arxiv: v2 [cs.lg] 14 Sep 2017

Similar documents
Variance-Reduced and Projection-Free Stochastic Optimization

arxiv: v1 [math.oc] 1 Jul 2016

arxiv: v1 [math.oc] 10 Oct 2018

Large-scale Stochastic Optimization

Projection-free Distributed Online Learning in Networks

Stochastic Dual Coordinate Ascent Methods for Regularized Loss Minimization

SVRG++ with Non-uniform Sampling

Making Gradient Descent Optimal for Strongly Convex Stochastic Optimization

Stochastic and online algorithms

Tutorial: PART 2. Optimization for Machine Learning. Elad Hazan Princeton University. + help from Sanjeev Arora & Yoram Singer

Modern Stochastic Methods. Ryan Tibshirani (notes by Sashank Reddi and Ryan Tibshirani) Convex Optimization

Stochastic Optimization

Course Notes for EE227C (Spring 2018): Convex Optimization and Approximation

Accelerating Stochastic Optimization

Linear Convergence under the Polyak-Łojasiewicz Inequality

arxiv: v4 [math.oc] 5 Jan 2016

Stochastic Gradient Descent with Variance Reduction

On Acceleration with Noise-Corrupted Gradients. + m k 1 (x). By the definition of Bregman divergence:

Full-information Online Learning

Lecture: Adaptive Filtering

A Greedy Framework for First-Order Optimization

Convex Optimization Lecture 16

arxiv: v1 [math.oc] 18 Mar 2016

References. --- a tentative list of papers to be mentioned in the ICML 2017 tutorial. Recent Advances in Stochastic Convex and Non-Convex Optimization

Frank-Wolfe Method. Ryan Tibshirani Convex Optimization

Linear Convergence under the Polyak-Łojasiewicz Inequality

Stochastic Gradient Descent with Only One Projection

Finite-sum Composition Optimization via Variance Reduced Gradient Descent

Barzilai-Borwein Step Size for Stochastic Gradient Descent

Accelerating SVRG via second-order information

Comparison of Modern Stochastic Optimization Algorithms

Contents. 1 Introduction. 1.1 History of Optimization ALG-ML SEMINAR LISSA: LINEAR TIME SECOND-ORDER STOCHASTIC ALGORITHM FEBRUARY 23, 2016

min f(x). (2.1) Objectives consisting of a smooth convex term plus a nonconvex regularization term;

NOTES ON FIRST-ORDER METHODS FOR MINIMIZING SMOOTH FUNCTIONS. 1. Introduction. We consider first-order methods for smooth, unconstrained

Optimization in the Big Data Regime 2: SVRG & Tradeoffs in Large Scale Learning. Sham M. Kakade

Stochastic Optimization Algorithms Beyond SG

OLSO. Online Learning and Stochastic Optimization. Yoram Singer August 10, Google Research

ONLINE VARIANCE-REDUCING OPTIMIZATION

STA141C: Big Data & High Performance Statistical Computing

Conditional Gradient (Frank-Wolfe) Method

ECS289: Scalable Machine Learning

Accelerated Randomized Primal-Dual Coordinate Method for Empirical Risk Minimization

Stochastic Gradient Descent. Ryan Tibshirani Convex Optimization

Coordinate Descent and Ascent Methods

Improved Optimization of Finite Sums with Minibatch Stochastic Variance Reduced Proximal Iterations

Adaptive Online Gradient Descent

First Efficient Convergence for Streaming k-pca: a Global, Gap-Free, and Near-Optimal Rate

Big Data Analytics: Optimization and Randomization

The FTRL Algorithm with Strongly Convex Regularizers

CSCI 1951-G Optimization Methods in Finance Part 12: Variants of Gradient Descent

Minimizing Finite Sums with the Stochastic Average Gradient Algorithm

Fast Asynchronous Parallel Stochastic Gradient Descent: A Lock-Free Approach with Convergence Guarantee

arxiv: v2 [math.oc] 29 Jul 2016

A Universal Catalyst for Gradient-Based Optimization

Stochastic Variance Reduction for Nonconvex Optimization. Barnabás Póczos

Complexity bounds for primal-dual methods minimizing the model of objective function

On the Iteration Complexity of Oblivious First-Order Optimization Algorithms

Parallel and Distributed Stochastic Learning -Towards Scalable Learning for Big Data Intelligence

Tutorial: PART 2. Online Convex Optimization, A Game- Theoretic Approach to Learning

ECS171: Machine Learning

Logarithmic Regret Algorithms for Strongly Convex Repeated Games

Simple Optimization, Bigger Models, and Faster Learning. Niao He

Stochastic Quasi-Newton Methods

Perceptron Mistake Bounds

Barzilai-Borwein Step Size for Stochastic Gradient Descent

Online Convex Optimization

An Optimal Affine Invariant Smooth Minimization Algorithm.

Machine Learning. Support Vector Machines. Fabio Vandin November 20, 2017

Ad Placement Strategies

On the Generalization Ability of Online Strongly Convex Programming Algorithms

Fast Stochastic Optimization Algorithms for ML

Proximal and First-Order Methods for Convex Optimization

Optimizing Nonconvex Finite Sums by a Proximal Primal-Dual Method

An Evolving Gradient Resampling Method for Machine Learning. Jorge Nocedal

CS260: Machine Learning Algorithms

IFT Lecture 6 Nesterov s Accelerated Gradient, Stochastic Gradient Descent

Generalized Conditional Gradient and Its Applications

Lower and Upper Bounds on the Generalization of Stochastic Exponentially Concave Optimization

Lecture 9: September 28

arxiv: v2 [stat.ml] 16 Jun 2015

CSE 417T: Introduction to Machine Learning. Lecture 11: Review. Henry Chai 10/02/18

An Extended Frank-Wolfe Method, with Application to Low-Rank Matrix Completion

Online Passive-Aggressive Algorithms

Nonlinear Optimization Methods for Machine Learning

Online Convex Optimization. Gautam Goel, Milan Cvitkovic, and Ellen Feldman CS 159 4/5/2016

Optimization for Machine Learning

Warm up. Regrade requests submitted directly in Gradescope, do not instructors.

