nips nips2013 nips2013-111 knowledge-graph by maker-knowledge-mining
Source: pdf
Author: John Duchi, Michael Jordan, Brendan McMahan
Abstract: We study stochastic optimization problems when the data is sparse, which is in a sense dual to current perspectives on high-dimensional statistical learning and optimization. We highlight both the difficulties—in terms of increased sample complexity that sparse data necessitates—and the potential benefits, in terms of allowing parallelism and asynchrony in the design of algorithms. Concretely, we derive matching upper and lower bounds on the minimax rate for optimization and learning with sparse data, and we exhibit algorithms achieving these rates. We also show how leveraging sparsity leads to (still minimax optimal) parallel and asynchronous algorithms, providing experimental evidence complementing our theoretical results on several medium to large-scale learning tasks. 1 Introduction and problem setting In this paper, we investigate stochastic optimization problems in which the data is sparse. Formally, let {F (·; ξ), ξ ∈ Ξ} be a collection of real-valued convex functions, each of whose domains contains the convex set X ⊂ Rd . For a probability distribution P on Ξ, we consider the following optimization problem: minimize f (x) := E[F (x; ξ)] = x∈X F (x; ξ)dP (ξ). (1) Ξ By data sparsity, we mean the samples ξ are sparse: assuming that samples ξ lie in Rd , and defining the support supp(x) of a vector x to the set of indices of its non-zero components, we assume supp F (x; ξ) ⊂ supp ξ. (2) The sparsity condition (2) means that F (x; ξ) does not “depend” on the values of xj for indices j such that ξj = 0.1 This type of data sparsity is prevalent in statistical optimization problems and machine learning applications; in spite of its prevalence, study of such problems has been limited. As a motivating example, consider a text classification problem: data ξ ∈ Rd represents words appearing in a document, and we wish to minimize a logistic loss F (x; ξ) = log(1 + exp( ξ, x )) on the data (we encode the label implicitly with the sign of ξ). Such generalized linear models satisfy the sparsity condition (2), and while instances are of very high dimension, in any given instance, very few entries of ξ are non-zero [8]. From a modelling perspective, it thus makes sense to allow a dense predictor x: any non-zero entry of ξ is potentially relevant and important. In a sense, this is dual to the standard approaches to high-dimensional problems; one usually assumes that the data ξ may be dense, but there are only a few relevant features, and thus a parsimonious model x is desirous [2]. So 1 Formally, if πξ denotes the coordinate projection zeroing all indices j of its argument where ξj = 0, then F (πξ (x); ξ) = F (x; ξ) for all x, ξ. This follows from the first-order conditions for convexity [6]. 1 while such sparse data problems are prevalent—natural language processing, information retrieval, and other large data settings all have significant data sparsity—they do not appear to have attracted as much study as their high-dimensional “duals” of dense data and sparse predictors. In this paper, we investigate algorithms and their inherent limitations for solving problem (1) under natural conditions on the data generating distribution. Recent work in the optimization and machine learning communities has shown that data sparsity can be leveraged to develop parallel (and even asynchronous [12]) optimization algorithms [13, 14], but this work does not consider the statistical effects of data sparsity. In another line of research, Duchi et al. [4] and McMahan and Streeter [9] develop “adaptive” stochastic gradient algorithms to address problems in sparse data regimes (2). These algorithms exhibit excellent practical performance and have theoretical guarantees on their convergence, but it is not clear if they are optimal—in that no algorithm can attain better statistical performance—or whether they can leverage parallel computing as in the papers [12, 14]. In this paper, we take a two-pronged approach. First, we investigate the fundamental limits of optimization and learning algorithms in sparse data regimes. In doing so, we derive lower bounds on the optimization error of any algorithm for problems of the form (1) with sparsity condition (2). These results have two main implications. They show that in some scenarios, learning with sparse data is quite difficult, as essentially each coordinate j ∈ [d] can be relevant and must be optimized for. In spite of this seemingly negative result, we are also able to show that the A DAG RAD algorithms of [4, 9] are optimal, and we show examples in which their dependence on the dimension d can be made exponentially better than standard gradient methods. As the second facet of our two-pronged approach, we study how sparsity may be leveraged in parallel computing frameworks to give substantially faster algorithms that still achieve optimal sample complexity in terms of the number of samples ξ used. We develop two new algorithms, asynchronous dual averaging (A SYNC DA) and asynchronous A DAG RAD (A SYNC A DAG RAD), which allow asynchronous parallel solution of the problem (1) for general convex f and X . Combining insights of Niu et al.’s H OGWILD ! [12] with a new analysis, we prove our algorithms achieve linear speedup in the number of processors while maintaining optimal statistical guarantees. We also give experiments on text-classification and web-advertising tasks to illustrate the benefits of the new algorithms. 2 Minimax rates for sparse optimization We begin our study of sparse optimization problems by establishing their fundamental statistical and optimization-theoretic properties. To do this, we derive bounds on the minimax convergence rate of any algorithm for such problems. Formally, let x denote any estimator for a minimizer of the objective (1). We define the optimality gap N for the estimator x based on N samples ξ 1 , . . . , ξ N from the distribution P as N (x, F, X , P ) := f (x) − inf f (x) = EP [F (x; ξ)] − inf EP [F (x; ξ)] . x∈X x∈X This quantity is a random variable, since x is a random variable (it is a function of ξ 1 , . . . , ξ N ). To define the minimax error, we thus take expectations of the quantity N , though we require a bit more than simply E[ N ]. We let P denote a collection of probability distributions, and we consider a collection of loss functions F specified by a collection F of convex losses F : X × ξ → R. We can then define the minimax error for the family of losses F and distributions P as ∗ N (X , P, F) := inf sup sup EP [ x P ∈P F ∈F N (x(ξ 1:N ), F, X , P )], (3) where the infimum is taken over all possible estimators x (an estimator is an optimization scheme, or a measurable mapping x : ΞN → X ) . 2.1 Minimax lower bounds Let us now give a more precise characterization of the (natural) set of sparse optimization problems we consider to provide the lower bound. For the next proposition, we let P consist of distributions supported on Ξ = {−1, 0, 1}d , and we let pj := P (ξj = 0) be the marginal probability of appearance of feature j ∈ {1, . . . , d}. For our class of functions, we set F to consist of functions F satisfying the sparsity condition (2) and with the additional constraint that for g ∈ ∂x F (x; ξ), we have that the jth coordinate |gj | ≤ Mj for a constant Mj < ∞. We obtain 2 Proposition 1. Let the conditions of the preceding paragraph hold. Let R be a constant such that X ⊃ [−R, R]d . Then √ d pj 1 ∗ . Mj min pj , √ N (X , P, F) ≥ R 8 j=1 N log 3 We provide the proof of Proposition 1 in the supplement A.1 in the full version of the paper, providing a few remarks here. We begin by giving a corollary to Proposition 1 that follows when the data ξ obeys a type of power law: let p0 ∈ [0, 1], and assume that P (ξj = 0) = p0 j −α . We have Corollary 2. Let α ≥ 0. Let the conditions of Proposition 1 hold with Mj ≡ M for all j, and assume the power law condition P (ξj = 0) = p0 j −α on coordinate appearance probabilities. Then (1) If d > (p0 N )1/α , ∗ N (X , P, F) ≥ 2−α 1−α p0 p0 (p0 N ) 2α − 1 + d1−α − (p0 N ) α N 1−α 2 MR 8 2−α (2) If d ≤ (p0 N )1/α , ∗ N (X , P, F) ≥ MR 8 p0 N α 1 1 d1− 2 − 1 − α/2 1 − α/2 . . Expanding Corollary 2 slightly, for simplicity assume the number of samples is large enough that d ≤ (p0 N )1/α . Then we find that the lower bound on optimization error is of order p0 1− α p0 p0 d 2 when α < 2, M R log d when α → 2, and M R when α > 2. (4) N N N These results beg the question of tightness: are they improvable? As we see presently, they are not. MR 2.2 Algorithms for attaining the minimax rate To show that the lower bounds of Proposition 1 and its subsequent specializations are sharp, we review a few stochastic gradient algorithms. We begin with stochastic gradient descent (SGD): SGD repeatedly samples ξ ∼ P , computes g ∈ ∂x F (x; ξ), then performs the update x ← ΠX (x − ηg), where η is a stepsize parameter and ΠX denotes Euclidean projection onto X . Standard analyses of stochastic gradient descent [10] show that after N samples ξ i , the SGD estimator x(N ) satisfies R2 M ( d j=1 1 pj ) 2 √ , (5) N where R2 denotes the 2 -radius of X . Dual averaging, due to Nesterov [11] (sometimes called “follow the regularized leader” [5]) is a more recent algorithm. In dual averaging, one again samples g ∈ ∂x F (x; ξ), but instead of updating the parameter vector x one updates a dual vector z by z ← z + g, then computes 1 x ← argmin z, x + ψ(x) , η x∈X E[f (x(N ))] − inf f (x) ≤ O(1) x∈X 2 1 where ψ(x) is a strongly convex function defined over X (often one takes ψ(x) = 2 x 2 ). As we discuss presently, the dual averaging algorithm is somewhat more natural in asynchronous and parallel computing environments, and it enjoys the same type of convergence guarantees (5) as SGD. The A DAG RAD algorithm [4, 9] is an extension of the preceding stochastic gradient methods. It maintains a diagonal matrix S, where upon receiving a new sample ξ, A DAG RAD performs the following: it computes g ∈ ∂x F (x; ξ), then updates 2 Sj ← Sj + gj for j ∈ [d]. The dual averaging variant of A DAG RAD updates the usual dual vector z ← z + g; the update to x is based on S and a stepsize η and computes x ← argmin z, x + x ∈X 3 1 1 x ,S2x 2η . After N samples ξ, the averaged parameter x(N ) returned by A DAG RAD satisfies R∞ M E[f (x(N ))] − inf f (x) ≤ O(1) √ x∈X N d √ pj , (6) j=1 where R∞ denotes the ∞ -radius of X (cf. [4, Section 1.3 and Theorem 5]). By inspection, the A DAG RAD rate (6) matches the lower bound in Proposition 1 and is thus optimal. It is interesting to note, though, that in the power law setting of Corollary 2 (recall the error order (4)), a calculation √ shows that the multiplier for the SGD guarantee (5) becomes R∞ d max{d(1−α)/2 , 1}, while A DA G RAD attains rate at worst R∞ max{d1−α/2 , log d}. For α > 1, the A DAG RAD rate is no worse, √ and for α ≥ 2, is more than d/ log d better—an exponential improvement in the dimension. 3 Parallel and asynchronous optimization with sparsity As we note in the introduction, recent works [12, 14] have suggested that sparsity can yield benefits in our ability to parallelize stochastic gradient-type algorithms. Given the optimality of A DAG RADtype algorithms, it is natural to focus on their parallelization in the hope that we can leverage their ability to “adapt” to sparsity in the data. To provide the setting for our further algorithms, we first revisit Niu et al.’s H OGWILD ! [12]. H OGWILD ! is an asynchronous (parallelized) stochastic gradient algorithm for optimization over product-space domains, meaning that X in problem (1) decomposes as X = X1 × · · · × Xd , where Xj ⊂ R. Fix a stepsize η > 0. A pool of independently running processors then performs the following updates asynchronously to a centralized vector x: 1. Sample ξ ∼ P 2. Read x and compute g ∈ ∂x F (x; ξ) 3. For each j s.t. gj = 0, update xj ← ΠXj (xj − ηgj ). Here ΠXj denotes projection onto the jth coordinate of the domain X . The key of H OGWILD ! is that in step 2, the parameter x is allowed to be inconsistent—it may have received partial gradient updates from many processors—and for appropriate problems, this inconsistency is negligible. Indeed, Niu et al. [12] show linear speedup in optimization time as the number of processors grow; they show this empirically in many scenarios, providing a proof under the somewhat restrictive assumptions that there is at most one non-zero entry in any gradient g and that f has Lipschitz gradients. 3.1 Asynchronous dual averaging A weakness of H OGWILD ! is that it appears only applicable to problems for which the domain X is a product space, and its analysis assumes g 0 = 1 for all gradients g. In effort to alleviate these difficulties, we now develop and present our asynchronous dual averaging algorithm, A SYNC DA. A SYNC DA maintains and upates a centralized dual vector z instead of a parameter x, and a pool of processors perform asynchronous updates to z, where each processor independently iterates: 1. Read z and compute x := argminx∈X 1 z, x + η ψ(x) // Implicitly increment “time” counter t and let x(t) = x 2. Sample ξ ∼ P and let g ∈ ∂x F (x; ξ) // Let g(t) = g. 3. For j ∈ [d] such that gj = 0, update zj ← zj + gj . Because the actual computation of the vector x in A SYNC DA is performed locally on each processor in step 1 of the algorithm, the algorithm can be executed with any proximal function ψ and domain X . The only communication point between any of the processors is the addition operation in step 3. Since addition is commutative and associative, forcing all asynchrony to this point of the algorithm is a natural strategy for avoiding synchronization problems. In our analysis of A SYNC DA, and in our subsequent analysis of the adaptive methods, we require a measurement of time elapsed. With that in mind, we let t denote a time index that exists (roughly) behind-the-scenes. We let x(t) denote the vector x ∈ X computed in the tth step 1 of the A SYNC DA 4 algorithm, that is, whichever is the tth x actually computed by any of the processors. This quantity t exists and is recoverable from the algorithm, and it is possible to track the running sum τ =1 x(τ ). Additionally, we state two assumptions encapsulating the conditions underlying our analysis. Assumption A. There is an upper bound m on the delay of any processor. In addition, for each j ∈ [d] there is a constant pj ∈ [0, 1] such that P (ξj = 0) ≤ pj . We also require certain continuity (Lipschitzian) properties of the loss functions; these amount to a second moment constraint on the instantaneous ∂F and a rough measure of gradient sparsity. Assumption B. There exist constants M and (Mj )d such that the following bounds hold for all j=1 2 x ∈ X : E[ ∂x F (x; ξ) 2 ] ≤ M2 and for each j ∈ [d] we have E[|∂xj F (x; ξ)|] ≤ pj Mj . With these definitions, we have the following theorem, which captures the convergence behavior of A SYNC DA under the assumption that X is a Cartesian product, meaning that X = X1 × · · · × Xd , 2 where Xj ⊂ R, and that ψ(x) = 1 x 2 . Note the algorithm itself can still be efficiently parallelized 2 for more general convex X , even if the theorem does not apply. Theorem 3. Let Assumptions A and B and the conditions in the preceding paragraph hold. Then T E t=1 F (x(t); ξ t ) − F (x∗ ; ξ t ) ≤ 1 x∗ 2η d 2 2 η 2 p2 Mj . + T M2 + ηT m j 2 j=1 We now provide a few remarks to explain and simplify the result. Under the more stringent condition 2 d 2 that |∂xj F (x; ξ)| ≤ Mj , Assumption A implies E[ ∂x F (x; ξ) 2 ] ≤ j=1 pj Mj . Thus, for the d 2 remainder of this section we take M2 = j=1 pj Mj , which upper bounds the Lipschitz continuity constant of the objective function f . We then obtain the following corollary. √ T 1 Corollary 4. Define x(T ) = T t=1 x(t), and set η = x∗ 2 /M T . Then E[f (x(T )) − f (x∗ )] ≤ M x∗ √ T 2 +m x∗ 2 √ 2M T d 2 p2 M j . j j=1 Corollary 4 is nearly immediate: since ξ t is independent of x(t), we have E[F (x(t); ξ t ) | x(t)] = f (x(t)); applying Jensen’s inequality to f (x(T )) and performing an algebraic manipulation give the result. If the data is suitably sparse, meaning that pj ≤ 1/m, the bound in Corollary 4 simplifies to 3 M x∗ √ E[f (x(T )) − f (x )] ≤ 2 T ∗ 2 3 = 2 d j=1 2 pj M j x ∗ √ T 2 , (7) which is the convergence rate of stochastic gradient descent even in centralized settings (5). The √ convergence guarantee (7) shows that after T timesteps, the error scales as 1/ T ; however, if we have k processors, updates occur roughly k times as quickly, as they are asynchronous, and in time scaling as N/k, we can evaluate N gradient samples: a linear speedup. 3.2 Asynchronous AdaGrad We now turn to extending A DAG RAD to asynchronous settings, developing A SYNC A DAG RAD (asynchronous A DAG RAD). As in the A SYNC DA algorithm, A SYNC A DAG RAD maintains a shared dual vector z (the sum of gradients) and the shared matrix S, which is the diagonal sum of squares of gradient entries (recall Section 2.2). The matrix S is initialized as diag(δ 2 ), where δj ≥ 0 is an initial value. Each processor asynchronously performs the following iterations: 1 1 1. Read S and z and set G = S 2 . Compute x := argminx∈X { z, x + 2η x, Gx } increment “time” counter t and let x(t) = x, S(t) = S 2. Sample ξ ∼ P and let g ∈ ∂F (x; ξ) 2 3. For j ∈ [d] such that gj = 0, update Sj ← Sj + gj and zj ← zj + gj . 5 // Implicitly As in the description of A SYNC DA, we note that x(t) is the vector x ∈ X computed in the tth “step” of the algorithm (step 1), and similarly associate ξ t with x(t). To analyze A SYNC A DAG RAD, we make a somewhat stronger assumption on the sparsity properties of the losses F than Assumption B. 2 Assumption C. There exist constants (Mj )d such that E[(∂xj F (x; ξ))2 | ξj = 0] ≤ Mj for all j=1 x ∈ X. 2 Indeed, taking M2 = j pj Mj shows that Assumption C implies Assumption B with specific constants. We then have the following convergence result. Theorem 5. In addition to the conditions of Theorem 3, let Assumption C hold. Assume that for all 2 j we have δ 2 ≥ Mj m and X ⊂ [−R∞ , R∞ ]d . Then T t=1 E F (x(t); ξ t ) − F (x∗ ; ξ t ) d ≤ min j=1 T 1 2 R E η ∞ 2 δ + gj (t) 2 1 2 T + ηE gj (t) t=1 2 1 2 (1 + pj m), Mj R∞ pj T . t=1 It is possible to relax the condition on the initial constant diagonal term; we defer this to the full version of the paper. It is natural to ask in which situations the bound provided by Theorem 5 is optimal. We note that, as in the case with Theorem 3, we may obtain a convergence rate for f (x(T )) − f (x∗ ) using convexity, T 1 where x(T ) = T t=1 x(t). By Jensen’s inequality, we have for any δ that T E 2 δ + gj (t) 2 1 2 t=1 T ≤ 2 2 E[gj (t) ] δ + t=1 1 2 ≤ 2 δ 2 + T pj Mj . For interpretation, let us now make a few assumptions on the probabilities pj . If we assume that pj ≤ c/m for a universal (numerical) constant c, then Theorem 5 guarantees that d log(T )/T + pj √ (8) , pj , T j=1 √ which is the convergence rate of A DAG RAD except for a small factor of min{ log T /T, pj } in addition to the usual pj /T rate. In particular, optimizing by choosing η = R∞ , and assuming 1 pj T log T , we have convergence guarantee √ d pj E[f (x(T )) − f (x∗ )] ≤ O(1)R∞ Mj min √ , pj , T j=1 E[f (x(T )) − f (x∗ )] ≤ O(1) 1 2 R +η η ∞ Mj min which is minimax optimal by Proposition 1. In fact, however, the bounds of Theorem 5 are somewhat stronger: they provide bounds using the expectation of the squared gradients gj (t) rather than the maximal value Mj , though the bounds are perhaps clearer in the form (8). We note also that our analysis applies to more adversarial settings than stochastic optimization (e.g., to online convex optimization [5]). Specifically, an adversary may choose an arbitrary sequence of functions subject to the random data sparsity constraint (2), and our results provide an expected regret bound, which is strictly stronger than the stochastic convergence guarantees provided (and guarantees high-probability convergence in stochastic settings [3]). Moreover, our comments in Section 2 about the relative optimality of A DAG RAD versus standard gradient methods apply. When the data is sparse, we indeed should use asynchronous algorithms, but using adaptive methods yields even more improvement than simple gradient-based methods. 4 Experiments In this section, we give experimental validation of our theoretical results on A SYNC A DAG RAD and A SYNC DA, giving results on two datasets selected for their high-dimensional sparsity.2 2 In our experiments, A SYNC DA and H OGWILD ! had effectively identical performance. 6 8 0.07 6 5 4 0.024 Test error Training loss Speedup 0.025 0.065 7 0.06 0.055 0.05 0.045 0.04 0.023 0.022 0.021 0.02 0.035 3 0.019 2 1 2 4 0.03 A-A DAG RAD A SYNC DA Number of workers 6 8 10 12 14 0.018 0.025 0.02 16 2 4 6 8 10 12 14 Number of workers 0.017 16 2 4 6 8 10 12 14 Number of workers 16 Figure 1. Experiments with URL data. Left: speedup relative to one processor. Middle: training dataset loss versus number of processors. Right: test set error rate versus number of processors. AA DAG RAD abbreviates A SYNC A DAG RAD. 1.03 1.02 1.01 1.00 1.0 1 2 4 8 16 64 256 number of passes A-AdaGrad, η = 0.008 L2 = 0 A-AdaGrad, η = 0.008 L2 = 80 A-DA, η = 0.8 L2 = 0 A-DA, η = 0.8 L2 = 80 1.00 1.01 1.4 1.02 1.03 1.04 Impact of L2 regularizaton on test error 1.04 Fixed stepsizes, test data, L2=0 1.2 relative log-loss 1.6 1.8 Fixed stepsizes, training data, L2=0 A-AdaGrad η = 0.002 A-AdaGrad η = 0.004 A-AdaGrad η = 0.008 A-AdaGrad η = 0.016 A-DA η = 0.800 A-DA η = 1.600 A-DA η = 3.200 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 2: Relative accuracy for various stepsize choices on an click-through rate prediction dataset. 4.1 Malicious URL detection For our first set of experiments, we consider the speedup attainable by applying A SYNC A DAG RAD and A SYNC DA, investigating the performance of each algorithm on a malicious URL prediction task [7]. The dataset in this case consists of an anonymized collection of URLs labeled as malicious (e.g., spam, phishing, etc.) or benign over a span of 120 days. The data in this case consists of 2.4 · 106 examples with dimension d = 3.2 · 106 (sparse) features. We perform several experiments, randomly dividing the dataset into 1.2 · 106 training and test samples for each experiment. In Figure 1 we compare the performance of A SYNC A DAG RAD and A SYNC DA after doing after single pass through the training dataset. (For each algorithm, we choose the stepsize η for optimal training set performance.) We perform the experiments on a single machine running Ubuntu Linux with six cores (with two-way hyperthreading) and 32Gb of RAM. From the left-most plot in Fig. 1, we see that up to six processors, both A SYNC DA and A SYNC A DAG RAD enjoy the expected linear speedup, and from 6 to 12, they continue to enjoy a speedup that is linear in the number of processors though at a lesser slope (this is the effect of hyperthreading). For more than 12 processors, there is no further benefit to parallelism on this machine. The two right plots in Figure 1 plot performance of the different methods (with standard errors) versus the number of worker threads used. Both are essentially flat; increasing the amount of parallelism does nothing to the average training loss or the test error rate for either method. It is clear, however, that for this dataset, the adaptive A SYNC A DAG RAD algorithm provides substantial performance benefits over A SYNC DA. 4.2 Click-through-rate prediction experiments We also experiment on a proprietary datasets consisting of search ad impressions. Each example corresponds to showing a search-engine user a particular text ad in response to a query string. From this, we construct a very sparse feature vector based on the text of the ad displayed and the query string (no user-specific data is used). The target label is 1 if the user clicked the ad and -1 otherwise. 