black-box evasion and poisoning attacks - arxiv › pdf › 1909.13806.pdfmin-max optimization...

31
Min-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks Sijia Liu *1 Songtao Lu *2 Xiangyi Chen *3 Yao Feng *4 Kaidi Xu *5 Abdullah Al-Dujaili *61 Mingyi Hong 3 Una-May O’Reilly 61 Abstract In this paper, we study the problem of constrained min-max optimization in a black-box setting, where the desired optimizer cannot access the gra- dients of the objective function but may query its values. We present a principled optimiza- tion framework, integrating a zeroth-order (ZO) gradient estimator with an alternating projected stochastic gradient descent-ascent method, where the former only requires a small number of func- tion queries and the later needs just one-step de- scent/ascent update. We show that the proposed framework, referred to as ZO-Min-Max, has a sub- linear convergence rate under mild conditions and scales gracefully with problem size. We also ex- plore a promising connection between black-box min-max optimization and black-box evasion and poisoning attacks in adversarial machine learn- ing (ML). Our empirical evaluations on these use cases demonstrate the effectiveness of our approach and its scalability to dimensions that prohibit using recent black-box solvers. 1. Introduction Min-max optimization problems have been studied for mul- tiple decades (Wald, 1945), and the majority of the proposed methods assume access to first-order (FO) information, i.e. gradients, to find or approximate robust solutions (Nesterov, 2007; Gidel et al., 2017; Hamedani et al., 2018; Qian et al., 2019; Rafique et al., 2018; Sanjabi et al., 2018b; Lu et al., 2019; Nouiehed et al., 2019; Lu et al., 2019; Jin et al., 2019). Different from standard optimization, min-max optimiza- * Equal contribution 1 MIT-IBM Watson AI Lab, IBM Research 2 Thomas J. Watson Research Center, IBM Research 3 ECE, Uni- versity of Minnesota 4 Tsinghua University 5 ECE, Northeastern University 6 CSAIL, MIT. Correspondence to: Sijia Liu <Si- [email protected]>. Proceedings of the 37 th International Conference on Machine Learning, Vienna, Austria, PMLR 108, 2020. Copyright 2020 by the author(s). tion tackles a composition of an inner maximization problem and an outer minimization problem. It can be used in many real-world applications, which are faced with various forms of uncertainty or adversary. For instance, when training a ML model on user-provided data, malicious users can carry out a data poisoning attack: providing false data with the aim of corrupting the learned model (Steinhardt et al., 2017; Tran et al., 2018; Jagielski et al., 2018). At inference time, malicious users can evade detection of multiple models in the form of adversarial example attacks (Goodfellow et al., 2014; Liu et al., 2016; 2018a). Our study is particularly motivated by the design of data poisoning and evasion ad- versarial attacks from black-box machine learning (ML) or deep learning (DL) systems, whose internal configuration and operating mechanism are unknown to adversaries. We propose zeroth-order (gradient-free) min-max optimization methods, where gradients are neither symbolically nor nu- merically available, or they are tedious to compute. Recently, zeroth-order (ZO) optimization has attracted in- creasing attention in solving ML/DL problems, where FO gradients (or stochastic gradients) are approximated based only on the function values. For example, ZO optimization serves as a powerful and practical tool for generation of adversarial example to evaluate the adversarial robustness of black-box ML/DL models (Chen et al., 2017; Ilyas et al., 2018; Tu et al., 2018; Ilyas et al., 2019; Chen et al., 2018; Li et al., 2019). ZO optimization can also help to solve au- tomated ML problems, where the gradients with respect to ML pipeline configuration parameters are intractable (Aggarwal et al., 2019; Wang & Wu, 2019). Furthermore, ZO optimization provides computationally-efficient alterna- tives of high-order optimization methods for solving com- plex ML/DL tasks, e.g., robust training by leveraging input gradient or curvature regularization (Finlay & Oberman, 2019; Moosavi-Dezfooli et al., 2019), network control and management (Chen & Giannakis, 2018; Liu et al., 2018b), and data processing in high dimension (Liu et al., 2018b; Golovin et al., 2019). Other recent applications include generating model-agnostic contrastive explanations (Dhu- randhar et al., 2019) and escaping saddle points (Flokas et al., 2019). arXiv:1909.13806v3 [cs.LG] 17 Jun 2020

Upload: others

Post on 25-Jun-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients: Convergence and Applications toBlack-Box Evasion and Poisoning Attacks

Sijia Liu * 1 Songtao Lu * 2 Xiangyi Chen * 3 Yao Feng * 4 Kaidi Xu * 5 Abdullah Al-Dujaili * 6 1 Mingyi Hong 3

Una-May O’Reilly 6 1

Abstract

In this paper, we study the problem of constrainedmin-max optimization in a black-box setting,where the desired optimizer cannot access the gra-dients of the objective function but may queryits values. We present a principled optimiza-tion framework, integrating a zeroth-order (ZO)gradient estimator with an alternating projectedstochastic gradient descent-ascent method, wherethe former only requires a small number of func-tion queries and the later needs just one-step de-scent/ascent update. We show that the proposedframework, referred to as ZO-Min-Max, has a sub-linear convergence rate under mild conditions andscales gracefully with problem size. We also ex-plore a promising connection between black-boxmin-max optimization and black-box evasion andpoisoning attacks in adversarial machine learn-ing (ML). Our empirical evaluations on theseuse cases demonstrate the effectiveness of ourapproach and its scalability to dimensions thatprohibit using recent black-box solvers.

1. IntroductionMin-max optimization problems have been studied for mul-tiple decades (Wald, 1945), and the majority of the proposedmethods assume access to first-order (FO) information, i.e.gradients, to find or approximate robust solutions (Nesterov,2007; Gidel et al., 2017; Hamedani et al., 2018; Qian et al.,2019; Rafique et al., 2018; Sanjabi et al., 2018b; Lu et al.,2019; Nouiehed et al., 2019; Lu et al., 2019; Jin et al., 2019).Different from standard optimization, min-max optimiza-

*Equal contribution 1MIT-IBM Watson AI Lab, IBM Research2Thomas J. Watson Research Center, IBM Research 3ECE, Uni-versity of Minnesota 4Tsinghua University 5ECE, NortheasternUniversity 6CSAIL, MIT. Correspondence to: Sijia Liu <[email protected]>.

Proceedings of the 37 th International Conference on MachineLearning, Vienna, Austria, PMLR 108, 2020. Copyright 2020 bythe author(s).

tion tackles a composition of an inner maximization problemand an outer minimization problem. It can be used in manyreal-world applications, which are faced with various formsof uncertainty or adversary. For instance, when training aML model on user-provided data, malicious users can carryout a data poisoning attack: providing false data with theaim of corrupting the learned model (Steinhardt et al., 2017;Tran et al., 2018; Jagielski et al., 2018). At inference time,malicious users can evade detection of multiple models inthe form of adversarial example attacks (Goodfellow et al.,2014; Liu et al., 2016; 2018a). Our study is particularlymotivated by the design of data poisoning and evasion ad-versarial attacks from black-box machine learning (ML) ordeep learning (DL) systems, whose internal configurationand operating mechanism are unknown to adversaries. Wepropose zeroth-order (gradient-free) min-max optimizationmethods, where gradients are neither symbolically nor nu-merically available, or they are tedious to compute.

Recently, zeroth-order (ZO) optimization has attracted in-creasing attention in solving ML/DL problems, where FOgradients (or stochastic gradients) are approximated basedonly on the function values. For example, ZO optimizationserves as a powerful and practical tool for generation ofadversarial example to evaluate the adversarial robustnessof black-box ML/DL models (Chen et al., 2017; Ilyas et al.,2018; Tu et al., 2018; Ilyas et al., 2019; Chen et al., 2018;Li et al., 2019). ZO optimization can also help to solve au-tomated ML problems, where the gradients with respectto ML pipeline configuration parameters are intractable(Aggarwal et al., 2019; Wang & Wu, 2019). Furthermore,ZO optimization provides computationally-efficient alterna-tives of high-order optimization methods for solving com-plex ML/DL tasks, e.g., robust training by leveraging inputgradient or curvature regularization (Finlay & Oberman,2019; Moosavi-Dezfooli et al., 2019), network control andmanagement (Chen & Giannakis, 2018; Liu et al., 2018b),and data processing in high dimension (Liu et al., 2018b;Golovin et al., 2019). Other recent applications includegenerating model-agnostic contrastive explanations (Dhu-randhar et al., 2019) and escaping saddle points (Flokaset al., 2019).

arX

iv:1

909.

1380

6v3

[cs

.LG

] 1

7 Ju

n 20

20

Page 2: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Current studies (Ghadimi & Lan, 2013; Nesterov &Spokoiny, 2015; Duchi et al., 2015; Ghadimi et al., 2016;Shamir, 2017; Balasubramanian & Ghadimi, 2018; Liu et al.,2018c; 2019) suggested that ZO methods for solving single-objective optimization problems typically agree with theiteration complexity of FO methods but encounter a slow-down factor up to a small-degree polynomial of the problemdimensionality. To the best of our knowledge, it was anopen question whether any convergence rate analysis canbe established for black-box min-max (bi-level) optimiza-tion. In this paper, we develop a provable and scalableblack-box min-max stochastic optimization method by inte-grating a query-efficient randomized ZO gradient estimatorwith a computation-efficient alternating gradient descent-ascent framework. Here the former requires a small numberof function queries, and the latter needs just one-step de-scent/ascent update.

Contribution. We summarize our contributions as fol-lows. (i) We identify a class of black-box attack problemswhich turn out to be min-max black-box optimization prob-lems. (ii) We propose a scalable and principled framework(ZO-Min-Max) for solving constrained min-max saddlepoint problems under both one-sided and two-sided black-box objective functions. Here the one-sided setting refersto the scenario where only the outer minimization problemis black-box. (iii) We provide a novel convergence analy-sis characterizing the number of objective function evalua-tions required to attain locally robust solution to black-boxmin-max problems (structured by nonconvex outer mini-mization and strongly concave inner maximization). Ouranalysis handles stochasticity in both objective function andZO gradient estimator, and shows that ZO-Min-Max yieldsO(1/T + 1/b+ d/q) convergence rate, where T is numberof iterations, b is mini-batch size, q is number of randomdirection vectors used in ZO gradient estimation, and d isnumber of optimization variables. (iv) We demonstrate theeffectiveness of our proposal in practical data poisoning andevasion attack generation problems.1

2. Related WorkFO min-max optimization. Gradient-based methodshave been applied with celebrated success to solve min-maxproblems such as robust learning (Qian et al., 2019), gener-ative adversarial networks (GANs) (Sanjabi et al., 2018a),adversarial training (Al-Dujaili et al., 2018b; Madry et al.,2018), and robust adversarial attack generation (Wang et al.,2019b). Some of FO methods are motivated by theoret-ical justifications based on Danskin’s theorem (Danskin,1966), which implies that the negative of the gradient of theouter minimization problem at inner maximizer is a descent

1Source code will be released.

direction (Madry et al., 2018). Convergence analysis ofother FO min-max methods has been studied under differentproblem settings, e.g., (Lu et al., 2019; Qian et al., 2019;Rafique et al., 2018; Sanjabi et al., 2018b; Nouiehed et al.,2019). It was shown in (Lu et al., 2019) that a determinis-tic FO min-max algorithm has O(1/T ) convergence rate.In (Qian et al., 2019; Rafique et al., 2018), stochastic FOmin-max methods have also been proposed, which yield theconvergence rate in the order of O(1/

√T ) and O(1/T 1/4),

respectively. However, these works were restricted to uncon-strained optimization at the minimization side. In (Sanjabiet al., 2018b), noncovnex-concave min-max problems werestudied, but the proposed analysis requires solving the maxi-mization problem only up to some small error. In (Nouiehedet al., 2019), the O(1/T ) convergence rate was proved fornonconvex-nonconcave min-max problems under Polyak-Łojasiewicz conditions. Different from the aforementionedFO settings, ZO min-max stochastic optimization suffersrandomness from both stochastic sampling in objective func-tion and ZO gradient estimation, and this randomness wouldbe coupled in alternating gradient descent-descent steps andthus make it more challenging in convergence analysis.

Gradient-free min-max optimization. In the black-boxsetup, coevolutionary algorithms were used extensively tosolve min-max problems (Herrmann, 1999; Schmiedlechneret al., 2018). However, they may oscillate and never con-verge to a solution due to pathological behaviors such asfocusing and relativism (Watson & Pollack, 2001). Fixes tothese issues have been proposed and analyzed—e.g., asym-metric fitness (Jensen, 2003; Branke & Rosenbusch, 2008).In (Al-Dujaili et al., 2018c), the authors employed an evo-lution strategy as an unbiased approximate for the descentdirection of the outer minimization problem and showedempirical gains over coevlutionary techniques, albeit with-out any theoretical guarantees. Min-max black-box prob-lems can also be addressed by resorting to direct search andmodel-based descent and trust region methods (Audet &Hare, 2017; Larson et al., 2019; Rios & Sahinidis, 2013).However, these methods lack convergence rate analysis andare difficult to scale to high-dimensional problems. For ex-ample, the off-the-shelf model-based solver COBYLA onlysupports problems with 216 variables at maximum in SciPyPython library (Jones et al., 2001), which is even smallerthan the size of a single ImageNet image.

The recent work (Bogunovic et al., 2018) proposed a robustBayesian optimization (BO) algorithm and established a the-oretical lower bound on the required number of the min-maxobjective evaluations to find a near-optimal point. However,BO approaches are often tailored to low-dimensional prob-lems and its computational complexity prohibits scalableapplication. From a game theory perspective, the min-maxsolution for some problems correspond to the Nash equi-

Page 3: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

librium between the outer minimizer and the inner maxi-mizer, and hence black-box Nash equilibria solvers can beused (Picheny et al., 2019; Al-Dujaili et al., 2018a). Thissetup, however, does not always hold in general. Our workcontrasts with the above lines of work in designing and ana-lyzing black-box min-max techniques that are both scalableand theoretically grounded.