Sub-Sampled Newton Methods

Lecture 3: Minimizing Large Sums. Peter Richtárik

Linearly-Convergent Stochastic-Gradient Methods

Computational and Statistical Learning Theory

Proximal Minimization by Incremental Surrogate Optimization (MISO)

FAST DISTRIBUTED COORDINATE DESCENT FOR NON-STRONGLY CONVEX LOSSES. Olivier Fercoq Zheng Qu Peter Richtárik Martin Takáč

Selected Topics in Optimization. Some slides borrowed from

Logistic Regression. Stochastic Gradient Descent

Fast Asynchronous Parallel Stochastic Gradient Descent: A Lock-Free Approach with Convergence Guarantee

Support Vector Machines: Training with Stochastic Gradient Descent. Machine Learning Fall 2017

Second-Order Stochastic Optimization for Machine Learning in Linear Time

Beyond the regret minimization barrier: an optimal algorithm for stochastic strongly-convex optimization

Faster Machine Learning via Low-Precision Communication & Computation. Dan Alistarh (IST Austria & ETH Zurich), Hantian Zhang (ETH Zurich)

Transcription:

Elad Hazan Princeton University, Princeton, NJ 08540, USA Haipeng Luo Princeton University, Princeton, NJ 08540, USA EHAZAN@CS.PRINCETON.EDU HAIPENGL@CS.PRINCETON.EDU arxiv:160.0101v [cs.lg] 14 Sep 017 Abstract The Frank-Wolfe optimization algorithm has recently regained popularity for machine learning applications due to its projection-free property and its ability to handle structured constraints. However, in the stochastic learning setting, it is still relatively understudied compared to the gradient descent counterpart. In this work, leveraging a recent variance reduction technique, we propose two stochastic Frank-Wolfe variants which substantially improve previous results in terms of the number of stochastic gradient evaluations needed to achieve 1 accuracy. For example, we improve from O 1 to Oln 1 if the objective function is smooth and strongly convex, and from O 1 to O 1 if the objective 1.5 function is smooth and Lipschitz. The theoretical improvement is also observed in experiments on real-world datasets for a multiclass classification application. 1. Introduction We consider the following optimization problem 1 n min fw = min f i w w Ω w Ω n i=1 which is an extremely common objective in machine learning. We are interested in the case where 1 n, usually corresponding to the number of training examples, is very large and therefore stochastic optimization is much more efficient; and the domain Ω admits fast linear optimization, while projecting onto it is much slower, necessitating projection-free optimization algorithms. Examples of such problem include multiclass classification, multitask learning, recommendation systems, matrix learning and many Proceedings of the 33 rd International Conference on Machine Learning, New York, NY, USA, 016. JMLR: W&CP volume 48. Copyright 016 by the authors. more see for example Hazan & Kale, 01; Hazan et al., 01; Jaggi, 013; Dudik et al., 01; Zhang et al., 01; Harchaoui et al., 015. The Frank-Wolfe algorithm Frank & Wolfe, 1956 also known as conditional gradient and it variants are natural candidates for solving these problems, due to its projectionfree property and its ability to handle structured constraints. However, despite gaining more popularity recently, its applicability and efficiency in the stochastic learning setting, where computing stochastic gradients is much faster than computing exact gradients, is still relatively understudied compared to variants of projected gradient descent methods. In this work, we thus try to answer the following question: what running time can a projection-free algorithm achieve in terms of the number of stochastic gradient evaluations and the number of linear optimizations needed to achieve a certain accuracy? Utilizing Nesterov s acceleration technique Nesterov, 1983 and the recent variance reduction idea Johnson & Zhang, 013; Mahdavi et al., 013, we propose two new algorithms that are substantially faster than previous work. Specifically, to achieve 1 accuracy, while the number of linear optimization is the same as previous work, the improvement of the number of stochastic gradient evaluations is summarized in Table 1: previous work this work Smooth O 1 O 1 1.5 Smooth and Strongly Convex O 1 Oln 1 Table 1: Comparisons of number of stochastic gradients The extra overhead of our algorithms is computing at most Oln 1 exact gradients, which is computationally insignificant compared to the other operations. A more detailed comparisons to previous work is included in Table, which will be further explained in Section. While the idea of our algorithms is quite straightforward,

