b

DiscoverSearch
About
My stuff
Halpern Iteration for Near-Optimal and Parameter-Free Monotone Inclusion and Strong Solutions to Variational Inequalities
2020·arXiv
Abstract
Abstract

We leverage the connections between nonexpansive maps, monotone Lipschitz operators, and proximal mappings to obtain near-optimal (i.e., optimal up to poly-log factors in terms of iteration complexity) and parameter-free methods for solving monotone inclusion problems. These results immediately translate into near-optimal guarantees for approximating strong solutions to variational inequality problems, approximating convex-concave min-max optimization problems, and minimizing the norm of the gradient in min-max optimization problems. Our analysis is based on a novel and simple potential-based proof of convergence of Halpern iteration, a classical iteration for finding fixed points of nonexpansive maps. Additionally, we provide a series of algorithmic reductions that highlight connections between different problem classes and lead to lower bounds that certify near-optimality of the studied methods.

Given a closed convex set  U ⊆ Rd and a single-valued monotone operator  F : Rd → Rd, i.e., an operator that maps each vector to another vector and satisfies:

image

the monotone inclusion problem consists in finding a point  u∗ that satisfies:

image

is the indicator function of the set  U ⊆ Rd and ∂IU(·)denotes the subdifferential operator (the set of all subgradients at the argument point) of  IU.

Monotone inclusion is a fundamental problem in continuous optimization that is closely related to variational inequalities (VIs) with monotone operators, which model a plethora of problems in mathematical programming, game theory, engineering, and finance [Facchinei and Pang, 2003, Section 1.4]. Within machine learning, VIs with monotone operators and associated monotone inclusion problems arise, for example, as an abstraction of convex-concave min-max optimization problems, which naturally model adversarial training [Madry et al., 2018, Arjovsky et al., 2017, Arjovsky and Bottou, 2017, Goodfellow et al., 2014].

When it comes to convex-concave min-max optimization, approximating the associated VI leads to guarantees in terms of the optimality gap. Such guarantees are generally possible only when the feasible set

U is bounded; a simple example that demonstrates this fact is  Φ(x, y) = ⟨x, y⟩with the feasible set  x, y ∈Rd.The only (min-max or saddle-point) solution in this case is obtained when both x and y are the all-zeros vectors. However, if either  x ̸= 0 or y ̸= 0, then the optimality gap  maxy′∈Rd Φ(x, y′)−minx′∈Rd Φ(x′, y)is infinite. On the other hand, approximate monotone inclusion is well-defined even for unbounded feasible sets. In the context of min-max optimization, it corresponds to guarantees in terms of stationarity. Specifically, in the unconstrained setting, solving monotone inclusion corresponds to minimizing the norm of the gradient of Φ.Note that even in the special setting of convex optimization, convergence in norm of the gradient is much less understood than convergence in optimality gap [Nesterov, 2012, Kim and Fessler, 2018]. Further, unlike classical results for VIs that provide convergence guarantees for approximating weak solutions [Nemirovski, 2004, Nesterov, 2007], approximations to monotone inclusion lead to approximations to strong solutions (see Section 1.2 for definitions of weak and strong solutions and their relationship to monotone inclusion). We leverage the connections between nonexpansive maps, structured monotone operators, and proximal maps to obtain near-optimal algorithms for solving monotone inclusion over different classes of problems with Lipschitz-continuous operators. In particular, we make use of the classical Halpern iteration, which is defined by [Halpern, 1967]:

image

where  T : Rd → Rd is a nonexpansive map, i.e.,  ∀u, v ∈ Rd : ∥T(u) − T(v)∥ ≤ ∥u − v∥.

In addition to its simplicity, Halpern iteration is particularly relevant to machine learning applications, as it is an implicitly regularized method with the following property: if the set of fixed points of T is nonempty, then Halpern iteration (Hal) started at a point  u0and applied with any choice of step sizes  {λk}k≥1that satisfy all of the following conditions:

image

converges to the fixed point of T with the minimum  ℓ2distance to  u0.This result was proved by Wittmann [1992], who extended a similar though less general result previously obtained by Browder [1967]. The result of Wittmann [1992] has since been extended to various other settings [Bauschke, 1996, Xu, 2002, Kohlenbach, 2011, K¨ornlein, 2015, Lieder, 2017, and references therein].