7 (B) A-AdaGrad speedup (D) Impact of training data ordering 1.004 1.005 1.006 1.007 1.008 1 2 4 8 16 32 number of passes 64 128 256 1.000 1 2 A-DA base η = 1.600 A-AdaGrad base η = 0.023 0 1.005 relative stepsize (C) Optimal stepsize scaling relative log-loss 1.003 target relative log-loss 1.005 1.010 1.002 1.010 1.015 8 4 0 speedup A-DA η = 1.600 A-AdaGrad η = 0.016 1.001 1.000 relative log-loss 1.015 A-DA, L2=80 A-AdaGrad, L2=80 12 (A) Optimized stepsize for each number of passes 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 3. (A) Relative test-set log-loss for A SYNC DA and A SYNC A DAG RAD, choosing the best stepsize (within a factor of about 1.4×) individually for each number of passes. (B) Effective speedup for A SYNC A DAG RAD. (C) The best stepsize η, expressed as a scaling factor on the stepsize used for one pass. (D) Five runs with different random seeds for each algorithm (with 2 penalty 80). We fit logistic regression models using both A SYNC DA and A SYNC A DAG RAD. We run extensive experiments on a moderate-sized dataset (about 107 examples, split between training and testing), which allows thorough investigation of the impact of the stepsize η, the number of training passes,3 and 2 -regularization on accuracy. For these experiments we used 32 threads on 16 core machines for each run, as A SYNC A DAG RAD and A SYNC DA achieve similar speedups from parallelization. On this dataset, A SYNC A DAG RAD typically achieves an effective additional speedup over A SYNC DA of 4× or more. That is, to reach a given level of accuracy, A SYNC DA generally needs four times as many effective passes over the dataset. We measure accuracy with log-loss (the logistic loss) averaged over five runs using different random seeds (which control the order in which the algorithms sample examples during training). We report relative values in Figures 2 and 3, that is, the ratio of the mean loss for the given datapoint to the lowest (best) mean loss obtained. Our results are not particularly sensitive to the choice of relative log-loss as the metric of interest; we also considered AUC (the area under the ROC curve) and observed similar results. Figure 2 shows relative log-loss as a function of the number of training passes for various stepsizes. Without regularization, A SYNC A DAG RAD is prone to overfitting: it achieves significantly higher accuracy on the training data (Fig. 2 (left)), but unless the stepsize is tuned carefully to the number of passes, it will overfit (Fig. 2 (middle)). Fortunately, the addition of 2 regularization largely solves this problem. Indeed, Figure 2 (right) shows that while adding an 2 penalty of 80 has very little impact on A SYNC DA, it effectively prevents the overfitting of A SYNC A DAG RAD.4 Fixing √ regularization multiplier to 80, we varied the stepsize η over a multiplicative grid with res2 olution 2 for each number of passes and for each algorithm. Figure 3 reports the results obtained by selecting the best stepsize in terms of test set log-loss for each number of passes. Figure 3(A) shows relative log-loss of the best stepsize for each algorithm; 3(B) shows the relative time A SYNC DA requires with respect to A SYNC A DAG RAD to achieve a given loss. Specifically, Fig. 3(B) shows the ratio of the number of passes the algorithms require to achieve a fixed loss, which gives a broader estimate of the speedup obtained by using A SYNC A DAG RAD; speedups range from 3.6× to 12×. Figure 3(C) shows the optimal stepsizes as a function of the best setting for one pass. The optimal stepsize decreases moderately for A SYNC A DAG RAD, but are somewhat noisy for A SYNC DA. It is interesting to note that A SYNC A DAG RAD’s accuracy is largely independent of the ordering of the training data, while A SYNC DA shows significant variability. This can be seen both in the error bars on Figure 3(A), and explicitly in Figure 3(D), where we plot one line for each of the five random seeds used. Thus, while on the one hand A SYNC DA requires somewhat less tuning of the stepsize and 2 parameter, tuning A SYNC A DAG RAD is much easier because of its predictable response. 3 Here “number of passes” more precisely means the expected number of times each example in the dataset is trained on. That is, each worker thread randomly selects a training example from the dataset for each update, and we continued making updates until (dataset size) × (number of passes) updates have been processed. 4 For both algorithms, this is accomplished by adding the term η80 x 2 to the ψ function. We can achieve 2 slightly better results for A SYNC A DAG RAD by varying the 2 penalty with the number of passes. 8 References [1] P. Auer and C. Gentile. Adaptive and self-confident online learning algorithms. In Proceedings of the Thirteenth Annual Conference on Computational Learning Theory, 2000. [2] P. B¨ hlmann and S. van de Geer. Statistics for High-Dimensional Data: Methods, Theory and u Applications. Springer, 2011. [3] N. Cesa-Bianchi, A. Conconi, and C. Gentile. On the generalization ability of on-line learning algorithms. IEEE Transactions on Information Theory, 50(9):2050–2057, September 2004. [4] J. C. Duchi, E. Hazan, and Y. Singer. Adaptive subgradient methods for online learning and stochastic optimization. Journal of Machine Learning Research, 12:2121–2159, 2011. [5] E. Hazan. The convex optimization approach to regret minimization. In Optimization for Machine Learning, chapter 10. MIT Press, 2012. [6] J. Hiriart-Urruty and C. Lemar´ chal. Convex Analysis and Minimization Algorithms I & II. e Springer, New York, 1996. [7] J. Ma, L. K. Saul, S. Savage, and G. M. Voelker. Identifying malicious urls: An application of large-scale online learning. In Proceedings of the 26th International Conference on Machine Learning, 2009. [8] C. Manning and H. Sch¨ tze. Foundations of Statistical Natural Language Processing. MIT u Press, 1999. [9] B. McMahan and M. Streeter. Adaptive bound optimization for online convex optimization. In Proceedings of the Twenty Third Annual Conference on Computational Learning Theory, 2010. [10] A. Nemirovski, A. Juditsky, G. Lan, and A. Shapiro. Robust stochastic approximation approach to stochastic programming. SIAM Journal on Optimization, 19(4):1574–1609, 2009. [11] Y. Nesterov. Primal-dual subgradient methods for convex problems. Mathematical Programming, 120(1):261–283, 2009. [12] F. Niu, B. Recht, C. R´ , and S. Wright. Hogwild: a lock-free approach to parallelizing stochase tic gradient descent. In Advances in Neural Information Processing Systems 24, 2011. [13] P. Richt´ rik and M. Tak´ c. Parallel coordinate descent methods for big data optimization. a aˇ arXiv:1212.0873 [math.OC], 2012. URL http://arxiv.org/abs/1212.0873. [14] M. Tak´ c, A. Bijral, P. Richt´ rik, and N. Srebro. Mini-batch primal and dual methods for aˇ a SVMs. In Proceedings of the 30th International Conference on Machine Learning, 2013. 9
Reference: text
sentIndex sentText sentNum sentScore
1 com Abstract We study stochastic optimization problems when the data is sparse, which is in a sense dual to current perspectives on high-dimensional statistical learning and optimization. [sent-8, score-0.127]
2 We highlight both the difficulties—in terms of increased sample complexity that sparse data necessitates—and the potential benefits, in terms of allowing parallelism and asynchrony in the design of algorithms. [sent-9, score-0.096]
3 Concretely, we derive matching upper and lower bounds on the minimax rate for optimization and learning with sparse data, and we exhibit algorithms achieving these rates. [sent-10, score-0.155]
4 We also show how leveraging sparsity leads to (still minimax optimal) parallel and asynchronous algorithms, providing experimental evidence complementing our theoretical results on several medium to large-scale learning tasks. [sent-11, score-0.28]
5 1 Introduction and problem setting In this paper, we investigate stochastic optimization problems in which the data is sparse. [sent-12, score-0.062]
6 Formally, let {F (·; ξ), ξ ∈ Ξ} be a collection of real-valued convex functions, each of whose domains contains the convex set X ⊂ Rd . [sent-13, score-0.068]
7 (1) Ξ By data sparsity, we mean the samples ξ are sparse: assuming that samples ξ lie in Rd , and defining the support supp(x) of a vector x to the set of indices of its non-zero components, we assume supp F (x; ξ) ⊂ supp ξ. [sent-15, score-0.103]
8 (2) The sparsity condition (2) means that F (x; ξ) does not “depend” on the values of xj for indices j such that ξj = 0. [sent-16, score-0.122]
9 1 This type of data sparsity is prevalent in statistical optimization problems and machine learning applications; in spite of its prevalence, study of such problems has been limited. [sent-17, score-0.122]
10 Such generalized linear models satisfy the sparsity condition (2), and while instances are of very high dimension, in any given instance, very few entries of ξ are non-zero [8]. [sent-19, score-0.068]
11 In a sense, this is dual to the standard approaches to high-dimensional problems; one usually assumes that the data ξ may be dense, but there are only a few relevant features, and thus a parsimonious model x is desirous [2]. [sent-21, score-0.065]
12 1 while such sparse data problems are prevalent—natural language processing, information retrieval, and other large data settings all have significant data sparsity—they do not appear to have attracted as much study as their high-dimensional “duals” of dense data and sparse predictors. [sent-24, score-0.075]
13 Recent work in the optimization and machine learning communities has shown that data sparsity can be leveraged to develop parallel (and even asynchronous [12]) optimization algorithms [13, 14], but this work does not consider the statistical effects of data sparsity. [sent-26, score-0.305]
14 [4] and McMahan and Streeter [9] develop “adaptive” stochastic gradient algorithms to address problems in sparse data regimes (2). [sent-28, score-0.1]
15 First, we investigate the fundamental limits of optimization and learning algorithms in sparse data regimes. [sent-31, score-0.06]
16 In doing so, we derive lower bounds on the optimization error of any algorithm for problems of the form (1) with sparsity condition (2). [sent-32, score-0.118]
17 In spite of this seemingly negative result, we are also able to show that the A DAG RAD algorithms of [4, 9] are optimal, and we show examples in which their dependence on the dimension d can be made exponentially better than standard gradient methods. [sent-35, score-0.059]
18 As the second facet of our two-pronged approach, we study how sparsity may be leveraged in parallel computing frameworks to give substantially faster algorithms that still achieve optimal sample complexity in terms of the number of samples ξ used. [sent-36, score-0.12]
19 We develop two new algorithms, asynchronous dual averaging (A SYNC DA) and asynchronous A DAG RAD (A SYNC A DAG RAD), which allow asynchronous parallel solution of the problem (1) for general convex f and X . [sent-37, score-0.581]
20 [12] with a new analysis, we prove our algorithms achieve linear speedup in the number of processors while maintaining optimal statistical guarantees. [sent-40, score-0.18]
21 2 Minimax rates for sparse optimization We begin our study of sparse optimization problems by establishing their fundamental statistical and optimization-theoretic properties. [sent-42, score-0.12]
22 To do this, we derive bounds on the minimax convergence rate of any algorithm for such problems. [sent-43, score-0.116]
23 We let P denote a collection of probability distributions, and we consider a collection of loss functions F specified by a collection F of convex losses F : X × ξ → R. [sent-54, score-0.12]
24 1 Minimax lower bounds Let us now give a more precise characterization of the (natural) set of sparse optimization problems we consider to provide the lower bound. [sent-57, score-0.08]
25 For the next proposition, we let P consist of distributions supported on Ξ = {−1, 0, 1}d , and we let pj := P (ξj = 0) be the marginal probability of appearance of feature j ∈ {1, . [sent-58, score-0.222]
26 For our class of functions, we set F to consist of functions F satisfying the sparsity condition (2) and with the additional constraint that for g ∈ ∂x F (x; ξ), we have that the jth coordinate |gj | ≤ Mj for a constant Mj < ∞. [sent-62, score-0.092]
27 Mj min pj , √ N (X , P, F) ≥ R 8 j=1 N log 3 We provide the proof of Proposition 1 in the supplement A. [sent-67, score-0.222]
28 Let the conditions of Proposition 1 hold with Mj ≡ M for all j, and assume the power law condition P (ξj = 0) = p0 j −α on coordinate appearance probabilities. [sent-72, score-0.08]
29 2 Algorithms for attaining the minimax rate To show that the lower bounds of Proposition 1 and its subsequent specializations are sharp, we review a few stochastic gradient algorithms. [sent-80, score-0.165]
30 We begin with stochastic gradient descent (SGD): SGD repeatedly samples ξ ∼ P , computes g ∈ ∂x F (x; ξ), then performs the update x ← ΠX (x − ηg), where η is a stepsize parameter and ΠX denotes Euclidean projection onto X . [sent-81, score-0.294]
31 Standard analyses of stochastic gradient descent [10] show that after N samples ξ i , the SGD estimator x(N ) satisfies R2 M ( d j=1 1 pj ) 2 √ , (5) N where R2 denotes the 2 -radius of X . [sent-82, score-0.309]
32 As we discuss presently, the dual averaging algorithm is somewhat more natural in asynchronous and parallel computing environments, and it enjoys the same type of convergence guarantees (5) as SGD. [sent-85, score-0.326]
33 The A DAG RAD algorithm [4, 9] is an extension of the preceding stochastic gradient methods. [sent-86, score-0.092]
34 It maintains a diagonal matrix S, where upon receiving a new sample ξ, A DAG RAD performs the following: it computes g ∈ ∂x F (x; ξ), then updates 2 Sj ← Sj + gj for j ∈ [d]. [sent-87, score-0.198]
35 The dual averaging variant of A DAG RAD updates the usual dual vector z ← z + g; the update to x is based on S and a stepsize η and computes x ← argmin z, x + x ∈X 3 1 1 x ,S2x 2η . [sent-88, score-0.401]
36 After N samples ξ, the averaged parameter x(N ) returned by A DAG RAD satisfies R∞ M E[f (x(N ))] − inf f (x) ≤ O(1) √ x∈X N d √ pj , (6) j=1 where R∞ denotes the ∞ -radius of X (cf. [sent-89, score-0.267]
37 It is interesting to note, though, that in the power law setting of Corollary 2 (recall the error order (4)), a calculation √ shows that the multiplier for the SGD guarantee (5) becomes R∞ d max{d(1−α)/2 , 1}, while A DA G RAD attains rate at worst R∞ max{d1−α/2 , log d}. [sent-93, score-0.059]
38 3 Parallel and asynchronous optimization with sparsity As we note in the introduction, recent works [12, 14] have suggested that sparsity can yield benefits in our ability to parallelize stochastic gradient-type algorithms. [sent-95, score-0.306]
39 Given the optimality of A DAG RADtype algorithms, it is natural to focus on their parallelization in the hope that we can leverage their ability to “adapt” to sparsity in the data. [sent-96, score-0.066]
40 is an asynchronous (parallelized) stochastic gradient algorithm for optimization over product-space domains, meaning that X in problem (1) decomposes as X = X1 × · · · × Xd , where Xj ⊂ R. [sent-101, score-0.242]
41 A pool of independently running processors then performs the following updates asynchronously to a centralized vector x: 1. [sent-103, score-0.176]
42 is that in step 2, the parameter x is allowed to be inconsistent—it may have received partial gradient updates from many processors—and for appropriate problems, this inconsistency is negligible. [sent-111, score-0.07]
43 [12] show linear speedup in optimization time as the number of processors grow; they show this empirically in many scenarios, providing a proof under the somewhat restrictive assumptions that there is at most one non-zero entry in any gradient g and that f has Lipschitz gradients. [sent-113, score-0.281]
44 1 Asynchronous dual averaging A weakness of H OGWILD ! [sent-115, score-0.097]
45 In effort to alleviate these difficulties, we now develop and present our asynchronous dual averaging algorithm, A SYNC DA. [sent-117, score-0.239]
46 A SYNC DA maintains and upates a centralized dual vector z instead of a parameter x, and a pool of processors perform asynchronous updates to z, where each processor independently iterates: 1. [sent-118, score-0.405]
47 For j ∈ [d] such that gj = 0, update zj ← zj + gj . [sent-122, score-0.32]
48 The only communication point between any of the processors is the addition operation in step 3. [sent-124, score-0.093]
49 We let x(t) denote the vector x ∈ X computed in the tth step 1 of the A SYNC DA 4 algorithm, that is, whichever is the tth x actually computed by any of the processors. [sent-128, score-0.058]
50 In addition, for each j ∈ [d] there is a constant pj ∈ [0, 1] such that P (ξj = 0) ≤ pj . [sent-133, score-0.444]
51 We also require certain continuity (Lipschitzian) properties of the loss functions; these amount to a second moment constraint on the instantaneous ∂F and a rough measure of gradient sparsity. [sent-134, score-0.076]
52 There exist constants M and (Mj )d such that the following bounds hold for all j=1 2 x ∈ X : E[ ∂x F (x; ξ) 2 ] ≤ M2 and for each j ∈ [d] we have E[|∂xj F (x; ξ)|] ≤ pj Mj . [sent-136, score-0.242]
53 Let Assumptions A and B and the conditions in the preceding paragraph hold. [sent-140, score-0.062]
54 Under the more stringent condition 2 d 2 that |∂xj F (x; ξ)| ≤ Mj , Assumption A implies E[ ∂x F (x; ξ) 2 ] ≤ j=1 pj Mj . [sent-143, score-0.239]
55 Thus, for the d 2 remainder of this section we take M2 = j=1 pj Mj , which upper bounds the Lipschitz continuity constant of the objective function f . [sent-144, score-0.26]
56 If the data is suitably sparse, meaning that pj ≤ 1/m, the bound in Corollary 4 simplifies to 3 M x∗ √ E[f (x(T )) − f (x )] ≤ 2 T ∗ 2 3 = 2 d j=1 2 pj M j x ∗ √ T 2 , (7) which is the convergence rate of stochastic gradient descent even in centralized settings (5). [sent-150, score-0.585]
57 The √ convergence guarantee (7) shows that after T timesteps, the error scales as 1/ T ; however, if we have k processors, updates occur roughly k times as quickly, as they are asynchronous, and in time scaling as N/k, we can evaluate N gradient samples: a linear speedup. [sent-151, score-0.091]
58 2 Asynchronous AdaGrad We now turn to extending A DAG RAD to asynchronous settings, developing A SYNC A DAG RAD (asynchronous A DAG RAD). [sent-153, score-0.142]
59 As in the A SYNC DA algorithm, A SYNC A DAG RAD maintains a shared dual vector z (the sum of gradients) and the shared matrix S, which is the diagonal sum of squares of gradient entries (recall Section 2. [sent-154, score-0.121]
60 For j ∈ [d] such that gj = 0, update Sj ← Sj + gj and zj ← zj + gj . [sent-161, score-0.451]
61 To analyze A SYNC A DAG RAD, we make a somewhat stronger assumption on the sparsity properties of the losses F than Assumption B. [sent-163, score-0.123]
62 2 Indeed, taking M2 = j pj Mj shows that Assumption C implies Assumption B with specific constants. [sent-166, score-0.222]
63 Then T t=1 E F (x(t); ξ t ) − F (x∗ ; ξ t ) d ≤ min j=1 T 1 2 R E η ∞ 2 δ + gj (t) 2 1 2 T + ηE gj (t) t=1 2 1 2 (1 + pj m), Mj R∞ pj T . [sent-171, score-0.706]
64 By Jensen’s inequality, we have for any δ that T E 2 δ + gj (t) 2 1 2 t=1 T ≤ 2 2 E[gj (t) ] δ + t=1 1 2 ≤ 2 δ 2 + T pj Mj . [sent-175, score-0.353]
65 For interpretation, let us now make a few assumptions on the probabilities pj . [sent-176, score-0.222]
66 If we assume that pj ≤ c/m for a universal (numerical) constant c, then Theorem 5 guarantees that d log(T )/T + pj √ (8) , pj , T j=1 √ which is the convergence rate of A DAG RAD except for a small factor of min{ log T /T, pj } in addition to the usual pj /T rate. [sent-177, score-1.152]
67 In particular, optimizing by choosing η = R∞ , and assuming 1 pj T log T , we have convergence guarantee √ d pj E[f (x(T )) − f (x∗ )] ≤ O(1)R∞ Mj min √ , pj , T j=1 E[f (x(T )) − f (x∗ )] ≤ O(1) 1 2 R +η η ∞ Mj min which is minimax optimal by Proposition 1. [sent-178, score-0.741]
68 In fact, however, the bounds of Theorem 5 are somewhat stronger: they provide bounds using the expectation of the squared gradients gj (t) rather than the maximal value Mj , though the bounds are perhaps clearer in the form (8). [sent-179, score-0.224]
69 We note also that our analysis applies to more adversarial settings than stochastic optimization (e. [sent-180, score-0.062]
70 Moreover, our comments in Section 2 about the relative optimality of A DAG RAD versus standard gradient methods apply. [sent-184, score-0.106]
71 When the data is sparse, we indeed should use asynchronous algorithms, but using adaptive methods yields even more improvement than simple gradient-based methods. [sent-185, score-0.165]
72 0 1 2 4 8 16 64 256 number of passes A-AdaGrad, η = 0. [sent-220, score-0.106]
73 200 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 2: Relative accuracy for various stepsize choices on an click-through rate prediction dataset. [sent-241, score-0.423]
74 1 Malicious URL detection For our first set of experiments, we consider the speedup attainable by applying A SYNC A DAG RAD and A SYNC DA, investigating the performance of each algorithm on a malicious URL prediction task [7]. [sent-243, score-0.138]
75 The dataset in this case consists of an anonymized collection of URLs labeled as malicious (e. [sent-244, score-0.086]
76 (For each algorithm, we choose the stepsize η for optimal training set performance. [sent-254, score-0.217]
77 1, we see that up to six processors, both A SYNC DA and A SYNC A DAG RAD enjoy the expected linear speedup, and from 6 to 12, they continue to enjoy a speedup that is linear in the number of processors though at a lesser slope (this is the effect of hyperthreading). [sent-257, score-0.214]
78 The two right plots in Figure 1 plot performance of the different methods (with standard errors) versus the number of worker threads used. [sent-259, score-0.059]
79 Both are essentially flat; increasing the amount of parallelism does nothing to the average training loss or the test error rate for either method. [sent-260, score-0.106]
80 From this, we construct a very sparse feature vector based on the text of the ad displayed and the query string (no user-specific data is used). [sent-265, score-0.062]
81 7 (B) A-AdaGrad speedup (D) Impact of training data ordering 1. [sent-267, score-0.114]
82 008 1 2 4 8 16 32 number of passes 64 128 256 1. [sent-272, score-0.106]
83 005 relative stepsize (C) Optimal stepsize scaling relative log-loss 1. [sent-276, score-0.452]
84 015 A-DA, L2=80 A-AdaGrad, L2=80 12 (A) Optimized stepsize for each number of passes 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 3. [sent-287, score-0.508]
85 (A) Relative test-set log-loss for A SYNC DA and A SYNC A DAG RAD, choosing the best stepsize (within a factor of about 1. [sent-288, score-0.19]
86 (C) The best stepsize η, expressed as a scaling factor on the stepsize used for one pass. [sent-291, score-0.38]
87 We run extensive experiments on a moderate-sized dataset (about 107 examples, split between training and testing), which allows thorough investigation of the impact of the stepsize η, the number of training passes,3 and 2 -regularization on accuracy. [sent-294, score-0.283]
88 On this dataset, A SYNC A DAG RAD typically achieves an effective additional speedup over A SYNC DA of 4× or more. [sent-296, score-0.087]
89 That is, to reach a given level of accuracy, A SYNC DA generally needs four times as many effective passes over the dataset. [sent-297, score-0.106]
90 We report relative values in Figures 2 and 3, that is, the ratio of the mean loss for the given datapoint to the lowest (best) mean loss obtained. [sent-299, score-0.076]
91 Figure 2 shows relative log-loss as a function of the number of training passes for various stepsizes. [sent-301, score-0.169]
92 2 (left)), but unless the stepsize is tuned carefully to the number of passes, it will overfit (Fig. [sent-303, score-0.19]
93 4 Fixing √ regularization multiplier to 80, we varied the stepsize η over a multiplicative grid with res2 olution 2 for each number of passes and for each algorithm. [sent-307, score-0.313]
94 Figure 3 reports the results obtained by selecting the best stepsize in terms of test set log-loss for each number of passes. [sent-308, score-0.19]
95 Figure 3(A) shows relative log-loss of the best stepsize for each algorithm; 3(B) shows the relative time A SYNC DA requires with respect to A SYNC A DAG RAD to achieve a given loss. [sent-309, score-0.262]
96 3(B) shows the ratio of the number of passes the algorithms require to achieve a fixed loss, which gives a broader estimate of the speedup obtained by using A SYNC A DAG RAD; speedups range from 3. [sent-311, score-0.211]
97 The optimal stepsize decreases moderately for A SYNC A DAG RAD, but are somewhat noisy for A SYNC DA. [sent-314, score-0.223]
98 Thus, while on the one hand A SYNC DA requires somewhat less tuning of the stepsize and 2 parameter, tuning A SYNC A DAG RAD is much easier because of its predictable response. [sent-317, score-0.223]
99 That is, each worker thread randomly selects a training example from the dataset for each update, and we continued making updates until (dataset size) × (number of passes) updates have been processed. [sent-319, score-0.13]
100 Mini-batch primal and dual methods for aˇ a SVMs. [sent-407, score-0.065]
wordName wordTfidf (topN-words)