3. Problem SetupIn this section, we define the black-box min-max problemand briefly motivate its applications. By min-max, we meanthat the problem is a composition of inner maximization andouter minimization of the objective function f . By black-box, we mean that the objective function f is only accessiblevia functional evaluations. Mathematically, we have

minx∈X

maxy∈Y

f(x,y) (1)

where x and y are optimization variables, f is a differen-tiable objective function, and X ⊂ Rdx and Y ⊂ Rdy arecompact convex sets. For ease of notation, let dx = dy = d.In (1), the objective function f could represent either a deter-ministic loss or stochastic loss f(x,y) = Eξ∼p [f(x,y; ξ)],where ξ is a random variable following the distribution p. Inthis paper, we cover the stochastic variant in (1). We focuson two black-box scenarios. (a) One-sided black-box: f isa white box w.r.t. y but a black box w.r.t. x. (b) Two-sidedblack-box: f is a black box w.r.t. both x and y.

Motivation of setup (a) and (b). The formulation of theone-sided black-box min-max problem can be used to designthe black-box ensemble evasion attack, where the attackergenerates adversarial examples (i.e., crafted examples withslight perturbations for misclassification at the testing phase)and optimizes its worst-case performance against an ensem-ble of black-box classifiers and/or example classes. Theformulation of two-sided black-box min-max problem rep-resents another type of attack at the training phase, knownas poisoning attack, where the attacker deliberately influ-ences the training data (by injecting poisoned samples) tomanipulate the results of a black-box predictive model. Al-though problems of designing ensemble evasion attack (Liuet al., 2016; 2018a; Wang et al., 2019b) and data poison-ing attack (Jagielski et al., 2018; Wang et al., 2019a) havebeen studied in the literature, most of them assumed that theadversary has the full knowledge of the target ML model,leading to an impractical white-box attack setting. By con-trast, we provide a solution to min-max attack generationunder black-box ML models. We refer readers to Section 6for further discussion and demonstration of our frameworkon these problems.

4. ZO-Min-Max: A Framework forBlack-Box Min-Max Optimization

Our interest is in a scalable and theoretically principledframework for solving black-box min-max problems of theform (1). To this end, we first introduce a randomizedgradient estimator that only requires a few number of point-wise function evaluations. Based on that, we then propose aZO alternating projected gradient method to solve (1) underboth one-sided and two-sided black-box setups.

Randomized gradient estimator. In the ZO setting, weadopt a randomized gradient estimator to estimate the gradi-ent of a function with the generic form h(x) := Eξ[h(x; ξ)](Gao et al., 2014; Berahas et al., 2019),

∇xh(x) =1

bq

∑j∈I

q∑i=1

d[h(x + µui; ξj)− h(x; ξj)]

µui, (2)

where d is number of variables, I denotes the mini-batch setof b i.i.d. stochastic samples ξjbj=1, uiqi=1 are q i.i.d.random direction vectors drawn uniformly from the unitsphere, and µ > 0 is a smoothing parameter. We note thatthe ZO gradient estimator (2) involves randomness fromboth stochastic sampling w.r.t. ui as well as the randomdirection sampling w.r.t. ξj . It is known from (Gao et al.,2014, Lemma 2) that ∇xh(x) provides an unbiased estimateof the gradient of the smoothing function of h rather thanthe true gradient of h. Here the smoothing function of h isdefined by hµ(x) = Ev[h(x + µv)], where v follows theuniform distribution over the unit Euclidean ball. Besidesthe bias, we provide an upper bound on the variance of (2)in Lemma 1.Lemma 1. Suppose that for all ξ, h(x; ξ) has Lh Lips-chitz continuous gradients and the gradient of h(x; ξ) isupper bounded as ‖∇xh(x; ξ)‖22 ≤ η2 at x ∈ Rd. Then

E[∇xh(x)

]= ∇xhµ(x), and

E[‖∇xh(x)−∇xhµ(x)‖22

]≤ σ2(Lh, µ, b, q, d), (3)

where the expectation is taken over all randomness, andσ2(Lh, µ, b, q, d) = 2η2

b +4dη2+µ2L2

hd2

q .

Proof: See Appendix B.2.

In Lemma 1, if we choose µ ≤ 1/√d, then the variance

bound is given by O(1/b + d/q). In our problem setting(1), the ZO gradients ∇xf(x,y) and ∇yf(x,y) follow thegeneric form of (2) by fixing y and letting h(·) := f(·,y)or by fixing x and letting h(·) := f(x, ·), respectively.

Algorithmic framework. To solve problem (1), we al-ternatingly perform ZO projected gradient descent/ascentmethod for updating x and y. Specifically, the ZO projectedgradient descent (ZO-PGD) is conducted over x

x(t) = projX

(x(t−1) − α∇xf

(x(t−1),y(t−1)

)), (4)

Page 4: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

where t is the iteration index, ∇xf denotes the ZO gradientestimate of f w.r.t. x, α > 0 is the learning rate at the x-minimization step, and projX (a) signifies the projection ofa ontoX , given by the solution to the problem minx∈X ‖x−a‖22. For one-sided ZO min-max optimization, besides (4),we perform FO projected gradient ascent (PGA) over y.And for two-sided ZO min-max optimization, our update ony obeys ZO-PGA

y(t) = projY

(y(t−1) + β∇yf

(x(t),y(t−1)

)), (5)

where β > 0 is the learning rate at the y-maximization step,and ∇yf denotes the ZO gradient estimate of f w.r.t. y.The proposed method is named as ZO-Min-Max; see thepseudocode in Appendix A.

Why estimates gradient rather than distribution of func-tion values? Besides ZO optimization using random gra-dient estimates, the black-box min-max problem (1) can alsobe solved using the Bayesian optimization (BO) approach,e.g., (Bogunovic et al., 2018; Al-Dujaili et al., 2018a). Thecore idea of BO is to approximate the objective functionas a Gaussian process (GP) learnt from the history of func-tion values at queried points. Based on GP, the solution toproblem (1) is then updated by maximizing a certain rewardfunction, known as acquisition function. The advantageof BO is its mild requirements on the setting of black-boxproblems, e.g., at the absence of differentiability. How-ever, BO usually does not scale beyond low-dimensionalproblems since learning the accurate GP model and solvingthe acquisition problem takes intensive computation costper iteration. By contrast, our proposed method is moreefficient, and mimics the first-order method by just usingthe random gradient estimate (2) as the descent/ascent di-rection. In Figure A1, we compare ZO-Min-Max with theBO based STABLEOPT algorithm proposed by (Bogunovicet al., 2018) through a toy example shown in Appendix Cand a poisoning attack generation example in Sec. 6.2. Wecan see that ZO-Min-Max not only achieves more accuratesolution but also requires less computation time.

Why is difficult to analyze the convergence of ZO-Min-Max? The convergence analysis of ZO-Min-Max is morechallenging than the case of FO min-max algorithms. Thestochasticity of the gradient estimator makes the conver-gence analysis sufficiently different from the FO determinis-tic case (Lu et al., 2019; Qian et al., 2019), since the errorsin minimization and maximization are coupled as the algo-rithm proceeds.

Moreover, the conventioanl analysis of ZO optimizationfor single-objective problems cannot directly be applied toZO-Min-Max. Even at the one-sided black-box setting, ZO-Min-Max conducts alternating optimization using one-stepZO-PGD and PGA with respect to x and y respectively.This is different from a reduced ZO optimization problem

with respect to x only by solving problem minx∈X h(x),where h(x) := maxy∈Y f(x,y). Here acquiring the solu-tion to the inner maximization problem could be non-trivialand computationally intensive. Furthermore, the alternatingalgorithmic structure leads to opposite optimization direc-tions (minimization vs maximization) over variables x andy respectively. Even applying ZO optimization only to oneside, it needs to quantify the effect of ZO gradient estima-tion on the descent over both x and y. We provide a detailedconvergence analysis of ZO-Min-Max in Sec. 5.

5. Convergence Rate AnalysisWe first elaborate on assumptions and notations used inanalyzing the convergence of ZO-Min-Max (Algorithm A1).

A1: In (1), f(x,y) is continuously differentiable, and isstrongly concave w.r.t. y with parameter γ > 0, namely,given x ∈ X , f(x,y1) ≤ f(x,y2) +∇yf(x,y2)T (y1 −y2) − γ

2 ‖y1 − y2‖2 for all points y1,y2 ∈ Y . And fis lower bounded by a finite number f∗ and has boundedgradients ‖∇xf(x,y; ξ)‖ ≤ η2 and ‖∇yf(x,y; ξ)‖ ≤ η2

for stochastic optimization with ξ ∼ p. Here ‖·‖ denotes the`2 norm. The constraint sets X ,Y are convex and boundedwith diameter R.

A2: f(x,y) has Lipschitz continuous gradients, i.e., thereexistsLx, Ly > 0 such that ‖∇xf(x1,y)−∇xf(x2,y)‖ ≤Lx‖x1−x2 ‖ for ∀x1,x2 ∈ X , and ‖∇yf(x1,y) −∇yf(x2,y)‖ ≤ Ly‖x1−x2 ‖ and ‖∇yf(x,y1) −∇yf(x,y2)‖ ≤ Ly‖y1−y2 ‖ for ∀y1,y2 ∈ Y .

We remark that A1 and A2 are required for analyzing theconvergence of ZO-Min-Max. They were used even forthe analysis of first-order optimization methods with sin-gle rather than bi-level objective functions (Chen et al.,2019; Ward et al., 2019). In A1, the strongly concavity off(x,y) with respect to y holds for applications such as ro-bust learning over multiple domains (Qian et al., 2019), andadversarial attack generation shown in Section 6. In A2, theassumption of smoothness (namely, Lipschitz continuousgradient) is required to quantify the descent of the alter-nating projected stochastic gradient descent-ascent method.For clarity, we also summarize the problem and algorithmicparameters used in our convergence analysis in Table A1 ofAppendix B.1.

We measure the convergence of ZO-Min-Max by the proxi-mal gradient (Lu et al., 2019; Ghadimi et al., 2016),

G(x,y) =

[(1/α) (x−projX (x−α∇xf(x,y)))(1/β)

(y−projY(y+β∇yf(x,y))

) ] , (6)

where (x,y) is a first-order stationary point of (1) iff‖G(x,y)‖ = 0.

Descent property in x-minimization. In what follows,

Page 5: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

we delve into our convergence analysis. Since ZO-Min-Max (Algorithm A1) calls for ZO-PGD for solving bothone-sided and two-sided black-box optimization problems,we first show the descent property at the x-minimizationstep of ZO-Min-Max in Lemma 2.Lemma 2. (Descent lemma in minimization) Under A1-A2,let (x(t),y(t)) be a sequence generated by ZO-Min-Max.When f(x,y) is black-box w.r.t. x, then we have followingdescent property w.r.t. x:

E[f(x(t+1),y(t))] ≤E[f(x(t),y(t))]−(

1

α− Lx

2

)E‖∆(t+1)

x ‖2

+ ασ2x + Lxµ

2 (7)

where ∆(t)x := x(t)−x(t−1), and σ2

x := σ2(Lx, µ, b, q, d)defined in (3).

Proof: See Appendix B.3.1. It is clear fromLemma 2 that updating x leads to the reduced objectivevalue when choosing a small learning rate α. However, ZOgradient estimation brings in additional errors in terms ofασ2

x and Lxµ2, where the former is induced by the varianceof gradient estimates in (3) and the latter is originated frombounding the distance between f and its smoothing version;see (24) in Appendix B.3.

Convergence rate of ZO-Min-Max by performing PGA.We next investigate the convergence of ZO-Min-Max whenPGA is used at the y-maximization step (Line 8 of Al-gorithm A1) for solving one-sided black-box optimizationproblems.Lemma 3. (Descent lemma in maximization) Under A1-A2, let (x(t),y(t)) be a sequence generated by AlgorithmA1 and define the potential function as

P(x(t),y(t),∆(t)y ) =E[f(x(t),y(t))]

+4 + 4β2L2

y − 7βγ

2β2γE‖∆(t)

y ‖2, (8)

where ∆(t)y := y(t)−y(t−1). When f(x,y) is black-box

w.r.t. x and white-box w.r.t. y, then we have the followingdescent property w.r.t. y:

P(x(t+1),y(t+1),∆(t+1)y ) ≤ P(x(t+1),y(t),∆(t)

y )

−(

1

2β−

2L2y

γ

)E‖∆(t+1)

y ‖2 +

(2

γ2β+β

2

)L2xE‖∆(t+1)

x ‖2.

(9)

Proof: See Appendix B.3.2.

It is shown from (9) that when β is small enough, then theterm (1/(2β)−2L2

y/γ)E‖∆(t+1)y ‖2 will give some descent

of the potential function after PGA, while the last term in(9) will give some ascent to the potential function. How-ever, such a quantity will be compensated by the descent

of the objective function in the minimization step shown byLemma 2. Combining Lemma 2 and Lemma 3, we obtainthe convergence rate of ZO-Min-Max in Theorem 1.Theorem 1. Suppose that A1-A2 hold, the sequence(x(t),y(t)) over T iterations is generated by AlgorithmA1 in which learning rates satisfy β < 1/(4L2

y) andα ≤ min1/Lx, 1/(Lx/2+2L2

x/(γ2β)+βL2

x/2). Whenf(x,y) is black-box w.r.t. x and white-box w.r.t. y, theconvergence rate of ZO-Min-Max under a uniformly andrandomly picked (x(r),y(r)) from (x(t),y(t))Tt=1 is givenby

E‖G(x(r),y(r))‖2 ≤ c

ζ

(P1 − f∗ − νR2)

T+cασ2

x

ζ+cLxµ

2

ζ(10)

where ζ is a constant independent on the parameters µ,b, q, d and T , Pt := P(x(t),y(t),∆

(t)y ) given by (8),

c = maxLx + 3/α, 3/β, ν = min4 + 4β2L2y −

7βγ, 0/(2β2γ), σ2x is variance bound of ZO gradient es-

timate given in (7), and f∗, R, γ, Lx and Ly have beendefined in A1-A2.

