the original distances and the Hamming distances of the corresponding binary embeddings. We develop a scalable coordinate-descent algorithm for our proposed hashing objective that is able to efficiently learn hash functions in a variety of settings. Unlike existing methods such as semantic hashing and spectral hashing, our method is easily kernelized and does not require restrictive assumptions about the underlying distribution of the data. We present results over several domains to demonstrate that our method outperforms existing state-of-the-art techniques. 1

3 0.41694596 198 nips-2009-Rank-Approximate Nearest Neighbor Search: Retaining Meaning and Speed in High Dimensions

Author: Parikshit Ram, Dongryeol Lee, Hua Ouyang, Alexander G. Gray

Abstract: The long-standing problem of efďŹ cient nearest-neighbor (NN) search has ubiquitous applications ranging from astrophysics to MP3 ďŹ ngerprinting to bioinformatics to movie recommendations. As the dimensionality of the dataset increases, exact NN search becomes computationally prohibitive; (1+đ?œ–) distance-approximate NN search can provide large speedups but risks losing the meaning of NN search present in the ranks (ordering) of the distances. This paper presents a simple, practical algorithm allowing the user to, for the ďŹ rst time, directly control the true accuracy of NN search (in terms of ranks) while still achieving the large speedups over exact NN. Experiments on high-dimensional datasets show that our algorithm often achieves faster and more accurate results than the best-known distance-approximate method, with much more stable behavior. 1

4 0.37170056 91 nips-2009-Fast, smooth and adaptive regression in metric spaces

Author: Samory Kpotufe

Abstract: It was recently shown that certain nonparametric regressors can escape the curse of dimensionality when the intrinsic dimension of data is low ([1, 2]). We prove some stronger results in more general settings. In particular, we consider a regressor which, by combining aspects of both tree-based regression and kernel regression, adapts to intrinsic dimension, operates on general metrics, yields a smooth function, and evaluates in time O(log n). We derive a tight convergence rate of the form n−2/(2+d) where d is the Assouad dimension of the input space. 1

5 0.367127 143 nips-2009-Localizing Bugs in Program Executions with Graphical Models

Author: Laura Dietz, Valentin Dallmeier, Andreas Zeller, Tobias Scheffer

Abstract: We devise a graphical model that supports the process of debugging software by guiding developers to code that is likely to contain defects. The model is trained using execution traces of passing test runs; it reflects the distribution over transitional patterns of code positions. Given a failing test case, the model determines the least likely transitional pattern in the execution trace. The model is designed such that Bayesian inference has a closed-form solution. We evaluate the Bernoulli graph model on data of the software projects AspectJ and Rhino. 1

6 0.3424868 34 nips-2009-Anomaly Detection with Score functions based on Nearest Neighbor Graphs

7 0.32593989 238 nips-2009-Submanifold density estimation

8 0.32121566 182 nips-2009-Optimal Scoring for Unsupervised Learning

9 0.30925167 32 nips-2009-An Online Algorithm for Large Scale Image Similarity Learning

10 0.30659434 77 nips-2009-Efficient Match Kernel between Sets of Features for Visual Recognition

11 0.30517533 139 nips-2009-Linear-time Algorithms for Pairwise Statistical Problems

12 0.30484667 220 nips-2009-Slow Learners are Fast

13 0.30000913 190 nips-2009-Polynomial Semantic Indexing

14 0.29293185 74 nips-2009-Efficient Bregman Range Search

15 0.27363548 30 nips-2009-An Integer Projected Fixed Point Method for Graph Matching and MAP Inference

16 0.26460195 33 nips-2009-Analysis of SVM with Indefinite Kernels

17 0.26377806 227 nips-2009-Speaker Comparison with Inner Product Discriminant Functions

18 0.25600556 8 nips-2009-A Fast, Consistent Kernel Two-Sample Test

19 0.25371638 234 nips-2009-Streaming k-means approximation

20 0.24877836 116 nips-2009-Information-theoretic lower bounds on the oracle complexity of convex optimization


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(24, 0.072), (25, 0.055), (35, 0.053), (36, 0.094), (39, 0.04), (58, 0.09), (61, 0.013), (71, 0.056), (81, 0.025), (86, 0.147), (89, 0.216), (91, 0.051)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.87023681 142 nips-2009-Locality-sensitive binary codes from shift-invariant kernels

Author: Maxim Raginsky, Svetlana Lazebnik

Abstract: This paper addresses the problem of designing binary codes for high-dimensional data such that vectors that are similar in the original space map to similar binary strings. We introduce a simple distribution-free encoding scheme based on random projections, such that the expected Hamming distance between the binary codes of two vectors is related to the value of a shift-invariant kernel (e.g., a Gaussian kernel) between the vectors. We present a full theoretical analysis of the convergence properties of the proposed scheme, and report favorable experimental performance as compared to a recent state-of-the-art method, spectral hashing.