[('sync', 0.628), ('rad', 0.389), ('dag', 0.33), ('da', 0.237), ('pj', 0.222), ('stepsize', 0.19), ('mj', 0.173), ('asynchronous', 0.142), ('gj', 0.131), ('passes', 0.106), ('ogwild', 0.095), ('processors', 0.093), ('speedup', 0.087), ('dual', 0.065), ('minimax', 0.054), ('sparsity', 0.051), ('malicious', 0.051), ('stepsizes', 0.047), ('niu', 0.041), ('xj', 0.039), ('parallelism', 0.038), ('gradient', 0.038), ('mcmahan', 0.036), ('seeds', 0.036), ('relative', 0.036), ('proposition', 0.036), ('workers', 0.034), ('somewhat', 0.033), ('parallel', 0.033), ('url', 0.033), ('ad', 0.032), ('sgd', 0.032), ('averaging', 0.032), ('updates', 0.032), ('stochastic', 0.032), ('hyperthreading', 0.032), ('sparse', 0.03), ('optimization', 0.03), ('centralized', 0.029), ('mr', 0.029), ('tth', 0.029), ('zj', 0.029), ('inf', 0.028), ('corollary', 0.028), ('asynchrony', 0.028), ('supp', 0.027), ('training', 0.027), ('processor', 0.026), ('sj', 0.026), ('urls', 0.026), ('tak', 0.026), ('convex', 0.025), ('coordinate', 0.024), ('richt', 0.024), ('presently', 0.024), ('rik', 0.023), ('read', 0.023), ('adaptive', 0.023), ('preceding', 0.022), ('ep', 0.022), ('paragraph', 0.022), ('asynchronously', 0.022), ('worker', 0.022), ('impact', 0.022), ('convergence', 0.021), ('rate', 0.021), ('losses', 0.021), ('spite', 0.021), ('argminx', 0.021), ('law', 0.021), ('increment', 0.02), ('bounds', 0.02), ('loss', 0.02), ('prevalent', 0.02), ('threads', 0.02), ('counter', 0.02), ('leveraged', 0.019), ('maintains', 0.018), ('collection', 0.018), ('continuity', 0.018), ('speedups', 0.018), ('assumption', 0.018), ('conditions', 0.018), ('dataset', 0.017), ('samples', 0.017), ('multiplier', 0.017), ('enjoy', 0.017), ('condition', 0.017), ('computes', 0.017), ('versus', 0.017), ('remarks', 0.017), ('jensen', 0.017), ('ts', 0.017), ('implicitly', 0.017), ('fixed', 0.016), ('xd', 0.016), ('parallelized', 0.016), ('dense', 0.015), ('optimality', 0.015), ('indices', 0.015)]
simIndex simValue paperId paperTitle
same-paper 1 1.0000002 111 nips-2013-Estimation, Optimization, and Parallelism when Data is Sparse
Author: John Duchi, Michael Jordan, Brendan McMahan
Abstract: We study stochastic optimization problems when the data is sparse, which is in a sense dual to current perspectives on high-dimensional statistical learning and optimization. We highlight both the difficulties—in terms of increased sample complexity that sparse data necessitates—and the potential benefits, in terms of allowing parallelism and asynchrony in the design of algorithms. Concretely, we derive matching upper and lower bounds on the minimax rate for optimization and learning with sparse data, and we exhibit algorithms achieving these rates. We also show how leveraging sparsity leads to (still minimax optimal) parallel and asynchronous algorithms, providing experimental evidence complementing our theoretical results on several medium to large-scale learning tasks. 1 Introduction and problem setting In this paper, we investigate stochastic optimization problems in which the data is sparse. Formally, let {F (·; ξ), ξ ∈ Ξ} be a collection of real-valued convex functions, each of whose domains contains the convex set X ⊂ Rd . For a probability distribution P on Ξ, we consider the following optimization problem: minimize f (x) := E[F (x; ξ)] = x∈X F (x; ξ)dP (ξ). (1) Ξ By data sparsity, we mean the samples ξ are sparse: assuming that samples ξ lie in Rd , and defining the support supp(x) of a vector x to the set of indices of its non-zero components, we assume supp F (x; ξ) ⊂ supp ξ. (2) The sparsity condition (2) means that F (x; ξ) does not “depend” on the values of xj for indices j such that ξj = 0.1 This type of data sparsity is prevalent in statistical optimization problems and machine learning applications; in spite of its prevalence, study of such problems has been limited. As a motivating example, consider a text classification problem: data ξ ∈ Rd represents words appearing in a document, and we wish to minimize a logistic loss F (x; ξ) = log(1 + exp( ξ, x )) on the data (we encode the label implicitly with the sign of ξ). Such generalized linear models satisfy the sparsity condition (2), and while instances are of very high dimension, in any given instance, very few entries of ξ are non-zero [8]. From a modelling perspective, it thus makes sense to allow a dense predictor x: any non-zero entry of ξ is potentially relevant and important. In a sense, this is dual to the standard approaches to high-dimensional problems; one usually assumes that the data ξ may be dense, but there are only a few relevant features, and thus a parsimonious model x is desirous [2]. So 1 Formally, if πξ denotes the coordinate projection zeroing all indices j of its argument where ξj = 0, then F (πξ (x); ξ) = F (x; ξ) for all x, ξ. This follows from the first-order conditions for convexity [6]. 1 while such sparse data problems are prevalent—natural language processing, information retrieval, and other large data settings all have significant data sparsity—they do not appear to have attracted as much study as their high-dimensional “duals” of dense data and sparse predictors. In this paper, we investigate algorithms and their inherent limitations for solving problem (1) under natural conditions on the data generating distribution. Recent work in the optimization and machine learning communities has shown that data sparsity can be leveraged to develop parallel (and even asynchronous [12]) optimization algorithms [13, 14], but this work does not consider the statistical effects of data sparsity. In another line of research, Duchi et al. [4] and McMahan and Streeter [9] develop “adaptive” stochastic gradient algorithms to address problems in sparse data regimes (2). These algorithms exhibit excellent practical performance and have theoretical guarantees on their convergence, but it is not clear if they are optimal—in that no algorithm can attain better statistical performance—or whether they can leverage parallel computing as in the papers [12, 14]. In this paper, we take a two-pronged approach. First, we investigate the fundamental limits of optimization and learning algorithms in sparse data regimes. In doing so, we derive lower bounds on the optimization error of any algorithm for problems of the form (1) with sparsity condition (2). These results have two main implications. They show that in some scenarios, learning with sparse data is quite difficult, as essentially each coordinate j ∈ [d] can be relevant and must be optimized for. In spite of this seemingly negative result, we are also able to show that the A DAG RAD algorithms of [4, 9] are optimal, and we show examples in which their dependence on the dimension d can be made exponentially better than standard gradient methods. As the second facet of our two-pronged approach, we study how sparsity may be leveraged in parallel computing frameworks to give substantially faster algorithms that still achieve optimal sample complexity in terms of the number of samples ξ used. We develop two new algorithms, asynchronous dual averaging (A SYNC DA) and asynchronous A DAG RAD (A SYNC A DAG RAD), which allow asynchronous parallel solution of the problem (1) for general convex f and X . Combining insights of Niu et al.’s H OGWILD ! [12] with a new analysis, we prove our algorithms achieve linear speedup in the number of processors while maintaining optimal statistical guarantees. We also give experiments on text-classification and web-advertising tasks to illustrate the benefits of the new algorithms. 2 Minimax rates for sparse optimization We begin our study of sparse optimization problems by establishing their fundamental statistical and optimization-theoretic properties. To do this, we derive bounds on the minimax convergence rate of any algorithm for such problems. Formally, let x denote any estimator for a minimizer of the objective (1). We define the optimality gap N for the estimator x based on N samples ξ 1 , . . . , ξ N from the distribution P as N (x, F, X , P ) := f (x) − inf f (x) = EP [F (x; ξ)] − inf EP [F (x; ξ)] . x∈X x∈X This quantity is a random variable, since x is a random variable (it is a function of ξ 1 , . . . , ξ N ). To define the minimax error, we thus take expectations of the quantity N , though we require a bit more than simply E[ N ]. We let P denote a collection of probability distributions, and we consider a collection of loss functions F specified by a collection F of convex losses F : X × ξ → R. We can then define the minimax error for the family of losses F and distributions P as ∗ N (X , P, F) := inf sup sup EP [ x P ∈P F ∈F N (x(ξ 1:N ), F, X , P )], (3) where the infimum is taken over all possible estimators x (an estimator is an optimization scheme, or a measurable mapping x : ΞN → X ) . 2.1 Minimax lower bounds Let us now give a more precise characterization of the (natural) set of sparse optimization problems we consider to provide the lower bound. For the next proposition, we let P consist of distributions supported on Ξ = {−1, 0, 1}d , and we let pj := P (ξj = 0) be the marginal probability of appearance of feature j ∈ {1, . . . , d}. For our class of functions, we set F to consist of functions F satisfying the sparsity condition (2) and with the additional constraint that for g ∈ ∂x F (x; ξ), we have that the jth coordinate |gj | ≤ Mj for a constant Mj < ∞. We obtain 2 Proposition 1. Let the conditions of the preceding paragraph hold. Let R be a constant such that X ⊃ [−R, R]d . Then √ d pj 1 ∗ . Mj min pj , √ N (X , P, F) ≥ R 8 j=1 N log 3 We provide the proof of Proposition 1 in the supplement A.1 in the full version of the paper, providing a few remarks here. We begin by giving a corollary to Proposition 1 that follows when the data ξ obeys a type of power law: let p0 ∈ [0, 1], and assume that P (ξj = 0) = p0 j −α . We have Corollary 2. Let α ≥ 0. Let the conditions of Proposition 1 hold with Mj ≡ M for all j, and assume the power law condition P (ξj = 0) = p0 j −α on coordinate appearance probabilities. Then (1) If d > (p0 N )1/α , ∗ N (X , P, F) ≥ 2−α 1−α p0 p0 (p0 N ) 2α − 1 + d1−α − (p0 N ) α N 1−α 2 MR 8 2−α (2) If d ≤ (p0 N )1/α , ∗ N (X , P, F) ≥ MR 8 p0 N α 1 1 d1− 2 − 1 − α/2 1 − α/2 . . Expanding Corollary 2 slightly, for simplicity assume the number of samples is large enough that d ≤ (p0 N )1/α . Then we find that the lower bound on optimization error is of order p0 1− α p0 p0 d 2 when α < 2, M R log d when α → 2, and M R when α > 2. (4) N N N These results beg the question of tightness: are they improvable? As we see presently, they are not. MR 2.2 Algorithms for attaining the minimax rate To show that the lower bounds of Proposition 1 and its subsequent specializations are sharp, we review a few stochastic gradient algorithms. We begin with stochastic gradient descent (SGD): SGD repeatedly samples ξ ∼ P , computes g ∈ ∂x F (x; ξ), then performs the update x ← ΠX (x − ηg), where η is a stepsize parameter and ΠX denotes Euclidean projection onto X . Standard analyses of stochastic gradient descent [10] show that after N samples ξ i , the SGD estimator x(N ) satisfies R2 M ( d j=1 1 pj ) 2 √ , (5) N where R2 denotes the 2 -radius of X . Dual averaging, due to Nesterov [11] (sometimes called “follow the regularized leader” [5]) is a more recent algorithm. In dual averaging, one again samples g ∈ ∂x F (x; ξ), but instead of updating the parameter vector x one updates a dual vector z by z ← z + g, then computes 1 x ← argmin z, x + ψ(x) , η x∈X E[f (x(N ))] − inf f (x) ≤ O(1) x∈X 2 1 where ψ(x) is a strongly convex function defined over X (often one takes ψ(x) = 2 x 2 ). As we discuss presently, the dual averaging algorithm is somewhat more natural in asynchronous and parallel computing environments, and it enjoys the same type of convergence guarantees (5) as SGD. The A DAG RAD algorithm [4, 9] is an extension of the preceding stochastic gradient methods. It maintains a diagonal matrix S, where upon receiving a new sample ξ, A DAG RAD performs the following: it computes g ∈ ∂x F (x; ξ), then updates 2 Sj ← Sj + gj for j ∈ [d]. The dual averaging variant of A DAG RAD updates the usual dual vector z ← z + g; the update to x is based on S and a stepsize η and computes x ← argmin z, x + x ∈X 3 1 1 x ,S2x 2η . After N samples ξ, the averaged parameter x(N ) returned by A DAG RAD satisfies R∞ M E[f (x(N ))] − inf f (x) ≤ O(1) √ x∈X N d √ pj , (6) j=1 where R∞ denotes the ∞ -radius of X (cf. [4, Section 1.3 and Theorem 5]). By inspection, the A DAG RAD rate (6) matches the lower bound in Proposition 1 and is thus optimal. It is interesting to note, though, that in the power law setting of Corollary 2 (recall the error order (4)), a calculation √ shows that the multiplier for the SGD guarantee (5) becomes R∞ d max{d(1−α)/2 , 1}, while A DA G RAD attains rate at worst R∞ max{d1−α/2 , log d}. For α > 1, the A DAG RAD rate is no worse, √ and for α ≥ 2, is more than d/ log d better—an exponential improvement in the dimension. 3 Parallel and asynchronous optimization with sparsity As we note in the introduction, recent works [12, 14] have suggested that sparsity can yield benefits in our ability to parallelize stochastic gradient-type algorithms. Given the optimality of A DAG RADtype algorithms, it is natural to focus on their parallelization in the hope that we can leverage their ability to “adapt” to sparsity in the data. To provide the setting for our further algorithms, we first revisit Niu et al.’s H OGWILD ! [12]. H OGWILD ! is an asynchronous (parallelized) stochastic gradient algorithm for optimization over product-space domains, meaning that X in problem (1) decomposes as X = X1 × · · · × Xd , where Xj ⊂ R. Fix a stepsize η > 0. A pool of independently running processors then performs the following updates asynchronously to a centralized vector x: 1. Sample ξ ∼ P 2. Read x and compute g ∈ ∂x F (x; ξ) 3. For each j s.t. gj = 0, update xj ← ΠXj (xj − ηgj ). Here ΠXj denotes projection onto the jth coordinate of the domain X . The key of H OGWILD ! is that in step 2, the parameter x is allowed to be inconsistent—it may have received partial gradient updates from many processors—and for appropriate problems, this inconsistency is negligible. Indeed, Niu et al. [12] show linear speedup in optimization time as the number of processors grow; they show this empirically in many scenarios, providing a proof under the somewhat restrictive assumptions that there is at most one non-zero entry in any gradient g and that f has Lipschitz gradients. 3.1 Asynchronous dual averaging A weakness of H OGWILD ! is that it appears only applicable to problems for which the domain X is a product space, and its analysis assumes g 0 = 1 for all gradients g. In effort to alleviate these difficulties, we now develop and present our asynchronous dual averaging algorithm, A SYNC DA. A SYNC DA maintains and upates a centralized dual vector z instead of a parameter x, and a pool of processors perform asynchronous updates to z, where each processor independently iterates: 1. Read z and compute x := argminx∈X 1 z, x + η ψ(x) // Implicitly increment “time” counter t and let x(t) = x 2. Sample ξ ∼ P and let g ∈ ∂x F (x; ξ) // Let g(t) = g. 3. For j ∈ [d] such that gj = 0, update zj ← zj + gj . Because the actual computation of the vector x in A SYNC DA is performed locally on each processor in step 1 of the algorithm, the algorithm can be executed with any proximal function ψ and domain X . The only communication point between any of the processors is the addition operation in step 3. Since addition is commutative and associative, forcing all asynchrony to this point of the algorithm is a natural strategy for avoiding synchronization problems. In our analysis of A SYNC DA, and in our subsequent analysis of the adaptive methods, we require a measurement of time elapsed. With that in mind, we let t denote a time index that exists (roughly) behind-the-scenes. We let x(t) denote the vector x ∈ X computed in the tth step 1 of the A SYNC DA 4 algorithm, that is, whichever is the tth x actually computed by any of the processors. This quantity t exists and is recoverable from the algorithm, and it is possible to track the running sum τ =1 x(τ ). Additionally, we state two assumptions encapsulating the conditions underlying our analysis. Assumption A. There is an upper bound m on the delay of any processor. In addition, for each j ∈ [d] there is a constant pj ∈ [0, 1] such that P (ξj = 0) ≤ pj . We also require certain continuity (Lipschitzian) properties of the loss functions; these amount to a second moment constraint on the instantaneous ∂F and a rough measure of gradient sparsity. Assumption B. There exist constants M and (Mj )d such that the following bounds hold for all j=1 2 x ∈ X : E[ ∂x F (x; ξ) 2 ] ≤ M2 and for each j ∈ [d] we have E[|∂xj F (x; ξ)|] ≤ pj Mj . With these definitions, we have the following theorem, which captures the convergence behavior of A SYNC DA under the assumption that X is a Cartesian product, meaning that X = X1 × · · · × Xd , 2 where Xj ⊂ R, and that ψ(x) = 1 x 2 . Note the algorithm itself can still be efficiently parallelized 2 for more general convex X , even if the theorem does not apply. Theorem 3. Let Assumptions A and B and the conditions in the preceding paragraph hold. Then T E t=1 F (x(t); ξ t ) − F (x∗ ; ξ t ) ≤ 1 x∗ 2η d 2 2 η 2 p2 Mj . + T M2 + ηT m j 2 j=1 We now provide a few remarks to explain and simplify the result. Under the more stringent condition 2 d 2 that |∂xj F (x; ξ)| ≤ Mj , Assumption A implies E[ ∂x F (x; ξ) 2 ] ≤ j=1 pj Mj . Thus, for the d 2 remainder of this section we take M2 = j=1 pj Mj , which upper bounds the Lipschitz continuity constant of the objective function f . We then obtain the following corollary. √ T 1 Corollary 4. Define x(T ) = T t=1 x(t), and set η = x∗ 2 /M T . Then E[f (x(T )) − f (x∗ )] ≤ M x∗ √ T 2 +m x∗ 2 √ 2M T d 2 p2 M j . j j=1 Corollary 4 is nearly immediate: since ξ t is independent of x(t), we have E[F (x(t); ξ t ) | x(t)] = f (x(t)); applying Jensen’s inequality to f (x(T )) and performing an algebraic manipulation give the result. If the data is suitably sparse, meaning that pj ≤ 1/m, the bound in Corollary 4 simplifies to 3 M x∗ √ E[f (x(T )) − f (x )] ≤ 2 T ∗ 2 3 = 2 d j=1 2 pj M j x ∗ √ T 2 , (7) which is the convergence rate of stochastic gradient descent even in centralized settings (5). The √ convergence guarantee (7) shows that after T timesteps, the error scales as 1/ T ; however, if we have k processors, updates occur roughly k times as quickly, as they are asynchronous, and in time scaling as N/k, we can evaluate N gradient samples: a linear speedup. 3.2 Asynchronous AdaGrad We now turn to extending A DAG RAD to asynchronous settings, developing A SYNC A DAG RAD (asynchronous A DAG RAD). As in the A SYNC DA algorithm, A SYNC A DAG RAD maintains a shared dual vector z (the sum of gradients) and the shared matrix S, which is the diagonal sum of squares of gradient entries (recall Section 2.2). The matrix S is initialized as diag(δ 2 ), where δj ≥ 0 is an initial value. Each processor asynchronously performs the following iterations: 1 1 1. Read S and z and set G = S 2 . Compute x := argminx∈X { z, x + 2η x, Gx } increment “time” counter t and let x(t) = x, S(t) = S 2. Sample ξ ∼ P and let g ∈ ∂F (x; ξ) 2 3. For j ∈ [d] such that gj = 0, update Sj ← Sj + gj and zj ← zj + gj . 5 // Implicitly As in the description of A SYNC DA, we note that x(t) is the vector x ∈ X computed in the tth “step” of the algorithm (step 1), and similarly associate ξ t with x(t). To analyze A SYNC A DAG RAD, we make a somewhat stronger assumption on the sparsity properties of the losses F than Assumption B. 2 Assumption C. There exist constants (Mj )d such that E[(∂xj F (x; ξ))2 | ξj = 0] ≤ Mj for all j=1 x ∈ X. 2 Indeed, taking M2 = j pj Mj shows that Assumption C implies Assumption B with specific constants. We then have the following convergence result. Theorem 5. In addition to the conditions of Theorem 3, let Assumption C hold. Assume that for all 2 j we have δ 2 ≥ Mj m and X ⊂ [−R∞ , R∞ ]d . Then T t=1 E F (x(t); ξ t ) − F (x∗ ; ξ t ) d ≤ min j=1 T 1 2 R E η ∞ 2 δ + gj (t) 2 1 2 T + ηE gj (t) t=1 2 1 2 (1 + pj m), Mj R∞ pj T . t=1 It is possible to relax the condition on the initial constant diagonal term; we defer this to the full version of the paper. It is natural to ask in which situations the bound provided by Theorem 5 is optimal. We note that, as in the case with Theorem 3, we may obtain a convergence rate for f (x(T )) − f (x∗ ) using convexity, T 1 where x(T ) = T t=1 x(t). By Jensen’s inequality, we have for any δ that T E 2 δ + gj (t) 2 1 2 t=1 T ≤ 2 2 E[gj (t) ] δ + t=1 1 2 ≤ 2 δ 2 + T pj Mj . For interpretation, let us now make a few assumptions on the probabilities pj . If we assume that pj ≤ c/m for a universal (numerical) constant c, then Theorem 5 guarantees that d log(T )/T + pj √ (8) , pj , T j=1 √ which is the convergence rate of A DAG RAD except for a small factor of min{ log T /T, pj } in addition to the usual pj /T rate. In particular, optimizing by choosing η = R∞ , and assuming 1 pj T log T , we have convergence guarantee √ d pj E[f (x(T )) − f (x∗ )] ≤ O(1)R∞ Mj min √ , pj , T j=1 E[f (x(T )) − f (x∗ )] ≤ O(1) 1 2 R +η η ∞ Mj min which is minimax optimal by Proposition 1. In fact, however, the bounds of Theorem 5 are somewhat stronger: they provide bounds using the expectation of the squared gradients gj (t) rather than the maximal value Mj , though the bounds are perhaps clearer in the form (8). We note also that our analysis applies to more adversarial settings than stochastic optimization (e.g., to online convex optimization [5]). Specifically, an adversary may choose an arbitrary sequence of functions subject to the random data sparsity constraint (2), and our results provide an expected regret bound, which is strictly stronger than the stochastic convergence guarantees provided (and guarantees high-probability convergence in stochastic settings [3]). Moreover, our comments in Section 2 about the relative optimality of A DAG RAD versus standard gradient methods apply. When the data is sparse, we indeed should use asynchronous algorithms, but using adaptive methods yields even more improvement than simple gradient-based methods. 4 Experiments In this section, we give experimental validation of our theoretical results on A SYNC A DAG RAD and A SYNC DA, giving results on two datasets selected for their high-dimensional sparsity.2 2 In our experiments, A SYNC DA and H OGWILD ! had effectively identical performance. 6 8 0.07 6 5 4 0.024 Test error Training loss Speedup 0.025 0.065 7 0.06 0.055 0.05 0.045 0.04 0.023 0.022 0.021 0.02 0.035 3 0.019 2 1 2 4 0.03 A-A DAG RAD A SYNC DA Number of workers 6 8 10 12 14 0.018 0.025 0.02 16 2 4 6 8 10 12 14 Number of workers 0.017 16 2 4 6 8 10 12 14 Number of workers 16 Figure 1. Experiments with URL data. Left: speedup relative to one processor. Middle: training dataset loss versus number of processors. Right: test set error rate versus number of processors. AA DAG RAD abbreviates A SYNC A DAG RAD. 1.03 1.02 1.01 1.00 1.0 1 2 4 8 16 64 256 number of passes A-AdaGrad, η = 0.008 L2 = 0 A-AdaGrad, η = 0.008 L2 = 80 A-DA, η = 0.8 L2 = 0 A-DA, η = 0.8 L2 = 80 1.00 1.01 1.4 1.02 1.03 1.04 Impact of L2 regularizaton on test error 1.04 Fixed stepsizes, test data, L2=0 1.2 relative log-loss 1.6 1.8 Fixed stepsizes, training data, L2=0 A-AdaGrad η = 0.002 A-AdaGrad η = 0.004 A-AdaGrad η = 0.008 A-AdaGrad η = 0.016 A-DA η = 0.800 A-DA η = 1.600 A-DA η = 3.200 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 2: Relative accuracy for various stepsize choices on an click-through rate prediction dataset. 4.1 Malicious URL detection For our first set of experiments, we consider the speedup attainable by applying A SYNC A DAG RAD and A SYNC DA, investigating the performance of each algorithm on a malicious URL prediction task [7]. The dataset in this case consists of an anonymized collection of URLs labeled as malicious (e.g., spam, phishing, etc.) or benign over a span of 120 days. The data in this case consists of 2.4 · 106 examples with dimension d = 3.2 · 106 (sparse) features. We perform several experiments, randomly dividing the dataset into 1.2 · 106 training and test samples for each experiment. In Figure 1 we compare the performance of A SYNC A DAG RAD and A SYNC DA after doing after single pass through the training dataset. (For each algorithm, we choose the stepsize η for optimal training set performance.) We perform the experiments on a single machine running Ubuntu Linux with six cores (with two-way hyperthreading) and 32Gb of RAM. From the left-most plot in Fig. 1, we see that up to six processors, both A SYNC DA and A SYNC A DAG RAD enjoy the expected linear speedup, and from 6 to 12, they continue to enjoy a speedup that is linear in the number of processors though at a lesser slope (this is the effect of hyperthreading). For more than 12 processors, there is no further benefit to parallelism on this machine. The two right plots in Figure 1 plot performance of the different methods (with standard errors) versus the number of worker threads used. Both are essentially flat; increasing the amount of parallelism does nothing to the average training loss or the test error rate for either method. It is clear, however, that for this dataset, the adaptive A SYNC A DAG RAD algorithm provides substantial performance benefits over A SYNC DA. 4.2 Click-through-rate prediction experiments We also experiment on a proprietary datasets consisting of search ad impressions. Each example corresponds to showing a search-engine user a particular text ad in response to a query string. From this, we construct a very sparse feature vector based on the text of the ad displayed and the query string (no user-specific data is used). The target label is 1 if the user clicked the ad and -1 otherwise. 