we emphasize that our analysis is non-trivial, especially for the second algorithm where the convergence of a sequence of auxiliary points in Nesterov s algorithm needs to be shown. To support our theoretical results, we also conducted experiments on three large real-word datasets for a multiclass classification application. These experiments show significant improvement over both previous projection-free algorithms and algorithms such as projected stochastic gradient descent and its variance-reduced version. The rest of the paper is organized as follows: Section setups the problem more formally and discusses related work. Our two new algorithms are presented and analyzed in Section 3 and 4, followed by experiment details in Section 5.. Preliminary and Related Work We assume each function f i is convex and L-smooth in R d so that for any w, v R d, 1 f i v w v f i w f i v f i v w v + L w v. We will use two more important properties of smoothness. The first one is f i w f i v Lf i w f i v f i v w v proven in Appendix A for completeness, and the second one is f i λw + 1 λv 1 λf i w + 1 λf i v L λ1 λ w v for any w, v Ω and λ [0, 1]. Notice that f = 1 n n i=1 f i is also L-smooth since smoothness is preserved under convex combinations. For some cases, we also assume each f i is G-Lipschitz: f i w G for any w Ω, and f although not necessarily each f i is α-strongly convex, that is, fw fv fw w v α w v for any w, v Ω. As usual, µ = L α is called the condition number of f. We assume the domain Ω R d is a compact convex set with diameter D. We are interested in the case where linear optimization on Ω, formally argmin v Ω w v for any 1 We thank Sebastian Pokutta and Gábor Braun for pointing out that f i needs to be defined over R d, rather than only over Ω, in order for property 1 to hold. w R d, is much faster than projection onto Ω, formally argmin v Ω w v. Examples of such domains include the set of all bounded trace norm matrices, the convex hull of all rotation matrices, flow polytope and many more see for instance Hazan & Kale, 01..1. Example Application: Multiclass Classification Consider a multiclass classification problem where a set of training examples e i, y i i=1,...,n is given beforehand. Here e i R m is a feature vector and y i {1,..., h} is the label. Our goal is to find an accurate linear predictor, a matrix w = [w 1 ;..., w h ] Rh m that predicts argmax l w l e for any example e. Note that here the dimensionality d is hm. Previous work Dudik et al., 01; Zhang et al., 01 found that finding w by minimizing a regularized multivariate logistic loss gives a very accurate predictor in general. Specifically, the objective can be written in our notation with f i w = log 1 + expw l e i w y i e i l y i and Ω = {w R h m : w τ} where denotes the matrix trace norm. In this case, projecting onto Ω is equivalent to performing an SVD, which takes Ohm min{h, m} time, while linear optimization on Ω amounts to finding the top singular vector, which can be done in time linear to the number of non-zeros in the corresponding h by m matrix, and is thus much faster. One can also verify that each f i is smooth. The number of examples n can be prohibitively large for non-stochastic methods for instance, tens of millions for the ImageNet dataset Deng et al., 009, which makes stochastic optimization necessary... Detailed Efficiency Comparisons We call f i w a stochastic gradient for f at some w, where i is picked from {1,..., n} uniformly at random. Note that a stochastic gradient f i w is an unbiased estimator of the exact gradient fw. The efficiency of a projection-free algorithm is measured by how many numbers of exact gradient evaluations, stochastic gradient evaluations and linear optimizations respectively are needed to achieve 1 accuracy, that is, to output a point w Ω such that E[fw fw ] where w argmin w Ω fw is any optimum. In Table, we summarize the efficiency and extra assumptions needed beside convexity and smoothness of existing In general, condition G-Lipschitz in Table means each f i is G-Lipschitz, except for our STORC algorithm which only requires f being G-Lipschitz.

Algorithm Extra Conditions #Exact Gradients #Stochastic Gradients #Linear Optimizations Frank-Wolfe Garber & Hazan, 013 α-strongly convex Ω is polytope O LD 0 O LD Odµρ ln LD 0 Odµρ ln LD SFW G-Lipschitz 0 O G LD 4 3 O LD Online-FW Hazan & Kale, 01 SCGS Lan & Zhou, 014 SVRF this work STORC this work G-Lipschitz 0 O d LD +GD 4 O dld +GD 4 G-Lipschitz 0 O G4 D 4 O G4 D 4 4 4 L = allowed G-Lipschitz 0 O G D O LD G-Lipschitz α-strongly convex G-Lipschitz fw = 0 α-strongly convex 0 O G α LD O Oln LD O L D 4 O LD Oln LD O LD G 1.5 O LD Oln LD O LD O LD Oln LD Oµ ln LD O LD Table : Comparisons of different Frank-Wolfe variants see Section. for further explanations. algorithms in the literature as well as the two new algorithms we propose. Below we briefly explain these results from top to bottom. The standard Frank-Wolfe algorithm: v k = argmin fw k 1 v v Ω 3 w k = 1 γ k w k 1 + γ k v k for some appropriate chosen γ k requires O 1 iteration without additional conditions Frank & Wolfe, 1956; Jaggi, 013. In a recent paper, Garber & Hazan 013 give a variant that requires Odµρ ln 1 iterations when f is strongly convex and smooth, and Ω is a polytope 3. Although the dependence on is much better, the geometric constant ρ depends on the polyhedral set and can be very large. Moreover, each iteration of the algorithm requires further computation besides the linear optimization step. The most obvious way to obtain a stochastic Frank-Wolfe variant is to replace fw k 1 by some f i w k 1, or more generally the average of some number of iid samples of f i w k 1 mini-batch approach. We call this method SFW and include its analysis in Appendix B since we do not find it explicitly analyzed before. SFW needs O 1 3 stochastic gradients and O 1 linear optimization steps to reach an -approximate optimum. The work by Hazan & Kale 01 focuses on a online learning setting. One can extract two results from this work 015. 3 See also recent follow up work Lacoste-Julien & Jaggi, for the setting studied here. 4 In any case, the result is worse than SFW for both the number of stochastic gradients and the number of linear optimizations. Stochastic Condition Gradient Sliding SCGS, recently proposed by Lan & Zhou, 014, uses Nesterov s acceleration technique to speed up Frank-Wolfe. Without strong convexity, SCGS needs O 1 stochastic gradients, improving SFW. With strong convexity, this number can even be improved to O 1. In both cases, the number of linear optimization steps is O 1. The key idea of our algorithms is to combine the variance reduction technique proposed in Johnson & Zhang, 013; Mahdavi et al., 013 with some of the above-mentioned algorithms. For example, our algorithm SVRF combines this technique with SFW, also improving the number of stochastic gradients from O 1 3 to O 1, but without any extra conditions such as Lipschitzness required for SCGS. More importantly, despite having seemingly same convergence rate, SVRF substantially outperforms SCGS empirically see Section 5. On the other hand, our second algorithm STORC combines variance reduction with SCGS, providing even further improvements. Specifically, the number of stochastic gradients is improved to: O 1 when f is Lipschitz; O 1 1.5 when fw = 0; and finally Oln 1 when f is strongly convex. Note that the condition fw = 0 essentially 4 The first result comes from the setting where the online loss functions are stochastic, and the second one comes from a completely online setting with the standard online-to-batch conversion.