2 0.85246855 51 nips-2009-Clustering sequence sets for motif discovery

Author: Jong K. Kim, Seungjin Choi

Abstract: Most of existing methods for DNA motif discovery consider only a single set of sequences to find an over-represented motif. In contrast, we consider multiple sets of sequences where we group sets associated with the same motif into a cluster, assuming that each set involves a single motif. Clustering sets of sequences yields clusters of coherent motifs, improving signal-to-noise ratio or enabling us to identify multiple motifs. We present a probabilistic model for DNA motif discovery where we identify multiple motifs through searching for patterns which are shared across multiple sets of sequences. Our model infers cluster-indicating latent variables and learns motifs simultaneously, where these two tasks interact with each other. We show that our model can handle various motif discovery problems, depending on how to construct multiple sets of sequences. Experiments on three different problems for discovering DNA motifs emphasize the useful behavior and confirm the substantial gains over existing methods where only a single set of sequences is considered.

3 0.83556098 55 nips-2009-Compressed Least-Squares Regression

Author: Odalric Maillard, Rémi Munos

Abstract: We consider the problem of learning, from K data, a regression function in a linear space of high dimension N using projections onto a random subspace of lower dimension M . From any algorithm minimizing the (possibly penalized) empirical risk, we provide bounds on the excess risk of the estimate computed in the projected subspace (compressed domain) in terms of the excess risk of the estimate built in the high-dimensional space (initial domain). We show that solving the problem in the compressed domain instead of the initial domain reduces the estimation error at the price of an increased (but controlled) approximation error. We apply the analysis to Least-Squares (LS) regression and discuss the excess risk and numerical complexity of the resulting “Compressed Least Squares Re√ gression” (CLSR) in terms of N , K, and M . When we choose M = O( K), we √ show that CLSR has an estimation error of order O(log K/ K). 1 Problem setting We consider a regression problem where we observe data DK = ({xk , yk }k≤K ) (where xk ∈ X and yk ∈ R) are assumed to be independently and identically distributed (i.i.d.) from some distribution P , where xk ∼ PX and yk = f ∗ (xk ) + ηk (xk ), where f ∗ is the (unknown) target function, and ηk a centered independent noise of variance σ 2 (xk ). For a given class of functions F, and f ∈ F, we define the empirical (quadratic) error def LK (f ) = 1 K K [yk − f (xk )]2 , k=1 and the generalization (quadratic) error def L(f ) = E(X,Y )∼P [(Y − f (X))2 ]. Our goal is to return a regression function f ∈ F with lowest possible generalization error L(f ). Notations: In the sequel we will make use of the following notations about norms: for h : X → R, we write ||h||P for the L2 norm of h with respect to (w.r.t.) the measure P , ||h||PK for the L2 norm n 2 1/2 of h w.r.t. the empirical measure PK , and for u ∈ Rn , ||u|| denotes by default . i=1 ui The measurable function minimizing the generalization error is f ∗ , but it may be the case that f ∗ ∈ F. For any regression function f , we define the excess risk / L(f ) − L(f ∗ ) = ||f − f ∗ ||2 , P which decomposes as the sum of the estimation error L(f ) − inf f ∈F L(f ) and the approximation error inf f ∈F L(f ) − L(f ∗ ) = inf f ∈F ||f − f ∗ ||2 which measures the distance between f ∗ and the P function space F. 1 In this paper we consider a class of linear functions FN defined as the span of a set of N functions def def N {ϕn }1≤n≤N called features. Thus: FN = {fα = n=1 αn ϕn , α ∈ RN }. When the number of data K is larger than the number of features N , the ordinary Least-Squares Regression (LSR) provides the LS solution fα which is the minimizer of the empirical risk LK (f ) b 1 in FN . Note that here LK (fα ) rewrites K ||Φα − Y ||K where Φ is the K × N matrix with elements (ϕn (xk ))1≤n≤N,1≤k≤K and Y the K-vector with components (yk )1≤k≤K . Usual results provide bound on the estimation error as a function of the capacity of the function space and the number of data. In the case of linear approximation, the capacity measures (such as covering numbers [23] or the pseudo-dimension [16]) depend on the number of features (for example the pseudo-dimension is at most N + 1). For example, let fα be a LS estimate (minimizer of LK b in FN ), then (a more precise statement will be stated later in Subsection 3) the expected estimation error is bounded as: N log K E L(fα ) − inf L(f ) ≤ cσ2 , (1) b f ∈FN K def where c is a universal constant, σ = supx∈X σ(x), and the expectation is taken with respect to P . Now, the excess risk is the sum of this estimation error and the approximation error inf f ∈FN ||f − f ∗ ||P of the class FN . Since the later usually decreases when the number of features N increases [13] (e.g. when N FN is dense in L2 (P )), we see the usual tradeoff between small estimation error (low N ) and small approximation error (large N ). In this paper we are interested in the setting when N is large so that the approximation error is