7 (B) A-AdaGrad speedup (D) Impact of training data ordering 1.004 1.005 1.006 1.007 1.008 1 2 4 8 16 32 number of passes 64 128 256 1.000 1 2 A-DA base η = 1.600 A-AdaGrad base η = 0.023 0 1.005 relative stepsize (C) Optimal stepsize scaling relative log-loss 1.003 target relative log-loss 1.005 1.010 1.002 1.010 1.015 8 4 0 speedup A-DA η = 1.600 A-AdaGrad η = 0.016 1.001 1.000 relative log-loss 1.015 A-DA, L2=80 A-AdaGrad, L2=80 12 (A) Optimized stepsize for each number of passes 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 3. (A) Relative test-set log-loss for A SYNC DA and A SYNC A DAG RAD, choosing the best stepsize (within a factor of about 1.4×) individually for each number of passes. (B) Effective speedup for A SYNC A DAG RAD. (C) The best stepsize η, expressed as a scaling factor on the stepsize used for one pass. (D) Five runs with different random seeds for each algorithm (with 2 penalty 80). We fit logistic regression models using both A SYNC DA and A SYNC A DAG RAD. We run extensive experiments on a moderate-sized dataset (about 107 examples, split between training and testing), which allows thorough investigation of the impact of the stepsize η, the number of training passes,3 and 2 -regularization on accuracy. For these experiments we used 32 threads on 16 core machines for each run, as A SYNC A DAG RAD and A SYNC DA achieve similar speedups from parallelization. On this dataset, A SYNC A DAG RAD typically achieves an effective additional speedup over A SYNC DA of 4× or more. That is, to reach a given level of accuracy, A SYNC DA generally needs four times as many effective passes over the dataset. We measure accuracy with log-loss (the logistic loss) averaged over five runs using different random seeds (which control the order in which the algorithms sample examples during training). We report relative values in Figures 2 and 3, that is, the ratio of the mean loss for the given datapoint to the lowest (best) mean loss obtained. Our results are not particularly sensitive to the choice of relative log-loss as the metric of interest; we also considered AUC (the area under the ROC curve) and observed similar results. Figure 2 shows relative log-loss as a function of the number of training passes for various stepsizes. Without regularization, A SYNC A DAG RAD is prone to overfitting: it achieves significantly higher accuracy on the training data (Fig. 2 (left)), but unless the stepsize is tuned carefully to the number of passes, it will overfit (Fig. 2 (middle)). Fortunately, the addition of 2 regularization largely solves this problem. Indeed, Figure 2 (right) shows that while adding an 2 penalty of 80 has very little impact on A SYNC DA, it effectively prevents the overfitting of A SYNC A DAG RAD.4 Fixing √ regularization multiplier to 80, we varied the stepsize η over a multiplicative grid with res2 olution 2 for each number of passes and for each algorithm. Figure 3 reports the results obtained by selecting the best stepsize in terms of test set log-loss for each number of passes. Figure 3(A) shows relative log-loss of the best stepsize for each algorithm; 3(B) shows the relative time A SYNC DA requires with respect to A SYNC A DAG RAD to achieve a given loss. Specifically, Fig. 3(B) shows the ratio of the number of passes the algorithms require to achieve a fixed loss, which gives a broader estimate of the speedup obtained by using A SYNC A DAG RAD; speedups range from 3.6× to 12×. Figure 3(C) shows the optimal stepsizes as a function of the best setting for one pass. The optimal stepsize decreases moderately for A SYNC A DAG RAD, but are somewhat noisy for A SYNC DA. It is interesting to note that A SYNC A DAG RAD’s accuracy is largely independent of the ordering of the training data, while A SYNC DA shows significant variability. This can be seen both in the error bars on Figure 3(A), and explicitly in Figure 3(D), where we plot one line for each of the five random seeds used. Thus, while on the one hand A SYNC DA requires somewhat less tuning of the stepsize and 2 parameter, tuning A SYNC A DAG RAD is much easier because of its predictable response. 3 Here “number of passes” more precisely means the expected number of times each example in the dataset is trained on. That is, each worker thread randomly selects a training example from the dataset for each update, and we continued making updates until (dataset size) × (number of passes) updates have been processed. 4 For both algorithms, this is accomplished by adding the term η80 x 2 to the ψ function. We can achieve 2 slightly better results for A SYNC A DAG RAD by varying the 2 penalty with the number of passes. 8 References [1] P. Auer and C. Gentile. Adaptive and self-confident online learning algorithms. In Proceedings of the Thirteenth Annual Conference on Computational Learning Theory, 2000. [2] P. B¨ hlmann and S. van de Geer. Statistics for High-Dimensional Data: Methods, Theory and u Applications. Springer, 2011. [3] N. Cesa-Bianchi, A. Conconi, and C. Gentile. On the generalization ability of on-line learning algorithms. IEEE Transactions on Information Theory, 50(9):2050–2057, September 2004. [4] J. C. Duchi, E. Hazan, and Y. Singer. Adaptive subgradient methods for online learning and stochastic optimization. Journal of Machine Learning Research, 12:2121–2159, 2011. [5] E. Hazan. The convex optimization approach to regret minimization. In Optimization for Machine Learning, chapter 10. MIT Press, 2012. [6] J. Hiriart-Urruty and C. Lemar´ chal. Convex Analysis and Minimization Algorithms I & II. e Springer, New York, 1996. [7] J. Ma, L. K. Saul, S. Savage, and G. M. Voelker. Identifying malicious urls: An application of large-scale online learning. In Proceedings of the 26th International Conference on Machine Learning, 2009. [8] C. Manning and H. Sch¨ tze. Foundations of Statistical Natural Language Processing. MIT u Press, 1999. [9] B. McMahan and M. Streeter. Adaptive bound optimization for online convex optimization. In Proceedings of the Twenty Third Annual Conference on Computational Learning Theory, 2010. [10] A. Nemirovski, A. Juditsky, G. Lan, and A. Shapiro. Robust stochastic approximation approach to stochastic programming. SIAM Journal on Optimization, 19(4):1574–1609, 2009. [11] Y. Nesterov. Primal-dual subgradient methods for convex problems. Mathematical Programming, 120(1):261–283, 2009. [12] F. Niu, B. Recht, C. R´ , and S. Wright. Hogwild: a lock-free approach to parallelizing stochase tic gradient descent. In Advances in Neural Information Processing Systems 24, 2011. [13] P. Richt´ rik and M. Tak´ c. Parallel coordinate descent methods for big data optimization. a aˇ arXiv:1212.0873 [math.OC], 2012. URL http://arxiv.org/abs/1212.0873. [14] M. Tak´ c, A. Bijral, P. Richt´ rik, and N. Srebro. Mini-batch primal and dual methods for aˇ a SVMs. In Proceedings of the 30th International Conference on Machine Learning, 2013. 9
2 0.088639431 193 nips-2013-Mixed Optimization for Smooth Functions
Author: Mehrdad Mahdavi, Lijun Zhang, Rong Jin
Abstract: It is well known that the optimal convergence rate for stochastic optimization of √ smooth functions is O(1/ T ), which is same as stochastic optimization of Lipschitz continuous convex functions. This is in contrast to optimizing smooth functions using full gradients, which yields a convergence rate of O(1/T 2 ). In this work, we consider a new setup for optimizing smooth functions, termed as Mixed Optimization, which allows to access both a stochastic oracle and a full gradient oracle. Our goal is to significantly improve the convergence rate of stochastic optimization of smooth functions by having an additional small number of accesses to the full gradient oracle. We show that, with an O(ln T ) calls to the full gradient oracle and an O(T ) calls to the stochastic oracle, the proposed mixed optimization algorithm is able to achieve an optimization error of O(1/T ). 1
3 0.083761126 174 nips-2013-Lexical and Hierarchical Topic Regression
Author: Viet-An Nguyen, Jordan Boyd-Graber, Philip Resnik
Abstract: Inspired by a two-level theory from political science that unifies agenda setting and ideological framing, we propose supervised hierarchical latent Dirichlet allocation (S H L DA), which jointly captures documents’ multi-level topic structure and their polar response variables. Our model extends the nested Chinese restaurant processes to discover tree-structured topic hierarchies and uses both per-topic hierarchical and per-word lexical regression parameters to model response variables. S H L DA improves prediction on political affiliation and sentiment tasks in addition to providing insight into how topics under discussion are framed. 1 Introduction: Agenda Setting and Framing in Hierarchical Models How do liberal-leaning bloggers talk about immigration in the US? What do conservative politicians have to say about education? How do Fox News and MSNBC differ in their language about the gun debate? Such questions concern not only what, but how things are talked about. In political communication, the question of “what” falls under the heading of agenda setting theory, which concerns the issues introduced into political discourse (e.g., by the mass media) and their influence over public priorities [1]. The question of “how” concerns framing: the way the presentation of an issue reflects or encourages a particular perspective or interpretation [2]. For example, the rise of the “innocence frame” in the death penalty debate, emphasizing the irreversible consequence of mistaken convictions, has led to a sharp decline in the use of capital punishment in the US [3]. In its concern with the subjects or issues under discussion in political discourse, agenda setting maps neatly to topic modeling [4] as a means of discovering and characterizing those issues [5]. Interestingly, one line of communication theory seeks to unify agenda setting and framing by viewing frames as a second-level kind of agenda [1]: just as agenda setting is about which objects of discussion are salient, framing is about the salience of attributes of those objects. The key is that what communications theorists consider an attribute in a discussion can itself be an object, as well. For example, “mistaken convictions” is one attribute of the death penalty discussion, but it can also be viewed as an object of discussion in its own right. This two-level view leads naturally to the idea of using a hierarchical topic model to formalize both agendas and frames within a uniform setting. In this paper, we introduce a new model to do exactly that. The model is predictive: it represents the idea of alternative or competing perspectives via a continuous-valued response variable. Although inspired by the study of political discourse, associating texts with “perspectives” is more general and has been studied in sentiment analysis, discovery of regional variation, and value-sensitive design. We show experimentally that the model’s hierarchical structure improves prediction of perspective in both a political domain and on sentiment analysis tasks, and we argue that the topic hierarchies exposed by the model are indeed capturing structure in line with the theory that motivated the work. 1 ߨ ݉ ߠௗ ߙ ߰ௗ ߛ ݐௗ௦ ݖௗ௦ ݓௗ௦ ܿௗ௧ ܰௗ௦ ∞ ߩ ܵௗ ݕௗ ܦ ߱ ߟ ߬௩ ܸ 1. For each node k ∈ [1, ∞) in the tree (a) Draw topic φk ∼ Dir(βk ) (b) Draw regression parameter ηk ∼ N (µ, σ) 2. For each word v ∈ [1, V ], draw τv ∼ Laplace(0, ω) 3. For each document d ∈ [1, D] (a) Draw level distribution θd ∼ GEM(m, π) (b) Draw table distribution ψd ∼ GEM(α) (c) For each table t ∈ [1, ∞), draw a path cd,t ∼ nCRP(γ) (d) For each sentence s ∈ [1, Sd ], draw a table indicator td,s ∼ Mult(ψd ) i. For each token n ∈ [1, Nd,s ] A. Draw level zd,s,n ∼ Mult(θd ) B. Draw word wd,s,n ∼ Mult(φcd,td,s ,zd,s,n ) ¯ ¯ (e) Draw response yd ∼ N (η T zd + τ T wd , ρ): ߶ ∞ ߤ i. zd,k = ¯ ߪ ߚ ii. wd,v = ¯ 1 Nd,· 1 Nd,· Sd s=1 Sd s=1 Nd,s n=1 I [kd,s,n = k] Nd,s n=1 I [wd,s,n = v] Figure 1: S H L DA’s generative process and plate diagram. Words w are explained by topic hierarchy φ, and response variables y are explained by per-topic regression coefficients η and global lexical coefficients τ . 2 S H L DA: Combining Supervision and Hierarchical Topic Structure Jointly capturing supervision and hierarchical topic structure falls under a class of models called supervised hierarchical latent Dirichlet allocation. These models take as input a set of D documents, each of which is associated with a response variable yd , and output a hierarchy of topics which is informed by yd . Zhang et al. [6] introduce the S H L DA family, focusing on a categorical response. In contrast, our novel model (which we call S H L DA for brevity), uses continuous responses. At its core, S H L DA’s document generative process resembles a combination of hierarchical latent Dirichlet allocation [7, HLDA] and the hierarchical Dirichlet process [8, HDP]. HLDA uses the nested Chinese restaurant process (nCRP(γ)), combined with an appropriate base distribution, to induce an unbounded tree-structured hierarchy of topics: general topics at the top, specific at the bottom. A document is generated by traversing this tree, at each level creating a new child (hence a new path) with probability proportional to γ or otherwise respecting the “rich-get-richer” property of a CRP. A drawback of HLDA, however, is that each document is restricted to only a single path in the tree. Recent work relaxes this restriction through different priors: nested HDP [9], nested Chinese franchises [10] or recursive CRPs [11]. In this paper, we address this problem by allowing documents to have multiple paths through the tree by leveraging information at the sentence level using the twolevel structure used in HDP. More specifically, in the HDP’s Chinese restaurant franchise metaphor, customers (i.e., tokens) are grouped by sitting at tables and each table takes a dish (i.e., topic) from a flat global menu. In our S H L DA, dishes are organized in a tree-structured global menu by using the nCRP as prior. Each path in the tree is a collection of L dishes (one for each level) and is called a combo. S H L DA groups sentences of a document by assigning them to tables and associates each table with a combo, and thus, models each document as a distribution over combos.1 In S H L DA’s metaphor, customers come in a restaurant and sit at a table in groups, where each group is a sentence. A sentence wd,s enters restaurant d and selects a table t (and its associated combo) with probability proportional to the number of sentences Sd,t at that table; or, it sits at a new table with probability proportional to α. After choosing the table (indexed by td,s ), if the table is new, the group will select a combo of dishes (i.e., a path, indexed by cd,t ) from the tree menu. Once a combo is in place, each token in the sentence chooses a “level” (indexed by zd,s,n ) in the combo, which specifies the topic (φkd,s,n ≡ φcd,td,s ,zd,s,n ) producing the associated observation (Figure 2). S H L DA also draws on supervised LDA [12, SLDA] associating each document d with an observable continuous response variable yd that represents the author’s perspective toward a topic, e.g., positive vs. negative sentiment, conservative vs. liberal ideology, etc. This lets us infer a multi-level topic structure informed by how topics are “framed” with respect to positions along the yd continuum. 1 We emphasize that, unlike in HDP where each table is assigned to a single dish, each table in our metaphor is associated with a combo–a collection of L dishes. We also use combo and path interchangeably. 2 Sd Sd,t ߶ଵ ߟଵ dish ߶ଵଵ ߟଵଵ ߶ଵଶ ߟଵଶ ߶ଵଵଵ ߟଵଵଵ ߶ଵଵଶ ߟଵଵଶ ߶ଵଶଵ ߟଵଶଵ ߶ଵଶଶ ߟଵଶଶ table ܿௗ௧ 1=ݐ 2=ݐ 1=ݐ 2=ݐ 3=ݐ 1=ݐ 2=ݐ ݐௗ௦ 2=ݏ 1=ݏ ܵ = ݏଵ 3=ݏ 2=ݏ 1=ݏ ݀=1 ݇ௗ௦ ܵ = ݏଶ ܵ = ݏ ݀=2 ߶ଵ ߟଵ ݀=ܦ customer group (token) (sentence) restaurant (document) ߶ଵଵ ߟଵଵ ݀=1 1=ݏ ߶ଵଵଵ ߟଵଵଵ combo (path) Nd,s Nd,·,l Nd,·,>l Nd,·,≥l Mc,l Cc,l,v Cd,x,l,v φk ηk τv cd,t td,s zd,s,n kd,s,n L C+ Figure 2: S H L DA’s restaurant franchise metaphor. # sentences in document d # groups (i.e. sentences) sitting at table t in restaurant d # tokens wd,s # tokens in wd assigned to level l # tokens in wd assigned to level > l ≡ Nd,·,l + Nd,·,>l # tables at level l on path c # word type v assigned to level l on path c # word type v in vd,x assigned to level l Topic at node k Regression parameter at node k Regression parameter of word type v Path assignment for table t in restaurant d Table assignment for group wd,s Level assignment for wd,s,n Node assignment for wd,s,n (i.e., node at level zd,s,n on path cd,td,s ) Height of the tree Set of all possible paths (including new ones) of the tree Table 1: Notation used in this paper Unlike SLDA, we model the response variables using a normal linear regression that contains both pertopic hierarchical and per-word lexical regression parameters. The hierarchical regression parameters are just like topics’ regression parameters in SLDA: each topic k (here, a tree node) has a parameter ηk , and the model uses the empirical distribution over the nodes that generated a document as the regressors. However, the hierarchy in S H L DA makes it possible to discover relationships between topics and the response variable that SLDA’s simple latent space obscures. Consider, for example, a topic model trained on Congressional debates. Vanilla LDA would likely discover a healthcare category. SLDA [12] could discover a pro-Obamacare topic and an anti-Obamacare topic. S H L DA could do that and capture the fact that there are alternative perspectives, i.e., that the healthcare issue is being discussed from two ideological perspectives, along with characterizing how the higher level topic is discussed by those on both sides of that ideological debate. Sometimes, of course, words are strongly associated with extremes on the response variable continuum regardless of underlying topic structure. Therefore, in addition to hierarchical regression parameters, we include global lexical regression parameters to model the interaction between specific words and response variables. We denote the regression parameter associated with a word type v in the vocabulary as τv , and use the normalized frequency of v in the documents to be its regressor. Including both hierarchical and lexical parameters is important. For detecting ideology in the US, “liberty” is an effective indicator of conservative speakers regardless of context; however, “cost” is a conservative-leaning indicator in discussions about environmental policy but liberal-leaning in debates about foreign policy. For sentiment, “wonderful” is globally a positive word; however, “unexpected” is a positive descriptor of books but a negative one of a car’s steering. S H L DA captures these properties in a single model. 3 Posterior Inference and Optimization Given documents with observed words w = {wd,s,n } and response variables y = {yd }, the inference task is to find the posterior distribution over: the tree structure including topic φk and regression parameter ηk for each node k, combo assignment cd,t for each table t in document d, table assignment td,s for each sentence s in a document d, and level assignment zd,s,n for each token wd,s,n . We approximate S H L DA’s posterior using stochastic EM, which alternates between a Gibbs sampling E-step and an optimization M-step. More specifically, in the E-step, we integrate out ψ, θ and φ to construct a Markov chain over (t, c, z) and alternate sampling each of them from their conditional distributions. In the M-step, we optimize the regression parameters η and τ using L-BFGS [13]. Before describing each step in detail, let us define the following probabilities. For more thorough derivations, please see the supplement. 3 • First, define vd,x as a set of tokens (e.g., a token, a sentence or a set of sentences) in document d. The conditional density of vd,x being assigned to path c given all other assignments is −d,x Γ(Cc,l,· + V βl ) L −d,x fc (vd,x ) = l=1 −d,x Γ(Cc,l,v + Cd,x,l,v + βl ) V −d,x Γ(Cc,l,· + Cd,x,l,· + V βl ) (1) −d,x Γ(Cc,l,v + βl ) v=1 where superscript −d,x denotes the same count excluding assignments of vd,x ; marginal counts −d,x are represented by ·’s. For a new path cnew , if the node does not exist, Ccnew ,l,v = 0 for all word types v. • Second, define the conditional density of the response variable yd of document d given vd,x being −d,x assigned to path c and all other assignments as gc (yd ) = 1 N Nd,· ηc,l · Cd,x,l,· + ηcd,td,s ,zd,s,n + wd,s,n ∈{wd \vd,x } Sd Nd,s L τwd,s,n , ρ (2) s=1 n=1 l=1 where Nd,· is the total number of tokens in document d. For a new node at level l on a new path cnew , we integrate over all possible values of ηcnew ,l . Sampling t: For each group wd,s we need to sample a table td,s . The conditional distribution of a table t given wd,s and other assignments is proportional to the number of sentences sitting at t times the probability of wd,s and yd being observed under this assignment. This is P (td,s = t | rest) ∝ P (td,s = t | t−s ) · P (wd,s , yd | td,s = t, w−d,s , t−d,s , z, c, η) d ∝ −d,s −d,s −d,s Sd,t · fcd,t (wd,s ) · gcd,t (yd ), for existing table t; (3) −d,s −d,s α · c∈C + P (cd,tnew = c | c−d,s ) · fc (wd,s ) · gc (yd ), for new table tnew . For a new table tnew , we need to sum over all possible paths C + of the tree, including new ones. For example, the set C + for the tree shown in Figure 2 consists of four existing paths (ending at one of the four leaf nodes) and three possible new paths (a new leaf off of one of the three internal nodes). The prior probability of path c is: P (cd,tnew = c | c−d,s ) ∝ L l=2 −d,s Mc,l −d,s Mc,l−1 + γl−1 γl∗ −d,s M ∗ cnew ,l∗ + γl , l∗ l=2 for an existing path c; (4) −d,s Mcnew ,l , for a new path cnew which consists of an existing path −d,s Mcnew ,l−1 + γl−1 from the root to a node at level l∗ and a new node. Sampling z: After assigning a sentence wd,s to a table, we assign each token wd,s,n to a level to choose a dish from the combo. The probability of assigning wd,s,n to level l is −s,n P (zd,s,n = l | rest) ∝ P (zd,s,n = l | zd )P (wd,s,n , yd | zd,s,n = l, w−d,s,n , z −d,s,n , t, c, η) (5) The first factor captures the probability that a customer in restaurant d is assigned to level l, conditioned on the level assignments of all other customers in restaurant d, and is equal to P (zd,s,n = −s,n l | zd ) = −d,s,n mπ + Nd,·,l −d,s,n π + Nd,·,≥l l−1 −d,s,n (1 − m)π + Nd,·,>j −d,s,n π + Nd,·,≥j j=1 , The second factor is the probability of observing wd,s,n and yd , given that wd,s,n is assigned to level −d,s,n −d,s,n l: P (wd,s,n , yd | zd,s,n = l, w−d,s,n , z −d,s,n , t, c, η) = fcd,t (wd,s,n ) · gcd,t (yd ). d,s d,s Sampling c: After assigning customers to tables and levels, we also sample path assignments for all tables. This is important since it can change the assignments of all customers sitting at a table, which leads to a well-mixed Markov chain and faster convergence. The probability of assigning table t in restaurant d to a path c is P (cd,t = c | rest) ∝ P (cd,t = c | c−d,t ) · P (wd,t , yd | cd,t = c, w−d,t , c−d,t , t, z, η) (6) where we slightly abuse the notation by using wd,t ≡ ∪{s|td,s =t} wd,s to denote the set of customers in all the groups sitting at table t in restaurant d. The first factor is the prior probability of a path given all tables’ path assignments c−d,t , excluding table t in restaurant d and is given in Equation 4. The second factor in Equation 6 is the probability of observing wd,t and yd given the new path −d,t −d,t assignments, P (wd,t , yd | cd,t = c, w−d,t , c−d,t , t, z, η) = fc (wd,t ) · gc (yd ). 4 Optimizing η and τ : We optimize the regression parameters η and τ via the likelihood, 1 L(η, τ ) = − 2ρ D 1 ¯ ¯ (yd − η zd − τ wd ) − 2σ T d=1 T K+ 2 (ηk − µ)2 − k=1 1 ω V |τv |, (7) v=1 where K + is the number of nodes in the tree.2 This maximization is performed using L-BFGS [13]. 