means that w is in the interior of Ω, but it is still an interesting case when the optimum is not unique and doing unconstraint optimization would not necessary return a point in Ω. Both of our algorithms require O 1 linear optimization steps as previous work, and overall require computing Oln LD exact gradients. However, we emphasize that this extra overhead is much more affordable compared to non-stochastic Frank-Wolfe that is, computing exact gradients every iteration since it does not have any polynomial dependence on parameters such as d, L or µ..3. Variance-Reduced Stochastic Gradients Originally proposed in Johnson & Zhang, 013 and independently in Mahdavi et al., 013, the idea of variancereduced stochastic gradients is proven to be highly useful and has been extended to various different algorithms such as Frostig et al., 015; Moritz et al., 016. A variance-reduced stochastic gradient at some point w Ω with some snapshot w 0 Ω is defined as fw; w 0 = f i w f i w 0 fw 0, where i is again picked from {1,..., n} uniformly at random. The snapshot w 0 is usually a decision point from some previous iteration of the algorithm and its exact gradient fw 0 has been pre-computed before, so that computing fw; w 0 only requires two standard stochastic gradient evaluations: f i w and f i w 0. A variance-reduced stochastic gradient is clearly also unbiased, that is, E[ fw; w 0 ] = fw. More importantly, the term f i w 0 fw 0 serves as a correction term to reduce the variance of the stochastic gradient. Formally, one can prove the following Lemma 1. For any w, w 0 Ω, we have E[ fw; w 0 fw ] 6LE[fw fw ] + E[fw 0 fw ]. In words, the variance of the variance-reduced stochastic gradient is bounded by how close the current point and the snapshot are to the optimum. The original work proves a bound on E[ fw; w 0 ] under the assumption fw = 0, which we do not require here. However, the main idea of the proof is similar and we defer it to Section 6. 3. Stochastic Variance-Reduced Frank-Wolfe With the previous discussion, our first algorithm is pretty straightforward: compared to the standard Frank-Wolfe, we simply replace the exact gradient with the average of Algorithm 1 Stochastic Variance-Reduced Frank-Wolfe SVRF 1: Input: Objective function f = 1 n n i=1 f i. : Input: Parameters γ k, m k and N k. 3: Initialize: w 0 = argmin w Ω fx w for some arbitrary x Ω. 4: for t = 1,,..., T do 5: Take snapshot: x 0 = w t 1 and compute fx 0. 6: for k = 1 to N t do 7: Compute k, the average of m k iid samples of fx k 1, x 0. 8: Compute v k = argmin v Ω k v. 9: Compute x k = 1 γ k x k 1 + γ k v k. 10: end for 11: Set w t = x Nt. 1: end for a mini-batch of variance-reduced stochastic gradients, and take snapshots every once in a while. We call this algorithm Stochastic Variance-Reduced Frank-Wolfe SVRF, whose pseudocode is presented in Alg 1. The convergence rate of this algorithm is shown in the following theorem. Theorem 1. With the following parameters, γ k = k + 1, m k = 96k + 1, N t = t+3, Algorithm 1 ensures E[fw t fw ] LD t+1 for any t. Before proving this theorem, we first show a direct implication of this convergence result. Corollary 1. To achieve 1 accuracy, Algorithm 1 requires Oln LD exact gradient evaluations, O L D 4 stochastic gradient evaluations and O LD linear optimizations. Proof. According to the algorithm and the choice of parameters, it is clear that these three numbers are T + 1, T Nt t=1 k=1 m k = O4 T and T t=1 N t = O T respectively. Theorem 1 implies that T should be of order LD Θlog. Plugging in all parameters concludes the proof. To prove Theorem 1, we first consider a fixed iteration t and prove the following lemma: Lemma. For any k, we have E[fx k fw ] 4LD k + if E[ s fx s 1 ] L D s+1 for all s k.