Proof: See Appendix B.3.3.

To better interpret Theorem 1, we begin by clarifying theparameters involved in our convergence rate (10). First,the parameter ζ in the denominator of the derived conver-gence error is non-trivially lower bounded given appropriatelearning rates α and β (as will be evident in Remark 1).Second, the parameter c is inversely proportional to α andβ. Thus, to guarantee the constant effect of the ratio c/ξ,it is better not to set these learning rates too small; seea specification in Remark 1-2. Third, the parameter ν isnon-negative and appears in terms of −νR2, thus, it willnot make convergence rate worse. Fourth, P1 is the initialvalue of the potential function (8). By setting an appropriatelearning rate β (e.g., following Remark 2), P1 is then upperbounded by a constant determined by the initial value ofthe objective function, the distance of the first two updates,Lipschitz constant Ly and strongly concave parameter γ.We next provide Remarks 1-3 on Theorem 1.

Remark 1. Recall that ζ = minc1, c2 (Appendix B.2.3),where c1 = 1/(2β) − 2L2

y/γ and c2 = 1α − (Lx

2 +2L2

x

γ2β +βL2

x

2 ). Given the fact that Lx and Ly are Lipschitz constantsand γ is the strongly concavity constant, a proper lowerbound of ζ thus relies on the choice of the learning rates αand β. By setting β ≤ γ

8L2y

and α ≤ 1/(Lx+4L2

x

γ2β +βL2x), it

is easy to verify that c1 ≥2L2

y

γ and c2 ≥ Lx

2 +2L2

x

γ2β +βL2

x

2 ≥Lx

2 +2L2

x

γ . Thus, we obtain that ζ ≥ min 2L2y

γ ,2L2

x

γ + Lx

2 .This justifies that ζ has a non-trivial lower bound, whichwill not make the convergence error bound (10) vacuous(although the bound has not been optimized over α and β).

Remark 2. It is not wise to set learning rates α and β to

Page 6: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

extremely small values since c is inversely proportional to αand β. We could choose β = γ

8L2y

and α = 1/(Lx +4L2

x

γ2β +

βL2x) in Remark 1 to guarantee the constant effect of c/ζ.

Remark 3. By setting µ ≤ min1/√d, 1/√T, we obtain

σ2x = O(1/b+d/q) from Lemma 1, and Theorem 1 implies

that ZO-Min-Max yieldsO(1/T + 1/b+ d/q) convergencerate for one-sided black-box optimization. Compared to theFO rateO(1/T ) (Lu et al., 2019; Sanjabi et al., 2018a), ZO-Min-Max converges only to a neighborhood of stationarypoints with O(1/T ) rate, where the size of the neighbor-hood is determined by the mini-batch size b and the numberof random direction vectors q used in ZO gradient estima-tion. It is also worth mentioning that such a stationary gapmay exist even in the FO/ZO projected stochastic gradientdescent for solving single-objective minimization problems(Ghadimi et al., 2016).

As shown in Remark 3, ZO-Min-Max could result in a sta-tionary gap. A large mini-batch size b or number of randomdirection vectors q can improve its iteration complexity.However, this requires O(bq) times more function queriesper iteration from (2). It implies the tradeoff between it-eration complexity and function query complexity in ZOoptimization.

Convergence rate of ZO-Min-Max by performing ZO-PGA. The previous convergence analysis of ZO-Min-Maxis served as a basis for analyzing the more general two-sided black-box case, where ZO-PGA is used at the y-maximization step (Line 6 of Algorithm A1). In Lemma 4,we examine the descent property in maximization by usingZO gradient estimation.Lemma 4. (Descent lemma in maximization) Under A1-A2, let (x(t),y(t)) be a sequence generated by AlgorithmA1 and define the potential function as

P ′(x(t),y(t),∆(t)y ) =E[f(x(t),y(t))]

+4 + 4(3L2

y + 2)β2 − 7βγ

β2γE‖∆(t)

y ‖2. (11)

When function f(x,y) is black-box w.r.t. both x and y, wehave the following descent w.r.t. y:

P ′(x(t+1),y(t+1),∆(t+1)y ) ≤ P ′(x(t+1),y(t),∆(t)

y )

−(

1

2β−

6L2y + 4

γ

)E‖∆(t+1)

y ‖2

+

(6L2

x

γ2β+

3βL2x

2

)E‖∆(t+1)

x ‖2 +7β2γ2 + 28βγ + 12

βγ2σ2y

+βγ + 4

4β2γµ2d2L2

y, (12)

where σ2y := σ2(Ly, µ, b, q, d) given in (3).

Proof: See Appendix B.4.1.

Lemma 4 is analogous to Lemma 3 by taking into accountthe effect of ZO gradient estimate ∇yf(x,y) on the poten-tial function (11). Such an effect is characterized by theterms related to σ2

y and µ2d2L2y in (12).

Theorem 2. Suppose that A1-A2 hold, the sequence(x(t),y(t)) over T iterations is generated by Algorithm A1in which learning rates satisfy β < γ/(4(3L2

y + 2)) andα ≤ minLx, 1/(Lx/2 + 6L2

x/(γ2β) + 3βL2

x/2). Whenf(x,y) is black-box w.r.t. both x and y, the convergencerate of ZO-Min-Max under a uniformly and randomly picked(x(r),y(r)) from (x(t),y(t))Tt=1 is given by

E‖G(x(r),y(r))‖2 ≤ c

ζ′P ′1 − f∗ − ν′R2

T+cα

ζ′σ2x

+

(cb1ζ′

+ d2L2y

)µ2 +

(cb2ζ′

+ 2

)σ2y,

where ζ ′ is a constant independent on the parameters µ,b, q, d and T , P ′t := P ′(x(t),y(t),∆

(t)y ) in (11), c has

been defined in (10) , ν′ =min4+4(3L2

y+2)β2−7βγ,0β2γ , b1 =

Lx+d2L2

y(4+βγ)

4β2γ and b2 = 7β2γ2+28βγ+12βγ2 , σ2

x and σ2y have

been introduced in (7) and (12), and f∗, R, γ, Lx and Lyhave been defined in A1-A2.

Proof: See Appendix B.4.2.

Following the similar argument in Remark 1 of Theorem 1,one can choose proper learning rates α and β to obtain validlower bound on ζ ′. However, different from Theorem 1,the convergence error shown by Theorem 2 involves anadditional error term related to σ2

y and has worse dimension-dependence on the term related to µ2. The latter yields amore restricted choice of the smoothing parameter µ: weobtain O(1/T + 1/b + d/q) convergence rate when µ ≤1/(d√T ).

6. Applications & Experiment ResultsIn what follows, we evaluate the empirical performance ofZO-Min-Max on two applications of adversarial exploration:a) design of black-box ensemble evasion attack against deepneural networks (DNNs), and b) design of black-box poi-soning attack against a logistic regression model.

6.1. Ensemble attack via universal perturbation

A black-box min-max problem formulation. We con-sider the scenario in which the attacker generates a universaladversarial perturbation over multiple input images againstan ensemble of multiple classifiers (Liu et al., 2016; 2018a).Considering I classes of images (the group of images withina class i is denoted by Ωi) and J network models, the goalof the adversary is to find the universal perturbation x addi-tive to I classes of images against J models. The proposedblack-box attack formulation mimics the white-box attack

Page 7: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

formulation (Wang et al., 2019b)

minimizex∈X

maximizew∈W

f(x,w) :=

J∑j=1

I∑i=1

[wijFij (x; Ωi)]

− λ‖w − 1/(IJ)‖22, (13)

where x ∈ Rd and w ∈ RIJ are optimization variables,wij denotes the (i, j)th entry of w corresponding to theimportance weight of attacking the set of images at class iunder the model j, X denotes the perturbation constraint,e.g., X = x | ‖x‖∞ ≤ ε,∀z ∈ ∪iΩi, andW is the proba-bilistic simplexW = w | 1Tw = 1,w ≥ 0. In problem(13), Fij (x; Ωi) is the attack loss for attacking images inΩi under model j, and λ > 0 is a regularization parameterto strike a balance between the worse-case attack loss andthe average loss (Wang et al., 2019b). We note that Fijare black-box functions w.r.t. x since the adversary hasno access to the internal configuration s of DNN modelsto be attacked. Accordingly, the input gradient cannot becomputed by back-propagation. This implies that problem(13) belongs to the one-sided black-box optimization prob-lem (white-box objective w.r.t. w). We refer readers toAppendix D for more details on (13).

Implementation details. We consider J = 2 DNN-basedclassifiers, Inception-V3 (Szegedy et al., 2016) and ResNet-50 (He et al., 2016), and I = 2 image classes, each ofwhich contains 20 images randomly selected from ImageNet(Deng et al., 2009).

In (13), we choose λ = 5. In Algorithm A1, we set thelearning rates by α = 0.05 and β = 0.01. Also, we usethe full batch of image samples and set q = 10 in gradientestimation (2), where we set µ = 5 × 10−3. The functionquery complexity is thus q (= 10) times more than thenumber of iterations.

Baseline methods for comparison. In experiments, wecompare ZO-Min-Max with (a) FO-Min-Max, (b) ZO-PGD(Ghadimi et al., 2016), and (c) ZO-Finite-Sum, where themethod (a) is the FO counterpart of Algorithm A1, themethod (b) performs single-objective ZO minimizationunder the equivalent form of (13), minx∈X h(x) whereh(x) = maxw∈W f(x,w), and the method (c) performsZO-PGD to minimize the finite-sum (average) loss ratherthan the worst-case (min-max) loss. We remark that thebaseline method (b) calls for the solution to the inner max-imization problem maxw∈W f(x,w), which is elaboratedon Appendix D. It is also worth mentioning that althoughZO-Finite-Sum tackles an objective function different from(13), it is motivated by the previous work on designingthe adversarial attack against model ensembles (Liu et al.,2018a), and we can fairly compare ZO-Min-Max with ZO-Finite-Sum in terms of the attack performance of obtaineduniversal adversarial perturbations.

0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00Number of iterations 1e4

10 4

10 3

10 2

10 1

100

101

Stat

iona

ry g

ap ZO-Min-MaxFO-Min-Max

(a)

0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00Number of iterations 1e4

0123456

Atta

ck lo

ss

ZO-Min-Max, M1C1 ZO-Min-Max, M1C2 ZO-Min-Max, M2C1 ZO-Min-Max, M2C2

FO-Min-Max, M1C1 FO-Min-Max, M1C2 FO-Min-Max, M2C1 FO-Min-Max, M2C2

(b)

Figure 1: ZO-Min-Max vs. FO-Min-Max in attack generation. (a)Stationary gap; (b) attack loss at each model-class pair.

0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00Number of iterations 1e4

0

1

2

3

4

5

6

Atta

ck lo

ss

for e

ach

mod

el-c

lass

pai

r

ZO-Min-Max, M1C1 ZO-Min-Max, M1C2 ZO-Min-Max, M2C1 ZO-Min-Max, M2C2

ZO-PGD, M1C1 ZO-PGD, M1C2 ZO-PGD, M2C1 ZO-PGD, M2C2

Figure 2: Comparison of attack losses achieved by ZO-Min-Maxand ZO-PGD.

Results. In Figure 1, we demonstrate the empirical con-vergence of ZO-Min-Max, in terms of the stationary gap‖G(x,y)‖2 given in (6) and the attack loss corresponding toeach model-class pair MjCi. Here M and C represents net-work model and image class, respectively. For comparison,we also present the convergence of FO-Min-Max. Figure 1-(a) shows that the stationary gap decreases as the iterationincreases, and converges to an iteration-independent biascompared with FO-Min-Max. In Figure 1-(b), we see thatZO-Min-Max yields slightly worse attack performance (interms of higher attack loss at each model-class pair) thanFO-Min-Max. However, it does not need to have access tothe configuration of the victim neural network models.

In Figure 2, we compare the attack loss of using ZO-Min-Max with that of using ZO-PGD, which solves problem(13) by calling for an internal maximization oracle in w(see Appendix D). As we can see, ZO-Min-Max converges

Page 8: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

100 101 102 103 104

Number of iterations

0

1

2

3

4

Stat

iona

ry g

ap

ZO-Min-Max: q=1ZO-Min-Max: q=5ZO-Min-Max: q=10ZO-Min-Max: q=20FO-Min-Max

100 101 102 103 104

Number of iterations

0.60

0.65

0.70

0.75

0.80

0.85

0.90

0.95

Test

ing

accu

racy

ZO-Min-MaxFO-Min-MaxNo Poison

100 101 102 103

Optimization time (seconds)

0.65

0.70

0.75

0.80

0.85

0.90

0.95

Test

ing

accu

racy

STABLEOPTZO-Min-Max

(c) (b) (c)

Figure 3: Empirical performance of ZO-Min-Max in design of poisoning attack: a) stationary gap versus iterations b) testing accuracyversus iterations (the shaded region represents variance of 10 random trials), and c) comparison between ZO-Min-Max and STABLEOPTon testing accuracy versus optimization time.

slower than ZO-PGD at early iterations. That is becausethe former only performs one-step PGA to update w, whilethe latter solves the w-maximization problem analytically.However, as the number of iterations increases, ZO-Min-Max achieves almost the same performance as ZO-PGD.

In Appendix D, we have provided additional comparisons onZO-Min-Max versus ZO-Finite-Sum, and versus per-imagePGD attack (Madry et al., 2018).