4 Data: Congress, Products, Films We conduct our experiments using three datasets: Congressional floor debates, Amazon product reviews, and movie reviews. For all datasets, we remove stopwords, add bigrams to the vocabulary, and filter the vocabulary using tf-idf.3 • U.S Congressional floor debates: We downloaded debates of the 109th US Congress from GovTrack4 and preprocessed them as in Thomas et al. [14]. To remove uninterestingly non-polarized debates, we ignore bills with less than 20% “Yea” votes or less than 20% “Nay” votes. Each document d is a turn (a continuous utterance by a single speaker, i.e. speech segment [14]), and its response variable yd is the first dimension of the speaker’s DW- NOMINATE score [15], which captures the traditional left-right political distinction.5 After processing, our corpus contains 5,201 turns in the House, 3,060 turns in the Senate, and 5,000 words in the vocabulary.6 • Amazon product reviews: From a set of Amazon reviews of manufactured products such as computers, MP 3 players, GPS devices, etc. [16], we focused on the 50 most frequently reviewed products. After filtering, this corpus contains 37,191 reviews with a vocabulary of 5,000 words. We use the rating associated with each review as the response variable yd .7 • Movie reviews: Our third corpus is a set of 5,006 reviews of movies [17], again using review ratings as the response variable yd , although in this corpus the ratings are normalized to the range from 0 to 1. After preprocessing, the vocabulary contains 5,000 words. 5 Evaluating Prediction S H L DA’s response variable predictions provide a formally rigorous way to assess whether it is an improvement over prior methods. We evaluate effectiveness in predicting values of the response variables for unseen documents in the three datasets. For comparison we consider these baselines: • Multiple linear regression (MLR) models the response variable as a linear function of multiple features (or regressors). Here, we consider two types of features: topic-based features and lexicallybased features. Topic-based MLR, denoted by MLR - LDA, uses the topic distributions learned by vanilla LDA as features [12], while lexically-based MLR, denoted by MLR - VOC, uses the frequencies of words in the vocabulary as features. MLR - LDA - VOC uses both features. • Support vector regression (SVM) is a discriminative method [18] that uses LDA topic distributions (SVM - LDA), word frequencies (SVM - VOC), and both (SVM - LDA - VOC) as features.8 • Supervised topic model (SLDA): we implemented SLDA using Gibbs sampling. The version of SLDA we use is slightly different from the original SLDA described in [12], in that we place a Gaussian prior N (0, 1) over the regression parameters to perform L2-norm regularization.9 For parametric models (LDA and SLDA), which require the number of topics K to be specified beforehand, we use K ∈ {10, 30, 50}. We use symmetric Dirichlet priors in both LDA and SLDA, initialize The superscript + is to denote that this number is unbounded and varies during the sampling process. To find bigrams, we begin with bigram candidates that occur at least 10 times in the corpus and use Pearson’s χ2 -test to filter out those that have χ2 -value less than 5, which corresponds to a significance level of 0.025. We then treat selected bigrams as single word types and add them to the vocabulary. 2 3 4 http://www.govtrack.us/data/us/109/ 5 Scores were downloaded from http://voteview.com/dwnomin_joint_house_and_senate.htm 6 Data will be available after blind review. 7 The ratings can range from 1 to 5, but skew positive. 8 9 http://svmlight.joachims.org/ This performs better than unregularized SLDA in our experiments. 5 Floor Debates House-Senate Senate-House PCC ↑ MSE ↓ PCC ↑ MSE ↓ Amazon Reviews PCC ↑ MSE ↓ Movie Reviews PCC ↑ MSE ↓ SVM - LDA 10 SVM - LDA 30 SVM - LDA 50 SVM - VOC SVM - LDA - VOC 0.173 0.172 0.169 0.336 0.256 0.861 0.840 0.832 1.549 0.784 0.08 0.155 0.215 0.131 0.246 1.247 1.183 1.135 1.467 1.101 0.157 0.277 0.245 0.373 0.371 1.241 1.091 1.130 0.972 0.965 0.327 0.365 0.395 0.584 0.585 0.970 0.938 0.906 0.681 0.678 MLR - LDA 10 MLR - LDA 30 MLR - LDA 50 MLR - VOC MLR - LDA - VOC 0.163 0.160 0.150 0.322 0.319 0.735 0.737 0.741 0.889 0.873 0.068 0.162 0.248 0.191 0.194 1.151 1.125 1.081 1.124 1.120 0.143 0.258 0.234 0.408 0.410 1.034 1.065 1.114 0.869 0.860 0.328 0.367 0.389 0.568 0.581 0.957 0.936 0.914 0.721 0.702 SLDA 10 SLDA 30 SLDA 50 0.154 0.174 0.254 0.729 0.793 0.897 0.090 0.128 0.245 1.145 1.188 1.184 0.270 0.357 0.241 1.113 1.146 1.939 0.383 0.433 0.503 0.953 0.852 0.772 S H L DA 0.356 0.753 0.303 1.076 0.413 0.891 0.597 0.673 Models Table 2: Regression results for Pearson’s correlation coefficient (PCC, higher is better (↑)) and mean squared error (MSE, lower is better (↓)). Results on Amazon product reviews and movie reviews are averaged over 5 folds. Subscripts denote the number of topics for parametric models. For SVM - LDA - VOC and MLR - LDA - VOC, only best results across K ∈ {10, 30, 50} are reported. Best results are in bold. the Dirichlet hyperparameters to 0.5, and use slice sampling [19] for updating hyperparameters. For SLDA , the variance of the regression is set to 0.5. For S H L DA , we use trees with maximum depth of three. We slice sample m, π, β and γ, and fix µ = 0, σ = 0.5, ω = 0.5 and ρ = 0.5. We found that the following set of initial hyperparameters works reasonably well for all the datasets in our experiments: m = 0.5, π = 100, β = (1.0, 0.5, 0.25), γ = (1, 1), α = 1. We also set the regression parameter of the root node to zero, which speeds inference (since it is associated with every document) and because it is reasonable to assume that it would not change the response variable. To compare the performance of different methods, we compute Pearson’s correlation coefficient (PCC) and mean squared error (MSE) between the true and predicted values of the response variables and average over 5 folds. For the Congressional debate corpus, following Yu et al. [20], we use documents in the House to train and test on documents in the Senate and vice versa. Results and analysis Table 2 shows the performance of all models on our three datasets. Methods that only use topic-based features such as SVM - LDA and MLR - LDA do poorly. Methods only based on lexical features like SVM - VOC and MLR - VOC outperform methods that are based only on topic features significantly for the two review datasets, but are comparable or worse on congressional debates. This suggests that reviews have more highly discriminative words than political speeches (Table 3). Combining topic-based and lexically-based features improves performance, which supports our choice of incorporating both per-topic and per-word regression parameters in S H L DA. In all cases, S H L DA achieves strong performance results. For the two cases where S H L DA was second best in MSE score (Amazon reviews and House-Senate), it outperforms other methods in PCC. Doing well in PCC for these two datasets is important since achieving low MSE is relatively easier due to the response variables’ bimodal distribution in the floor debates and positively-skewed distribution in Amazon reviews. For the floor debate dataset, the results of the House-Senate experiment are generally better than those of the Senate-House experiment, which is consistent with previous results [20] and is explained by the greater number of debates in the House. 6 Qualitative Analysis: Agendas and Framing/Perspective Although a formal coherence evaluation [21] remains a goal for future work, a qualitative look at the topic hierarchy uncovered by the model suggests that it is indeed capturing agenda/framing structure as discussed in Section 1. In Figure 3, a portion of the topic hierarchy induced from the Congressional debate corpus, Nodes A and B illustrate agendas—issues introduced into political discourse—associated with a particular ideology: Node A focuses on the hardships of the poorer victims of hurricane Katrina and is associated with Democrats, and text associated with Node E discusses a proposed constitutional amendment to ban flag burning and is associated with Republicans. Nodes C and D, children of a neutral “tax” topic, reveal how parties frame taxes as gains in terms of new social services (Democrats) and losses for job creators (Republicans). 6 E flag constitution freedom supreme_court elections rights continuity american_flag constitutional_amendm ent gses credit_rating fannie_mae regulator freddie_mac market financial_services agencies competition investors fannie bill speaker time amendment chairman people gentleman legislation congress support R:1.1 R:0 A minimum_wage commission independent_commissio n investigate hurricane_katrina increase investigation R:1.0 B percent tax economy estate_tax capital_gains money taxes businesses families tax_cuts pay tax_relief social_security affordable_housing housing manager fund activities funds organizations voter_registration faithbased nonprofits R:0.4 D:1.7 C death_tax jobs businesses business family_businesses equipment productivity repeal_permanency employees capital farms D REPUBLICAN billion budget children cuts debt tax_cuts child_support deficit education students health_care republicans national_debt R:4.3 D:2.2 DEMOCRAT D:4.5 Figure 3: Topics discovered from Congressional floor debates. Many first-level topics are bipartisan (purple), while lower level topics are associated with specific ideologies (Democrats blue, Republicans red). For example, the “tax” topic (B) is bipartisan, but its Democratic-leaning child (D) focuses on social goals supported by taxes (“children”, “education”, “health care”), while its Republican-leaning child (C) focuses on business implications (“death tax”, “jobs”, “businesses”). The number below each topic denotes the magnitude of the learned regression parameter associated with that topic. Colors and the numbers beneath each topic show the regression parameter η associated with the topic. Figure 4 shows the topic structure discovered by S H L DA in the review corpus. Nodes at higher levels are relatively neutral, with relatively small regression parameters.10 These nodes have general topics with no specific polarity. However, the bottom level clearly illustrates polarized positive/negative perspective. For example, Node A concerns washbasins for infants, and has two polarized children nodes: reviewers take a positive perspective when their children enjoy the product (Node B: “loves”, “splash”, “play”) but have negative reactions when it leaks (Node C: “leak(s/ed/ing)”). transmitter ipod car frequency iriver product transmitters live station presets itrip iriver_aft charges international_mode driving P:6.6 tried waste batteries tunecast rabbit_ears weak terrible antenna hear returned refund returning item junk return A D router setup network expander set signal wireless connect linksys connection house wireless_router laptop computer wre54g N:2.2 N:1.0 tivo adapter series adapters phone_line tivo_wireless transfer plugged wireless_adapter tivos plug dvr tivo_series tivo_box tivo_unit P:5.1 tub baby water bath sling son daughter sit bathtub sink newborn months bath_tub bathe bottom N:8.0 months loves hammock splash love baby drain eurobath hot fits wash play infant secure slip P:7.5 NEGATIVE N:0 N:2.7 B POSITIVE time bought product easy buy love using price lot able set found purchased money months transmitter car static ipod radio mp3_player signal station sound music sound_quality volume stations frequency frequencies C leaks leaked leak leaking hard waste snap suction_cups lock tabs difficult bottom tub_leaks properly ring N:8.9 monitor radio weather_radio night baby range alerts sound sony house interference channels receiver static alarm N:1.7 hear feature static monitors set live warning volume counties noise outside alert breathing rechargeable_battery alerts P:6.2 version hours phone F firmware told spent linksys tech_support technical_supportcusto mer_service range_expander support return N:10.6 E router firmware ddwrt wrt54gl version wrt54g tomato linksys linux routers flash versions browser dlink stable P:4.8 z22 palm pda palm_z22 calendar software screen contacts computer device sync information outlook data programs N:1.9 headphones sound pair bass headset sound_quality ear ears cord earbuds comfortable hear head earphones fit N:1.3 appointments organized phone lists handheld organizer photos etc pictures memos track bells books purse whistles P:5.8 noise_canceling noise sony exposed noise_cancellation stopped wires warranty noise_cancelling bud pay white_noise disappointed N:7.6 bottles bottle baby leak nipples nipple avent avent_bottles leaking son daughter formula leaks gas milk comfortable sound phones sennheiser bass px100 px100s phone headset highs portapros portapro price wear koss N:2.0 leak formula bottles_leak feeding leaked brown frustrating started clothes waste newborn playtex_ventaire soaked matter N:7.9 P:5.7 nipple breast nipples dishwasher ring sippy_cups tried breastfeed screwed breastfeeding nipple_confusion avent_system bottle P:6.4 Figure 4: Topics discovered from Amazon reviews. Higher topics are general, while lower topics are more specific. The polarity of the review is encoded in the color: red (negative) to blue (positive). Many of the firstlevel topics have no specific polarity and are associated with a broad class of products such as “routers” (Node D). However, the lowest topics in the hierarchy are often polarized; one child topic of “router” focuses on upgradable firmware such as “tomato” and “ddwrt” (Node E, positive) while another focuses on poor “tech support” and “customer service” (Node F, negative). The number below each topic is the regression parameter learned with that topic. In addition to the per-topic regression parameters, S H L DA also associates each word with a lexical regression parameter τ . Table 3 shows the top ten words with highest and lowest τ . The results are unsuprising, although the lexical regression for the Congressional debates is less clear-cut than other 10 All of the nodes at the second level have slightly negative values for the regression parameters mainly due to the very skewed distribution of the review ratings in Amazon. 7 datasets. As we saw in Section 5, for similar datasets, S H L DA’s context-specific regression is more useful when global lexical weights do not readily differentiate documents. Dataset Floor Debates Amazon Reviews Movie Reviews Top 10 words with positive weights bringing, private property, illegally, tax relief, regulation, mandates, constitutional, committee report, illegal alien highly recommend, pleased, love, loves, perfect, easy, excellent, amazing, glad, happy hilarious, fast, schindler, excellent, motion pictures, academy award, perfect, journey, fortunately, ability Top 10 words with negative weights bush administration, strong opposition, ranking, republicans, republican leadership, secret, discriminate, majority, undermine waste, returned, return, stopped, leak, junk, useless, returning, refund, terrible bad, unfortunately, supposed, waste, mess, worst, acceptable, awful, suppose, boring Table 3: Top words based on the global lexical regression coefficient, τ . For the floor debates, positive τ ’s are Republican-leaning while negative τ ’s are Democrat-leaning. 7 Related Work S H L DA joins a family of LDA extensions that introduce hierarchical topics, supervision, or both. Owing to limited space, we focus here on related work that combines the two. Petinot et al. [22] propose hierarchical Labeled LDA (hLLDA), which leverages an observed document ontology to learn topics in a tree structure; however, hLLDA assumes that the underlying tree structure is known a priori. SSHLDA [23] generalizes hLLDA by allowing the document hierarchy labels to be partially observed, with unobserved labels and topic tree structure then inferred from the data. Boyd-Graber and Resnik [24] used hierarchical distributions within topics to learn topics across languages. In addition to these “upstream” models [25], Perotte et al. [26] propose a “downstream” model called HSLDA , which jointly models documents’ hierarchy of labels and topics. HSLDA ’s topic structure is flat, however, and the response variable is a hierarchy of labels associated with each document, unlike S H L DA’s continuous response variable. Finally, another body related body of work includes models that jointly capture topics and other facets such as ideologies/perspectives [27, 28] and sentiments/opinions [29], albeit with discrete rather than continuously valued responses. Computational modeling of sentiment polarity is a voluminous field [30], and many computational political science models describe agendas [5] and ideology [31]. Looking at framing or bias at the sentence level, Greene and Resnik [32] investigate the role of syntactic structure in framing, Yano et al. [33] look at lexical indications of sentence-level bias, and Recasens et al. [34] develop linguistically informed sentence-level features for identifying bias-inducing words. 8 Conclusion We have introduced S H L DA, a model that associates a continuously valued response variable with hierarchical topics to capture both the issues under discussion and alternative perspectives on those issues. The two-level structure improves predictive performance over existing models on multiple datasets, while also adding potentially insightful hierarchical structure to the topic analysis. Based on a preliminary qualitative analysis, the topic hierarchy exposed by the model plausibly captures the idea of agenda setting, which is related to the issues that get discussed, and framing, which is related to authors’ perspectives on those issues. We plan to analyze the topic structure produced by S H L DA with political science collaborators and more generally to study how S H L DA and related models can help analyze and discover useful insights from political discourse. Acknowledgments This research was supported in part by NSF under grant #1211153 (Resnik) and #1018625 (BoydGraber and Resnik). Any opinions, findings, conclusions, or recommendations expressed here are those of the authors and do not necessarily reflect the view of the sponsor. 8 References [1] McCombs, M. The agenda-setting role of the mass media in the shaping of public opinion. North, 2009(05-12):21, 2002. [2] McCombs, M., S. Ghanem. The convergence of agenda setting and framing. In Framing public life. 2001. [3] Baumgartner, F. R., S. L. De Boef, A. E. Boydstun. The decline of the death penalty and the discovery of innocence. Cambridge University Press, 2008. [4] Blei, D. M., A. Ng, M. Jordan. Latent Dirichlet allocation. JMLR, 3, 2003. [5] Grimmer, J. A Bayesian hierarchical topic model for political texts: Measuring expressed agendas in Senate press releases. Political Analysis, 18(1):1–35, 2010. [6] Zhang, J. Explore objects and categories in unexplored environments based on multimodal data. Ph.D. thesis, University of Hamburg, 2012. [7] Blei, D. M., T. L. Griffiths, M. I. Jordan. The nested Chinese restaurant process and Bayesian nonparametric inference of topic hierarchies. J. ACM, 57(2), 2010. [8] Teh, Y. W., M. I. Jordan, M. J. Beal, et al. Hierarchical Dirichlet processes. JASA, 101(476), 2006. [9] Paisley, J. W., C. Wang, D. M. Blei, et al. Nested hierarchical Dirichlet processes. arXiv:1210.6738, 2012. [10] Ahmed, A., L. Hong, A. Smola. The nested Chinese restaurant franchise process: User tracking and document modeling. In ICML. 2013. [11] Kim, J. H., D. Kim, S. Kim, et al. Modeling topic hierarchies with the recursive Chinese restaurant process. In CIKM, pages 783–792. 2012. [12] Blei, D. M., J. D. McAuliffe. Supervised topic models. In NIPS. 2007. [13] Liu, D., J. Nocedal. On the limited memory BFGS method for large scale optimization. Math. Prog., 1989. [14] Thomas, M., B. Pang, L. Lee. Get out the vote: Determining support or opposition from Congressional floor-debate transcripts. In EMNLP. 2006. [15] Lewis, J. B., K. T. Poole. Measuring bias and uncertainty in ideal point estimates via the parametric bootstrap. Political Analysis, 12(2), 2004. [16] Jindal, N., B. Liu. Opinion spam and analysis. In WSDM. 2008. [17] Pang, B., L. Lee. Seeing stars: Exploiting class relationships for sentiment categorization with respect to rating scales. In ACL. 2005. [18] Joachims, T. Making large-scale SVM learning practical. In Adv. in Kernel Methods - SVM. 1999. [19] Neal, R. M. Slice sampling. Annals of Statistics, 31:705–767, 2003. [20] Yu, B., D. Diermeier, S. Kaufmann. Classifying party affiliation from political speech. JITP, 2008. [21] Chang, J., J. Boyd-Graber, C. Wang, et al. Reading tea leaves: How humans interpret topic models. In NIPS. 2009. [22] Petinot, Y., K. McKeown, K. Thadani. A hierarchical model of web summaries. In HLT. 2011. [23] Mao, X., Z. Ming, T.-S. Chua, et al. SSHLDA: A semi-supervised hierarchical topic model. In EMNLP. 2012. [24] Boyd-Graber, J., P. Resnik. Holistic sentiment analysis across languages: Multilingual supervised latent Dirichlet allocation. In EMNLP. 2010. [25] Mimno, D. M., A. McCallum. Topic models conditioned on arbitrary features with Dirichlet-multinomial regression. In UAI. 2008. [26] Perotte, A. J., F. Wood, N. Elhadad, et al. Hierarchically supervised latent Dirichlet allocation. In NIPS. 2011. [27] Ahmed, A., E. P. Xing. Staying informed: Supervised and semi-supervised multi-view topical analysis of ideological perspective. In EMNLP. 2010. [28] Eisenstein, J., A. Ahmed, E. P. Xing. Sparse additive generative models of text. In ICML. 2011. [29] Jo, Y., A. H. Oh. Aspect and sentiment unification model for online review analysis. In WSDM. 2011. [30] Pang, B., L. Lee. Opinion Mining and Sentiment Analysis. Now Publishers Inc, 2008. [31] Monroe, B. L., M. P. Colaresi, K. M. Quinn. Fightin’words: Lexical feature selection and evaluation for identifying the content of political conflict. Political Analysis, 16(4):372–403, 2008. [32] Greene, S., P. Resnik. More than words: Syntactic packaging and implicit sentiment. In NAACL. 2009. [33] Yano, T., P. Resnik, N. A. Smith. Shedding (a thousand points of) light on biased language. In NAACL-HLT Workshop on Creating Speech and Language Data with Amazon’s Mechanical Turk. 2010. [34] Recasens, M., C. Danescu-Niculescu-Mizil, D. Jurafsky. Linguistic models for analyzing and detecting biased language. In ACL. 2013. 9
4 0.080463082 153 nips-2013-Learning Feature Selection Dependencies in Multi-task Learning
Author: Daniel Hernández-Lobato, José Miguel Hernández-Lobato
Abstract: A probabilistic model based on the horseshoe prior is proposed for learning dependencies in the process of identifying relevant features for prediction. Exact inference is intractable in this model. However, expectation propagation offers an approximate alternative. Because the process of estimating feature selection dependencies may suffer from over-fitting in the model proposed, additional data from a multi-task learning scenario are considered for induction. The same model can be used in this setting with few modifications. Furthermore, the assumptions made are less restrictive than in other multi-task methods: The different tasks must share feature selection dependencies, but can have different relevant features and model coefficients. Experiments with real and synthetic data show that this model performs better than other multi-task alternatives from the literature. The experiments also show that the model is able to induce suitable feature selection dependencies for the problems considered, only from the training data. 1
5 0.07788039 317 nips-2013-Streaming Variational Bayes
Author: Tamara Broderick, Nicholas Boyd, Andre Wibisono, Ashia C. Wilson, Michael Jordan
Abstract: We present SDA-Bayes, a framework for (S)treaming, (D)istributed, (A)synchronous computation of a Bayesian posterior. The framework makes streaming updates to the estimated posterior according to a user-specified approximation batch primitive. We demonstrate the usefulness of our framework, with variational Bayes (VB) as the primitive, by fitting the latent Dirichlet allocation model to two large-scale document collections. We demonstrate the advantages of our algorithm over stochastic variational inference (SVI) by comparing the two after a single pass through a known amount of data—a case where SVI may be applied—and in the streaming setting, where SVI does not apply. 1
6 0.07178127 158 nips-2013-Learning Multiple Models via Regularized Weighting
7 0.066936463 82 nips-2013-Decision Jungles: Compact and Rich Models for Classification
8 0.061463758 142 nips-2013-Information-theoretic lower bounds for distributed statistical estimation with communication constraints
9 0.05338515 245 nips-2013-Pass-efficient unsupervised feature selection