We defer the proof of this lemma to Section 6 for coherence. With the help of Lemma, we are now ready to prove the main convergence result. Proof of Theorem 1. We prove by induction. For t = 0, by smoothness, the optimality of w 0 and convexity, we have fw 0 fx + fx w 0 x + L w 0 x fx + fx w x + LD fw + LD. Now assuming E[fw t 1 fw ] LD, we consider t iteration t of the algorithm and use another induction to show E[fx k fw ] 4LD k+ for any k N t. The base case is trivial since x 0 = w t 1. Suppose E[fx s 1 fw ] 4LD s+1 for any s k. Now because s is the average of m s iid samples of fx s 1 ; x 0, its variance is reduced by a factor of m s. That is, with Lemma 1 we have E[ s fx s 1 ] 6L E[fx s 1 fw ] + E[fx 0 fw ] m s 8LD 6L m s 6L m s s + 1 + LD t 8LD s + 1 + 8LD s + 1 = L D s + 1, where the last inequality is by the fact s N t = t+3 and the last equality is by plugging the choice of m s. Therefore the condition of Lemma is satisfied and the induction is completed. Finally with the choice of N t we thus prove E[fw t fw ] = E[fx Nt fw ] 4LD N = LD t+. t+1 We remark that in Alg 1, we essentially restart the algorithm that is, reseting k to 1 after taking a new snapshot. However, another option is to continue increasing k and never reset it. Although one can show that this only leads to constant speed up for the convergence, it provides more stable update and is thus what we implement in experiments. 4. Stochastic Variance-Reduced Conditional Gradient Sliding Our second algorithm applies variance reduction to the SCGS algorithm Lan & Zhou, 014. Again, the key difference is that we replace the stochastic gradients with the average of a mini-batch of variance-reduced stochastic gradients, and take snapshots every once in a while. See pseudocode in Alg for details. Algorithm STOchastic variance-reduced Conditional gradient sliding STORC 1: Input: Objective function f = 1 n n i=1 f i. : Input: Parameters γ k, β k, η t,k, m t,k and N t. 3: Initialize: w 0 = argmin w Ω fx w for some arbitrary x Ω. 4: for t = 1,,... do 5: Take snapshot: y 0 = w t 1 and compute fy 0. 6: Initialize x 0 = y 0. 7: for k = 1 to N t do 8: Compute z k = 1 γ k y k 1 + γ k x k 1. 9: Compute k, the average of m t,k iid samples of fz k ; y 0. 10: Let gx = β k x x k 1 + k x. 11: Compute x k, the output of using standard Frank- Wolfe to solve min x Ω gx until the duality gap is at most η t,k, that is, max gx k x k x η t,k. 4 x Ω 1: Compute y k = 1 γ k y k 1 + γ k x k. 13: end for 14: Set w t = y Nt. 15: end for The algorithm makes use of two auxiliary sequences x k and z k Line 8 and 1, which is standard for Nesterov s algorithm. x k is obtained by approximately solving a square norm regularized linear optimization so that it is close to x k 1 Line 11. Note that this step does not require computing any extra gradients of f or f i, and is done by performing the standard Frank-Wolfe algorithm Eq. 3 until the duality gap is at most a certain value η t,k. The duality gap is a certificate of approximate optimality see Jaggi, 013, and is a side product of the linear optimization performed at each step, requiring no extra cost. Also note that the stochastic gradients are computed at the sequence z k instead of y k, which is also standard in Nesterov s algorithm. However, according to Lemma 1, we thus need to show the convergence rate of the auxiliary sequence z k, which appears to be rarely studied previously to the best our knowledge. This is one of the key steps in our analysis. The main convergence result of STORC is the following: Theorem. With the following parameters where D t is defined later below: γ k = k + 1, β k = 3L k, η t,k = LD t N t k, Algorithm ensures E[fw t fw ] LD t+1 for any t if any of the following three cases holds:

a fw = 0 and D t = D, N t = t +, m t,k = 900N t. b f is G-Lipschitz and D t = D, N t = t +, m t,k = 700N t + 4NtGk+1 LD. c f is α-strongly convex and D t = µd t 1, N t = 3µ, m t,k = 5600N t µ where µ = L α. Again we first give a direct implication of the above result: Corollary. To achieve 1 accuracy, Algorithm requires Oln LD exact gradient evaluations and O LD linear optimizations. The numbers of stochastic gradient evaluations for Case a, b and c are respectively O LD, O LD + LD G and Oµ ln LD 1.5. Proof. Line 11 requires O β kd η t,k iterations of the standard Frank-Wolfe algorithm since gx is β k -smooth see e.g. Jaggi, 013, Theorem. So the numbers of exact gradient evaluations, stochastic gradient evaluations and linear optimizations are respectively T +1, T Nt t=1 k=1 m t,k and Nt k=1 β k D η t,k O T t=1 LD be of order Θlog the corollary.. Theorem implies that T should. Plugging in all parameters proves To prove Theorem, we again first consider a fixed iteration t and use the following lemma, which is essentially proven in Lan & Zhou, 014. We include a distilled proof in Appendix C for completeness. Lemma 3. Suppose E[ y 0 w ] D t holds for some positive constant D t D. Then for any k, we have E[fy k fw ] if E[ s fz s ] 8LD t kk + 1 L D t N ts+1 for all s k. Proof of Theorem. We prove by induction. The base case t = 0 holds by the exact same argument as in the proof of Theorem 1. Suppose E[fw t 1 fw ] LD and t consider iteration t. Below we use another induction to prove E[fy k fw ] 8LD t kk+1 for any 1 k N t, which will concludes the proof since for any of the three cases, we have E[fw t fw ] = E[fy Nt fw ] which is at most 8LD t N t LD t+1. We first show that the condition E[ y 0 w ] D t holds. This is trivial for Cases a and b when D t = D. For Case c, by strong convexity and the inductive assumption, we have E[ y 0 w ] α E[fy 0 fw ] LD α t 1 = D t. Next note that Lemma 1 implies that E[ s fz s ] 6L is at most m t,s E[fz s fw ] + E[fy 0 fw ]. So the key is to bound E[fz s fw ]. With z 1 = y 0 one can verify that E[ 1 fz 1 ] is at most 18L m t,1 E[fy 0 fw ] 18L D m t,1 for all three t L D t 4N t cases, and thus E[fy s fw ] 8LD t ss+1 holds for s = 1 by Lemma 3. Now suppose it holds for any s < k, below we discuss the three cases separately to show that it also holds for s = k. Case a. By smoothness, the condition fw = 0, the construction of z s, and Cauchy-Schwarz inequality, we have for any 1 < s k, fz s fy s 1 + fy s 1 fw z s y s 1 + L z s y s 1 = fy s 1 + γ s fy s 1 fw x s 1 y s 1 + Lγ s x s 1 y s 1 fy s 1 + γ s D fy s 1 fw + LD γs. Property 1 and the optimality of w implies: fy s 1 fw Lfy s 1 fw fw y s 1 w Lfy s 1 fw. So subtracting fw and taking expectation on both sides, and applying Jensen s inequality and the inductive assumption, we have E[fz s fw ] E[fy s 1 fw ] + γ s D LE[fy s 1 fw ] + LD s + 1 8LD s 1s + 8LD s + 1 s 1s + LD s + 1 < 55LD s + 1. On the other hand, we have E[fy 0 fw ] LD t 16LD N t 1 < 40LD N t+1 40LD s+1. So E[ s fz s 900L is at most D m t,ss+1, and the choice of m t,s ensures that L this bound is at most D N ts+1, satisfying the condition of Lemma 3 and thus completing the induction.