1.1 Contributions and Related Work

A special case of what is now known as the Halpern iteration (Hal) was introduced and its asymptotic convergence properties were analyzed by Halpern [1967] in the setting of  u0 = 0 and T : B2 → B2,where  B2is the unit Euclidean ball. Using the proof-theoretic techniques of Kohlenbach [2008], Leustean [2007] extracted from the asymptotic convergence result of Wittmann [1992] the rate at which Halpern iteration converges to a fixed point. The results obtained by Leustean [2007] are rather loose and provide guarantees of the form  ∥T(uk) − uk∥ = O( Mlog(k))in the best case (obtained for  λk = Θ( 1k)), whereM ≥ ∥u0∥+∥T(u0)∥+∥uk∥, ∀k.A tighter result that shows that  ∥T(uk)−uk∥decreases at rate that is at least as good as  1/√kwas obtained by Kohlenbach [2011]. The results of Leustean [2007] and Kohlenbach

[2011] apply to general normed spaces. The work of Kohlenbach [2011] also provided an explicit rate of metastability that characterizes the convergence of the sequence of iterates  {uk}in Hilbert spaces.

More recently, Lieder [2017] proved that under the standard assumption that T has a fixed point  u∗ andfor the step size  λk = 1k+1,Halpern iteration converges to a fixed point as  ∥T(uk) − uk∥ = 2∥u0−u∗∥k+1 . Asimilar result but for an alternative algorithm was recently obtained by Kim [2019]. These two results (as well as all the results from this paper) only apply to Hilbert spaces. Unlike Halpern iteration, the algorithm introduced by Kim [2019] is not known to possess the implicit regularization property discussed earlier in this paper. The results of Lieder [2017] and Kim [2019] can be used to obtain the same 1/k convergence rate for monotone inclusion with a cocoercive operator but only if the cocoercivity parameter is known, which is rarely the case in practice. Similarly, those results can also be extended to more general monotone Lipschitz operators but only if the proximal map (or resolvent) of F can be computed exactly, an assumption that can rarely be met (see Section 1.2 for definitions of cocoercive operators and proximal maps). We also note that the results of Lieder [2017] and Kim [2019] were obtained using the performance estimation (PEP) framework of Drori and Teboulle [2014]. The convergence proofs resulting from the use of PEP are computer-assisted: they are generated as solutions to large semidefinite programs, which typically makes them hard to interpret and generalize.

Our approach is arguably simpler, as it relies on the use of a potential function, which allows us to remove the assumptions about the knowledge of the problem parameters and availability of exact proximal maps. Our main contributions are summarized as follows:

Results for cocoercive operators. We introduce a new, potential-based, proof of convergence of Halpern iteration that applies to more general step sizes  λkthan handled by the analysis of Lieder [2017] (Section 2). The proof is simple and only requires elementary algebra. Further, the proof is derived for cocoercive operators and leads to a parameter-free algorithm for monotone inclusion. We also extend this parameter-free method to the constrained setting using the concept of gradient mapping generalized to monotone operators (Section 2.1). To the best of our knowledge, this is the first work to obtain the 1/k convergence rate with a parameter-free method.

Results for monotone Lipschitz operators. Up to a logarithmic factor, we obtain the same 1/k convergence rate for the parameter-free setting of the more general monotone Lipschitz operators (Section 2.2). The best known convergence rate established by previous work for the same setting was of the order 1/√k [Dang and Lan, 2015, Ryu et al., 2019]. We obtain the improved convergence rate through the use of the Halpern iteration with inexact proximal maps that can be implemented efficiently. The idea of coupling inexact proximal maps with another method is similar in spirit to the Catalyst framework [Lin et al., 2017] and other instantiations of the inexact proximal-point method, such as, e.g., in the work of Davis and Drusvyatskiy [2019], Asi and Duchi [2019], Lin et al. [2018]. However, we note that, unlike in the previous work, the coupling used here is with a method (Halpern iteration) whose convergence properties were not well-understood and for which no simple potential-based convergence proof existed prior to our work.

Results for strongly monotone Lipschitz operators. We show that a simple restarting-based approach applied to our method for operators that are only monotone and Lipschitz (described above) leads to a parameter-free method for strongly monotone and Lipschitz operators (Section 2.3). Under mild assumptions about the problem parameters and up to a poly-logarithmic factor, the resulting algorithm is iteration-complexity-optimal. To the best of our knowledge, this is the first near-optimal parameter-free method for the setting of strongly monotone Lipschitz operators and any of the associated problems – monotone inclusion, VIs, or convex-concave min-max optimization.