10 0.048002034 268 nips-2013-Reflection methods for user-friendly submodular optimization
11 0.046011828 146 nips-2013-Large Scale Distributed Sparse Precision Estimation
12 0.043082446 34 nips-2013-Analyzing Hogwild Parallel Gaussian Gibbs Sampling
13 0.04301415 313 nips-2013-Stochastic Majorization-Minimization Algorithms for Large-Scale Optimization
14 0.040685173 68 nips-2013-Confidence Intervals and Hypothesis Testing for High-Dimensional Statistical Models
15 0.038426761 177 nips-2013-Local Privacy and Minimax Bounds: Sharp Rates for Probability Estimation
16 0.038418338 19 nips-2013-Accelerated Mini-Batch Stochastic Dual Coordinate Ascent
17 0.038211994 192 nips-2013-Minimax Theory for High-dimensional Gaussian Mixtures with Sparse Mean Separation
18 0.037950005 125 nips-2013-From Bandits to Experts: A Tale of Domination and Independence
19 0.037889492 198 nips-2013-More Effective Distributed ML via a Stale Synchronous Parallel Parameter Server
20 0.037066683 91 nips-2013-Dirty Statistical Models
topicId topicWeight
[(0, 0.112), (1, 0.015), (2, 0.038), (3, -0.008), (4, 0.029), (5, 0.036), (6, -0.045), (7, 0.021), (8, 0.028), (9, 0.02), (10, 0.032), (11, 0.028), (12, 0.02), (13, -0.047), (14, -0.006), (15, -0.005), (16, 0.033), (17, 0.0), (18, -0.012), (19, 0.003), (20, -0.003), (21, -0.039), (22, 0.016), (23, 0.059), (24, -0.042), (25, -0.045), (26, 0.041), (27, -0.029), (28, -0.006), (29, -0.049), (30, 0.025), (31, 0.051), (32, -0.022), (33, 0.013), (34, 0.057), (35, 0.025), (36, -0.057), (37, 0.039), (38, -0.021), (39, -0.081), (40, -0.108), (41, -0.069), (42, 0.017), (43, 0.104), (44, 0.034), (45, 0.045), (46, -0.068), (47, -0.015), (48, -0.034), (49, -0.04)]
simIndex simValue paperId paperTitle
same-paper 1 0.87266129 111 nips-2013-Estimation, Optimization, and Parallelism when Data is Sparse
Author: John Duchi, Michael Jordan, Brendan McMahan
Abstract: We study stochastic optimization problems when the data is sparse, which is in a sense dual to current perspectives on high-dimensional statistical learning and optimization. We highlight both the difficulties—in terms of increased sample complexity that sparse data necessitates—and the potential benefits, in terms of allowing parallelism and asynchrony in the design of algorithms. Concretely, we derive matching upper and lower bounds on the minimax rate for optimization and learning with sparse data, and we exhibit algorithms achieving these rates. We also show how leveraging sparsity leads to (still minimax optimal) parallel and asynchronous algorithms, providing experimental evidence complementing our theoretical results on several medium to large-scale learning tasks. 1 Introduction and problem setting In this paper, we investigate stochastic optimization problems in which the data is sparse. Formally, let {F (·; ξ), ξ ∈ Ξ} be a collection of real-valued convex functions, each of whose domains contains the convex set X ⊂ Rd . For a probability distribution P on Ξ, we consider the following optimization problem: minimize f (x) := E[F (x; ξ)] = x∈X F (x; ξ)dP (ξ). (1) Ξ By data sparsity, we mean the samples ξ are sparse: assuming that samples ξ lie in Rd , and defining the support supp(x) of a vector x to the set of indices of its non-zero components, we assume supp F (x; ξ) ⊂ supp ξ. (2) The sparsity condition (2) means that F (x; ξ) does not “depend” on the values of xj for indices j such that ξj = 0.1 This type of data sparsity is prevalent in statistical optimization problems and machine learning applications; in spite of its prevalence, study of such problems has been limited. As a motivating example, consider a text classification problem: data ξ ∈ Rd represents words appearing in a document, and we wish to minimize a logistic loss F (x; ξ) = log(1 + exp( ξ, x )) on the data (we encode the label implicitly with the sign of ξ). Such generalized linear models satisfy the sparsity condition (2), and while instances are of very high dimension, in any given instance, very few entries of ξ are non-zero [8]. From a modelling perspective, it thus makes sense to allow a dense predictor x: any non-zero entry of ξ is potentially relevant and important. In a sense, this is dual to the standard approaches to high-dimensional problems; one usually assumes that the data ξ may be dense, but there are only a few relevant features, and thus a parsimonious model x is desirous [2]. So 1 Formally, if πξ denotes the coordinate projection zeroing all indices j of its argument where ξj = 0, then F (πξ (x); ξ) = F (x; ξ) for all x, ξ. This follows from the first-order conditions for convexity [6]. 1 while such sparse data problems are prevalent—natural language processing, information retrieval, and other large data settings all have significant data sparsity—they do not appear to have attracted as much study as their high-dimensional “duals” of dense data and sparse predictors. In this paper, we investigate algorithms and their inherent limitations for solving problem (1) under natural conditions on the data generating distribution. Recent work in the optimization and machine learning communities has shown that data sparsity can be leveraged to develop parallel (and even asynchronous [12]) optimization algorithms [13, 14], but this work does not consider the statistical effects of data sparsity. In another line of research, Duchi et al. [4] and McMahan and Streeter [9] develop “adaptive” stochastic gradient algorithms to address problems in sparse data regimes (2). These algorithms exhibit excellent practical performance and have theoretical guarantees on their convergence, but it is not clear if they are optimal—in that no algorithm can attain better statistical performance—or whether they can leverage parallel computing as in the papers [12, 14]. In this paper, we take a two-pronged approach. First, we investigate the fundamental limits of optimization and learning algorithms in sparse data regimes. In doing so, we derive lower bounds on the optimization error of any algorithm for problems of the form (1) with sparsity condition (2). These results have two main implications. They show that in some scenarios, learning with sparse data is quite difficult, as essentially each coordinate j ∈ [d] can be relevant and must be optimized for. In spite of this seemingly negative result, we are also able to show that the A DAG RAD algorithms of [4, 9] are optimal, and we show examples in which their dependence on the dimension d can be made exponentially better than standard gradient methods. As the second facet of our two-pronged approach, we study how sparsity may be leveraged in parallel computing frameworks to give substantially faster algorithms that still achieve optimal sample complexity in terms of the number of samples ξ used. We develop two new algorithms, asynchronous dual averaging (A SYNC DA) and asynchronous A DAG RAD (A SYNC A DAG RAD), which allow asynchronous parallel solution of the problem (1) for general convex f and X . Combining insights of Niu et al.’s H OGWILD ! [12] with a new analysis, we prove our algorithms achieve linear speedup in the number of processors while maintaining optimal statistical guarantees. We also give experiments on text-classification and web-advertising tasks to illustrate the benefits of the new algorithms. 2 Minimax rates for sparse optimization We begin our study of sparse optimization problems by establishing their fundamental statistical and optimization-theoretic properties. To do this, we derive bounds on the minimax convergence rate of any algorithm for such problems. Formally, let x denote any estimator for a minimizer of the objective (1). We define the optimality gap N for the estimator x based on N samples ξ 1 , . . . , ξ N from the distribution P as N (x, F, X , P ) := f (x) − inf f (x) = EP [F (x; ξ)] − inf EP [F (x; ξ)] . x∈X x∈X This quantity is a random variable, since x is a random variable (it is a function of ξ 1 , . . . , ξ N ). To define the minimax error, we thus take expectations of the quantity N , though we require a bit more than simply E[ N ]. We let P denote a collection of probability distributions, and we consider a collection of loss functions F specified by a collection F of convex losses F : X × ξ → R. We can then define the minimax error for the family of losses F and distributions P as ∗ N (X , P, F) := inf sup sup EP [ x P ∈P F ∈F N (x(ξ 1:N ), F, X , P )], (3) where the infimum is taken over all possible estimators x (an estimator is an optimization scheme, or a measurable mapping x : ΞN → X ) . 2.1 Minimax lower bounds Let us now give a more precise characterization of the (natural) set of sparse optimization problems we consider to provide the lower bound. For the next proposition, we let P consist of distributions supported on Ξ = {−1, 0, 1}d , and we let pj := P (ξj = 0) be the marginal probability of appearance of feature j ∈ {1, . . . , d}. For our class of functions, we set F to consist of functions F satisfying the sparsity condition (2) and with the additional constraint that for g ∈ ∂x F (x; ξ), we have that the jth coordinate |gj | ≤ Mj for a constant Mj < ∞. We obtain 2 Proposition 1. Let the conditions of the preceding paragraph hold. Let R be a constant such that X ⊃ [−R, R]d . Then √ d pj 1 ∗ . Mj min pj , √ N (X , P, F) ≥ R 8 j=1 N log 3 We provide the proof of Proposition 1 in the supplement A.1 in the full version of the paper, providing a few remarks here. We begin by giving a corollary to Proposition 1 that follows when the data ξ obeys a type of power law: let p0 ∈ [0, 1], and assume that P (ξj = 0) = p0 j −α . We have Corollary 2. Let α ≥ 0. Let the conditions of Proposition 1 hold with Mj ≡ M for all j, and assume the power law condition P (ξj = 0) = p0 j −α on coordinate appearance probabilities. Then (1) If d > (p0 N )1/α , ∗ N (X , P, F) ≥ 2−α 1−α p0 p0 (p0 N ) 2α − 1 + d1−α − (p0 N ) α N 1−α 2 MR 8 2−α (2) If d ≤ (p0 N )1/α , ∗ N (X , P, F) ≥ MR 8 p0 N α 1 1 d1− 2 − 1 − α/2 1 − α/2 . . Expanding Corollary 2 slightly, for simplicity assume the number of samples is large enough that d ≤ (p0 N )1/α . Then we find that the lower bound on optimization error is of order p0 1− α p0 p0 d 2 when α < 2, M R log d when α → 2, and M R when α > 2. (4) N N N These results beg the question of tightness: are they improvable? As we see presently, they are not. MR 2.2 Algorithms for attaining the minimax rate To show that the lower bounds of Proposition 1 and its subsequent specializations are sharp, we review a few stochastic gradient algorithms. We begin with stochastic gradient descent (SGD): SGD repeatedly samples ξ ∼ P , computes g ∈ ∂x F (x; ξ), then performs the update x ← ΠX (x − ηg), where η is a stepsize parameter and ΠX denotes Euclidean projection onto X . Standard analyses of stochastic gradient descent [10] show that after N samples ξ i , the SGD estimator x(N ) satisfies R2 M ( d j=1 1 pj ) 2 √ , (5) N where R2 denotes the 2 -radius of X . Dual averaging, due to Nesterov [11] (sometimes called “follow the regularized leader” [5]) is a more recent algorithm. In dual averaging, one again samples g ∈ ∂x F (x; ξ), but instead of updating the parameter vector x one updates a dual vector z by z ← z + g, then computes 1 x ← argmin z, x + ψ(x) , η x∈X E[f (x(N ))] − inf f (x) ≤ O(1) x∈X 2 1 where ψ(x) is a strongly convex function defined over X (often one takes ψ(x) = 2 x 2 ). As we discuss presently, the dual averaging algorithm is somewhat more natural in asynchronous and parallel computing environments, and it enjoys the same type of convergence guarantees (5) as SGD. The A DAG RAD algorithm [4, 9] is an extension of the preceding stochastic gradient methods. It maintains a diagonal matrix S, where upon receiving a new sample ξ, A DAG RAD performs the following: it computes g ∈ ∂x F (x; ξ), then updates 2 Sj ← Sj + gj for j ∈ [d]. The dual averaging variant of A DAG RAD updates the usual dual vector z ← z + g; the update to x is based on S and a stepsize η and computes x ← argmin z, x + x ∈X 3 1 1 x ,S2x 2η . After N samples ξ, the averaged parameter x(N ) returned by A DAG RAD satisfies R∞ M E[f (x(N ))] − inf f (x) ≤ O(1) √ x∈X N d √ pj , (6) j=1 where R∞ denotes the ∞ -radius of X (cf. [4, Section 1.3 and Theorem 5]). By inspection, the A DAG RAD rate (6) matches the lower bound in Proposition 1 and is thus optimal. It is interesting to note, though, that in the power law setting of Corollary 2 (recall the error order (4)), a calculation √ shows that the multiplier for the SGD guarantee (5) becomes R∞ d max{d(1−α)/2 , 1}, while A DA G RAD attains rate at worst R∞ max{d1−α/2 , log d}. For α > 1, the A DAG RAD rate is no worse, √ and for α ≥ 2, is more than d/ log d better—an exponential improvement in the dimension. 3 Parallel and asynchronous optimization with sparsity As we note in the introduction, recent works [12, 14] have suggested that sparsity can yield benefits in our ability to parallelize stochastic gradient-type algorithms. Given the optimality of A DAG RADtype algorithms, it is natural to focus on their parallelization in the hope that we can leverage their ability to “adapt” to sparsity in the data. To provide the setting for our further algorithms, we first revisit Niu et al.’s H OGWILD ! [12]. H OGWILD ! is an asynchronous (parallelized) stochastic gradient algorithm for optimization over product-space domains, meaning that X in problem (1) decomposes as X = X1 × · · · × Xd , where Xj ⊂ R. Fix a stepsize η > 0. A pool of independently running processors then performs the following updates asynchronously to a centralized vector x: 1. Sample ξ ∼ P 2. Read x and compute g ∈ ∂x F (x; ξ) 3. For each j s.t. gj = 0, update xj ← ΠXj (xj − ηgj ). Here ΠXj denotes projection onto the jth coordinate of the domain X . The key of H OGWILD ! is that in step 2, the parameter x is allowed to be inconsistent—it may have received partial gradient updates from many processors—and for appropriate problems, this inconsistency is negligible. Indeed, Niu et al. [12] show linear speedup in optimization time as the number of processors grow; they show this empirically in many scenarios, providing a proof under the somewhat restrictive assumptions that there is at most one non-zero entry in any gradient g and that f has Lipschitz gradients. 3.1 Asynchronous dual averaging A weakness of H OGWILD ! is that it appears only applicable to problems for which the domain X is a product space, and its analysis assumes g 0 = 1 for all gradients g. In effort to alleviate these difficulties, we now develop and present our asynchronous dual averaging algorithm, A SYNC DA. A SYNC DA maintains and upates a centralized dual vector z instead of a parameter x, and a pool of processors perform asynchronous updates to z, where each processor independently iterates: 1. Read z and compute x := argminx∈X 1 z, x + η ψ(x) // Implicitly increment “time” counter t and let x(t) = x 2. Sample ξ ∼ P and let g ∈ ∂x F (x; ξ) // Let g(t) = g. 3. For j ∈ [d] such that gj = 0, update zj ← zj + gj . Because the actual computation of the vector x in A SYNC DA is performed locally on each processor in step 1 of the algorithm, the algorithm can be executed with any proximal function ψ and domain X . The only communication point between any of the processors is the addition operation in step 3. Since addition is commutative and associative, forcing all asynchrony to this point of the algorithm is a natural strategy for avoiding synchronization problems. In our analysis of A SYNC DA, and in our subsequent analysis of the adaptive methods, we require a measurement of time elapsed. With that in mind, we let t denote a time index that exists (roughly) behind-the-scenes. We let x(t) denote the vector x ∈ X computed in the tth step 1 of the A SYNC DA 4 algorithm, that is, whichever is the tth x actually computed by any of the processors. This quantity t exists and is recoverable from the algorithm, and it is possible to track the running sum τ =1 x(τ ). Additionally, we state two assumptions encapsulating the conditions underlying our analysis. Assumption A. There is an upper bound m on the delay of any processor. In addition, for each j ∈ [d] there is a constant pj ∈ [0, 1] such that P (ξj = 0) ≤ pj . We also require certain continuity (Lipschitzian) properties of the loss functions; these amount to a second moment constraint on the instantaneous ∂F and a rough measure of gradient sparsity. Assumption B. There exist constants M and (Mj )d such that the following bounds hold for all j=1 2 x ∈ X : E[ ∂x F (x; ξ) 2 ] ≤ M2 and for each j ∈ [d] we have E[|∂xj F (x; ξ)|] ≤ pj Mj . With these definitions, we have the following theorem, which captures the convergence behavior of A SYNC DA under the assumption that X is a Cartesian product, meaning that X = X1 × · · · × Xd , 2 where Xj ⊂ R, and that ψ(x) = 1 x 2 . Note the algorithm itself can still be efficiently parallelized 2 for more general convex X , even if the theorem does not apply. Theorem 3. Let Assumptions A and B and the conditions in the preceding paragraph hold. Then T E t=1 F (x(t); ξ t ) − F (x∗ ; ξ t ) ≤ 1 x∗ 2η d 2 2 η 2 p2 Mj . + T M2 + ηT m j 2 j=1 We now provide a few remarks to explain and simplify the result. Under the more stringent condition 2 d 2 that |∂xj F (x; ξ)| ≤ Mj , Assumption A implies E[ ∂x F (x; ξ) 2 ] ≤ j=1 pj Mj . Thus, for the d 2 remainder of this section we take M2 = j=1 pj Mj , which upper bounds the Lipschitz continuity constant of the objective function f . We then obtain the following corollary. √ T 1 Corollary 4. Define x(T ) = T t=1 x(t), and set η = x∗ 2 /M T . Then E[f (x(T )) − f (x∗ )] ≤ M x∗ √ T 2 +m x∗ 2 √ 2M T d 2 p2 M j . j j=1 Corollary 4 is nearly immediate: since ξ t is independent of x(t), we have E[F (x(t); ξ t ) | x(t)] = f (x(t)); applying Jensen’s inequality to f (x(T )) and performing an algebraic manipulation give the result. If the data is suitably sparse, meaning that pj ≤ 1/m, the bound in Corollary 4 simplifies to 3 M x∗ √ E[f (x(T )) − f (x )] ≤ 2 T ∗ 2 3 = 2 d j=1 2 pj M j x ∗ √ T 2 , (7) which is the convergence rate of stochastic gradient descent even in centralized settings (5). The √ convergence guarantee (7) shows that after T timesteps, the error scales as 1/ T ; however, if we have k processors, updates occur roughly k times as quickly, as they are asynchronous, and in time scaling as N/k, we can evaluate N gradient samples: a linear speedup. 3.2 Asynchronous AdaGrad We now turn to extending A DAG RAD to asynchronous settings, developing A SYNC A DAG RAD (asynchronous A DAG RAD). As in the A SYNC DA algorithm, A SYNC A DAG RAD maintains a shared dual vector z (the sum of gradients) and the shared matrix S, which is the diagonal sum of squares of gradient entries (recall Section 2.2). The matrix S is initialized as diag(δ 2 ), where δj ≥ 0 is an initial value. Each processor asynchronously performs the following iterations: 1 1 1. Read S and z and set G = S 2 . Compute x := argminx∈X { z, x + 2η x, Gx } increment “time” counter t and let x(t) = x, S(t) = S 2. Sample ξ ∼ P and let g ∈ ∂F (x; ξ) 2 3. For j ∈ [d] such that gj = 0, update Sj ← Sj + gj and zj ← zj + gj . 5 // Implicitly As in the description of A SYNC DA, we note that x(t) is the vector x ∈ X computed in the tth “step” of the algorithm (step 1), and similarly associate ξ t with x(t). To analyze A SYNC A DAG RAD, we make a somewhat stronger assumption on the sparsity properties of the losses F than Assumption B. 2 Assumption C. There exist constants (Mj )d such that E[(∂xj F (x; ξ))2 | ξj = 0] ≤ Mj for all j=1 x ∈ X. 2 Indeed, taking M2 = j pj Mj shows that Assumption C implies Assumption B with specific constants. We then have the following convergence result. Theorem 5. In addition to the conditions of Theorem 3, let Assumption C hold. Assume that for all 2 j we have δ 2 ≥ Mj m and X ⊂ [−R∞ , R∞ ]d . Then T t=1 E F (x(t); ξ t ) − F (x∗ ; ξ t ) d ≤ min j=1 T 1 2 R E η ∞ 2 δ + gj (t) 2 1 2 T + ηE gj (t) t=1 2 1 2 (1 + pj m), Mj R∞ pj T . t=1 It is possible to relax the condition on the initial constant diagonal term; we defer this to the full version of the paper. It is natural to ask in which situations the bound provided by Theorem 5 is optimal. We note that, as in the case with Theorem 3, we may obtain a convergence rate for f (x(T )) − f (x∗ ) using convexity, T 1 where x(T ) = T t=1 x(t). By Jensen’s inequality, we have for any δ that T E 2 δ + gj (t) 2 1 2 t=1 T ≤ 2 2 E[gj (t) ] δ + t=1 1 2 ≤ 2 δ 2 + T pj Mj . For interpretation, let us now make a few assumptions on the probabilities pj . If we assume that pj ≤ c/m for a universal (numerical) constant c, then Theorem 5 guarantees that d log(T )/T + pj √ (8) , pj , T j=1 √ which is the convergence rate of A DAG RAD except for a small factor of min{ log T /T, pj } in addition to the usual pj /T rate. In particular, optimizing by choosing η = R∞ , and assuming 1 pj T log T , we have convergence guarantee √ d pj E[f (x(T )) − f (x∗ )] ≤ O(1)R∞ Mj min √ , pj , T j=1 E[f (x(T )) − f (x∗ )] ≤ O(1) 1 2 R +η η ∞ Mj min which is minimax optimal by Proposition 1. In fact, however, the bounds of Theorem 5 are somewhat stronger: they provide bounds using the expectation of the squared gradients gj (t) rather than the maximal value Mj , though the bounds are perhaps clearer in the form (8). We note also that our analysis applies to more adversarial settings than stochastic optimization (e.g., to online convex optimization [5]). Specifically, an adversary may choose an arbitrary sequence of functions subject to the random data sparsity constraint (2), and our results provide an expected regret bound, which is strictly stronger than the stochastic convergence guarantees provided (and guarantees high-probability convergence in stochastic settings [3]). Moreover, our comments in Section 2 about the relative optimality of A DAG RAD versus standard gradient methods apply. When the data is sparse, we indeed should use asynchronous algorithms, but using adaptive methods yields even more improvement than simple gradient-based methods. 4 Experiments In this section, we give experimental validation of our theoretical results on A SYNC A DAG RAD and A SYNC DA, giving results on two datasets selected for their high-dimensional sparsity.2 2 In our experiments, A SYNC DA and H OGWILD ! had effectively identical performance. 6 8 0.07 6 5 4 0.024 Test error Training loss Speedup 0.025 0.065 7 0.06 0.055 0.05 0.045 0.04 0.023 0.022 0.021 0.02 0.035 3 0.019 2 1 2 4 0.03 A-A DAG RAD A SYNC DA Number of workers 6 8 10 12 14 0.018 0.025 0.02 16 2 4 6 8 10 12 14 Number of workers 0.017 16 2 4 6 8 10 12 14 Number of workers 16 Figure 1. Experiments with URL data. Left: speedup relative to one processor. Middle: training dataset loss versus number of processors. Right: test set error rate versus number of processors. AA DAG RAD abbreviates A SYNC A DAG RAD. 1.03 1.02 1.01 1.00 1.0 1 2 4 8 16 64 256 number of passes A-AdaGrad, η = 0.008 L2 = 0 A-AdaGrad, η = 0.008 L2 = 80 A-DA, η = 0.8 L2 = 0 A-DA, η = 0.8 L2 = 80 1.00 1.01 1.4 1.02 1.03 1.04 Impact of L2 regularizaton on test error 1.04 Fixed stepsizes, test data, L2=0 1.2 relative log-loss 1.6 1.8 Fixed stepsizes, training data, L2=0 A-AdaGrad η = 0.002 A-AdaGrad η = 0.004 A-AdaGrad η = 0.008 A-AdaGrad η = 0.016 A-DA η = 0.800 A-DA η = 1.600 A-DA η = 3.200 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 2: Relative accuracy for various stepsize choices on an click-through rate prediction dataset. 4.1 Malicious URL detection For our first set of experiments, we consider the speedup attainable by applying A SYNC A DAG RAD and A SYNC DA, investigating the performance of each algorithm on a malicious URL prediction task [7]. The dataset in this case consists of an anonymized collection of URLs labeled as malicious (e.g., spam, phishing, etc.) or benign over a span of 120 days. The data in this case consists of 2.4 · 106 examples with dimension d = 3.2 · 106 (sparse) features. We perform several experiments, randomly dividing the dataset into 1.2 · 106 training and test samples for each experiment. In Figure 1 we compare the performance of A SYNC A DAG RAD and A SYNC DA after doing after single pass through the training dataset. (For each algorithm, we choose the stepsize η for optimal training set performance.) We perform the experiments on a single machine running Ubuntu Linux with six cores (with two-way hyperthreading) and 32Gb of RAM. From the left-most plot in Fig. 1, we see that up to six processors, both A SYNC DA and A SYNC A DAG RAD enjoy the expected linear speedup, and from 6 to 12, they continue to enjoy a speedup that is linear in the number of processors though at a lesser slope (this is the effect of hyperthreading). For more than 12 processors, there is no further benefit to parallelism on this machine. The two right plots in Figure 1 plot performance of the different methods (with standard errors) versus the number of worker threads used. Both are essentially flat; increasing the amount of parallelism does nothing to the average training loss or the test error rate for either method. It is clear, however, that for this dataset, the adaptive A SYNC A DAG RAD algorithm provides substantial performance benefits over A SYNC DA. 4.2 Click-through-rate prediction experiments We also experiment on a proprietary datasets consisting of search ad impressions. Each example corresponds to showing a search-engine user a particular text ad in response to a query string. From this, we construct a very sparse feature vector based on the text of the ad displayed and the query string (no user-specific data is used). The target label is 1 if the user clicked the ad and -1 otherwise. 