6.2. Black-box data poisoning attack

Data poisoning against logistic regression. Let D =zi, tini=1 denote the training dataset, among which n′ n samples are corrupted by a perturbation vector x, leadingto poisoned training data zi + x towards breaking the train-ing process and thus the prediction accuracy. The poisoningattack problem is then formulated as

maximize‖x‖∞≤ε

minimizeθ

g(x,θ) := Ftr(x,θ;D0) + λ‖θ‖22, (14)

where x and θ are optimization variables, Ftr(x,θ;D0)denotes the training loss over model parameters θ at thepresence of data poison x, and λ > 0 is a regularizationparameter. Note that problem (14) can be written in the formof (1), minx maxθ −g(x,θ). Clearly, if Ftr is a convexloss, e.g., logistic regression or linear regression (Jagielskiet al., 2018)), then −g is strongly concave in θ. Since theadversary has no knowledge on the training objective andthe benign training data, g(x,θ) is a two-sided black-boxfunction in both x and θ from the adversary’s perspective.

Implementation & Baseline. In problem (14), we set thepoisoning ratio n′/n = 15% and λ = 10−3 for problem(14), where the sensitivity of λ is studied in Figure A4. Moredetails on the specification of problem (14) are provided inAppendix E. In Algorithm A1, unless specified otherwisewe choose b = 100, q = 5, α = 0.02, β = 0.05, and T =50000. We report the empirical results averaged over 10independent trials with random initialization. We compare

our method with FO-Min-Max and the state-of-the-art BO-based method STABLEOPT (Bogunovic et al., 2018).

In Figure 3, we present the convergence of ZO-Min-Maxto generate data poisoning attack and validate the result-ing attack performance in terms of testing accuracy of thelogistic regression model trained on the poisoned dataset.Figure 3-(a) shows the stationary gap of ZO-Min-Max underdifferent number of random direction vectors in gradientestimation (2). As we can see, a moderate choice of q (e.g.,q ≥ 5 in our example) is sufficient to achieve near-optimalsolution compared with FO-Min-Max. However, it suf-fers from a convergence bias, consistent with Theorem 2.Figure 3-(b) demonstrates the testing accuracy (against it-erations) of the model learnt from poisoned training data,where the poisoning attack is generated by ZO-Min-Max(black-box attack) and FO-Min-Max (white-box attack). Aswe can see, ZO-Min-Max yields promising attacking perfor-mance comparable to FO-Min-Max. We can also see that bycontrast with the testing accuracy of the clean model (94%without poison), the poisoning attack eventually reduces thetesting accuracy (below 70%). Furthermore In Figure 3-(c),we compare ZO-Min-Max with STABLEOPT (Bogunovicet al., 2018) in terms of testing accuracy versus computa-tion time, where the lower the accuracy is, the stronger thegenerated attack is. We observe that STABLEOPT has apoorer scalability while our method reaches a data poison-ing attack that induces much lower testing accuracy within500 seconds. In Appendix E, we provide additional resultson the model learnt under different data poisoning ratios.

7. ConclusionThis paper addresses black-box robust optimization prob-lems given a finite number of function evaluations. In partic-ular, we present ZO-Min-Max: a framework of alternating,randomized gradient estimation based ZO optimization algo-rithm to find a first-order stationary solution to the black-boxmin-max problem. Under mild assumptions, ZO-Min-Max

Page 9: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

enjoys a sub-linear convergence rate. It scales to dimen-sions that are infeasible for recent robust solvers based onBayesian optimization. Furthermore, we experimentallydemonstrate the potential application of the framework ongenerating black-box evasion and data poisoning attacks.

ReferencesAggarwal, C., Bouneffouf, D., Samulowitz, H., Buesser, B.,

Hoang, T., Khurana, U., Liu, S., Pedapati, T., Ram, P.,Rawat, A., et al. How can ai automate end-to-end datascience? arXiv preprint arXiv:1910.14436, 2019.

Al-Dujaili, A., Hemberg, E., and O’Reilly, U.-M. Approxi-mating nash equilibria for black-box games: A bayesianoptimization approach. arXiv preprint arXiv:1804.10586,2018a.

Al-Dujaili, A., Huang, A., Hemberg, E., and OReilly, U.-M.Adversarial deep learning for robust detection of binaryencoded malware. In 2018 IEEE Security and PrivacyWorkshops (SPW), pp. 76–82. IEEE, 2018b.

Al-Dujaili, A., Srikant, S., Hemberg, E., and O’Reilly, U.-M.On the application of danskin’s theorem to derivative-freeminimax optimization. arXiv preprint arXiv:1805.06322,2018c.

Audet, C. and Hare, W. Derivative-free and blackbox opti-mization. Springer, 2017.

Balasubramanian, K. and Ghadimi, S. Zeroth-order (non)-convex stochastic optimization via conditional gradientand gradient updates. In Advances in Neural InformationProcessing Systems, pp. 3455–3464, 2018.

Berahas, A. S., Cao, L., Choromanski, K., and Scheinberg,K. A theoretical and empirical comparison of gradientapproximations in derivative-free optimization. arXivpreprint arXiv:1905.01332, 2019.

Bogunovic, I., Scarlett, J., Jegelka, S., and Cevher, V. Ad-versarially robust optimization with gaussian processes.In Proc. of Advances in Neural Information ProcessingSystems, pp. 5765–5775, 2018.

Boyd, S. and Vandenberghe, L. Convex optimization. Cam-bridge university press, 2004.

Branke, J. and Rosenbusch, J. New approaches to coevo-lutionary worst-case optimization. In International Con-ference on Parallel Problem Solving from Nature, pp.144–153. Springer, 2008.

Carlini, N. and Wagner, D. Towards evaluating the robust-ness of neural networks. In Security and Privacy (SP),2017 IEEE Symposium on, pp. 39–57. IEEE, 2017.

Chen, J., Yi, J., and Gu, Q. A frank-wolfe framework forefficient and effective adversarial attacks. arXiv preprintarXiv:1811.10828, 2018.

Chen, P.-Y., Zhang, H., Sharma, Y., Yi, J., and Hsieh, C.-J.Zoo: Zeroth order optimization based black-box attacksto deep neural networks without training substitute mod-els. In Proceedings of the 10th ACM Workshop on Artifi-cial Intelligence and Security, pp. 15–26. ACM, 2017.

Chen, T. and Giannakis, G. B. Bandit convex optimizationfor scalable and dynamic IoT management. IEEE Internetof Things Journal, 2018.

Chen, X., Liu, S., Sun, R., and Hong, M. On the conver-gence of a class of adam-type algorithms for non-convexoptimization. International Conference on Learning Rep-resentations, 2019.

Danskin, J. M. The theory of max-min, with applications.SIAM Journal on Applied Mathematics, 14(4):641–664,1966.

Deng, J., Dong, W., Socher, R., Li, L.-J., Li, K., and Fei-Fei,L. Imagenet: A large-scale hierarchical image database.In Computer Vision and Pattern Recognition, 2009. CVPR2009. IEEE Conference on, pp. 248–255. IEEE, 2009.

Dhurandhar, A., Pedapati, T., Balakrishnan, A., Chen, P.-Y., Shanmugam, K., and Puri, R. Model agnostic con-trastive explanations for structured data. arXiv preprintarXiv:1906.00117, 2019.

Duchi, J. C., Jordan, M. I., Wainwright, M. J., and Wibisono,A. Optimal rates for zero-order convex optimization: Thepower of two function evaluations. IEEE Transactionson Information Theory, 61(5):2788–2806, 2015.

Finlay, C. and Oberman, A. M. Scaleable input gradientregularization for adversarial robustness. arXiv preprintarXiv:1905.11468, 2019.

Flokas, L., Vlatakis-Gkaragkounis, E.-V., and Piliouras,G. Efficiently avoiding saddle points with zero or-der methods: No gradients required. arXiv preprintarXiv:1910.13021, 2019.

Gao, X., Jiang, B., and Zhang, S. On the information-adaptive variants of the ADMM: an iteration complexityperspective. Optimization Online, 12, 2014.

Ghadimi, S. and Lan, G. Stochastic first-and zeroth-ordermethods for nonconvex stochastic programming. SIAMJournal on Optimization, 23(4):2341–2368, 2013.

Ghadimi, S., Lan, G., and Zhang, H. Mini-batch stochasticapproximation methods for nonconvex stochastic com-posite optimization. Mathematical Programming, 155(1-2):267–305, 2016.

Page 10: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Gidel, G., Jebara, T., and Lacoste-Julien, S. Frank-WolfeAlgorithms for Saddle Point Problems. In Proceedingsof the 20th International Conference on Artificial Intel-ligence and Statistics, volume 54, pp. 362–371. PMLR,20–22 Apr 2017.

Golovin, D., Karro, J., Kochanski, G., Lee, C., Song, X.,et al. Gradientless descent: High-dimensional zeroth-order optimization. arXiv preprint arXiv:1911.06317,2019.

Goodfellow, I. J., Shlens, J., and Szegedy, C. Explain-ing and harnessing adversarial examples. arXiv preprintarXiv:1412.6572, 2014.

Hamedani, E. Y., Jalilzadeh, A., Aybat, N. S., and Shanbhag,U. V. Iteration complexity of randomized primal-dualmethods for convex-concave saddle point problems.arXiv preprint arXiv:1806.04118, 2018.

He, K., Zhang, X., Ren, S., and Sun, J. Deep residual learn-ing for image recognition. In Proceedings of the IEEEconference on computer vision and pattern recognition,pp. 770–778, 2016.

Herrmann, J. W. A genetic algorithm for minimax optimiza-tion problems. In CEC, volume 2, pp. 1099–1103. IEEE,1999.

Ilyas, A., Engstrom, L., Athalye, A., and Lin, J. Black-boxadversarial attacks with limited queries and information.arXiv preprint arXiv:1804.08598, 2018.

Ilyas, A., Engstrom, L., and Madry, A. Prior convictions:Black-box adversarial attacks with bandits and priors. InInternational Conference on Learning Representations,2019. URL https://openreview.net/forum?id=BkMiWhR5K7.

Jagielski, M., Oprea, A., Biggio, B., Liu, C., Nita-Rotaru,C., and Li, B. Manipulating machine learning: Poisoningattacks and countermeasures for regression learning. In2018 IEEE Symposium on Security and Privacy (SP), pp.19–35. IEEE, 2018.

Jensen, M. T. A new look at solving minimax problems withcoevolutionary genetic algorithms. In Metaheuristics:computer decision-making, pp. 369–384. Springer, 2003.

Jin, C., Netrapalli, P., and Jordan, M. I. Minmax optimiza-tion: Stable limit points of gradient descent ascent arelocally optimal. arXiv preprint arXiv:1902.00618, 2019.

Jones, E., Oliphant, T., Peterson, P., et al. SciPy: Opensource scientific tools for Python, 2001. URL http://www.scipy.org/.

Larson, J., Menickelly, M., and Wild, S. M. Derivative-free optimization methods. Acta Numerica, 28:287–404,2019.

Li, Y., Li, L., Wang, L., Zhang, T., and Gong, B. Nattack:Learning the distributions of adversarial examples foran improved black-box attack on deep neural networks.arXiv preprint arXiv:1905.00441, 2019.

Liu, J., Zhang, W., and Yu, N. Caad 2018: Iterative ensem-ble adversarial attack. arXiv preprint arXiv:1811.03456,2018a.

Liu, S., Chen, J., Chen, P.-Y., and Hero, A. O. Zeroth-orderonline admm: Convergence analysis and applications. InProceedings of the Twenty-First International Conferenceon Artificial Intelligence and Statistics, volume 84, pp.288–297, April 2018b.

Liu, S., Kailkhura, B., Chen, P.-Y., Ting, P., Chang, S., andAmini, L. Zeroth-order stochastic variance reduction fornonconvex optimization. In Proc. of Advances in NeuralInformation Processing Systems, 2018c.

Liu, S., Chen, P.-Y., Chen, X., and Hong, M. signSGD viazeroth-order oracle. In Proc. of International Conferenceon Learning Representations, 2019. URL https://openreview.net/forum?id=BJe-DsC5Fm.

Liu, Y., Chen, X., Liu, C., and Song, D. Delving intotransferable adversarial examples and black-box attacks.arXiv preprint arXiv:1611.02770, 2016.

Lu, S., Tsaknakis, I., and Hong, M. Block alternating opti-mization for non-convex min-max problems: Algorithmsand applications in signal processing and communica-tions. In Proc. of IEEE International Conference onAcoustics, Speech and Signal Processing, pp. 4754–4758,May 2019.

Lu, S., Tsaknakis, I., Hong, M., and Chen, Y. Hybridblock successive approximation for one-sided non-convexmin-max problems: Algorithms and applications. arXivpreprint arXiv:1902.08294, 2019.

Madry, A., Makelov, A., Schmidt, L., Tsipras, D., andVladu, A. Towards deep learning models resistant toadversarial attacks. ICLR, 2018.

Moosavi-Dezfooli, S.-M., Fawzi, A., Uesato, J., andFrossard, P. Robustness via curvature regularization,and vice versa. In Proceedings of the IEEE Conferenceon Computer Vision and Pattern Recognition, pp. 9078–9086, 2019.

Nesterov, Y. Dual extrapolation and its applications to solv-ing variational inequalities and related problems. Mathe-matical Programming, 109(2-3):319–344, 2007.

Page 11: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Nesterov, Y. and Spokoiny, V. Random gradient-free mini-mization of convex functions. Foundations of Computa-tional Mathematics, 2(17):527–566, 2015.

Nouiehed, M., Sanjabi, M., Lee, J. D., and Razaviyayn, M.Solving a class of non-convex min-max games using itera-tive first order methods. arXiv preprint arXiv:1902.08297,2019.

Parikh, N., Boyd, S., et al. Proximal algorithms. Founda-tions and Trends R© in Optimization, 1(3):127–239, 2014.

Picheny, V., Binois, M., and Habbal, A. A bayesian op-timization approach to find nash equilibria. Journal ofGlobal Optimization, 73(1):171–192, 2019.