small. Whenever N is larger than K we face the overfitting problem since there are more parameters than actual data (more variables than constraints), which is illustrated in the bound (1) which provides no information about the generalization ability of any LS estimate. In addition, there are many minimizers (in fact a vector space of same dimension as the null space of ΦT Φ) of the empirical risk. To overcome the problem, several approaches have been proposed in the literature: • LS solution with minimal norm: The solution is the minimizer of the empirical error with minimal (l1 or l2 )-norm: α = arg minΦα=Y ||α||1 or 2 , (or a robust solution arg min||Φα−Y ||2 ≤ε ||α||1 ). The choice of 2 -norm yields the ordinary LS solution. The choice of 1 -norm has been used for generating sparse solutions (e.g. the Basis Pursuit [10]), and assuming that the target function admits a sparse decomposition, the field of Compressed Sensing [9, 21] provides sufficient conditions for recovering the exact solution. However, such conditions (e.g. that Φ possesses a Restricted Isometric Property (RIP)) does not hold in general in this regression setting. On another aspect, solving these problems (both for l1 or l2 -norm) when N is large is numerically expensive. • Regularization. The solution is the minimizer of the empirical error plus a penalty term, for example f = arg min LK (f ) + λ||f ||p , for p = 1 or 2. p f ∈FN where λ is a parameter and usual choices for the norm are 2 (ridge-regression [20]) and 1 (LASSO [19]). A close alternative is the Dantzig selector [8, 5] which solves: α = arg min||α||1 ≤λ ||ΦT (Y − Φα)||∞ . The numerical complexity and generalization bounds of those methods depend on the sparsity of the target function decomposition in FN . Now if we possess a sequence of function classes (FN )N ≥1 with increasing capacity, we may perform structural risk minimization [22] by solving in each model the empirical risk penalized by a term that depends on the size of the model: fN = arg minf ∈FN ,N ≥1 LK (f ) + pen(N, K), where the penalty term measures the capacity of the function space. In this paper we follow another approach where instead of searching in the large space FN (where N > K) for a solution that minimizes the empirical error plus a penalty term, we simply search for the empirical error minimizer in a (randomly generated) lower dimensional subspace GM ⊂ FN (where M < K). Our contribution: We consider a set of M random linear combinations of the initial N features and perform our favorite LS regression algorithm (possibly regularized) using those “compressed 2 features”. This is equivalent to projecting the K points {ϕ(xk ) ∈ RN , k = 1..K} from the initial domain (of size N ) onto a random subspace of dimension M , and then performing the regression in the “compressed domain” (i.e. span of the compressed features). This is made possible because random projections approximately preserve inner products between vectors (by a variant of the Johnson-Lindenstrauss Lemma stated in Proposition 1. Our main result is a bound on the excess risk of a linear estimator built in the compressed domain in terms of the excess risk of the linear estimator built in the initial domain (Section 2). We further detail the case of ordinary Least-Squares Regression (Section 3) and discuss, in terms of M , N , K, the different tradeoffs concerning the excess risk (reduced estimation error in the compressed domain versus increased approximation error introduced by the random projection) and the numerical complexity (reduced complexity of solving the LSR in the compressed domain versus the additional load of performing the projection). √ As a consequence, we show that by choosing M = O( K) projections we define a Compressed Least-Squares Regression which uses O(N K 3/2 ) elementary operations to compute a regression √ function with estimation error (relatively to the initial function space FN ) of order log K/ K up to a multiplicative factor which depends on the best approximation of f ∗ in FN . This is competitive with the best methods, up to our knowledge. Related works: Using dimension reduction and random projections in various learning areas has received considerable interest over the past few years. In [7], the authors use a SVM algorithm in a compressed space for the purpose of classification and show that their resulting algorithm has good generalization properties. In [25], the authors consider a notion of compressed linear regression. For data Y = Xβ + ε, where β is the target and ε a standard noise, they use compression of the set of data, thus considering AY = AXβ + Aε, where A has a Restricted Isometric Property. They provide an analysis of the LASSO estimator built from these compressed data, and discuss a property called sparsistency, i.e. the number of random projections needed to recover β (with high probability) when it is sparse. These works differ from our approach in the fact that we do not consider a compressed (input and/or output) data space but a compressed feature space instead. In [11], the authors discuss how compressed measurements may be useful to solve many detection, classification and estimation problems without having to reconstruct the signal ever. Interestingly, they make no assumption about the signal being sparse, like in our work. In [6, 17], the authors show how to map a kernel k(x, y) = ϕ(x) · ϕ(y) into a low-dimensional space, while still approximately preserving the inner products. Thus they build a low-dimensional feature space specific for (translation invariant) kernels. 