Lower bounds. To certify near-optimality of the analyzed methods, we provide lower bounds that rely on algorithmic reductions between different problem classes and highlight connections between them (Section 3). The lower bounds are derived by leveraging the recent lower bound of Ouyang and Xu [2019] for approximating the optimality gap in convex-concave min-max optimization.

1.2 Notation and Preliminaries

Let  (E, ∥ · ∥) be a real d-dimensional Hilbert space, with norm  ∥ · ∥ =�⟨·, ·⟩, where ⟨·, ·⟩denotes the inner product. In particular, one may consider the Euclidean space  (Rd, ∥·∥2).Definitions that were already introduced at the beginning of the paper easily generalize from  (Rd, ∥·∥2) to (E, ∥·∥), and are not repeated here for space considerations.

Variational Inequalities and Monotone Operators. Let  U ⊆ Ebe closed and convex, and let  F : E →E be an L-Lipschitz-continuous operator defined on U. Namely, we assume that:

image

The definition of monotonicity was already provided in Eq. (1.1), and easily specializes to monotonicity on the set U by restricting u, v to be from U. Further, F is said to be:

1. strongly monotone (or coercive) on U with parameter m, if:

image

2. cocoercive on U with parameter  γ, if:

image

It is immediate from the definition of cocoercivity that every  γ-cocoercive operator is monotone and  1/γ-Lipschitz. The latter follows by applying the Cauchy-Schwarz inequality to the left-hand side of Eq. (1.5) and then dividing both sides by  γ∥F(u) − F(v)∥.

Examples of monotone operators include the gradient of a convex function and appropriately modified gradient of a convex-concave function. Namely, if a function  Φ(x, y)is convex in x and concave in y, then F([xy]) = [ ∇xΦ(x,y)−∇yΦ(x,y)]is monotone.

The Stampacchia Variational Inequality (SVI) problem consists in finding  u∗ ∈ Usuch that:

image

In this case,  u∗ is also referred to as a strong solution to the variational inequality (VI) corresponding to F and U. The Minty Variational Inequality (MVI) problem consists in finding  u∗ such that:

image

in which case  u∗ is referred to as a weak solution to the variational inequality corresponding to F and U. In general, if F is continuous, then the solutions to (MVI) are a subset of the solutions to (SVI). If we assume that F is monotone, then (1.1) implies that every solution to (SVI) is also a solution to (MVI), and thus the two solution sets are equivalent. The solution set to monotone inclusion is the same as the solution set to (SVI).

Approximate versions of variational inequality problems (SVI) and (MVI) are defined as follows: Given ǫ > 0, find an ǫ-approximate solution  u∗ǫ ∈ U,which is a solution that satisfies:

image

Clearly, when F is monotone, an  ǫ-approximate solution to (SVI) is also an  ǫ-approximate solution to (MVI); the reverse does not hold in general.

Similarly,  ǫ-approximate monotone inclusion can be defined as fidning  u∗ǫ that satisfies:

image

where  B(ǫ)is the ball w.r.t.  ∥ · ∥, centered at 0 and of radius  ǫ.We will sometimes write Eq. (1.6) in the equivalent form  −F(u∗ǫ) ∈ ∂IU(u∗ǫ) + B(ǫ).The following fact is immediate from Eq. (1.6).

Fact 1.1. Given  F and U, let u∗ǫ satisfy Eq. (1.6). Then:

image

Thus, when the diameter D is bounded, any ǫD-approximate solution to monotone inclusion is an  ǫ-approximate solution to (SVI) (and thus also to (MVI)); the converse does not hold in general. Recall that when D is unbounded, neither (SVI) nor (MVI) can be approximated.

We assume throughout the paper that a solution to monotone inclusion (MI) exists. This assumption implies that solutions to both (SVI) and (MVI) exist as well. Existence of solutions follows from standard results and is guaranteed whenever e.g., U is compact, or, if there exists a compact set  U′ such that Id − 1LFmaps  U′ to itself [Facchinei and Pang, 2003].

Nonexpansive Maps. Let  T : E → E. We say that T is nonexpansive on  U ⊆ E, if ∀u, v ∈ U :

image

Nonexpansive maps are closely related to cocoercive operators, and here we summarize some of the basic properties that are used in our analysis. More information can be found in, e.g., the book by Bauschke and Combettes [2011].

Fact 1.2. T is nonexpansive if and only if  Id − T is 12-cocoercive, where Id is the identity map.

T is said to be firmly nonexpansive or averaged, if  ∀u, v ∈ U :

image

Useful properties of firmly nonexpansive maps are summarized in the following fact.