Case b. With the G-Lipschitz condition we proceed similarly and bound fz s by fy s 1 + fy s 1 z s y s 1 + L z s y s 1 = fy s 1 + γ s fy s 1 x s 1 y s 1 + LD γ s fy s 1 + γ s GD + LD γs. So using bounds derived previously and the choice of m t,s, we bound E[ s fz s as follows: 6L 16LD m t,s s 1s + 4GD s + 1 + 4LD s + 1 + 40LD s + 1 < 6L 4GD m t,s s + 1 + 116LD s + 1 < L D N t s + 1, again completing the induction. Case c. Using the definition of z s and y s and direct calcalution, one can remove the dependence of x s and verify y s 1 = s + 1 s 1 z s + s s 1 y s for any s. Now we apply Property with λ = s+1 s 1 : fy s 1 s + 1 s 1 fz s + s s 1 fy s L s + 1s s 1 z s y s = fw + s + 1 s 1 fz s fw + s s 1 fy s fw Ls s + 1 y s 1 y s fw + 1 fz s fw L y s 1 y s, where the equality is by adding and subtracting fw and the fact y s 1 y s = s+1 s 1 z s y s, and the last inequality is by fy s fw and trivial relaxations. Rearranging gives fz s fw fy s 1 fw + L y s 1 y s. Applying Cauchy-Schwarz inequality, strong convexity and the fact µ 1, we continue with fz s fw fy s 1 fw + L y s 1 w + y s w fy s 1 fw + 4µfy s 1 fw + fy s fw 6µfy s 1 fw + 4µfy s fw, For s 3, we use the inductive assumption to show E[fz s fw ] 48µLD t s 1s + 3µLD t s s 1 448µLD t s+1. dataset #features #categories #examples news0 6,061 0 15,935 rcv1 47,36 53 15,564 aloi 18 1,000 108,000 Table 3: Summary of datasets The case for s = can be verified similarly using the bound on E[fy 0 fw ] and E[fy 1 fw ] base case. Finally we bound the term E[fy 0 fw ] LD = t LD t µ 3LD t N t+1 3LD t s+1, and conclude that the variance E[ s fz s is at most 6L m t,s 896µLD t s+1 L D t N ts+1, completing the induction by Lemma 3. 5. Experiments + 3LD t s+1 To support our theory, we conduct experiments in the multiclass classification problem mentioned in Sec.1. Three datasets are selected from the LIBSVM repository 5 with relatively large number of features, categories and examples, summarized in the Table 3. Recall that the loss function is multivariate logistic loss and Ω is the set of matrices with bounded trace norm τ. We focus on how fast the loss decreases instead of the final test error rate so that the tuning of τ is less important, and is fixed to 50 throughout. We compare six algorithms. Four of them SFW, SCGS, SVRF, STORC are projection-free as discussed, and the other two are standard projected stochastic gradient descent SGD and its variance-reduced version SVRG Johnson & Zhang, 013, both of which require expensive projection. For most of the parameters in these algorithms, we roughly follow what the theory suggests. For example, the size of mini-batch of stochastic gradients at round k is set to k, k 3 and k respectively for SFW, SCGS and SVRF, and is fixed to 100 for the other three. The number of iterations between taking two snapshots for variance-reduced methods SVRG, SVRF and STORC are fixed to 50. The learning rate is set to the typical decaying sequence c/ k for SGD and a constant c for SVRG as the original work suggests for some best tuned c and c. Since the complexity of computing gradients, performing linear optimization and projecting are very different, we measure the actual running time of the algorithms and see how fast the loss decreases. Results can be found in Figure 1, where one can clearly observe that for all datasets, 5 https://www.csie.ntu.edu.tw/ cjlin/ libsvmtools/datasets/