2 Linear regression in the compressed domain We remind that the initial set of features is {ϕn : X → def N FN = {fα = n=1 αn ϕn , α ∈ components (ϕn (x))n≤N . Let us R, 1 ≤ n ≤ N } and the initial domain R } is the span of those features. We write ϕ(x) the N -vector of N now define the random projection. Let A be a M × N matrix of i.i.d. elements drawn for some distribution ρ. Examples of distributions are: • Gaussian random variables N (0, 1/M ), √ • ± Bernoulli distributions, i.e. which takes values ±1/ M with equal probability 1/2, • Distribution taking values ± 3/M with probability 1/6 and 0 with probability 2/3. The following result (proof in the supplementary material) states the property that inner-product are approximately preserved through random projections (this is a simple consequence of the JohnsonLindenstrauss Lemma): Proposition 1 Let (uk )1≤k≤K and v be vectors of RN . Let A be a M × N matrix of i.i.d. elements drawn from one of the previously defined distributions. For any ε > 0, δ > 0, for M ≥ ε2 1 ε3 log 4K , we have, with probability at least 1 − δ, for all k ≤ K, δ 4 − 6 |Auk · Av − uk · v| ≤ ε||uk || ||v||. 3 def We now introduce the set of M compressed features (ψm )1≤m≤M such that ψm (x) = N We also write ψ(x) the M -vector of components (ψm (x))m≤M . Thus n=1 Am,n ϕn (x). ψ(x) = Aϕ(x). We define the compressed domain GM = {gβ = m=1 βm ψm , β ∈ RM } the span of the compressed features (vector space of dimension at most M ). Note that each ψm ∈ FN , thus GM is a subspace of FN . def 2.1 M Approximation error We now compare the approximation error assessed in the compressed domain GM versus in the initial space FN . This applies to the linear algorithms mentioned in the introduction such as ordinary LS regression (analyzed in details in Section 3), but also its penalized versions, e.g. LASSO and ridge regression. Define α+ = arg minα∈RN L(fα ) − L(f ∗ ) the parameter of the best regression function in FN . Theorem 1 For any δ > 0, any M ≥ 15 log(8K/δ), let A be a random M × N matrix defined like in Proposition 1, and GM be the compressed domain resulting from this choice of A. Then with probability at least 1 − δ, inf ||g−f ∗ ||2 ≤ P g∈GM 8 log(8K/δ) + 2 ||α || M E ||ϕ(X)||2 +2 sup ||ϕ(x)||2 x∈X log 4/δ + inf ||f −f ∗ ||2 . P f ∈FN 2K (2) This theorem shows the tradeoff in terms of estimation and approximation errors for an estimator g obtained in the compressed domain compared to an estimator f obtained in the initial domain: • Bounds on the estimation error of g in GM are usually smaller than that of f in FN when M < N (since the capacity of FN is larger than that of GM ). • Theorem 1 says that the approximation error assessed in GM increases by at most O( log(K/δ) )||α+ ||2 E||ϕ(X)||2 compared to that in FN . M def def Proof: Let us write f + = fα+ = arg minf ∈FN ||f − f ∗ ||P and g + = gAα+ . The approximation error assessed in the compressed domain GM is bounded as inf ||g − f ∗ ||2 P g∈GM ≤ ||g + − f ∗ ||2 = ||g + − f + ||2 + ||f + − f ∗ ||2 , P P P (3) since f + is the orthogonal projection of f ∗ on FN and g + belongs to FN . We now bound ||g + − def def f + ||2 using concentration inequalities. Define Z(x) = Aα+ · Aϕ(x) − α+ · ϕ(x). Define ε2 = P log(8K/δ) 8 M log(8K/δ). For M ≥ 15 log(8K/δ) we have ε < 3/4 thus M ≥ ε2 /4−ε3 /6 . Proposition 1 applies and says that on an event E of probability at least 1 − δ/2, we have for all k ≤ K, def |Z(xk )| ≤ ε||α+ || ||ϕ(xk )|| ≤ ε||α+ || sup ||ϕ(x)|| = C (4) x∈X On the event E, we have with probability at least 1 − δ , ||g + − f + ||2 P = ≤ ≤ EX∼PX |Z(X)|2 ≤ ε2 ||α+ ||2 ε2 ||α+ ||2 1 K 1 K K |Z(xk )|2 + C 2 k=1 K ||ϕ(xk )||2 + sup ||ϕ(x)||2 x∈X k=1 E ||ϕ(X)||2 + 2 sup ||ϕ(x)||2 x∈X log(2/δ ) 2K log(2/δ ) 2K log(2/δ ) . 2K where we applied two times Chernoff-Hoeffding’s inequality. Combining with (3), unconditioning, and setting δ = δ/2 then with probability at least (1 − δ/2)(1 − δ ) ≥ 1 − δ we have (2). 4 2.2 Computational issues We now discuss the relative computational costs of a given algorithm applied either in the initial or in the compressed domain. Let us write Cx(DK , FN , P ) the complexity (e.g. number of elementary operations) of an algorithm A to compute the regression function f when provided with the data DK and function space FN . We plot in the table below, both for the initial and the compressed versions of the algorithm A, the order of complexity for (i) the cost for building the feature matrix, (ii) the cost for computing the estimator, (iii) the cost for making one prediction (i.e. computing f (x) for any x): Construction of the feature matrix Computing the regression function Making one prediction Initial domain NK Cx(DK , FN , P ) N Compressed domain N KM Cx(DK , GM , P ) NM Note that the values mentioned for the compressed domain are upper-bounds on the real complexity and do not take into account the possible sparsity of the projection matrix A (which would speed up matrix computations, see e.g. [2, 1]). 