7 (B) A-AdaGrad speedup (D) Impact of training data ordering 1.004 1.005 1.006 1.007 1.008 1 2 4 8 16 32 number of passes 64 128 256 1.000 1 2 A-DA base η = 1.600 A-AdaGrad base η = 0.023 0 1.005 relative stepsize (C) Optimal stepsize scaling relative log-loss 1.003 target relative log-loss 1.005 1.010 1.002 1.010 1.015 8 4 0 speedup A-DA η = 1.600 A-AdaGrad η = 0.016 1.001 1.000 relative log-loss 1.015 A-DA, L2=80 A-AdaGrad, L2=80 12 (A) Optimized stepsize for each number of passes 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 3. (A) Relative test-set log-loss for A SYNC DA and A SYNC A DAG RAD, choosing the best stepsize (within a factor of about 1.4×) individually for each number of passes. (B) Effective speedup for A SYNC A DAG RAD. (C) The best stepsize η, expressed as a scaling factor on the stepsize used for one pass. (D) Five runs with different random seeds for each algorithm (with 2 penalty 80). We fit logistic regression models using both A SYNC DA and A SYNC A DAG RAD. We run extensive experiments on a moderate-sized dataset (about 107 examples, split between training and testing), which allows thorough investigation of the impact of the stepsize η, the number of training passes,3 and 2 -regularization on accuracy. For these experiments we used 32 threads on 16 core machines for each run, as A SYNC A DAG RAD and A SYNC DA achieve similar speedups from parallelization. On this dataset, A SYNC A DAG RAD typically achieves an effective additional speedup over A SYNC DA of 4× or more. That is, to reach a given level of accuracy, A SYNC DA generally needs four times as many effective passes over the dataset. We measure accuracy with log-loss (the logistic loss) averaged over five runs using different random seeds (which control the order in which the algorithms sample examples during training). We report relative values in Figures 2 and 3, that is, the ratio of the mean loss for the given datapoint to the lowest (best) mean loss obtained. Our results are not particularly sensitive to the choice of relative log-loss as the metric of interest; we also considered AUC (the area under the ROC curve) and observed similar results. Figure 2 shows relative log-loss as a function of the number of training passes for various stepsizes. Without regularization, A SYNC A DAG RAD is prone to overfitting: it achieves significantly higher accuracy on the training data (Fig. 2 (left)), but unless the stepsize is tuned carefully to the number of passes, it will overfit (Fig. 2 (middle)). Fortunately, the addition of 2 regularization largely solves this problem. Indeed, Figure 2 (right) shows that while adding an 2 penalty of 80 has very little impact on A SYNC DA, it effectively prevents the overfitting of A SYNC A DAG RAD.4 Fixing √ regularization multiplier to 80, we varied the stepsize η over a multiplicative grid with res2 olution 2 for each number of passes and for each algorithm. Figure 3 reports the results obtained by selecting the best stepsize in terms of test set log-loss for each number of passes. Figure 3(A) shows relative log-loss of the best stepsize for each algorithm; 3(B) shows the relative time A SYNC DA requires with respect to A SYNC A DAG RAD to achieve a given loss. Specifically, Fig. 3(B) shows the ratio of the number of passes the algorithms require to achieve a fixed loss, which gives a broader estimate of the speedup obtained by using A SYNC A DAG RAD; speedups range from 3.6× to 12×. Figure 3(C) shows the optimal stepsizes as a function of the best setting for one pass. The optimal stepsize decreases moderately for A SYNC A DAG RAD, but are somewhat noisy for A SYNC DA. It is interesting to note that A SYNC A DAG RAD’s accuracy is largely independent of the ordering of the training data, while A SYNC DA shows significant variability. This can be seen both in the error bars on Figure 3(A), and explicitly in Figure 3(D), where we plot one line for each of the five random seeds used. Thus, while on the one hand A SYNC DA requires somewhat less tuning of the stepsize and 2 parameter, tuning A SYNC A DAG RAD is much easier because of its predictable response. 3 Here “number of passes” more precisely means the expected number of times each example in the dataset is trained on. That is, each worker thread randomly selects a training example from the dataset for each update, and we continued making updates until (dataset size) × (number of passes) updates have been processed. 4 For both algorithms, this is accomplished by adding the term η80 x 2 to the ψ function. We can achieve 2 slightly better results for A SYNC A DAG RAD by varying the 2 penalty with the number of passes. 8 References [1] P. Auer and C. Gentile. Adaptive and self-confident online learning algorithms. In Proceedings of the Thirteenth Annual Conference on Computational Learning Theory, 2000. [2] P. B¨ hlmann and S. van de Geer. Statistics for High-Dimensional Data: Methods, Theory and u Applications. Springer, 2011. [3] N. Cesa-Bianchi, A. Conconi, and C. Gentile. On the generalization ability of on-line learning algorithms. IEEE Transactions on Information Theory, 50(9):2050–2057, September 2004. [4] J. C. Duchi, E. Hazan, and Y. Singer. Adaptive subgradient methods for online learning and stochastic optimization. Journal of Machine Learning Research, 12:2121–2159, 2011. [5] E. Hazan. The convex optimization approach to regret minimization. In Optimization for Machine Learning, chapter 10. MIT Press, 2012. [6] J. Hiriart-Urruty and C. Lemar´ chal. Convex Analysis and Minimization Algorithms I & II. e Springer, New York, 1996. [7] J. Ma, L. K. Saul, S. Savage, and G. M. Voelker. Identifying malicious urls: An application of large-scale online learning. In Proceedings of the 26th International Conference on Machine Learning, 2009. [8] C. Manning and H. Sch¨ tze. Foundations of Statistical Natural Language Processing. MIT u Press, 1999. [9] B. McMahan and M. Streeter. Adaptive bound optimization for online convex optimization. In Proceedings of the Twenty Third Annual Conference on Computational Learning Theory, 2010. [10] A. Nemirovski, A. Juditsky, G. Lan, and A. Shapiro. Robust stochastic approximation approach to stochastic programming. SIAM Journal on Optimization, 19(4):1574–1609, 2009. [11] Y. Nesterov. Primal-dual subgradient methods for convex problems. Mathematical Programming, 120(1):261–283, 2009. [12] F. Niu, B. Recht, C. R´ , and S. Wright. Hogwild: a lock-free approach to parallelizing stochase tic gradient descent. In Advances in Neural Information Processing Systems 24, 2011. [13] P. Richt´ rik and M. Tak´ c. Parallel coordinate descent methods for big data optimization. a aˇ arXiv:1212.0873 [math.OC], 2012. URL http://arxiv.org/abs/1212.0873. [14] M. Tak´ c, A. Bijral, P. Richt´ rik, and N. Srebro. Mini-batch primal and dual methods for aˇ a SVMs. In Proceedings of the 30th International Conference on Machine Learning, 2013. 9
2 0.69153881 19 nips-2013-Accelerated Mini-Batch Stochastic Dual Coordinate Ascent
Author: Shai Shalev-Shwartz, Tong Zhang
Abstract: Stochastic dual coordinate ascent (SDCA) is an effective technique for solving regularized loss minimization problems in machine learning. This paper considers an extension of SDCA under the mini-batch setting that is often used in practice. Our main contribution is to introduce an accelerated mini-batch version of SDCA and prove a fast convergence rate for this method. We discuss an implementation of our method over a parallel computing system, and compare the results to both the vanilla stochastic dual coordinate ascent and to the accelerated deterministic gradient descent method of Nesterov [2007]. 1
3 0.67256474 198 nips-2013-More Effective Distributed ML via a Stale Synchronous Parallel Parameter Server
Author: Qirong Ho, James Cipar, Henggang Cui, Seunghak Lee, Jin Kyu Kim, Phillip B. Gibbons, Garth A. Gibson, Greg Ganger, Eric Xing
Abstract: We propose a parameter server system for distributed ML, which follows a Stale Synchronous Parallel (SSP) model of computation that maximizes the time computational workers spend doing useful work on ML algorithms, while still providing correctness guarantees. The parameter server provides an easy-to-use shared interface for read/write access to an ML model’s values (parameters and variables), and the SSP model allows distributed workers to read older, stale versions of these values from a local cache, instead of waiting to get them from a central storage. This significantly increases the proportion of time workers spend computing, as opposed to waiting. Furthermore, the SSP model ensures ML algorithm correctness by limiting the maximum age of the stale values. We provide a proof of correctness under SSP, as well as empirical results demonstrating that the SSP model achieves faster algorithm convergence on several different ML problems, compared to fully-synchronous and asynchronous schemes. 1
4 0.59437966 333 nips-2013-Trading Computation for Communication: Distributed Stochastic Dual Coordinate Ascent
Author: Tianbao Yang
Abstract: We present and study a distributed optimization algorithm by employing a stochastic dual coordinate ascent method. Stochastic dual coordinate ascent methods enjoy strong theoretical guarantees and often have better performances than stochastic gradient descent methods in optimizing regularized loss minimization problems. It still lacks of efforts in studying them in a distributed framework. We make a progress along the line by presenting a distributed stochastic dual coordinate ascent algorithm in a star network, with an analysis of the tradeoff between computation and communication. We verify our analysis by experiments on real data sets. Moreover, we compare the proposed algorithm with distributed stochastic gradient descent methods and distributed alternating direction methods of multipliers for optimizing SVMs in the same distributed framework, and observe competitive performances. 1
5 0.57264978 245 nips-2013-Pass-efficient unsupervised feature selection
Author: Crystal Maung, Haim Schweitzer
Abstract: The goal of unsupervised feature selection is to identify a small number of important features that can represent the data. We propose a new algorithm, a modification of the classical pivoted QR algorithm of Businger and Golub, that requires a small number of passes over the data. The improvements are based on two ideas: keeping track of multiple features in each pass, and skipping calculations that can be shown not to affect the final selection. Our algorithm selects the exact same features as the classical pivoted QR algorithm, and has the same favorable numerical stability. We describe experiments on real-world datasets which sometimes show improvements of several orders of magnitude over the classical algorithm. These results appear to be competitive with recently proposed randomized algorithms in terms of pass efficiency and run time. On the other hand, the randomized algorithms may produce more accurate features, at the cost of small probability of failure. 1
6 0.53147632 345 nips-2013-Variance Reduction for Stochastic Gradient Optimization
7 0.51327395 20 nips-2013-Accelerating Stochastic Gradient Descent using Predictive Variance Reduction
8 0.50017715 313 nips-2013-Stochastic Majorization-Minimization Algorithms for Large-Scale Optimization
10 0.46399498 318 nips-2013-Structured Learning via Logistic Regression
11 0.4590165 317 nips-2013-Streaming Variational Bayes
12 0.44225666 45 nips-2013-BIG & QUIC: Sparse Inverse Covariance Estimation for a Million Variables
13 0.43701214 214 nips-2013-On Algorithms for Sparse Multi-factor NMF
14 0.42953983 134 nips-2013-Graphical Models for Inference with Missing Data
15 0.4250333 94 nips-2013-Distributed $k$-means and $k$-median Clustering on General Topologies
16 0.41686863 90 nips-2013-Direct 0-1 Loss Minimization and Margin Maximization with Boosting
17 0.41501474 275 nips-2013-Reservoir Boosting : Between Online and Offline Ensemble Learning
18 0.41334546 265 nips-2013-Reconciling "priors" & "priors" without prejudice?
19 0.41167381 146 nips-2013-Large Scale Distributed Sparse Precision Estimation
20 0.41129276 82 nips-2013-Decision Jungles: Compact and Rich Models for Classification
topicId topicWeight
[(2, 0.029), (16, 0.058), (33, 0.153), (34, 0.062), (41, 0.031), (49, 0.027), (56, 0.123), (70, 0.032), (85, 0.032), (88, 0.235), (89, 0.032), (93, 0.06), (95, 0.021)]
simIndex simValue paperId paperTitle
same-paper 1 0.80353099 111 nips-2013-Estimation, Optimization, and Parallelism when Data is Sparse
Author: John Duchi, Michael Jordan, Brendan McMahan
Abstract: We study stochastic optimization problems when the data is sparse, which is in a sense dual to current perspectives on high-dimensional statistical learning and optimization. We highlight both the difficulties—in terms of increased sample complexity that sparse data necessitates—and the potential benefits, in terms of allowing parallelism and asynchrony in the design of algorithms. Concretely, we derive matching upper and lower bounds on the minimax rate for optimization and learning with sparse data, and we exhibit algorithms achieving these rates. We also show how leveraging sparsity leads to (still minimax optimal) parallel and asynchronous algorithms, providing experimental evidence complementing our theoretical results on several medium to large-scale learning tasks. 1 Introduction and problem setting In this paper, we investigate stochastic optimization problems in which the data is sparse. Formally, let {F (·; ξ), ξ ∈ Ξ} be a collection of real-valued convex functions, each of whose domains contains the convex set X ⊂ Rd . For a probability distribution P on Ξ, we consider the following optimization problem: minimize f (x) := E[F (x; ξ)] = x∈X F (x; ξ)dP (ξ). (1) Ξ By data sparsity, we mean the samples ξ are sparse: assuming that samples ξ lie in Rd , and defining the support supp(x) of a vector x to the set of indices of its non-zero components, we assume supp F (x; ξ) ⊂ supp ξ. (2) The sparsity condition (2) means that F (x; ξ) does not “depend” on the values of xj for indices j such that ξj = 0.1 This type of data sparsity is prevalent in statistical optimization problems and machine learning applications; in spite of its prevalence, study of such problems has been limited. As a motivating example, consider a text classification problem: data ξ ∈ Rd represents words appearing in a document, and we wish to minimize a logistic loss F (x; ξ) = log(1 + exp( ξ, x )) on the data (we encode the label implicitly with the sign of ξ). Such generalized linear models satisfy the sparsity condition (2), and while instances are of very high dimension, in any given instance, very few entries of ξ are non-zero [8]. From a modelling perspective, it thus makes sense to allow a dense predictor x: any non-zero entry of ξ is potentially relevant and important. In a sense, this is dual to the standard approaches to high-dimensional problems; one usually assumes that the data ξ may be dense, but there are only a few relevant features, and thus a parsimonious model x is desirous [2]. So 1 Formally, if πξ denotes the coordinate projection zeroing all indices j of its argument where ξj = 0, then F (πξ (x); ξ) = F (x; ξ) for all x, ξ. This follows from the first-order conditions for convexity [6]. 1 while such sparse data problems are prevalent—natural language processing, information retrieval, and other large data settings all have significant data sparsity—they do not appear to have attracted as much study as their high-dimensional “duals” of dense data and sparse predictors. In this paper, we investigate algorithms and their inherent limitations for solving problem (1) under natural conditions on the data generating distribution. Recent work in the optimization and machine learning communities has shown that data sparsity can be leveraged to develop parallel (and even asynchronous [12]) optimization algorithms [13, 14], but this work does not consider the statistical effects of data sparsity. In another line of research, Duchi et al. [4] and McMahan and Streeter [9] develop “adaptive” stochastic gradient algorithms to address problems in sparse data regimes (2). These algorithms exhibit excellent practical performance and have theoretical guarantees on their convergence, but it is not clear if they are optimal—in that no algorithm can attain better statistical performance—or whether they can leverage parallel computing as in the papers [12, 14]. In this paper, we take a two-pronged approach. First, we investigate the fundamental limits of optimization and learning algorithms in sparse data regimes. In doing so, we derive lower bounds on the optimization error of any algorithm for problems of the form (1) with sparsity condition (2). These results have two main implications. They show that in some scenarios, learning with sparse data is quite difficult, as essentially each coordinate j ∈ [d] can be relevant and must be optimized for. In spite of this seemingly negative result, we are also able to show that the A DAG RAD algorithms of [4, 9] are optimal, and we show examples in which their dependence on the dimension d can be made exponentially better than standard gradient methods. As the second facet of our two-pronged approach, we study how sparsity may be leveraged in parallel computing frameworks to give substantially faster algorithms that still achieve optimal sample complexity in terms of the number of samples ξ used. We develop two new algorithms, asynchronous dual averaging (A SYNC DA) and asynchronous A DAG RAD (A SYNC A DAG RAD), which allow asynchronous parallel solution of the problem (1) for general convex f and X . Combining insights of Niu et al.’s H OGWILD ! [12] with a new analysis, we prove our algorithms achieve linear speedup in the number of processors while maintaining optimal statistical guarantees. We also give experiments on text-classification and web-advertising tasks to illustrate the benefits of the new algorithms. 2 Minimax rates for sparse optimization We begin our study of sparse optimization problems by establishing their fundamental statistical and optimization-theoretic properties. To do this, we derive bounds on the minimax convergence rate of any algorithm for such problems. Formally, let x denote any estimator for a minimizer of the objective (1). We define the optimality gap N for the estimator x based on N samples ξ 1 , . . . , ξ N from the distribution P as N (x, F, X , P ) := f (x) − inf f (x) = EP [F (x; ξ)] − inf EP [F (x; ξ)] . x∈X x∈X This quantity is a random variable, since x is a random variable (it is a function of ξ 1 , . . . , ξ N ). To define the minimax error, we thus take expectations of the quantity N , though we require a bit more than simply E[ N ]. We let P denote a collection of probability distributions, and we consider a collection of loss functions F specified by a collection F of convex losses F : X × ξ → R. We can then define the minimax error for the family of losses F and distributions P as ∗ N (X , P, F) := inf sup sup EP [ x P ∈P F ∈F N (x(ξ 1:N ), F, X , P )], (3) where the infimum is taken over all possible estimators x (an estimator is an optimization scheme, or a measurable mapping x : ΞN → X ) . 2.1 Minimax lower bounds Let us now give a more precise characterization of the (natural) set of sparse optimization problems we consider to provide the lower bound. For the next proposition, we let P consist of distributions supported on Ξ = {−1, 0, 1}d , and we let pj := P (ξj = 0) be the marginal probability of appearance of feature j ∈ {1, . . . , d}. For our class of functions, we set F to consist of functions F satisfying the sparsity condition (2) and with the additional constraint that for g ∈ ∂x F (x; ξ), we have that the jth coordinate |gj | ≤ Mj for a constant Mj < ∞. We obtain 2 Proposition 1. Let the conditions of the preceding paragraph hold. Let R be a constant such that X ⊃ [−R, R]d . Then √ d pj 1 ∗ . Mj min pj , √ N (X , P, F) ≥ R 8 j=1 N log 3 We provide the proof of Proposition 1 in the supplement A.1 in the full version of the paper, providing a few remarks here. We begin by giving a corollary to Proposition 1 that follows when the data ξ obeys a type of power law: let p0 ∈ [0, 1], and assume that P (ξj = 0) = p0 j −α . We have Corollary 2. Let α ≥ 0. Let the conditions of Proposition 1 hold with Mj ≡ M for all j, and assume the power law condition P (ξj = 0) = p0 j −α on coordinate appearance probabilities. Then (1) If d > (p0 N )1/α , ∗ N (X , P, F) ≥ 2−α 1−α p0 p0 (p0 N ) 2α − 1 + d1−α − (p0 N ) α N 1−α 2 MR 8 2−α (2) If d ≤ (p0 N )1/α , ∗ N (X , P, F) ≥ MR 8 p0 N α 1 1 d1− 2 − 1 − α/2 1 − α/2 . . Expanding Corollary 2 slightly, for simplicity assume the number of samples is large enough that d ≤ (p0 N )1/α . Then we find that the lower bound on optimization error is of order p0 1− α p0 p0 d 2 when α < 2, M R log d when α → 2, and M R when α > 2. (4) N N N These results beg the question of tightness: are they improvable? As we see presently, they are not. MR 2.2 Algorithms for attaining the minimax rate To show that the lower bounds of Proposition 1 and its subsequent specializations are sharp, we review a few stochastic gradient algorithms. We begin with stochastic gradient descent (SGD): SGD repeatedly samples ξ ∼ P , computes g ∈ ∂x F (x; ξ), then performs the update x ← ΠX (x − ηg), where η is a stepsize parameter and ΠX denotes Euclidean projection onto X . Standard analyses of stochastic gradient descent [10] show that after N samples ξ i , the SGD estimator x(N ) satisfies R2 M ( d j=1 1 pj ) 2 √ , (5) N where R2 denotes the 2 -radius of X . Dual averaging, due to Nesterov [11] (sometimes called “follow the regularized leader” [5]) is a more recent algorithm. In dual averaging, one again samples g ∈ ∂x F (x; ξ), but instead of updating the parameter vector x one updates a dual vector z by z ← z + g, then computes 1 x ← argmin z, x + ψ(x) , η x∈X E[f (x(N ))] − inf f (x) ≤ O(1) x∈X 2 1 where ψ(x) is a strongly convex function defined over X (often one takes ψ(x) = 2 x 2 ). As we discuss presently, the dual averaging algorithm is somewhat more natural in asynchronous and parallel computing environments, and it enjoys the same type of convergence guarantees (5) as SGD. The A DAG RAD algorithm [4, 9] is an extension of the preceding stochastic gradient methods. It maintains a diagonal matrix S, where upon receiving a new sample ξ, A DAG RAD performs the following: it computes g ∈ ∂x F (x; ξ), then updates 2 Sj ← Sj + gj for j ∈ [d]. The dual averaging variant of A DAG RAD updates the usual dual vector z ← z + g; the update to x is based on S and a stepsize η and computes x ← argmin z, x + x ∈X 3 1 1 x ,S2x 2η . After N samples ξ, the averaged parameter x(N ) returned by A DAG RAD satisfies R∞ M E[f (x(N ))] − inf f (x) ≤ O(1) √ x∈X N d √ pj , (6) j=1 where R∞ denotes the ∞ -radius of X (cf. [4, Section 1.3 and Theorem 5]). By inspection, the A DAG RAD rate (6) matches the lower bound in Proposition 1 and is thus optimal. It is interesting to note, though, that in the power law setting of Corollary 2 (recall the error order (4)), a calculation √ shows that the multiplier for the SGD guarantee (5) becomes R∞ d max{d(1−α)/2 , 1}, while A DA G RAD attains rate at worst R∞ max{d1−α/2 , log d}. For α > 1, the A DAG RAD rate is no worse, √ and for α ≥ 2, is more than d/ log d better—an exponential improvement in the dimension. 