Qian, Q., Zhu, S., Tang, J., Jin, R., Sun, B., and Li, H.Robust optimization over multiple domains. In Proceed-ings of the AAAI Conference on Artificial Intelligence,volume 33, pp. 4739–4746, 2019.

Rafique, H., Liu, M., Lin, Q., and Yang, T. Non-convex min-max optimization: Provable algorithms and applicationsin machine learning. arXiv preprint arXiv:1810.02060,2018.

Rios, L. M. and Sahinidis, N. V. Derivative-free optimiza-tion: a review of algorithms and comparison of softwareimplementations. Journal of Global Optimization, 56(3):1247–1293, 2013.

Sanjabi, M., Ba, J., Razaviyayn, M., and Lee, J. D. On theconvergence and robustness of training gans with regu-larized optimal transport. In Proceedings of the 32NdInternational Conference on Neural Information Process-ing Systems, pp. 7091–7101, 2018a.

Sanjabi, M., Ba, J., Razaviyayn, M., and Lee, J. D. Onthe convergence and robustness of training gans withregularized optimal transport. In Advances in NeuralInformation Processing Systems, pp. 7091–7101, 2018b.

Schmiedlechner, T., Al-Dujaili, A., Hemberg, E., andO’Reilly, U.-M. Towards distributed coevolutionary gans.arXiv preprint arXiv:1807.08194, 2018.

Shamir, O. An optimal algorithm for bandit and zero-orderconvex optimization with two-point feedback. Journal ofMachine Learning Research, 18(52):1–11, 2017.

Steinhardt, J., Koh, P. W. W., and Liang, P. S. Certifieddefenses for data poisoning attacks. In Advances in neuralinformation processing systems, pp. 3517–3529, 2017.

Szegedy, C., Vanhoucke, V., Ioffe, S., Shlens, J., and Wojna,Z. Rethinking the inception architecture for computervision. In IEEE Conference on Computer Vision andPattern Recognition (CVPR), pp. 2818–2826, 2016.

Tran, B., Li, J., and Madry, A. Spectral signatures in back-door attacks. In Advances in Neural Information Process-ing Systems, pp. 8000–8010, 2018.

Tu, C.-C., Ting, P., Chen, P.-Y., Liu, S., Zhang, H.,Yi, J., Hsieh, C.-J., and Cheng, S.-M. Autozoom:Autoencoder-based zeroth order optimization methodfor attacking black-box neural networks. arXiv preprintarXiv:1805.11770, 2018.

Wald, A. Statistical decision functions which minimize themaximum risk. Annals of Mathematics, pp. 265–280,1945.

Wang, B., Yao, Y., Shan, S., Li, H., Viswanath, B., Zheng,H., and Zhao, B. Y. Neural cleanse: Identifying andmitigating backdoor attacks in neural networks. NeuralCleanse: Identifying and Mitigating Backdoor Attacks inNeural Networks, pp. 0, 2019a.

Wang, C. and Wu, Q. Flo: Fast and lightweight hy-perparameter optimization for automl. arXiv preprintarXiv:1911.04706, 2019.

Wang, J., Zhang, T., Liu, S., Chen, P.-Y., Xu, J., Fardad,M., and Li, B. Towards a unified min-max framework foradversarial exploration and robustness, 2019b.

Ward, R., Wu, X., and Bottou, L. AdaGrad stepsizes: Sharpconvergence over nonconvex landscapes. In Proceed-ings of the 36th International Conference on MachineLearning, pp. 6677–6686, 2019.

Watson, R. A. and Pollack, J. B. Coevolutionary dynamicsin a minimal substrate. In Proceedings of the 3rd AnnualConference on Genetic and Evolutionary Computation,pp. 702–709. Morgan Kaufmann Publishers Inc., 2001.

Page 12: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Appendix

A. The ZO-Min-Max Algorithm

Algorithm A1 ZO-Min-Max to solve problem (1)

1: Input: given x(0) and y(0), learning rates α and β, the number of random directions q, and the possible mini-batch sizeb for stochastic optimization

2: for t = 1, 2, . . . , T do3: x-step: perform ZO-PGD (4)4: y-step:5: if f(x(t),y) is black box w.r.t. y then6: perform ZO-PGA (5)7: else8: perform PGA using∇yf(x(t),y(t−1)) like (5)9: end if

10: end for

B. Detailed Convergence AnalysisB.1. Table of Parameters

In Table A1, we summarize the problem and algorithmic parameters used in our convergence analysis.

Table A1: Summary of problem and algorithmic parameters and their descriptions.

parameter descriptiond # of optimization variablesb mini-batch sizeq # of random direction vectors used in ZO gradient estimationα learning rate for ZO-PGDβ learning rate for ZO-PGAγ strongly concavity parameter of f(x,y) with respect to yη upper bound on the gradient norm, implying Lipschitz continuity

Lx, Ly Lipschitz continuous gradient constant of f(x,y) with respect to x and y respectivelyR diameter of the compact convex set X or Yf∗ lower bound on the function value, implying feasibility

σ2x, σ2

y variances of ZO gradient estimator for variables x and y, bounded by (3)

B.2. Proof of Lemma 1

Before going into the proof, let’s review some preliminaries and give some definitions. Define hµ(x, ξ) to be the smoothedversion of h(x, ξ) and since ξ models a subsampling process over a finite number of candidate functions, we can furtherhave hµ(x) , Eξ[hµ(x, ξ)] and∇xhµ(x) = Eξ[∇xhµ(x, ξ)]

Recall that in the finite sum setting when ξj parameterizes the jth function, the gradient estimator is given by

∇xh(x) =1

bq

∑j∈I

q∑i=1

d[h(x + µui; ξj)− h(x; ξj)]

µui. (15)

where I is a set with b elements, containing the indices of functions selected for gradient evaluation.

From standard result of the zeroth order gradient estimator, we know

EI[Eui,i∈[q]

[∇xh(x)

] ∣∣I] = EI

1

b

∑j∈I∇xfµ(x, ξj)

= ∇xhµ(x). (16)

Page 13: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Now let’s go into the proof. First, we have

E[‖∇xh(x)−∇xhµ(x)‖22

]=EI

Eui,i∈[q]

∥∥∥∥∥∥∇xh(x)− 1

b

∑j∈I∇xfµ(x, ξj) +

1

b

∑j∈I∇xfµ(x, ξj)−∇xhµ(x)

∥∥∥∥∥∥2

2

∣∣∣∣I

≤2EI

Eui,i∈[q]

∥∥∥∥∥∥∇xh(x)− 1

b

∑j∈I∇xfµ(x, ξj)

∥∥∥∥∥∥2

2

+

∥∥∥∥∥∥1

b

∑j∈I∇xfµ(x, ξj)−∇xhµ(x)

∥∥∥∥∥∥2

2

∣∣∣∣I . (17)

Further, by definition, given I, ∇xh(x) is the average of ZO gradient estimates under q i.i.d. random directions, each ofwhich has the mean 1

b

∑j∈I ∇xfµ(x, ξj).

Thus for the first term at the right-hand-side (RHS) of the above inequality, we have

Eui,i∈[q]

∥∥∥∥∥∥∇xh(x)− 1

b

∑j∈I∇xfµ(x, ξj)

∥∥∥∥∥∥2

2

∣∣∣∣I ≤1

q

2d

∥∥∥∥∥∥1

b

∑j∈I∇xf(x, ξj)

∥∥∥∥∥∥2

+µ2L2

hd2

2

≤1

q

(2dη2 +

µ2L2hd

2

2

)(18)

where the first inequality is by the standard bound of the variance of zeroth order estimator and the second inequality is bythe assumption that ‖∇xh(x; ξ)‖2 ≤ η2 and thus ‖ 1b

∑j∈I ∇xf(x, ξj)‖2 ≤ η2.

In addition, we have

EI

Eui,i∈[q]

∥∥∥∥∥∥1

b

∑j∈I∇xfµ(x, ξj)−∇xhµ(x)

∥∥∥∥∥∥2

2

∣∣∣∣I

=EI

∥∥∥∥∥∥1

b

∑j∈I∇xfµ(x, ξj)−∇xhµ(x)

∥∥∥∥∥∥2

2

=

1

bEξ[‖∇xfµ(x, ξ)−∇xhµ(x)‖22

]≤ η2

b(19)

where the second equality is because ξj are i.i.d. draws from the same distribution as ξ and E[∇xfµ(x, ξ)] = ∇xhµ(x), thelast inequality is because ‖∇xfµ(x, ξ)‖22 ≤ η2 by assumption. Substituting (18) and (19) into (17) finishes the proof.

B.3. Convergence Analysis of ZO-Min-Max by Performing PGA

In this section, we will provide the details of the proofs. Before proceeding, we have the following illustration, which willbe useful in the proof.

The order of taking expectation: Since iterates x(t),y(t),∀t are random variables, we need to define

F (t) = x(t),y(t),x(t−1),y(t−1), . . . ,x(1),y(1) (20)

as the history of the iterates. Throughout the theoretical analysis, taking expectation means that we take expectation overrandom variable at the tth iteration conditioned on F (t−1) and then take expectation over F (t−1).

Page 14: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Subproblem: Also, it is worthy noting that performing (4) and (5) are equivalent to the following optimization problem:

x(t) = minx∈X

⟨∇xf(x(t−1),y(t−1)),x−x(t−1)

⟩+

1

2α‖x−x(t−1) ‖2, (21)

y(t) = maxy∈Y

⟨∇yf(x(t),y(t−1)),y−y(t−1)

⟩− 1

2β‖y−y(t−1) ‖2. (22)

When f(x,y) is white-box w.r.t. y, (22) becomes

y(t) = maxy∈Y

⟨∇yf(x(t),y(t−1)),y−y(t−1)

⟩− 1

2β‖y−y(t−1) ‖2. (23)

In the proof of ZO-Min-Max, we will use the optimality condition of these two problems to derive the descent lemmas.

Relationship with smoothing function We denote by fµ,x(x,y) the smoothing version of f w.r.t. x with parameterµ > 0. The similar definition holds for fµ,y(x,y). By taking fµ,x(x,y) as an example, under A2 f and fµ,x has thefollowing relationship (?)Lemma 4.1]gao2014information:

|fµ,x(x,y)− f(x,y))| ≤ Lxµ2

2and ‖∇xfµ,x(x,y)−∇xf(x,y)‖22 ≤

µ2d2L2x

4, (24)

|fµ,y(x,y)− f(x,y))| ≤ Lyµ2

2and ‖∇yfµ,y(x,y)−∇yf(x,y)‖22 ≤

µ2d2L2y

4. (25)

First, we will show the descent lemma in minimization as follows.

B.3.1. PROOF OF LEMMA 2

Proof: Since f(x,y) has Lx Lipschtiz continuous gradients with respect to x, we have

fµ(x(t+1),y(t)) ≤fµ(x(t),y(t)) + 〈∇xfµ(x(t),y(t)),x(t+1)−x(t)〉+Lx2‖x(t+1)−x(t) ‖2

=fµ(x(t),y(t)) + 〈∇xf(x(t),y(t)),x(t+1)−x(t)〉+Lx2‖x(t+1)−x(t) ‖2

+ 〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)),x(t+1)−x(t)〉. (26)

Recall that

x(t+1) = projX (x(t) − α∇xf(x(t),y(t))), (27)

From the optimality condition of x-subproblem (21), we have

〈∇xf(x(t),y(t)),x(t+1)−x(t)〉 ≤ − 1

α‖x(t+1)−x(t) ‖2. (28)

Here we use the fact that the optimality condition of problem (21) at the solution x(t+1) yields〈∇xf(x(t),y(t)) + (x(t+1)−x(t))/α,x(t+1)−x〉 ≤ 0 for any x ∈ X . By setting x = x(t), we obtain (28).

In addition, we define another iterate generated by∇xfµ(x(t),y(t))

x(t+1) = projX (x(t) − α∇xfµ(x(t),y(t))). (29)

Then, we can have

〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)),x(t+1)−x(t)〉

=〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)),x(t+1)−x(t)−(x(t+1) − x(t))〉

+ 〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)), x(t+1) − x(t)〉. (30)

Page 15: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Due to the fact that Eu[∇xf(x(t),y(t))] = ∇xfµ(x(t),y(t)), we further have

Eu[〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)), x(t+1) − x(t)〉] = 0. (31)

Finally, we also have

〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)),x(t+1)−x(t)−(x(t+1) − x(t))〉

≤α2‖∇xfµ(x(t),y(t))− ∇xf(x(t),y(t))‖2 +

1

2α‖x(t+1)−x(t)−(x(t+1) − x(t))‖2

≤α‖∇xfµ(x(t),y(t))− ∇xf(x(t),y(t))‖2 (32)

where the first inequality is due to Young’s inequality, the second inequality is due to non-expansiveness of the projectionoperator. Thus

Eu[〈∇xfµ(x(t),y(t))− ∇xf(x(t),y(t)),x(t+1)−x(t)−(x(t+1) − x(t))〉]

≤Eu[α‖∇xfµ(x(t),y(t))− ∇xf(x(t),y(t))‖2] ≤ ασ2x (33)

where σ2x := σ2(Lx, b, q, d) which was defined in (3).

Combining all above, we have

E[fµ(x(t+1),y(t))] ≤E[fµ(x(t),y(t))]−(

1

α− Lx

2

)‖x(t+1)−x(t) ‖2 + ασ2, (34)

and we request α ≤ 1/Lx, which completes the proof.

Using |fµ,x(x,y)− f(x,y))| ≤ Lxµ2

2 , we can get

E[f(x(t+1),y(t))]− Lxµ2

2≤ E[fµ(x(t+1),y(t))] ≤ E[f(x(t+1),y(t))] +

Lxµ2

2, (35)

so we are able to obtain from (3)

E[f(x(t+1),y(t))] ≤ E[f(x(t),y(t))]−(

1

α− Lx

2

)‖x(t+1)−x(t) ‖2 + ασ2

x + Lxµ2. (36)

Corollary 1.E⟨∇f(x(t),y(t−1))−∇fµ(x(t),y(t−1)),y(t)−y(t−1)

⟩≤ βσ2

y (37)

σ2y := σ2(Ly, b, q, d) which was defined in (3).