3 Compressed Least-Squares Regression We now analyze the specific case of Least-Squares Regression. 3.1 Excess risk of ordinary Least Squares regression In order to bound the estimation error, we follow the approach of [13] which truncates (up to the level ±L where L is a bound, assumed to be known, on ||f ∗ ||∞ ) the prediction of the LS regression function. The ordinary LS regression provides the regression function fα where b α= argmin α∈argminα ∈ RN ||α||. ||Y −Φα || Note that ΦΦT α = ΦT Y , hence α = Φ† Y ∈ RN where Φ† is the Penrose pseudo-inverse of Φ1 . def Then the truncated predictor is: fL (x) = TL [fα (x)], where b def TL (u) = u if |u| ≤ L, L sign(u) otherwise. Truncation after the computation of the parameter α ∈ RN , which is the solution of an unconstrained optimization problem, is easier than solving an optimization problem under the constraint that ||α|| is small (which is the approach followed in [23]) and allows for consistency results and prediction bounds. Indeed, the excess risk of fL is bounded as 1 + log K E(||f − f ∗ ||2 ) ≤ c max{σ2 , L2 } N + 8 inf ||f − f ∗ ||2 (5) P P f ∈FN K where a bound on c is 9216 (see [13]). We have a simpler bound when we consider the expectation EY conditionally on the input data: N EY (||f − f ∗ ||2 K ) ≤ σ2 + inf ||f − f ∗ ||2 K (6) P P K f ∈F Remark: Note that because we use the quadratic loss function, by following the analysis in [3], or by deriving tight bounds on the Rademacher complexity [14] and following Theorem 5.2 of Koltchinskii’s Saint Flour course, it is actually possible to state assumptions under which we can remove the log K term in (5). We will not further detail such bounds since our motivation here is not to provide the tightest possible bounds, but rather to show how the excess risk bound for LS regression in the initial domain extends to the compressed domain. 1 In the full rank case, Φ† = (ΦT Φ)−1 ΦT when K ≥ N and Φ† = ΦT (ΦΦT )−1 when K ≤ N 5 3.2 Compressed Least-Squares Regression (CLSR) CLSR is defined as the ordinary LSR in the compressed domain. Let β = Ψ† Y ∈ RM , where Ψ is the K × M matrix with elements (ψm (xk ))1≤m≤M,1≤k≤K . The CLSR estimate is defined as def gL (x) = TL [gβ (x)]. From Theorem 1, (5) and (6), we deduce the following excess risk bounds for b the CLSR estimate: √ ||α+ || E||ϕ(X)||2 K log(8K/δ) Corollary 1 For any δ > 0, set M = 8 max(σ,L) c (1+log K) . Then whenever M ≥ 15 log(8K/δ), with probability at least 1 − δ, the expected excess risk of the CLSR estimate is bounded as √ E(||gL − f ∗ ||2 ) ≤ 16 c max{σ, L}||α+ || E||ϕ(X)||2 P × 1+ supx ||ϕ(x)||2 E||ϕ(X)||2 (1 + log K) log(8K/δ) K log 4/δ + 8 inf ||f − f ∗ ||2 . P f ∈FN 2K (7) √ ||α+ || E||ϕ(X)||2 Now set M = 8K log(8K/δ). Assume N > K and that the features (ϕk )1≤k≤K σ are linearly independent. Then whenever M ≥ 15 log(8K/δ), with probability at least 1 − δ, the expected excess risk of the CLSR estimate conditionally on the input samples is upper bounded as 2 log(8K/δ) supx ||ϕ(x)||2 1+ K E||ϕ(X)||2 EY (||gL − f ∗ ||2 K ) ≤ 4σ||α+ || E||ϕ(X)||2 P log 4/δ . 2K Proof: Whenever M ≥ 15 log(8K/δ) we deduce from Theorem 1 and (5) that the excess risk of gL is bounded as E(||gL − f ∗ ||2 ) ≤ c max{σ2 , L2 } P +8 8 log(8K/δ) + 2 ||α || M 1 + log K M K E||ϕ(X)||2 + 2 sup ||ϕ(x)||2 x log 4/δ + inf ||f − f ∗ ||2 . P f ∈FN 2K By optimizing on M , we deduce (7). Similarly, using (6) we deduce the following bound on EY (||gL − f ∗ ||2 K ): P σ2 8 M + log(8K/δ)||α+ ||2 K M E||ϕ(X)||2 + 2 sup ||ϕ(x)||2 x log 4/δ + inf ||f − f ∗ ||2 K . P f ∈FN 2K By optimizing on M and noticing that inf f ∈FN ||f − f ∗ ||2 K = 0 whenever N > K and the features P (ϕk )1≤k≤K are linearly independent, we deduce the second result. Remark 1 Note that the second term in the parenthesis of (7) is negligible whenever K Thus we have the expected excess risk log K/δ + inf ||f − f ∗ ||2 . P f ∈FN K E(||gL − f ∗ ||2 ) = O ||α+ || E||ϕ(X)||2 √ P log 1/δ. (8) The choice of M in the previous corollary depends on ||α+ || and E||ϕ(X)|| which are a priori unknown (since f ∗ and PX are unknown). If we set M independently of ||α+ ||, then an additional multiplicative factor of ||α+ || appears in the bound, and if we replace E||ϕ(X)|| by its bound supx ||ϕ(x)|| (which is known) then this latter factor will appear instead of the former in the bound. Complexity of CLSR: The complexity of LSR for computing the regression function in the compressed domain only depends on M and K, and is (see e.g. [4]) Cx(DK , GM , P ) = O(M K 2 ) which √ is of order O(K 5/2 ) when we choose the optimized number of projections M = O( K). However the leading term when using CLSR is the cost for building the Ψ matrix: O(N K 3/2 ). 