Fact 1.3. For any firmly nonexpansive operator  T, Id−Tis also firmly non-expansive, and, moreover, both T and Id − Tare 1-cocoercive.

Halpern iteration is typically stated for nonexpansive maps T as in (Hal). Because our interest is in cocoercive operators F with the unknown parameter 1/L, we instead work with the following version of the Halpern iteration:

image

where  Lk ∈ (0, ∞), ∀k. If Lwas known, we could simply set  Lk+1 = L,in which case (H) would be equivalent to the standard Halpern iteration, due to Fact 1.2. We assume throughout that  λ1 = 12.

We start with the assumption that the setting is unconstrained:  U ≡ E.We will see in Section 2.1 how the result can be extended to the constrained case. Section 2.2 will consider the case of operators that are monotone and Lipschitz, while Section 2.3 will deal with the strongly monotone and Lipschitz case. Some of the proofs are omitted and are instead provided in Appendix A.

To analyze the convergence of (H) for the appropriate choices of sequences  {λi}i≥1 and {Li}i≥1, wemake use of the following potential function:

image

Let us first show that if  AkCkis non-increasing with k for an appropriately chosen sequence of positive numbers  {Ak}k≥1,then we can deduce a property that, under suitable conditions on  {λi}i≥1 and {Li}i≥1,implies a convergence rate for (H).

Lemma 2.1. Let  Ckbe defined as in Eq. (2.1) and let  u∗ be the solution to (MI) that minimizes  ∥u0 − u∗∥.Assume further that  ⟨F(u1) − F(u0), u1 − u0⟩ ≥ 1L1∥F(u1) − F(u0)∥2. If Ak+1Ck+1 ≤ AkCk, ∀k ≥ 1,where  {Ai}i≥1is a sequence of positive numbers that satisfies  A1 = 1, then:

image

Using Lemma 2.1, our goal is now to show that we can choose  Lk = O(L) and λk = O( 1k), whichin turn would imply the desired 1/k convergence rate:  ∥F(uk)∥ = O(L∥u0−u∗∥k ).The following lemma provides sufficient conditions for  {Ai}i≥1, {λi}i≥1, and {Li}i≥1to ensure that  Ak+1Ck+1 ≤ AkCk, ∀k ≥ 1,so that Lemma 2.1 applies.

Lemma 2.2. Let  Ckbe defined as in Eq. (2.1). Let  {Ai}i≥1be defined recursively as  A1 = 1 andAk+1 = Ak λk(1−λk)λk+1 for k ≥ 1.Assume that  {λi}i≥1is chosen so that  λ1 = 12 and for k ≥ 1 : λk+11−2λk+1 ≥

image

Observe first the following. If we knew  L and set Lk = L, λk = 1k+1, and Ak = k(k + 1)/2,then all of the conditions from Lemma 2.2 would be satisfied, and Lemma 2.1 would then imply  ∥F(uk)∥ ≤ L∥u0−u∗∥k ,which recovers the result of Lieder [2017]. The choice  λk = 1k+1 is also the tightest possible that satisfies the conditions Lemma 2.2 – the inequality relating  λk+1 and λkis satisfied with equality. This result is in line with the numerical observations made by Lieder [2017], who observed that the convergence of Halpern iteration is fastest for  λk = 1k+1.

To construct a parameter-free method, we use that F is L-cocoercive; namely, that there exists a constant L < ∞ such that Fsatisfies Eq. (1.5) with  γ = 1/L. The idea is to start to with a “guess” of  L (e.g., L0 = 1)and double the guess  Lkas long as  ⟨F(uk) − F(uk−1), uk − uk−1⟩ < 1Lk ∥F(uk) − F(uk−1)∥2. The totalnumber of times that the guess can be doubled is bounded above by  max{0, log2(2L/L0)}. Parameter λk issimply chosen to satisfy the condition from Lemma 2.2. The algorithm pseudocode is stated in Algorithm 1 for a given accuracy specified at the input.

We now prove the first of our main results. Note that the total number of arithmetic operations in Algorithm 1 is of the order of the number of oracle queries to F multiplied by the complexity of evaluating F at a point. The same will be true for all the algorithms stated in this paper, except that the complexity of evaluating F may be replaced by the complexity of projections onto U.