3 4 7 Loss.9.8.7.6 Loss 3.8 3.6 3.4 3. 3.8 Loss 6.8 6.6 6.4 6. SGD SVRG SCGS STORC SFW SVRF.5 0 100 00 300 400 500 Time s a news0.6 0 100 00 300 400 500 Time s b rcv1 6 0 50 100 150 Time s c aloi Figure 1: Comparison of six algorithms on three multiclass datasets best viewed in color SGD and SVRG are significantly slower compared to the others, due to the expensive projection step, highlighting the usefulness of projection-free algorithms. Moreover, we also observe large improvement gained from the variance reduction technique, especially when comparing SCGS and STORC, as well as SFW and SVRF on the aloi dataset. Interestingly, even though the STORC algorithm gives the best theoretical results, empirically the simpler algorithms SFW and SVRF tend to have consistent better performance. 6. Omitted Proofs Proof of Lemma 1. Let E i denotes the conditional expectation given all the past except the realization of i. We have E i [ fw; w 0 fw ] = E i [ f i w f i w 0 + fw 0 fw ] = E i [ f i w f i w f i w 0 f i w + fw 0 fw fw fw ] 3E i [ f i w f i w + f i w 0 f i w fw 0 fw + fw fw ] 3E i [ f i w f i w + f i w 0 f i w + fw fw ] where the first inequality is Cauchy-Schwarz inequality, and the second one is by the fact E i [ f i w 0 f i w ] = fw 0 fw and that the variance of a random variable is bounded by its second moment. We now apply Property 1 to bound each of the three terms above. For example, E i f i w f i w LE i [f i w f i w f i w w w ] = Lfw fw fw w w, which is at most Lfw fw by the optimality of w. Proceeding similarly for the other two terms concludes the proof. Proof of Lemma. For any s k, by smoothness we have fx s fx s 1 + fx s 1 x s x s 1 + L x s x s 1. Plugging in x s = 1 γ s x s 1 + γ s v s gives fx s fx s 1 + γ s fx s 1 v s x s 1 + Lγ s v s x s 1. Rewriting and using the fact that v s x s 1 D leads to fx s fx s 1 + γ s s v s x s 1 + γ s fx s 1 s v s x s 1 + LD γs. The optimality of v s implies s v s s w. So with further rewriting we arrive at fx s fx s 1 + γ s fx s 1 w x s 1 + γ s fx s 1 s v s w + LD γs. By convexity, term fx s 1 w x s 1 is bounded by fw fx s 1, and by Cauchy-Schwarz inequality, term fx s 1 s v s w is bounded by D s fx s 1, which in expectation is at most LD s+1 by the condition on E[ s fx s 1 ] and Jensen s inequality. Therefore we can bound E[fx s fw ] by 1 γ s E[fx s 1 fw ] + LD γ s s + 1 + LD γs = 1 γ s E[fx s 1 fw ] + LD γs. Finally we prove E[fx k fw ] 4LD k+ by induction. The base case is trival: E[fx 1 fw ] is bounded by 1 γ 1 E[fx 0 fw ]+LD γ1 = LD since γ 1 = 1. Suppose E[fx s 1 fw ] 4LD s+1 then with γ s = s+1 we bound E[fx s fw ] by 4LD 1 s + 1 s + 1 + 1 4LD s + 1 s +, completing the induction. 7. Conclusion and Open Problems We conclude that the variance reduction technique, previously shown to be highly useful for gradient descent variants, can also be very helpful in speeding up projection-free

algorithms. The main open question is, in the strongly convex case, whether the number of stochastic gradients for STORC can be improved from Oµ ln 1 to Oµ ln 1, which is typical for gradient descent methods, and whether the number of linear optimizations can be improved from O 1 to Oln 1. Acknowledgements The authors acknowledge support from the National Science Foundation grant IIS-153815 and a Google research award. References Deng, Jia, Dong, Wei, Socher, Richard, Li, Li-Jia, Li, Kai, and Fei-Fei, Li. Imagenet: A large-scale hierarchical image database. In Computer Vision and Pattern Recognition, 009. CVPR 009. IEEE Conference on, pp. 48 55. IEEE, 009. Dudik, Miro, Harchaoui, Zaid, and Malick, Jérôme. Lifted coordinate descent for learning with trace-norm regularization. In Proceedings of the Fifteenth International Conference on Artificial Intelligence and Statistics, volume, pp. 37 336, 01. Frank, Marguerite and Wolfe, Philip. An algorithm for quadratic programming. Naval research logistics quarterly, 31-:95 110, 1956. Frostig, Roy, Ge, Rong, Kakade, Sham M, and Sidford, Aaron. Competing with the empirical risk minimizer in a single pass. In Proceedings of the 8th Annual Conference on Learning Theory, 015. Garber, Dan and Hazan, Elad. A linearly convergent conditional gradient algorithm with applications to online and stochastic optimization. arxiv preprint arxiv:1301.4666, 013. Harchaoui, Zaid, Juditsky, Anatoli, and Nemirovski, Arkadi. Conditional gradient algorithms for normregularized smooth convex optimization. Mathematical Programming, 151-:75 11, 015. Hazan, Elad and Kale, Satyen. Projection-free online learning. In Proceedings of the 9th International Conference on Machine Learning, 01. Hazan, Elad, Kale, Satyen, and Shalev-Shwartz, Shai. Near-optimal algorithms for online matrix prediction. In COLT 01 - The 5th Annual Conference on Learning Theory, June 5-7, 01, Edinburgh, Scotland, pp. 38.1 38.13, 01. Jaggi, Martin. Revisiting frank-wolfe: Projection-free sparse convex optimization. In Proceedings of the 30th International Conference on Machine Learning, pp. 47 435, 013. Johnson, Rie and Zhang, Tong. Accelerating stochastic gradient descent using predictive variance reduction. In Advances in Neural Information Processing Systems 7, pp. 315 33, 013. Lacoste-Julien, Simon and Jaggi, Martin. On the global linear convergence of frank-wolfe optimization variants. In Advances in Neural Information Processing Systems 9, pp. 496 504, 015. Lan, Guanghui and Zhou, Yi. Conditional gradient sliding for convex optimization. Optimization-Online preprint 4605, 014. Mahdavi, Mehrdad, Zhang, Lijun, and Jin, Rong. Mixed optimization for smooth functions. In Advances in Neural Information Processing Systems, pp. 674 68, 013. Moritz, Philipp, Nishihara, Robert, and Jordan, Michael I. A linearly-convergent stochastic l-bfgs algorithm. In Proceedings of the Nineteenth International Conference on Artificial Intelligence and Statistics, 016. Nesterov, YU. E. A method of solving a convex programming problem with convergence rate o1/k. In Soviet Mathematics Doklady, volume 7, pp. 37 376, 1983. Zhang, Xinhua, Schuurmans, Dale, and Yu, Yao-liang. Accelerated training for matrix-norm regularization: A boosting approach. In Advances in Neural Information Processing Systems 6, pp. 906 914, 01.