6 4 4.1 Discussion The factor ||α+ || E||ϕ(X)||2 In light of Corollary 1, the important factor which will determine whether the CLSR provides low generalization error or not is ||α+ || E||ϕ(X)||2 . This factor indicates that a good set of features (for CLSR) should be such that the norm of those features as well as the norm of the parameter α+ of the projection of f ∗ onto the span of those features should be small. A natural question is whether this product can be made small for appropriate choices of features. We now provide two specific cases for which this is actually the case: (1) when the features are rescaled orthonormal basis functions, and (2) when the features are specific wavelet functions. In both cases, we relate the bound to an assumption of regularity on the function f ∗ , and show that the dependency w.r.t. N decreases when the regularity increases, and may even vanish. Rescaled Orthonormal Features: Consider a set of orthonormal functions (ηi )i≥1 w.r.t a measure µ, i.e. ηi , ηj µ = δi,j . In addition we assume that the law of the input data is dominated by µ, i.e. PX ≤ Cµ where C is a constant. For instance, this is the case when the set X is compact, µ is the uniform measure and PX has bounded density. def We define the set of N features as: ϕi = ci ηi , where ci > 0, for i ∈ {1, . . . , N }. Then any f ∈ FN decomposes as f = 2 we have: ||α|| = ||α+ ||2 E||ϕ||2 ≤ C N bi 2 i=1 ( ci ) N bi 2 i=1 ( ci ) and N i=1 N bi i=1 ci ϕi , where N 2 2 i=1 ci X ηi (x)dPX (x) f, ηi ηi = E||ϕ|| = 2 def bi = f, ηi . Thus ≤ C N 2 i=1 ci . Thus N 2 i=1 ci . Now, linear approximation theory (Jackson-type theorems) tells us that assuming a function f ∗ ∈ L2 (µ) is smooth, it may be decomposed onto the span of the N first (ηi )i∈{1,...,N } functions with decreasing coefficients |bi | ≤ i−λ for some λ ≥ 0 that depends on the smoothness of f ∗ . For example the class of functions with bounded total variation may be decomposed with Fourier basis (in dimension 1) with coefficients |bi | ≤ ||f ||V /(2πi). Thus here λ = 1. Other classes (such as Sobolev spaces) lead to larger values of λ related to the order of differentiability. √ N By choosing ci = i−λ/2 , we have ||α+ || E||ϕ||2 ≤ C i=1 i−λ . Thus if λ > 1, then this term is bounded by a constant that does not depend on N . If λ = 1 then it is bounded by O(log N ), and if 0 < λ < 1, then it is bounded by O(N 1−λ ). However any orthonormal basis, even rescaled, would not necessarily yield a small ||α+ || E||ϕ||2 term (this is all the more true when the dimension of X is large). The desired property that the coefficients (α+ )i of the decomposition of f ∗ rapidly decrease to 0 indicates that hierarchical bases, such as wavelets, that would decompose the function at different scales, may be interesting. Wavelets: Consider an infinite family of wavelets in [0, 1]: (ϕ0 ) = (ϕ0 ) (indexed by n ≥ 1 or n h,l equivalently by the scale h ≥ 0 and translation 0 ≤ l ≤ 2h − 1) where ϕ0 (x) = 2h/2 ϕ0 (2h x − l) h,l and ϕ0 is the mother wavelet. Then consider N = 2H features (ϕh,l )1≤h≤H defined as the rescaled def wavelets ϕh,l = ch 2−h/2 ϕ0 , where ch > 0 are some coefficients. Assume the mother wavelet h,l is C p (for p ≥ 1), has at least p vanishing moments, and that for all h ≥ 0, supx l ϕ0 (2h x − l)2 ≤ 1. Then the following result (proof in the supplementary material) provides a bound on supx∈X ||ϕ(x)||2 (thus on E||ϕ(X)||2 ) by a constant independent of N : Proposition 2 Assume that f ∗ is (L, γ)-Lipschitz (i.e. for all v ∈ X there exists a polynomial pv of degree γ such that for all u ∈ X , |f (u) − pv (u)| ≤ L|u − v|γ ) with 1/2 < γ ≤ p. Then setting γ 1 ch = 2h(1−2γ)/4 , we have ||α+ || supx ||ϕ(x)|| ≤ L 1−22 |ϕ0 |, which is independent of N . 1/2−γ 0 Notice that the Haar walevets has p = 1 vanishing moment but is not C 1 , thus the Proposition does not apply directly. However direct computations show that if f ∗ is L-Lipschitz (i.e. γ = 1) then L 0 αh,l ≤ L2−3h/2−2 , and thus ||α+ || supx ||ϕ(x)|| ≤ 4(1−2−1/2 ) with ch = 2−h/4 . 7 4.2 Comparison with other methods In the case when the factor ||α+ || E||ϕ(X)||2 does not depend on N (such as in the previous example), the bound (8) on the excess risk of CLSR states that the estimation error (assessed in √ √ terms of FN ) of CLSR is O(log K/ K). It is clear that whenever N > K (which is the case of interest here), this is better than the ordinary LSR in the initial domain, whose estimation error is O(N log K/K). It is difficult to compare this result with LASSO (or the Dantzig selector that has similar properties [5]) for which an important aspect is to design sparse regression functions or to recover a solution assumed to be sparse. From [12, 15, 24] one deduces that under some assumptions, the estimation error of LASSO is of order S log N where S is the sparsity (number of non-zero coefficients) of the K√ best regressor f + in FN . If S < K then LASSO is more interesting than CLSR in terms of excess risk. Otherwise CLSR may be an interesting alternative although this method does not make any assumption about the sparsity of f + and its goal is not to recover a possible sparse f + but only to make good predictions. However, in some sense our method finds a sparse solution in the fact that the regression function gL lies in a space GM of small dimension M N and can thus be expressed using only M coefficients. Now in terms of numerical complexity, CLSR requires O(N K 3/2 ) operations to build the matrix and compute the regression function, whereas according to [18], the (heuristical) complexity of the LASSO algorithm is O(N K 2 ) in the best cases (assuming that the number of steps required for convergence is O(K), which is not proved theoretically). Thus CLSR seems to be a good and simple competitor to LASSO. 