3 Parallel and asynchronous optimization with sparsity As we note in the introduction, recent works [12, 14] have suggested that sparsity can yield benefits in our ability to parallelize stochastic gradient-type algorithms. Given the optimality of A DAG RADtype algorithms, it is natural to focus on their parallelization in the hope that we can leverage their ability to “adapt” to sparsity in the data. To provide the setting for our further algorithms, we first revisit Niu et al.’s H OGWILD ! [12]. H OGWILD ! is an asynchronous (parallelized) stochastic gradient algorithm for optimization over product-space domains, meaning that X in problem (1) decomposes as X = X1 × · · · × Xd , where Xj ⊂ R. Fix a stepsize η > 0. A pool of independently running processors then performs the following updates asynchronously to a centralized vector x: 1. Sample ξ ∼ P 2. Read x and compute g ∈ ∂x F (x; ξ) 3. For each j s.t. gj = 0, update xj ← ΠXj (xj − ηgj ). Here ΠXj denotes projection onto the jth coordinate of the domain X . The key of H OGWILD ! is that in step 2, the parameter x is allowed to be inconsistent—it may have received partial gradient updates from many processors—and for appropriate problems, this inconsistency is negligible. Indeed, Niu et al. [12] show linear speedup in optimization time as the number of processors grow; they show this empirically in many scenarios, providing a proof under the somewhat restrictive assumptions that there is at most one non-zero entry in any gradient g and that f has Lipschitz gradients. 3.1 Asynchronous dual averaging A weakness of H OGWILD ! is that it appears only applicable to problems for which the domain X is a product space, and its analysis assumes g 0 = 1 for all gradients g. In effort to alleviate these difficulties, we now develop and present our asynchronous dual averaging algorithm, A SYNC DA. A SYNC DA maintains and upates a centralized dual vector z instead of a parameter x, and a pool of processors perform asynchronous updates to z, where each processor independently iterates: 1. Read z and compute x := argminx∈X 1 z, x + η ψ(x) // Implicitly increment “time” counter t and let x(t) = x 2. Sample ξ ∼ P and let g ∈ ∂x F (x; ξ) // Let g(t) = g. 3. For j ∈ [d] such that gj = 0, update zj ← zj + gj . Because the actual computation of the vector x in A SYNC DA is performed locally on each processor in step 1 of the algorithm, the algorithm can be executed with any proximal function ψ and domain X . The only communication point between any of the processors is the addition operation in step 3. Since addition is commutative and associative, forcing all asynchrony to this point of the algorithm is a natural strategy for avoiding synchronization problems. In our analysis of A SYNC DA, and in our subsequent analysis of the adaptive methods, we require a measurement of time elapsed. With that in mind, we let t denote a time index that exists (roughly) behind-the-scenes. We let x(t) denote the vector x ∈ X computed in the tth step 1 of the A SYNC DA 4 algorithm, that is, whichever is the tth x actually computed by any of the processors. This quantity t exists and is recoverable from the algorithm, and it is possible to track the running sum τ =1 x(τ ). Additionally, we state two assumptions encapsulating the conditions underlying our analysis. Assumption A. There is an upper bound m on the delay of any processor. In addition, for each j ∈ [d] there is a constant pj ∈ [0, 1] such that P (ξj = 0) ≤ pj . We also require certain continuity (Lipschitzian) properties of the loss functions; these amount to a second moment constraint on the instantaneous ∂F and a rough measure of gradient sparsity. Assumption B. There exist constants M and (Mj )d such that the following bounds hold for all j=1 2 x ∈ X : E[ ∂x F (x; ξ) 2 ] ≤ M2 and for each j ∈ [d] we have E[|∂xj F (x; ξ)|] ≤ pj Mj . With these definitions, we have the following theorem, which captures the convergence behavior of A SYNC DA under the assumption that X is a Cartesian product, meaning that X = X1 × · · · × Xd , 2 where Xj ⊂ R, and that ψ(x) = 1 x 2 . Note the algorithm itself can still be efficiently parallelized 2 for more general convex X , even if the theorem does not apply. Theorem 3. Let Assumptions A and B and the conditions in the preceding paragraph hold. Then T E t=1 F (x(t); ξ t ) − F (x∗ ; ξ t ) ≤ 1 x∗ 2η d 2 2 η 2 p2 Mj . + T M2 + ηT m j 2 j=1 We now provide a few remarks to explain and simplify the result. Under the more stringent condition 2 d 2 that |∂xj F (x; ξ)| ≤ Mj , Assumption A implies E[ ∂x F (x; ξ) 2 ] ≤ j=1 pj Mj . Thus, for the d 2 remainder of this section we take M2 = j=1 pj Mj , which upper bounds the Lipschitz continuity constant of the objective function f . We then obtain the following corollary. √ T 1 Corollary 4. Define x(T ) = T t=1 x(t), and set η = x∗ 2 /M T . Then E[f (x(T )) − f (x∗ )] ≤ M x∗ √ T 2 +m x∗ 2 √ 2M T d 2 p2 M j . j j=1 Corollary 4 is nearly immediate: since ξ t is independent of x(t), we have E[F (x(t); ξ t ) | x(t)] = f (x(t)); applying Jensen’s inequality to f (x(T )) and performing an algebraic manipulation give the result. If the data is suitably sparse, meaning that pj ≤ 1/m, the bound in Corollary 4 simplifies to 3 M x∗ √ E[f (x(T )) − f (x )] ≤ 2 T ∗ 2 3 = 2 d j=1 2 pj M j x ∗ √ T 2 , (7) which is the convergence rate of stochastic gradient descent even in centralized settings (5). The √ convergence guarantee (7) shows that after T timesteps, the error scales as 1/ T ; however, if we have k processors, updates occur roughly k times as quickly, as they are asynchronous, and in time scaling as N/k, we can evaluate N gradient samples: a linear speedup. 3.2 Asynchronous AdaGrad We now turn to extending A DAG RAD to asynchronous settings, developing A SYNC A DAG RAD (asynchronous A DAG RAD). As in the A SYNC DA algorithm, A SYNC A DAG RAD maintains a shared dual vector z (the sum of gradients) and the shared matrix S, which is the diagonal sum of squares of gradient entries (recall Section 2.2). The matrix S is initialized as diag(δ 2 ), where δj ≥ 0 is an initial value. Each processor asynchronously performs the following iterations: 1 1 1. Read S and z and set G = S 2 . Compute x := argminx∈X { z, x + 2η x, Gx } increment “time” counter t and let x(t) = x, S(t) = S 2. Sample ξ ∼ P and let g ∈ ∂F (x; ξ) 2 3. For j ∈ [d] such that gj = 0, update Sj ← Sj + gj and zj ← zj + gj . 5 // Implicitly As in the description of A SYNC DA, we note that x(t) is the vector x ∈ X computed in the tth “step” of the algorithm (step 1), and similarly associate ξ t with x(t). To analyze A SYNC A DAG RAD, we make a somewhat stronger assumption on the sparsity properties of the losses F than Assumption B. 2 Assumption C. There exist constants (Mj )d such that E[(∂xj F (x; ξ))2 | ξj = 0] ≤ Mj for all j=1 x ∈ X. 2 Indeed, taking M2 = j pj Mj shows that Assumption C implies Assumption B with specific constants. We then have the following convergence result. Theorem 5. In addition to the conditions of Theorem 3, let Assumption C hold. Assume that for all 2 j we have δ 2 ≥ Mj m and X ⊂ [−R∞ , R∞ ]d . Then T t=1 E F (x(t); ξ t ) − F (x∗ ; ξ t ) d ≤ min j=1 T 1 2 R E η ∞ 2 δ + gj (t) 2 1 2 T + ηE gj (t) t=1 2 1 2 (1 + pj m), Mj R∞ pj T . t=1 It is possible to relax the condition on the initial constant diagonal term; we defer this to the full version of the paper. It is natural to ask in which situations the bound provided by Theorem 5 is optimal. We note that, as in the case with Theorem 3, we may obtain a convergence rate for f (x(T )) − f (x∗ ) using convexity, T 1 where x(T ) = T t=1 x(t). By Jensen’s inequality, we have for any δ that T E 2 δ + gj (t) 2 1 2 t=1 T ≤ 2 2 E[gj (t) ] δ + t=1 1 2 ≤ 2 δ 2 + T pj Mj . For interpretation, let us now make a few assumptions on the probabilities pj . If we assume that pj ≤ c/m for a universal (numerical) constant c, then Theorem 5 guarantees that d log(T )/T + pj √ (8) , pj , T j=1 √ which is the convergence rate of A DAG RAD except for a small factor of min{ log T /T, pj } in addition to the usual pj /T rate. In particular, optimizing by choosing η = R∞ , and assuming 1 pj T log T , we have convergence guarantee √ d pj E[f (x(T )) − f (x∗ )] ≤ O(1)R∞ Mj min √ , pj , T j=1 E[f (x(T )) − f (x∗ )] ≤ O(1) 1 2 R +η η ∞ Mj min which is minimax optimal by Proposition 1. In fact, however, the bounds of Theorem 5 are somewhat stronger: they provide bounds using the expectation of the squared gradients gj (t) rather than the maximal value Mj , though the bounds are perhaps clearer in the form (8). We note also that our analysis applies to more adversarial settings than stochastic optimization (e.g., to online convex optimization [5]). Specifically, an adversary may choose an arbitrary sequence of functions subject to the random data sparsity constraint (2), and our results provide an expected regret bound, which is strictly stronger than the stochastic convergence guarantees provided (and guarantees high-probability convergence in stochastic settings [3]). Moreover, our comments in Section 2 about the relative optimality of A DAG RAD versus standard gradient methods apply. When the data is sparse, we indeed should use asynchronous algorithms, but using adaptive methods yields even more improvement than simple gradient-based methods. 4 Experiments In this section, we give experimental validation of our theoretical results on A SYNC A DAG RAD and A SYNC DA, giving results on two datasets selected for their high-dimensional sparsity.2 2 In our experiments, A SYNC DA and H OGWILD ! had effectively identical performance. 6 8 0.07 6 5 4 0.024 Test error Training loss Speedup 0.025 0.065 7 0.06 0.055 0.05 0.045 0.04 0.023 0.022 0.021 0.02 0.035 3 0.019 2 1 2 4 0.03 A-A DAG RAD A SYNC DA Number of workers 6 8 10 12 14 0.018 0.025 0.02 16 2 4 6 8 10 12 14 Number of workers 0.017 16 2 4 6 8 10 12 14 Number of workers 16 Figure 1. Experiments with URL data. Left: speedup relative to one processor. Middle: training dataset loss versus number of processors. Right: test set error rate versus number of processors. AA DAG RAD abbreviates A SYNC A DAG RAD. 1.03 1.02 1.01 1.00 1.0 1 2 4 8 16 64 256 number of passes A-AdaGrad, η = 0.008 L2 = 0 A-AdaGrad, η = 0.008 L2 = 80 A-DA, η = 0.8 L2 = 0 A-DA, η = 0.8 L2 = 80 1.00 1.01 1.4 1.02 1.03 1.04 Impact of L2 regularizaton on test error 1.04 Fixed stepsizes, test data, L2=0 1.2 relative log-loss 1.6 1.8 Fixed stepsizes, training data, L2=0 A-AdaGrad η = 0.002 A-AdaGrad η = 0.004 A-AdaGrad η = 0.008 A-AdaGrad η = 0.016 A-DA η = 0.800 A-DA η = 1.600 A-DA η = 3.200 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 2: Relative accuracy for various stepsize choices on an click-through rate prediction dataset. 4.1 Malicious URL detection For our first set of experiments, we consider the speedup attainable by applying A SYNC A DAG RAD and A SYNC DA, investigating the performance of each algorithm on a malicious URL prediction task [7]. The dataset in this case consists of an anonymized collection of URLs labeled as malicious (e.g., spam, phishing, etc.) or benign over a span of 120 days. The data in this case consists of 2.4 · 106 examples with dimension d = 3.2 · 106 (sparse) features. We perform several experiments, randomly dividing the dataset into 1.2 · 106 training and test samples for each experiment. In Figure 1 we compare the performance of A SYNC A DAG RAD and A SYNC DA after doing after single pass through the training dataset. (For each algorithm, we choose the stepsize η for optimal training set performance.) We perform the experiments on a single machine running Ubuntu Linux with six cores (with two-way hyperthreading) and 32Gb of RAM. From the left-most plot in Fig. 1, we see that up to six processors, both A SYNC DA and A SYNC A DAG RAD enjoy the expected linear speedup, and from 6 to 12, they continue to enjoy a speedup that is linear in the number of processors though at a lesser slope (this is the effect of hyperthreading). For more than 12 processors, there is no further benefit to parallelism on this machine. The two right plots in Figure 1 plot performance of the different methods (with standard errors) versus the number of worker threads used. Both are essentially flat; increasing the amount of parallelism does nothing to the average training loss or the test error rate for either method. It is clear, however, that for this dataset, the adaptive A SYNC A DAG RAD algorithm provides substantial performance benefits over A SYNC DA. 4.2 Click-through-rate prediction experiments We also experiment on a proprietary datasets consisting of search ad impressions. Each example corresponds to showing a search-engine user a particular text ad in response to a query string. From this, we construct a very sparse feature vector based on the text of the ad displayed and the query string (no user-specific data is used). The target label is 1 if the user clicked the ad and -1 otherwise. 7 (B) A-AdaGrad speedup (D) Impact of training data ordering 1.004 1.005 1.006 1.007 1.008 1 2 4 8 16 32 number of passes 64 128 256 1.000 1 2 A-DA base η = 1.600 A-AdaGrad base η = 0.023 0 1.005 relative stepsize (C) Optimal stepsize scaling relative log-loss 1.003 target relative log-loss 1.005 1.010 1.002 1.010 1.015 8 4 0 speedup A-DA η = 1.600 A-AdaGrad η = 0.016 1.001 1.000 relative log-loss 1.015 A-DA, L2=80 A-AdaGrad, L2=80 12 (A) Optimized stepsize for each number of passes 1 2 4 8 16 32 number of passes 64 128 256 1 2 4 8 16 32 64 128 256 number of passes Figure 3. (A) Relative test-set log-loss for A SYNC DA and A SYNC A DAG RAD, choosing the best stepsize (within a factor of about 1.4×) individually for each number of passes. (B) Effective speedup for A SYNC A DAG RAD. (C) The best stepsize η, expressed as a scaling factor on the stepsize used for one pass. (D) Five runs with different random seeds for each algorithm (with 2 penalty 80). We fit logistic regression models using both A SYNC DA and A SYNC A DAG RAD. We run extensive experiments on a moderate-sized dataset (about 107 examples, split between training and testing), which allows thorough investigation of the impact of the stepsize η, the number of training passes,3 and 2 -regularization on accuracy. For these experiments we used 32 threads on 16 core machines for each run, as A SYNC A DAG RAD and A SYNC DA achieve similar speedups from parallelization. On this dataset, A SYNC A DAG RAD typically achieves an effective additional speedup over A SYNC DA of 4× or more. That is, to reach a given level of accuracy, A SYNC DA generally needs four times as many effective passes over the dataset. We measure accuracy with log-loss (the logistic loss) averaged over five runs using different random seeds (which control the order in which the algorithms sample examples during training). We report relative values in Figures 2 and 3, that is, the ratio of the mean loss for the given datapoint to the lowest (best) mean loss obtained. Our results are not particularly sensitive to the choice of relative log-loss as the metric of interest; we also considered AUC (the area under the ROC curve) and observed similar results. Figure 2 shows relative log-loss as a function of the number of training passes for various stepsizes. Without regularization, A SYNC A DAG RAD is prone to overfitting: it achieves significantly higher accuracy on the training data (Fig. 2 (left)), but unless the stepsize is tuned carefully to the number of passes, it will overfit (Fig. 2 (middle)). Fortunately, the addition of 2 regularization largely solves this problem. Indeed, Figure 2 (right) shows that while adding an 2 penalty of 80 has very little impact on A SYNC DA, it effectively prevents the overfitting of A SYNC A DAG RAD.4 Fixing √ regularization multiplier to 80, we varied the stepsize η over a multiplicative grid with res2 olution 2 for each number of passes and for each algorithm. Figure 3 reports the results obtained by selecting the best stepsize in terms of test set log-loss for each number of passes. Figure 3(A) shows relative log-loss of the best stepsize for each algorithm; 3(B) shows the relative time A SYNC DA requires with respect to A SYNC A DAG RAD to achieve a given loss. Specifically, Fig. 3(B) shows the ratio of the number of passes the algorithms require to achieve a fixed loss, which gives a broader estimate of the speedup obtained by using A SYNC A DAG RAD; speedups range from 3.6× to 12×. Figure 3(C) shows the optimal stepsizes as a function of the best setting for one pass. The optimal stepsize decreases moderately for A SYNC A DAG RAD, but are somewhat noisy for A SYNC DA. It is interesting to note that A SYNC A DAG RAD’s accuracy is largely independent of the ordering of the training data, while A SYNC DA shows significant variability. This can be seen both in the error bars on Figure 3(A), and explicitly in Figure 3(D), where we plot one line for each of the five random seeds used. Thus, while on the one hand A SYNC DA requires somewhat less tuning of the stepsize and 2 parameter, tuning A SYNC A DAG RAD is much easier because of its predictable response. 3 Here “number of passes” more precisely means the expected number of times each example in the dataset is trained on. That is, each worker thread randomly selects a training example from the dataset for each update, and we continued making updates until (dataset size) × (number of passes) updates have been processed. 4 For both algorithms, this is accomplished by adding the term η80 x 2 to the ψ function. We can achieve 2 slightly better results for A SYNC A DAG RAD by varying the 2 penalty with the number of passes. 8 References [1] P. Auer and C. Gentile. Adaptive and self-confident online learning algorithms. In Proceedings of the Thirteenth Annual Conference on Computational Learning Theory, 2000. [2] P. B¨ hlmann and S. van de Geer. Statistics for High-Dimensional Data: Methods, Theory and u Applications. Springer, 2011. [3] N. Cesa-Bianchi, A. Conconi, and C. Gentile. On the generalization ability of on-line learning algorithms. IEEE Transactions on Information Theory, 50(9):2050–2057, September 2004. [4] J. C. Duchi, E. Hazan, and Y. Singer. Adaptive subgradient methods for online learning and stochastic optimization. Journal of Machine Learning Research, 12:2121–2159, 2011. [5] E. Hazan. The convex optimization approach to regret minimization. In Optimization for Machine Learning, chapter 10. MIT Press, 2012. [6] J. Hiriart-Urruty and C. Lemar´ chal. Convex Analysis and Minimization Algorithms I & II. e Springer, New York, 1996. [7] J. Ma, L. K. Saul, S. Savage, and G. M. Voelker. Identifying malicious urls: An application of large-scale online learning. In Proceedings of the 26th International Conference on Machine Learning, 2009. [8] C. Manning and H. Sch¨ tze. Foundations of Statistical Natural Language Processing. MIT u Press, 1999. [9] B. McMahan and M. Streeter. Adaptive bound optimization for online convex optimization. In Proceedings of the Twenty Third Annual Conference on Computational Learning Theory, 2010. [10] A. Nemirovski, A. Juditsky, G. Lan, and A. Shapiro. Robust stochastic approximation approach to stochastic programming. SIAM Journal on Optimization, 19(4):1574–1609, 2009. [11] Y. Nesterov. Primal-dual subgradient methods for convex problems. Mathematical Programming, 120(1):261–283, 2009. [12] F. Niu, B. Recht, C. R´ , and S. Wright. Hogwild: a lock-free approach to parallelizing stochase tic gradient descent. In Advances in Neural Information Processing Systems 24, 2011. [13] P. Richt´ rik and M. Tak´ c. Parallel coordinate descent methods for big data optimization. a aˇ arXiv:1212.0873 [math.OC], 2012. URL http://arxiv.org/abs/1212.0873. [14] M. Tak´ c, A. Bijral, P. Richt´ rik, and N. Srebro. Mini-batch primal and dual methods for aˇ a SVMs. In Proceedings of the 30th International Conference on Machine Learning, 2013. 9
2 0.77036124 79 nips-2013-DESPOT: Online POMDP Planning with Regularization
Author: Adhiraj Somani, Nan Ye, David Hsu, Wee Sun Lee
Abstract: POMDPs provide a principled framework for planning under uncertainty, but are computationally intractable, due to the “curse of dimensionality” and the “curse of history”. This paper presents an online POMDP algorithm that alleviates these difficulties by focusing the search on a set of randomly sampled scenarios. A Determinized Sparse Partially Observable Tree (DESPOT) compactly captures the execution of all policies on these scenarios. Our Regularized DESPOT (R-DESPOT) algorithm searches the DESPOT for a policy, while optimally balancing the size of the policy and its estimated value obtained under the sampled scenarios. We give an output-sensitive performance bound for all policies derived from a DESPOT, and show that R-DESPOT works well if a small optimal policy exists. We also give an anytime algorithm that approximates R-DESPOT. Experiments show strong results, compared with two of the fastest online POMDP algorithms. Source code along with experimental settings are available at http://bigbird.comp. nus.edu.sg/pmwiki/farm/appl/. 1
3 0.76629114 50 nips-2013-Bayesian Mixture Modelling and Inference based Thompson Sampling in Monte-Carlo Tree Search
Author: Aijun Bai, Feng Wu, Xiaoping Chen
Abstract: Monte-Carlo tree search (MCTS) has been drawing great interest in recent years for planning and learning under uncertainty. One of the key challenges is the trade-off between exploration and exploitation. To address this, we present a novel approach for MCTS using Bayesian mixture modeling and inference based Thompson sampling and apply it to the problem of online planning in MDPs. Our algorithm, named Dirichlet-NormalGamma MCTS (DNG-MCTS), models the uncertainty of the accumulated reward for actions in the search tree as a mixture of Normal distributions. We perform inferences on the mixture in Bayesian settings by choosing conjugate priors in the form of combinations of Dirichlet and NormalGamma distributions and select the best action at each decision node using Thompson sampling. Experimental results confirm that our algorithm advances the state-of-the-art UCT approach with better values on several benchmark problems. 1
Author: Harikrishna Narasimhan, Shivani Agarwal
Abstract: We investigate the relationship between three fundamental problems in machine learning: binary classification, bipartite ranking, and binary class probability estimation (CPE). It is known that a good binary CPE model can be used to obtain a good binary classification model (by thresholding at 0.5), and also to obtain a good bipartite ranking model (by using the CPE model directly as a ranking model); it is also known that a binary classification model does not necessarily yield a CPE model. However, not much is known about other directions. Formally, these relationships involve regret transfer bounds. In this paper, we introduce the notion of weak regret transfer bounds, where the mapping needed to transform a model from one problem to another depends on the underlying probability distribution (and in practice, must be estimated from data). We then show that, in this weaker sense, a good bipartite ranking model can be used to construct a good classification model (by thresholding at a suitable point), and more surprisingly, also to construct a good binary CPE model (by calibrating the scores of the ranking model). 1
5 0.72684503 281 nips-2013-Robust Low Rank Kernel Embeddings of Multivariate Distributions
Author: Le Song, Bo Dai
Abstract: Kernel embedding of distributions has led to many recent advances in machine learning. However, latent and low rank structures prevalent in real world distributions have rarely been taken into account in this setting. Furthermore, no prior work in kernel embedding literature has addressed the issue of robust embedding when the latent and low rank information are misspecified. In this paper, we propose a hierarchical low rank decomposition of kernels embeddings which can exploit such low rank structures in data while being robust to model misspecification. We also illustrate with empirical evidence that the estimated low rank embeddings lead to improved performance in density estimation. 1
7 0.67820817 355 nips-2013-Which Space Partitioning Tree to Use for Search?
8 0.67801756 45 nips-2013-BIG & QUIC: Sparse Inverse Covariance Estimation for a Million Variables
9 0.67723477 149 nips-2013-Latent Structured Active Learning
10 0.67684633 333 nips-2013-Trading Computation for Communication: Distributed Stochastic Dual Coordinate Ascent
11 0.67515576 233 nips-2013-Online Robust PCA via Stochastic Optimization
12 0.67492902 318 nips-2013-Structured Learning via Logistic Regression
13 0.67394215 311 nips-2013-Stochastic Convex Optimization with Multiple Objectives
14 0.67390019 74 nips-2013-Convex Tensor Decomposition via Structured Schatten Norm Regularization
15 0.67386526 304 nips-2013-Sparse nonnegative deconvolution for compressive calcium imaging: algorithms and phase transitions
16 0.67203742 110 nips-2013-Estimating the Unseen: Improved Estimators for Entropy and other Properties
17 0.67176998 33 nips-2013-An Approximate, Efficient LP Solver for LP Rounding
18 0.67127275 252 nips-2013-Predictive PAC Learning and Process Decompositions
19 0.67109513 19 nips-2013-Accelerated Mini-Batch Stochastic Dual Coordinate Ascent
20 0.67079937 249 nips-2013-Polar Operators for Structured Sparse Estimation