Proof:

Define

y(t) = projY(y(t) − β∇yfµ(x(t),y(t−1))), (38)

we have

〈∇yfµ(x(t),y(t−1))− ∇xf(x(t),y(t−1)),y(t)−y(t−1)〉

=〈∇yfµ(x(t),y(t−1))− ∇yf(x(t),y(t−1)),y(t)−y(t−1)−(y(t) − y(t−1))〉

+ 〈∇yfµ(x(t),y(t−1))− ∇yf(x(t),y(t−1)), y(t) − y(t−1)〉. (39)

Due to the fact that Eu[∇yf(x(t),y(t−1))] = ∇yfµ(x(t),y(t−1)), we further have

Eu[〈∇yfµ(x(t),y(t−1))− ∇yf(x(t),y(t−1)), y(t) − y(t−1)〉] = 0. (40)

Page 16: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Finally, we also have

Eu[〈∇yfµ(x(t),y(t−1))− ∇yf(x(t),y(t−1)),y(t)−y(t−1)−(y(t) − y(t−1))〉]

≤Eu[β

2‖〈∇yfµ(x(t),y(t−1))− ∇yf(x(t),y(t−1))‖2 +

1

2β‖y(t)−y(t−1)−(y(t) − y(t−1))‖2]

≤Eu[β‖∇yfµ(x(t),y(t−1))− ∇yf(x(t),y(t−1))‖2] ≤ βσ2y (41)

where σ2y := σ2(Ly, b, q, d) which was defined in (3).

Next, before showing the proof of Lemma 3, we need the following lemma to show the recurrence of the size of thesuccessive difference between two iterations.Lemma 5. Under assumption 1, assume iterates x(t),y(t) generated by algorithm A1. When f(x(t),y) is white-box, wehave

2

β2γE‖y(t+1)−y(t) ‖2 − 2

β2γE‖y(t)−y(t−1) ‖2 ≤ 2L2

x

βγ2E‖x(t+1)−x(t) ‖2

+2

βE‖y(t+1)−y(t) ‖2 −

(4

β−

2L2y

γ

)E‖y(t)−y(t−1) ‖2. (42)

Proof: from the optimality condition of y-subproblem (23) at iteration t and t− 1, we have the following two inequalities:

−〈∇yf(x(t+1),y(t))− 1

β(y(t+1)−y(t)),y(t+1)−y(t)〉 ≤0, (43)

〈∇yf(x(t),y(t−1))− 1

β(y(t)−y(t−1)),y(t+1)−y(t)〉 ≤0. (44)

Adding the above inequalities, we can get

1

β〈v(t+1),y(t+1)−y(t)〉 ≤

⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t)),y(t+1)−y(t)

⟩+⟨∇yf(x(t),y(t))−∇yf(x(t),y(t−1)),y(t+1)−y(t)

⟩(45)

where v(t+1) = y(t+1)−y(t)−(y(t)−y(t−1)).

According to the quadrilateral indentity, we know⟨v(t+1),y(t+1)−y(t)

⟩=

1

2

(‖y(t+1)−y(t) ‖2 + ‖v(t+1) ‖2 − ‖y(t)−y(t−1) ‖2

). (46)

Based on the definition of v(t+1), we substituting (46) into (45), which gives

1

2β‖y(t+1)−y(t) ‖2 ≤ 1

2β‖y(t)−y(t−1) ‖2 − 1

2β‖v(t+1) ‖2

+⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t)),y(t+1)−y(t)

⟩+⟨∇yf(x(t),y(t))−∇yf(x(t),y(t−1)),y(t+1)−y(t)

⟩(47)

(a)

≤ 1

2β‖y(t)−y(t−1) ‖2 +

⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t)),y(t+1)−y(t)

⟩+βL2

y

2‖y(t)−y(t−1) ‖2 − γ‖y(t)−y(t−1) ‖2

(b)

≤ 1

2β‖y(t)−y(t−1) ‖2 +

γ

2‖y(t+1)−y(t) ‖2

+L2x

2γ‖x(t+1)−x(t) ‖2 − (γ −

βL2y

2)‖y(t)−y(t−1) ‖2 (48)

Page 17: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

where in (a) we use the strong concavity of function f(x,y) in y (with parameter γ > 0) and Young’s inequality, i.e.,

〈∇yf(x(t),y(t))−∇yf(x(t),y(t−1)),y(t+1)−y(t)〉=〈∇yf(x(t),y(t))−∇yf(x(t),y(t−1)),v(t+1) +y(t)−y(t−1)〉

≤βL2

y

2‖y(t)−y(t−1) ‖2 +

1

2β‖v(t+1) ‖2 − γ‖y(t)−y(t−1) ‖2 (49)

and in (b) we apply the Young’s inequality, i.e.,

⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t)),y(t+1)−y(t)

⟩≤ L2

x

2γ‖x(t+1)−x(t) ‖2 +

γ

2‖y(t+1)−y(t) ‖2. (50)

Therefore, we have

1

2β‖y(t+1)−y(t) ‖2 ≤ 1

2β‖y(t)−y(t−1) ‖2 +

L2x

2γ‖x(t+1)−x(t) ‖2

2‖y(t+1)−y(t) ‖2 −

(γ −

βL2y

2

)‖y(t)−y(t−1) ‖2, (51)

which implies

2

β2γ‖y(t+1)−y(t) ‖2 ≤ 2

β2γ‖y(t)−y(t−1) ‖2 +

2L2x

βγ2‖x(t+1)−x(t) ‖2

+2

β‖y(t+1)−y(t) ‖2 −

(4

β−

2L2y

γ

)‖y(t)−y(t−1) ‖2. (52)

By taking the expectation on both sides of (52), we can get the results of Lemma 5.

Lemma 5 basically gives the recursion of ‖∆(t)y ‖2. It can be observed that term (4/β − 2L2

y/γ)‖∆(t)y ‖ provides the descent

of the recursion when β is small enough, which will take an important role in the proof of Lemma 3 when we quantify thedescent in maximization.

Then, we can quantify the descent of the objective value by the following descent lemma.

B.3.2. PROOF OF LEMMA 3

Proof: let f ′(x(t+1),y(t+1)) = f(x(t+1),y(t+1))− 1(y(t+1)) and 1(y) denote the indicator function with respect to theconstraint of y. From the optimality condition of sub-problem y in (22), we have

∇yf(x(t+1),y(t))− 1

β(y(t+1)−y(t))− ξ(t+1) = 0 (53)

where ξ(t) denote the subgradient of 1(y(t)). Since function f ′(x,y) is concave with respect to y, we have

f ′(x(t+1),y(t+1))− f ′(x(t+1),y(t)) ≤ 〈∇yf(x(t+1),y(t)),y(t+1)−y(t)〉 − 〈ξ(t),y(t+1)−y(t)〉(a)=

1

β‖y(t+1)−y(t) ‖2 − 〈ξ(t) − ξ(t+1),y(t+1)−y(t)〉

=1

β‖y(t+1)−y(t) ‖2 +

⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t−1)),y(t+1)−y(t)

⟩− 1

β

⟨v(t+1),y(t+1)−y(t)

⟩(54)

where in (a) we use ξ(t+1) = ∇yf(x(t+1),y(t))− 1β (y(t+1)−y(t)) .

Page 18: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

The last two terms of (54) is the same as the RHS of (45). We can apply the similar steps from (47) to (48). To be morespecific, the derivations are shown as follows: First, we know

f ′(x(t+1),y(t+1))− f ′(x(t+1),y(t)) ≤ 1

β‖y(t+1)−y(t) ‖2

+⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t−1)),y(t+1)−y(t)

⟩− 1

β

⟨v(t+1),y(t+1)−y(t)

⟩. (55)

Then, we move term 1/β〈v(t+1),y(t+1)−y(t)〉 to RHS of (54) and have

f(x(t+1),y(t+1))− f(x(t+1),y(t))

≤ 1

2β‖y(t+1)−y(t) ‖2 +

1

2β‖y(t)−y(t−1) ‖2 − 1

2β‖v(t+1) ‖2

+⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t)),y(t+1)−y(t)

⟩+⟨∇yf(x(t),y(t))−∇yf(x(t),y(t−1)),y(t+1)−y(t)

⟩≤ 1

2β‖y(t+1)−y(t) ‖2 +

⟨∇yf(x(t+1),y(t))−∇yf(x(t),y(t)),y(t+1)−y(t)

⟩+βL2

y

2‖y(t)−y(t−1) ‖2 − γ‖y(t)−y(t−1) ‖2

(a)

≤ 1

β‖y(t+1)−y(t) ‖2 +

1

2β‖y(t)−y(t−1) ‖2

+βL2

x

2‖x(t+1)−x(t) ‖2 − (γ −

βL2y

2)‖y(t)−y(t−1) ‖2 (56)

where in (a) we use

〈∇yf(x(t+1),y(t))−∇yf(x(t),y(t))〉 ≤ βL2x

2‖x(t+1)−x(t) ‖2 +

1

2β‖y(t+1)−y(t) ‖2 (57)

which is different from (50); also y(t),y(t+1) ∈ Y so have f ′(x(t+1),y(t+1)) = f(x(t+1),y(t+1)) and f ′(x(t+1),y(t)) =f(x(t+1),y(t)).

Combing (52), we have

f(x(t+1),y(t+1)) +

(2

β2γ+

1

)‖y(t+1)−y(t) ‖2 − 4

(1

β−L2y

)‖y(t+1)−y(t) ‖2

≤f(x(t+1),y(t)) +

(2

β2γ+

1

)‖y(t)−y(t−1) ‖2 − 4

(1

β−L2y

)‖y(t)−y(t−1) ‖2

(1

2β−

2L2y

γ

)‖y(t+1)−y(t) ‖2 +

(2L2

x

γ2β+βL2

x

2

)‖x(t+1)−x(t) ‖2. (58)

By taking the expectation on both sides of (52), we can get the results of Lemma 3.

Next, we use the following lemma to show the descent of the objective value after solving x-subproblem by (4).

B.3.3. PROOF OF THEOREM 1

Proof:

Page 19: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

From Lemma 3, we know

E[f(x(t+1),y(t+1))] +

(2

β2γ+

1

)E[‖y(t+1)−y(t) ‖2]

− 4

(1

β−L2y

)E[‖y(t+1)−y(t) ‖2] ≤ E[f(x(t+1),y(t))]

+

(2

β2γ+

1

)E[‖y(t)−y(t−1) ‖2]− 4

(1

β−L2y

)E[‖y(t)−y(t−1) ‖2]

(1

2β−

2L2y

γ

)E[‖y(t+1)−y(t) ‖2] +

(2L2

x

γ2β+βL2

x

2

)E[‖x(t+1)−x(t) ‖2]. (59)

Combining Lemma 2, we have

E[f(x(t+1),y(t+1))] +

(2

β2γ+

1

)E[‖y(t+1)−y(t) ‖2

]− 4

(1

β−L2y

)E[‖y(t+1)−y(t) ‖2

]≤ E[f(x(t),y(t))] +

(2

β2γ+

1

)E[‖y(t)−y(t−1) ‖2

]− 4

(1

β−L2y

)E[‖y(t)−y(t−1) ‖2

]−

(1

2β−

2L2y

γ

)︸ ︷︷ ︸

c1

E[‖y(t+1)−y(t) ‖2

]

−(

1

α−(Lx2

+2L2

x

γ2β+βL2

x

2

))︸ ︷︷ ︸

c2

E[‖x(t+1)−x(t) ‖2

]+ ασ2

x + Lxµ2. (60)

If

β <γ

4L2y

and α <1

Lx

2 +2L2

x

γ2β +βL2

x

2

, (61)

then we have that there exist positive constants c1 and c2 such that

P(x(t+1),y(t+1),∆(t+1)y )− P(x(t),y(t),∆(t)

y )

≤− c1E[‖y(t+1)−y(t) ‖2

]− c2E

[‖x(t+1)−x(t) ‖2

]+ ασ2

x + Lxµ2

≤− ζ(E[‖y(t+1)−y(t) ‖2

]+ E

[‖x(t+1)−x(t) ‖2

])+ ασ2

x + Lxµ2 (62)

where ζ = minc1, c2.

Page 20: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

From (6), we can have

‖G(x(t),y(t))‖

≤ 1

α‖x(t+1)−x(t) ‖+

1

α‖x(t+1)−projX (x(t)−α∇xf(x(t),y(t)))‖+

1

β‖y(t+1)−y(t) ‖

+1

β‖y(t+1)−projY(y(t) +β∇yf(x(t),y(t))‖

(a)

≤ 1

α‖x(t+1)−x(t) ‖

+1

α‖projX (x(t+1)−α(∇xf(x(t),y(t)) +

1

α(x(t+1)−x(t)))− projX (x(t)−α∇xf(x(t),y(t)))‖

+1

β‖y(t+1)−y(t) ‖

+1

β‖projY(y(t+1) +β(∇yf(x(t+1),y(t))− 1

β(y(t+1)−y(t)))− projY(y(t) +β∇yf(x(t),y(t)))‖

(b)

≤ 3

α‖x(t+1)−x(t) ‖+ ‖∇yf(x(t+1),y(t)))−∇yf(x(t),y(t)))‖+

3

β‖y(t+1)−y(t) ‖

(c)

≤(

3

α+ Lx

)‖x(t+1)−x(t) ‖+

3

β‖y(t+1)−y(t) ‖

where in (a) we use x(t+1) = projX (x(t+1)−α∇f(x(t+1),y(t))− (x(t+1)−x(t))); in (b) we use nonexpansiveness of theprojection operator; in (c) we apply the Lipschitz continuous of function f(x,y) with respect to x and y under assumptionA2.