5 Conclusion We considered the case when the number of features N is larger than the number of data K. The result stated in Theorem 1 enables to analyze the excess risk of any linear regression algorithm (LS or its penalized versions) performed in the compressed domain GM versus in the initial space FN . In the compressed domain the estimation error is reduced but an additional (controlled) approximation error (when compared to the best regressor in FN ) comes into the picture. In the case of LS regression, when the term ||α+ || E||ϕ(X)||2 has a mild dependency on N , then by choosing a √ random subspace of dimension M = O( K), CLSR has an estimation error (assessed in terms of √ FN ) bounded by O(log K/ K) and has numerical complexity O(N K 3/2 ). In short, CLSR provides an alternative to usual penalization techniques where one first selects a random subspace of lower dimension and then performs an empirical risk minimizer in this subspace. Further work needs to be done to provide additional settings (when the space X is of dimension > 1) for which the term ||α+ || E||ϕ(X)||2 is small. Acknowledgements: The authors wish to thank Laurent Jacques for numerous comments and Alessandro Lazaric and Mohammad Ghavamzadeh for exciting discussions. This work has been supported by French National Research Agency (ANR) through COSINUS program (project EXPLO-RA, ANR-08-COSI-004). References [1] Dimitris Achlioptas. Database-friendly random projections: Johnson-Lindenstrauss with binary coins. Journal of Computer and System Sciences, 66(4):671–687, June 2003. [2] Nir Ailon and Bernard Chazelle. Approximate nearest neighbors and the fast JohnsonLindenstrauss transform. In STOC ’06: Proceedings of the thirty-eighth annual ACM symposium on Theory of computing, pages 557–563, New York, NY, USA, 2006. ACM. [3] Jean-Yves Audibert and Olivier Catoni. Risk bounds in linear regression through pac-bayesian truncation. Technical Report HAL : hal-00360268, 2009. [4] David Bau III and Lloyd N. Trefethen. Numerical linear algebra. Philadelphia: Society for Industrial and Applied Mathematics, 1997. 8 [5] Peter J. Bickel, Ya’acov Ritov, and Alexandre B. Tsybakov. Simultaneous analysis of Lasso and Dantzig selector. To appear in Annals of Statistics, 2008. [6] Avrim Blum. Random projection, margins, kernels, and feature-selection. Subspace, Latent Structure and Feature Selection, pages 52–68, 2006. [7] Robert Calderbank, Sina Jafarpour, and Robert Schapire. Compressed learning: Universal sparse dimensionality reduction and learning in the measurement domain. Technical Report, 2009. [8] Emmanuel Candes and Terence Tao. The Dantzig selector: Statistical estimation when p is much larger than n. Annals of Statistics, 35:2313, 2007. [9] Emmanuel J. Candes and Justin K. Romberg. Signal recovery from random projections. volume 5674, pages 76–86. SPIE, 2005. [10] S. S. Chen, D. L. Donoho, and M. A. Saunders. Atomic decomposition by basis pursuit. SIAM Journal on Scientific Computing, 20:33–61, 1998. [11] Mark A. Davenport, Michael B. Wakin, and Richard G. Baraniuk. Detection and estimation with compressive measurements. Technical Report TREE 0610, Department of Electrical and Computer Engineering, Rice University, 2006. [12] E. Greenshtein and Y. Ritov. Persistency in high dimensional linear predictor-selection and the virtue of over-parametrization. Bernoulli, 10:971–988, 2004. [13] L. Gy¨ rfi, M. Kohler, A. Krzy˙ ak, and H. Walk. A distribution-free theory of nonparametric o z regression. Springer-Verlag, 2002. [14] Sham M. Kakade, Karthik Sridharan, and Ambuj Tewari. On the complexity of linear prediction: Risk bounds, margin bounds, and regularization. In Daphne Koller, Dale Schuurmans, Yoshua Bengio, and Leon Bottou, editors, Neural Information Processing Systems, pages 793– 800. MIT Press, 2008. [15] Yuval Nardi and Alessandro Rinaldo. On the asymptotic properties of the group Lasso estimator for linear models. Electron. J. Statist., 2:605–633, 2008. [16] D. Pollard. Convergence of Stochastic Processes. Springer Verlag, New York, 1984. [17] Ali Rahimi and Benjamin Recht. Random features for large-scale kernel machines. Neural Information Processing Systems, 2007. [18] Saharon Rosset and Ji Zhu. Piecewise linear regularized solution paths. Annals of Statistics, 35:1012, 2007. [19] Robert Tibshirani. Regression shrinkage and selection via the Lasso. Journal of the Royal Statistical Society, Series B, 58:267–288, 1994. [20] A. N. Tikhonov. Solution of incorrectly formulated problems and the regularization method. Soviet Math Dokl 4, pages 1035–1038, 1963. [21] Yaakov Tsaig and David L. Donoho. Compressed sensing. IEEE Trans. Inform. Theory, 52:1289–1306, 2006. [22] Vladimir N. Vapnik. The nature of statistical learning theory. Springer-Verlag New York, Inc., New York, NY, USA, 1995. [23] Tong Zhang. Covering number bounds of certain regularized linear function classes. Journal of Machine Learning Research, 2:527–550, 2002. [24] Tong Zhang. Some sharp performance bounds for least squares regression with L1 regularization. To appear in Annals of Statistics, 2009. [25] Shuheng Zhou, John D. Lafferty, and Larry A. Wasserman. Compressed regression. In John C. Platt, Daphne Koller, Yoram Singer, and Sam T. Roweis, editors, Neural Information Processing Systems. MIT Press, 2007. 9