Supplementary material for Variance-Reduced and Projection-Free Stochastic Optimization A. Proof of Property 1 Proof. We drop the subscript i for conciseness. Define gw = fw fv w, which is clearly also convex and L-smooth on Ω. Since gv = 0, v is one of the minimizers of gw. Therefore we have gv gw gw 1 gw gw L gw w 1 L gw w + L w 1 L gw w by smoothness of g = 1 L gw = 1 fw fv L Rearranging and plugging in the definition of g concludes the proof. B. Analysis for SFW The concrete update of SFW is v k = argmin v Ω k v w k = 1 γ k w k 1 + γ k v k where k is the average of m k iid samples of stochastic gradient f i w k 1. The convergence rate of SFW is presented below. Theorem 3. If each f i is G-Lipschitz, then with γ k = k+1 and m k = Gk+1, LD SFW ensures for any k, E[fw k fw ] 4LD k +. Proof. Similar to the proof of Lemma, we first proceed as follows, fw k fw k 1 + fw k 1 w k w k 1 + L w k w k 1 smoothness = fw k 1 + γ k fw k 1 v k w k 1 + Lγ k v k x k 1 w k w k 1 = γ k v k w k 1 fw k 1 + γ k k v k w k 1 + γ k fw k 1 k v k w k 1 + LD γ k fw k 1 + γ k k w w k 1 + γ k fw k 1 k v k w k 1 + LD γ k v k w k 1 D by optimality of v k = fw k 1 + γ k fw k 1 w w k 1 + γ k fw k 1 k v k w + LD γ k fw k 1 + γ k fw fw k 1 + γ k D k fw k 1 + LD γk, where the last step is by convexity and Cauchy-Schwarz inequality. Since f i is G-Lipschitz, with Jensen s inequality, we further have E[ k fw k 1 ] E[ k fw k 1 ] G mk, which is at most LDγ k with the choice of γ k and m k. So we arrive at E[fw k fw ] 1 γ k E[fw k 1 fw ] + LD γk. It remains to use a simple induction to conclude the proof. Now it is clear that to achieve 1 accuracy, SFW needs O LD iterations, and in total O G L D LD 3 = O G LD 4 3 stochastic gradients.

C. Proof of Lemma 3 Proof. Let δ s = s fz s. For any s k, we proceed as follows: fy s fz s + fz s y s z s + L y s z s by smoothness = 1 γ s fz s + fz s y s 1 z s + γ s fz s + fz s w z s + γ s fz s x s w + Lγ s x s x s 1 by definition of y s and z s 1 γ s fy s 1 + γ s fw + γ s fz s x s w + Lγ s x s x s 1 by convexity = 1 γ s fy s 1 + γ s fw + γ s s x s w + Lγ s x s x s 1 + γ s δ s w x s 1 γ s fy s 1 + γ s fw + γ s η t,s γ s β s x s x s 1 x s w + Lγ s x s x s 1 + γ s δ s w x s = 1 γ s fy s 1 + γ s fw + γ s η t,s + β sγ s x s 1 w x s w + γ s Lγ s β s x s x s 1 + δ s x s 1 x s + δ s w x s 1 1 γ s fy s 1 + γ s fw + γ s η t,s + β sγ s x s 1 w x s w + γ s where the last inequality is by the fact β s Lγ s and thus by Eq. 4 δ s + δ s w x s 1, β s Lγ s Lγ s β s x s x s 1 + δ s x s 1 x s = δ s β s Lγ s β s Lγ s x δ s s x s 1 β s Lγ s δ s. β s Lγ s Note that E[δ s w x s 1 ] = 0. So with the condition E[ δ s ] L D def t N ts+1 = σ s we arrive at E[fy s fw ] 1 γ s E[fy s 1 fw ]+γ s η t,s + β s E[ x s 1 w ] E[ x s w σ s ] +. β s Lγ s Now define Γ s = Γ s 1 1 γ s when s > 1 and Γ 1 = 1. By induction, one can verify Γ s = ss+1 and the following: which is at most k γ s Γ k Γ s=1 s E[fy k fw ] Γ k η s + k γ s Γ s=1 s σs + Γ k γ 1 β 1 β s Lγ s η t,s + β s E[ x s 1 w ] E[ x s w ] + Γ 1 E[ x 0 w ] + k γs β s s= Γ s σs, β s Lγ s γ s 1β s 1 E[ x s 1 w ]. Γ s 1 Finally plugging in the parameters γ s, β s, η t,s, Γ s and the bound E[ x 0 w ] D t concludes the proof: E[fy k fw ] kk + 1 k s=1 LD k t N t k + LDt + 3LD t N t k + 1 kk + 1 8LD t kk + 1.