Therefore, we can know that there exist a constant c = maxLx + 3α ,

3β such that

‖G(x(t),y(t))‖2 ≤ c(‖x(t+1)−x(t) ‖2 + ‖y(t+1)−y(t) ‖2

). (63)

After applying the telescope sum on (62) and taking expectation over (63), we have

1

T

T∑t=1

E‖G(x(t),y(t))‖2 ≤ c

ζ

(P1 − PT+1

T+ ασ2

x + Lxµ2

). (64)

Recall from A1 that f ≥ f∗ and Y is bounded with diameter R, therefore, Pt given by (8) yields

Pt ≥ f∗ +

(min4 + 4β2L2

y − 7βγ, 02β2γ

)R2, ∀t. (65)

And let (x(r),y(r)) be uniformly and randomly picked from (x(t),y(t))Tt=1, based on (64) and (65), we obtain

Er[E‖G(x(r),y(r))‖2] =1

T

T∑t=1

E‖G(x(t),y(t))‖2 ≤ c

ζ

(P1 − f∗ − νR2

T+ ασ2

x + Lxµ2

), (66)

where recall that ζ = minc1, c2, c = maxLx + 3α ,

3β and ν =

min4+4β2L2y−7βγ,0

2β2γ .

The proof is now complete.

B.4. Convergence Analysis of ZO-Min-Max by Performing ZO-PGA

Before showing the proof of Lemma 4, we first give the following lemma regarding to recursion of the difference betweentwo successive iterates of variable y.

Page 21: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Lemma 6. Under assumption 1, assume iterates x(t),y(t) generated by algorithm A1. When function f(x(t),y) is black-box,we have

2

β2γE‖y(t+1)−y(t) ‖2 ≤ 2

β2γE‖y(t)−y(t−1) ‖2 +

2

βE‖y(t+1)−y(t) ‖2

+6L2

y

βγ2E‖x(t+1)−x(t) ‖2 −

(4

β−

6L2y + 4

γ

)E‖y(t)−y(t−1) ‖2

+4σ2

y

βγ

(3

γ+ 4β

)+µ2d2L2

y

β2γ. (67)

From the optimality condition of y-subproblem in (22) at iteration t and t− 1, we have

−⟨∇yf(x(t+1),y(t))− 1

β(y(t+1)−y(t)),y(t+1)−y(t)

⟩≤0, (68)⟨

∇yf(x(t),y(t−1))− 1

β(y(t)−y(t−1)),y(t+1)−y(t)

⟩≤0. (69)

Adding the above inequalities and applying the definition of v(t+1), we can get

1

β〈v(t+1),y(t+1)−y(t)〉 ≤

⟨∇yf(x(t+1),y(t))− ∇yf(x(t),y(t)),y(t+1)−y(t)

⟩︸ ︷︷ ︸

I

+⟨∇yf(x(t),y(t))− ∇yf(x(t),y(t−1)),y(t+1)−y(t)

⟩︸ ︷︷ ︸

II

. (70)

Next, we will bound E[I] and E[II] separably as follows.

First, we give an upper bound of E[I] as the following,

E⟨∇yf(x(t+1),y(t))− ∇yf(x(t),y(t)),y(t+1)−y(t)

⟩≤ 3

2γE‖∇yf(x(t+1),y(t))−∇yfµ,y(x(t+1),y(t))‖2 +

γ

6E‖y(t+1)−y(t) ‖2

+3

2γE‖∇yfµ,y(x(t+1),y(t))−∇yfµ,y(x(t),y(t))‖2 +

γ

6E‖y(t+1)−y(t) ‖2

+3

2γE‖∇yfµ,y(x(t),y(t))− ∇fy(x(t),y(t))‖2 +

γ

6E‖y(t+1)−y(t) ‖2

≤3σ2

y

γ+

3L2x

2γE‖x(t+1)−x(t) ‖2 +

γ

2E‖y(t+1)−y(t) ‖2 (71)

where Lemma 1 is used.

Page 22: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Second, we need to give an upper bound of E[II] as follows:

〈∇f(x(t),y(t))− ∇f(x(t),y(t−1)),y(t+1)−y(t)〉

=〈∇f(x(t),y(t))− ∇f(x(t),y(t−1)),v(t+1) +y(t)−y(t−1)〉

=⟨∇f(x(t),y(t))−∇f(x(t),y(t−1)),y(t)−y(t−1)

⟩+⟨∇fµ,y(x(t),y(t))−∇f(x(t),y(t)),y(t)−y(t−1)

⟩+⟨∇f(x(t),y(t))−∇fµ,y(x(t),y(t)),y(t)−y(t−1)

⟩−⟨∇fµ,y(x(t),y(t−1))−∇f(x(t),y(t−1)),y(t)−y(t−1)

⟩−⟨∇f(x(t),y(t−1))−∇fµ,y(x(t),y(t−1)),y(t)−y(t−1)

⟩+ 〈∇f(x(t),y(t))− ∇f(x(t),y(t−1)),v(t+1)〉.

Next, we take expectation on both sides of the above equality and obtain

E〈∇f(x(t),y(t))− ∇f(x(t),y(t−1)),y(t+1)−y(t)〉(a)

(3βL2

y

2+ β

)‖y(t)−y(t−1) ‖2 +

1

2β‖v(t+1) ‖2 − γ‖y(t)−y(t−1) ‖2

+µ2d2L2

y

4β+ 4βσ2

y (72)

where in (a) we use the fact that 1) γ-strong concavity of f with respect to y:⟨∇f(x(t),y(t))−∇f(x(t),y(t−1)),y(t)−y(t−1)

⟩≤ −γ‖y(t)−y(t−1) ‖2; (73)

and the facts that 2) smoothing property (25) and Young’s inequality

E⟨∇fµ,y(x(t),y(t))−∇f(x(t),y(t)),y(t)−y(t−1)

⟩≤µ2d2L2

y

8β+β

2‖y(t)−y(t−1) ‖2; (74)

and the fact that 3) the ZO estimator is unbiased according to Lemma 1

E⟨∇f(x(t),y(t))−∇fµ,y(x(t),y(t)),y(t)−y(t−1)

⟩= 0; (75)

and

E⟨∇fµ,y(x(t),y(t−1))−∇f(x(t),y(t−1)),y(t)−y(t−1)

⟩≤µ2d2L2

y

8β+β

2‖y(t)−y(t−1) ‖2; (76)

and from Corollary 1 we have

E⟨∇f(x(t),y(t−1))−∇fµ,y(x(t),y(t−1)),y(t)−y(t−1)

⟩≤ βσ2

y; (77)

and

E〈∇f(x(t),y(t))− ∇f(x(t),y(t−1)),v(t+1)〉

≤3β

2E‖∇fµ,y(x(t),y(t))− ∇f(x(t),y(t))‖2 +

1

6β‖v(t+1) ‖2

+3β

2E‖∇fµ,y(x(t),y(t))−∇fµ,y(x(t),y(t−1))‖2 +

1

6β‖v(t+1) ‖2

+3β

2E‖∇fµ,y(x(t),y(t−1))− ∇f(x(t),y(t−1))‖2 +

1

6β‖v(t+1) ‖2

≤3βσ2y +

1

2β‖v(t+1) ‖2 +

3βL2y

2‖y(t)−y(t−1) ‖2. (78)

Page 23: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

Then, from (70), we can have

1

2βE‖y(t+1)−y(t) ‖2 ≤ 1

2βE‖y(t)−y(t−1) ‖2 − 1

2βE‖v(t+1) ‖2

+3σ2

y

γ+

3L2x

2γE‖x(t+1)−x(t) ‖2 +

γ

2E‖y(t+1)−y(t) ‖2

+⟨∇f(x(t),y(t))− ∇f(x(t),y(t−1)),y(t+1)−y(t)

⟩≤ 1

2βE‖y(t)−y(t−1) ‖2 +

γ

2E‖y(t+1)−y(t) ‖2

+3L2

y

2γE‖x(t+1)−x(t) ‖2 −

(γ −

(3βL2

y

2+ β

))E‖y(t)−y(t−1) ‖2

+3σ2

y

γ+ 4βσ2

y +µ2d2L2

y

4β, (79)

which implies

2

β2γE‖y(t+1)−y(t) ‖2 ≤ 2

β2γE‖y(t)−y(t−1) ‖2 +

2

βE‖y(t+1)−y(t) ‖2

+6L2

y

βγ2E‖x(t+1)−x(t) ‖2 −

(4

β−

6L2y + 4

γ

)E‖y(t)−y(t−1) ‖2

+4σ2

y

βγ

(3

γ+ 4β

)+µ2d2L2

y

β2γ. (80)

B.4.1. PROOF OF LEMMA 4

Proof: Similarly as B.3.2, let f ′(x(t+1),y(t+1)) = f(x(t+1),y(t+1))− 1(y(t+1)), 1(·) denotes the indicator function andξ(t) denote the subgradient of 1(y(t)). Since function f ′(x,y) is concave with respect to y, we have

f ′(x(t+1),y(t+1))− f ′(x(t+1),y(t)) ≤ 〈∇f(x(t+1),y(t)),y(t+1)−y(t)〉 − 〈ξ(t),y(t+1)−y(t)〉(a)=

1

β‖y(t+1)−y(t) ‖2 − 〈ξ(t) − ξ(t+1),y(t+1)−y(t)〉

=1

β‖y(t+1)−y(t) ‖2 +

⟨∇f(x(t+1),y(t))− ∇f(x(t),y(t−1)),y(t+1)−y(t)

⟩− 1

β

⟨v(t+1),y(t+1)−y(t)

⟩(81)

where in (a) we use ξ(t+1) = ∇f(x(t+1),y(t))− 1β (y(t+1)−y(t)). Then, we have

Ef(x(t+1),y(t+1))− Ef(x(t+1),y(t)) +1

β

⟨v(t+1),y(t+1)−y(t)

⟩≤ 1

β‖y(t+1)−y(t) ‖2 +

⟨∇f(x(t+1),y(t))− ∇f(x(t),y(t−1)),y(t+1)−y(t)

⟩.

Applying the steps from (72) to (79), we can have

Ef(x(t+1),y(t+1))− Ef(x(t+1),y(t))

≤ 1

βE‖y(t+1)−y(t) ‖2 +

1

2βE‖y(t)−y(t−1) ‖2 −

(γ −

(3βL2

y

2+ β

))‖y(t)−y(t−1) ‖2

+3βL2

x

2E‖x(t+1)−x(t) ‖2 + 7βσ2

y +µ2d2L2

y

4β(82)

Page 24: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

where we use

E⟨∇yf(x(t+1),y(t))− ∇yf(x(t),y(t)),y(t+1)−y(t)

⟩≤3βσ2

y +3βL2

x

2E‖x(t+1)−x(t) ‖2 +

1

2βE‖y(t+1)−y(t) ‖2. (83)

Combing (80), we have

Ef(x(t+1),y(t+1)) +

(2

β2γ+

1

)E‖y(t+1)−y(t) ‖2 −

(4

β−

6L2y + 4

γ

)E‖y(t+1)−y(t) ‖2

≤Ef(x(t+1),y(t)) +

(2

β2γ+

1

)E‖y(t)−y(t−1) ‖2 −

(4

β−

6L2y + 4

γ

)E‖y(t)−y(t−1) ‖2

(1

2β−

6L2y + 4

γ

)E‖y(t+1)−y(t) ‖2 +

(6L2

x

γ2β+

3βL2x

2

)E‖x(t+1)−x(t) ‖2.

+µ2d2L2

y

β

(1

4+

1

βγ

)+

(7β +

4

βγ

(3

γ+ 7β

))σ2y. (84)

B.4.2. PROOF OF THEOREM 2

Proof: From (36), we know the “descent” of the minimization step, i.e., the changes from P ′(x(t),y(t),∆(t)y ) to

P ′(x(t+1),y(t),∆(t)y ).

Combining the “descent” of the maximization step by Lemma 4 shown in (84), we can obtain the following:

P ′(x(t+1),y(t+1),∆(t+1)y )

≤P ′(x(t),y(t),∆(t)y )−

(1

2β−

6L2y + 4

γ

)︸ ︷︷ ︸

a1

E[‖y(t+1)−y(t) ‖2

](85)

−(

1

α−(Lx2

+6L2

x

γ2β+

3βL2x

2

))︸ ︷︷ ︸

a2

E[‖x(t+1)−x(t) ‖2

]

+ µ2

(Lx +

d2L2y

β

(1

4+

1

βγ

))︸ ︷︷ ︸

b1

+ασ2x +

(7β +

4

βγ

(3

γ+ 4β

))︸ ︷︷ ︸

b2

σ2y.

When β, α satisfy the following conditions:

β <γ

4(3L2y + 2)

, and α <1

Lx

2 +6L2

x

γ2β +3βL2

x

2

, (86)

we can conclude that there exist b1, b2 > 0 such that

P ′(x(t+1),y(t+1),∆(t+1)y )

≤P ′(x(t),y(t),∆(t)y )− a1E

[‖y(t+1)−y(t) ‖2

]− a2

[‖x(t+1)−x(t) ‖2

]+ b1µ

2 + ασ2x + b2σ

2y

≤− ζ ′E[‖y(t+1)−y(t) ‖2 + ‖x(t+1)−x(t) ‖2

]+ b1µ

2 + ασ2x + b2σ

2 (87)

Page 25: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

where ζ ′ = mina1, a2.