4 0.71010858 104 nips-2009-Group Sparse Coding

Author: Samy Bengio, Fernando Pereira, Yoram Singer, Dennis Strelow

Abstract: Bag-of-words document representations are often used in text, image and video processing. While it is relatively easy to determine a suitable word dictionary for text documents, there is no simple mapping from raw images or videos to dictionary terms. The classical approach builds a dictionary using vector quantization over a large set of useful visual descriptors extracted from a training set, and uses a nearest-neighbor algorithm to count the number of occurrences of each dictionary word in documents to be encoded. More robust approaches have been proposed recently that represent each visual descriptor as a sparse weighted combination of dictionary words. While favoring a sparse representation at the level of visual descriptors, those methods however do not ensure that images have sparse representation. In this work, we use mixed-norm regularization to achieve sparsity at the image level as well as a small overall dictionary. This approach can also be used to encourage using the same dictionary words for all the images in a class, providing a discriminative signal in the construction of image representations. Experimental results on a benchmark image classification dataset show that when compact image or dictionary representations are needed for computational efficiency, the proposed approach yields better mean average precision in classification. 1

5 0.70774645 135 nips-2009-Learning to Hash with Binary Reconstructive Embeddings

Author: Brian Kulis, Trevor Darrell

Abstract: Fast retrieval methods are increasingly critical for many large-scale analysis tasks, and there have been several recent methods that attempt to learn hash functions for fast and accurate nearest neighbor searches. In this paper, we develop an algorithm for learning hash functions based on explicitly minimizing the reconstruction error between the original distances and the Hamming distances of the corresponding binary embeddings. We develop a scalable coordinate-descent algorithm for our proposed hashing objective that is able to efficiently learn hash functions in a variety of settings. Unlike existing methods such as semantic hashing and spectral hashing, our method is easily kernelized and does not require restrictive assumptions about the underlying distribution of the data. We present results over several domains to demonstrate that our method outperforms existing state-of-the-art techniques. 1

6 0.70638633 169 nips-2009-Nonlinear Learning using Local Coordinate Coding

7 0.70574534 3 nips-2009-AUC optimization and the two-sample problem

8 0.70303059 119 nips-2009-Kernel Methods for Deep Learning

9 0.70143533 210 nips-2009-STDP enables spiking neurons to detect hidden causes of their inputs

10 0.69904315 137 nips-2009-Learning transport operators for image manifolds

11 0.69466519 87 nips-2009-Exponential Family Graph Matching and Ranking

12 0.69461668 151 nips-2009-Measuring Invariances in Deep Networks

13 0.69459021 32 nips-2009-An Online Algorithm for Large Scale Image Similarity Learning

14 0.69447625 213 nips-2009-Semi-supervised Learning using Sparse Eigenfunction Bases

15 0.69419104 199 nips-2009-Ranking Measures and Loss Functions in Learning to Rank

16 0.6899274 212 nips-2009-Semi-Supervised Learning in Gigantic Image Collections

17 0.68910426 203 nips-2009-Replacing supervised classification learning by Slow Feature Analysis in spiking neural networks

18 0.68815202 112 nips-2009-Human Rademacher Complexity

19 0.68794119 132 nips-2009-Learning in Markov Random Fields using Tempered Transitions

20 0.68678713 136 nips-2009-Learning to Rank by Optimizing NDCG Measure