From (6), we can have

E‖G(x(t),y(t))‖

≤ 1

αE‖x(t+1)−x(t) ‖+

1

αE‖x(t+1)−projX (x(t)−α∇xf(x(t),y(t)))‖

+1

βE‖y(t+1)−y(t) ‖+

1

βE‖y(t+1)−projY(y(t) +β∇yf(x(t),y(t))‖

(a)

≤ 1

αE‖x(t+1)−x(t) ‖+

1

βE‖y(t+1)−y(t) ‖

+1

αE‖projX (x(t+1)−α(∇xf(x(t),y(t)) +

1

α(x(t+1)−x(t)))− projX (x(t)−α∇xf(x(t),y(t)))‖

+1

βE‖projY(y(t+1) +β(∇yf(x(t+1),y(t))− 1

β(y(t+1)−y(t)))− projY(y(t) +β∇yf(x(t),y(t)))‖

(b)

≤ 3

αE‖x(t+1)−x(t) ‖+ E‖∇xf(x(t),y(t)))−∇xf(x(t),y(t)))‖

+3

βE‖y(t+1)−y(t) ‖+ E‖∇yf(x(t+1),y(t))−∇yf(x(t),y(t))‖

≤ 3

αE‖x(t+1)−x(t) ‖+ E‖∇xf(x(t),y(t)))−∇xfµ,y(x(t),y(t)))‖

+ E‖∇xfµ,y(x(t),y(t)))−∇xf(x(t),y(t)))‖

+3

βE‖y(t+1)−y(t) ‖+ E‖∇yf(x(t+1),y(t))−∇yfµ,y(x(t+1),y(t))‖

+ E‖∇yfµ,y(x(t+1),y(t))−∇yfµ,y(x(t),y(t))‖+ E‖∇yfµ,y(x(t),y(t))−∇yf(x(t),y(t))‖

(c)

≤(

3

α+ Lx

)E‖x(t+1)−x(t) ‖+

3

βE‖y(t+1)−y(t) ‖+ 2σ2

y + µ2d2L2y

where in (a) we use the optimality condition of x(t)-subproblem; in (b) we use nonexpansiveness of the projection operator;in (c) we apply the Lipschitz continuous of function f(x,y) under assumption A2.

Therefore, we can know that

E[‖G(x(t),y(t))‖2

]≤ c

(‖x(t+1)−x(t) ‖2 + ‖y(t+1)−y(t) ‖2

)+ 2σ2

y + µ2d2L2y. (88)

After applying the telescope sum on (87) and taking expectation over (88), we have

1

T

T∑t=1

E[‖G(x(t),y(t))‖2

]≤ c

ζ ′P1 − PT+1

T+cb1ζ ′µ2 +

cασ2x

ζ ′+cb2ζ ′σ2y + 2σ2

y + µ2d2L2y. (89)

Recall from A1 that f ≥ f∗ and Y is bounded with diameter R, therefore, Pt given by (11) yields

Pt ≥ f∗ +

(min4 + 4(3L2

y + 2)β2 − 7βγ, 0β2γ

)R2, ∀t. (90)

And let (x(r),y(r)) be uniformly and randomly picked from (x(t),y(t))Tt=1, based on (90) and (89), we obtain

Er[E[‖G(x(r),y(r))‖2

]]=

1

T

T∑t=1

E[‖G(x(t),y(t))‖2

]≤ c

ζ ′P1 − f∗ − ν′R2

T+cb1ζ ′µ2 +

cασ2x

ζ ′+cb2ζ ′σ2y + 2σ2

y + µ2d2L2y (91)

Page 26: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

where recall that ζ ′ = mina1, a2, c = maxLx + 3α ,

3β , and ν′ =

min4+4(3L2y+2)β2−7βγ,0β2γ .

The proof is now complete.

Page 27: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

C. Toy Example in (Bogunovic et al., 2018): ZO-Min-Max versus BOWe review the example in (Bogunovic et al., 2018) as below,

maximizex∈C

minimize‖δ‖2≤0.5

f(x− δ) := −2(x1 − δ1)6 + 12.2(x1 − δ1)5 − 21.2(x1 − δ1)4

−6.2(x1 − δ1) + 6.4(x1 − δ1)3 + 4.7(x1 − δ1)2 − (x2 − δ2)6

+11(x2 − δ2)5 − 43.3(x2 − δ2)4 + 10(x2 − δ2) + 74.8(x2 − δ2)3

−56.9(x2 − δ2)2 + 4.1(x1 − δ1)(x2 − δ2) + 0.1(x1 − δ1)2(x2 − δ2)2

−0.4(x2 − δ2)2(x1 − δ1)− 0.4(x1 − δ1)2(x2 − δ2),

(92)

where x ∈ R2, and C = x1 ∈ (−0.95, 3.2), x2 ∈ (−0.45, 4.4).

Problem (92) can be equivalently transformed to the min-max setting consistent with ours

minimizex∈C

maximize‖δ‖2≤0.5

−f(x− δ). (93)

The optimality of solving problem (92) is measured by regret versus iteration t,

Regret(t) = minimize‖δ‖2≤0.5

f(x∗ − δ)−minimize‖δ‖2≤0.5

f(x(t) − δ), (94)

where minimize‖δ‖2≤0.5 f(x∗ − δ) = −4.33 and x∗ = [−0.195, 0.284]T (Bogunovic et al., 2018).

In Figure A1, we compare the convergence performance and computation time of ZO-Min-Max with the BO based approachSTABLEOPT proposed in (Bogunovic et al., 2018). Here we choose the same initial point for both ZO-Min-Max andSTABLEOPT. And we set the same number of function queries per iteration for ZO-Min-Max (with q = 1) and STABLEOPT.We recall from (2) that the larger q is, the more queries ZO-Min-Max takes. In our experiments, we present the best achievedregret up to time t and report the average performance of each method over 5 random trials. As we can see, ZO-Min-Max ismore stable, with lower regret and less running time. Besides, as q becomes larger, ZO-Min-Max has a faster convergencerate. We remark that BO is slow since learning the accurate GP model and solving the acquisition problem takes intensivecomputation cost.

0 10 20 30 40 50Number of iterations

0

20

40

60

80

100

120

Regr

et

ZO-Min-Max q=1ZO-Min-Max q=5ZO-Min-Max q=10STABLEOPToptimal

0 10 20 30 40 50Number of iterations

100

101

102

103

104

Tota

l tim

e ZO-Min-Max q=1ZO-Min-Max q=5ZO-Min-Max q=10STABLEOPT

(a) (b)

Figure A1: Comparison of ZO-Min-Max against STABLEOPT (Bogunovic et al., 2018): a) Convergence performance; b) Computationtime (seconds).

Page 28: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

D. Additional Details on Ensemble Evasion AttackExperiment setup. We specify the attack loss Fij in (13) as the C&W untargeted attack loss (Carlini & Wagner, 2017),

Fij (x; Ωi) = (1/|Ωi|)∑z∈Ωi

maxgj(z + x)i −maxk 6=i

gj(z + x)k, 0, (95)

where |Ωi| is the cardinality of the set Ωi, gj(z + x)k denotes the prediction score of class k given the input z + xusing model j. In (13), the regularization parameter λ trikes a balance between the worse-case attack loss andthe average loss (Wang et al., 2019b). The rationale behind that is from two manifolds. First, as γ = 0, thenmaximizew∈W

∑Jj=1

∑Ii=1 [wijFij (x; Ωi)] = Fi∗j∗ (x; Ωi), where wi∗j∗ = 1 and 0s for (i, j) 6= (i∗, j∗), and

(i∗, j∗) = arg maxi,j Fij (x; Ωi) given x. On the other hand, as γ →∞, then w→ 1/(IJ).

Implementation of ZO-PGD for solving problem (13). To solve problem (13), the baseline method ZO-PGD performssingle-objective ZO minimization under the equivalent form of (13), minx∈X h(x), where h(x) = maxw∈W f(x,w). Itis worth noting that we report the best convergence performance of ZO-PGD by searching its learning rate over 5 gridpoints in [0.01, 0.05]. Also, when querying the function value of h (at a given point x), we need the solution to the innermaximization problem

maximizew∈W

J∑j=1

I∑i=1

[wijFij (x; Ωi)]− λ‖w − 1/(IJ)‖22. (96)

Problem (96) is equivalent to

minimizew

λ‖w − 1/(IJ)− (1/(2λ))f(x)‖22subject to 1Tw = 1,w ≥ 0

(97)

where f(x) := [F11(x), . . . , fIJ(x)]T . The solution is given by the projection of the point 1/(IJ) + (1/(2λ))f(x) on thethe probabilistic simplex (Parikh et al., 2014)

w∗ = [1/(IJ) + (1/(2λ))f(x)− µ1]+ , (98)

where [·]+ is element-wise non-negative operator, and µ is the root of the equation

1T [1/(IJ) + (1/(2λ))f(x)− µ1]+ =∑i

max0, 1/(IJ) + fi(x)/(2λ)− µ = 1. (99)

The above equation in µ can be solved using the bisection method at a given x (Boyd & Vandenberghe, 2004).

Additional results. In Figure A2-(a), We compare ZO-Min-Max with ZO-Finite-Sum, where the latter minimizes theaverage loss over all model-class combinations. As we can see, our approach significantly improves the worst-case attackperformance (corresponding to M1C1). Here the worst case represents the most robust model-class pair against the attack.This suggests that ZO-Min-Max takes into account different robustness levels of model-class pairs through the designof importance weights w. This can also be evidenced from Figure A2-(b): M1C1 has the largest weight while M2C2corresponds to the smallest weight.

In Figure A3, we contrast the success or failure (marked by blue or red in the plot) of attacking each image using the obtaineduniversal perturbation x with the attacking difficulty (in terms of required iterations for successful adversarial example) ofusing per-image non-universal PGD attack (Madry et al., 2018). We observe that the success rate of the ensemble universalattack is around 80% at each model-class pair, where the failed cases (red cross markers) also need a large amount ofiterations to succeed at the case of per-image PGD attack. And images that are difficult to attack keep consistent acrossmodels; see dash lines to associate the same images between two models in Figure A3.

Page 29: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00Number of iterations 1e4

0

1

2

3

4

5

6

Atta

ck lo

ss

for e

ach

mod

el-c

lass

pai

r

ZO-Min-Max, M1C1 ZO-Min-Max, M1C2 ZO-Min-Max, M2C1 ZO-Min-Max, M2C2

ZO-Finite-Sum, M1C1 ZO-Finite-Sum, M1C2 ZO-Finite-Sum, M2C1 ZO-Finite-Sum, M2C2

0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00Number of iterations 1e4

0.15

0.20

0.25

0.30

0.35

0.40

Wei

ghts

ZO-Min-Max, M1C1 ZO-Min-Max, M1C2 ZO-Min-Max, M2C1 ZO-Min-Max, M2C2

(a) (b)

Figure A2: Convergence performance of ZO-Min-Max in design of black-box ensemble attack. a) Attack loss of using ZO-Min-Max vs.ZO-Finite-Sum, and b) importance weights learnt from ZO-Min-Max.

M1C1 M2C1 M1C2 M2C2Model-class pairs

100

101

102

103

104

105

Num

ber o

f ite

ratio

ns

failure in ensemble attacksuccess in ensemble attack

Figure A3: Success or failure of our ensemble attack versus successful per-image PGD attack.

Page 30: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

E. Additional Details on Poisoning AttackExperiment setup. In our experiment, we generate a synthetic dataset that contains n = 1000 samples (zi, ti). Werandomly draw the feature vector zi ∈ R100 from N (0, I), and determine ti = 1 if 1/(1 + e−(z

Ti θ∗+νi)) > 0.5. Here

we choose θ∗ = 1 as the ground-truth model parameters, and νi ∈ N (0, 10−3) as random noise. We randomly split thegenerated dataset into the training dataset Dtr (70%) and the testing dataset Dte (30%). We specify our learning model asthe logistic regression model for binary classification. Thus, the loss function in problem (14) is chosen as Ftr(x,θ;Dtr) :=h(x,θ;Dtr,1) + h(0,θ;Dtr,2), where Dtr = Dtr,1 ∪ Dtr,2, Dtr,1 represents the subset of the training dataset that will bepoisoned, |Dtr,1|/|Dtr| denotes the poisoning ratio, h(x,θ;D) = −(1/|D|)

∑(zi,ti)∈D[ti log(h(x,θ; zi))+(1−ti) log(1−

h(x,θ; zi))], and h(x,θ; zi) = 1/(1 + e−(zi+x)T θ). In problem (14), we also set ε = 2 and λ = 10−3.

In Algorithm A1, unless specified otherwise we choose the the mini-batch size b = 100, the number of random directionvectors q = 5, the learning rate α = 0.02 and β = 0.05, and the total number of iterations T = 50000. We report theempirical results over 10 independent trials with random initialization.

Addition results. In Figure A4, we show the testing accuracy of the poisoned model as the regularization parameter λvaries. We observe that the poisoned model accuracy could be improved as λ increases, e.g., λ = 1. However, this leads to adecrease in clean model accuracy (below 90% at λ = 1). This implies a robustness-accuracy tradeoff. If λ continues toincrease, both the clean and poisoned accuracy will decrease dramatically as the training loss in (14) is less optimized.

In Figure A5, we present the testing accuracy of the learnt model under different data poisoning ratios. As we can see, only5% poisoned training data can significantly break the testing accuracy of a well-trained model.

10 4 10 2 100 102

Lambda

0.5

0.6

0.7

0.8

0.9

Test

ing

accu

racy

ZO-Min-MaxFO-Min-MaxNo Poison

Figure A4: Empirical performance of ZO-Min-Max in design of poisoning attack: Testing accuracy versus regularization parameter λ.

Page 31: Black-Box Evasion and Poisoning Attacks - arXiv › pdf › 1909.13806.pdfMin-Max Optimization without Gradients: Convergence and Applications to Black-Box Evasion and Poisoning Attacks

Min-Max Optimization without Gradients for ICML 2020

0 5 10 15 20Poisoning ratio (%)

0.65

0.70

0.75

0.80

0.85

0.90

0.95

Test

ing

accu

racy

ZO-Min-MaxFO-Min-Max

Figure A5: Testing accuracy of data poisoning attacks generated by ZO-Min-Max (vs. FO-Min-Max) for different data poisoning ratios.