Theorem 2.3. Given  u0 ∈ Uand an operator  F that is 1L-cocoercive on E, Algorithm 1 returns a point  uksuch that  ∥F(uk)∥ ≤ ǫafter at most max{2L,L0}∥u0−u∗∥ǫ + max{0, log2(2L/L0)}oracle queries to F.

image

Proof. As  F is 1L-cocoercive,  Lk ≤ max{2L, L0}and the total number of times that the algorithm enters the inner while loop is at most  max{0, log2(2L/L0)}.The parameters satisfy the assumptions of Lemmas 2.1 and 2.2, and, thus,  ∥F(uk)∥ ≤ Lk λk1−λk ∥u0 − u∗∥.Hence, we only need to show that  λk decreasessufficiently fast with  k. As Lkcan only be increased in any iteration, we have that

image

Hence, the total number of outer iterations is at most max{2L,L0}∥u0−u∗∥ǫ. Combining with the maximum total number of inner iterations from the beginning of the proof, the result follows.

2.1 Constrained Setups with Cocoercive Operators

Assume now that  U ⊆ E.We will make use of a counterpart to gradient mapping [Nesterov, 2018, Chapter 2] that we refer to as the operator mapping, defined as:

image

where  ΠU�u − 1ηF(u)�is the projection operator, namely:

image

Operator mapping generalizes a cocoercive operator to the constrained case: when  U ≡ E, Gη ≡ F.It is a well-known fact that the projection operator is firmly-nonexpansive [Bauschke and Combettes,

2011, Proposition 4.16]. Thus, Fact 1.3 can be used to show that, if  F is 1L-cocoercive and  η ≥ L, then Gηis 12η-cocoercive. This is shown in the following (simple) proposition.

Proposition 2.4. Let  F be an 1L-cocoercive operator and let  Gηbe defined as in Eq. (1.1), where  η ≥ L.Then  Gη is 12η-cocoercive.

image

max{0, log2(4L/L0)}oracle queries to F (as each computation of  Gηrequires one oracle query to F). To

image

complete this subsection, it remains to show that  Gηis a good surrogate for approximating (MI) (and (SVI)). This is indeed the case and it follows as a suitable generalization of Lemma 3 from Ghadimi and Lan [2016], which is provided here for completeness.

Lemma 2.5. Let  Gηbe defined as in Eq. (2.2). Denote  ¯u = ΠU(u − F(u)/η), so that Gη(u) = η(u − ¯u).If, for some  u ∈ U, ∥Gη(u)∥ ≤ ǫ, then

image

Proof. As, by definition,  ¯u = argminv∈U�⟨F(u), v⟩+ η2∥v−u∥2�,by first-order optimality of  ¯u, we have:0  ∈ F(u) + η(¯u − u) + ∂IU(¯u).Equivalently:  −F(¯u) ∈ F(u) − F(¯u) − Gη(u) + ∂IU(¯u).The rest of the proof follows simply by using  ∥Gη∥ ≤ ǫ and ∥F(u)−F(¯u)∥ = Lloc∥u− ¯u∥ = Llocη ∥Gη(u)∥ ≤ Llocη ǫ.

Lemma 2.5 implies that when the operator mapping is small in norm  ∥ · ∥, then ¯u = ΠU(u − F(u)/η)is an approximate solution to (MI) corresponding to F on U. We can now formally bound the number of oracle queries to F needed to approximate (MI) and (SVI).

image

2.  maxv∈U ⟨F(¯u), ¯u − v⟩ ≤ ǫafter at most

image

Further, every point  ukthat Algorithm 2 constructs is from the feasible set:  uk ∈ U, ∀k ≥ 0, and a simple modification to the algorithm takes at most max{4L,L0}∥u0−u∗∥ǫ + max{0, log2(4L/L0)}oracle queries to

F to construct a point such that  ∥GLk(uk)∥ ≤ ǫ.

Proof. By the definition of  Gη, if u0 ∈ U, then uk ∈ U, for all k.This follows simply as:

image

Observe that, due to Line 8 of Algorithm 2,  Lk ≥ ¯Lk.The rest of the proof follows using Lemma 2.5, Fact 1.1, and the same reasoning as in the proof of Theorem 2.3. Observe that if the goal is to only output a point  uk such that ∥GLk(uk)∥ ≤ ǫ, then computing  ¯uk and F(¯uk)is not needed, and the algorithm can instead use  ∥GLk(uk)∥ > ǫas the exit condition in the outer while loop.

2.2 Setups with non-Cocoercive Lipschitz Operators

We now consider the case in which F is not cocoercive, but only monotone and L-Lipschitz. To obtain the desired convergence result, we make use of the resolvent operator, defined as  JF +∂IU = (Id + F + ∂IU)−1.A useful property of the resolvent is that it is firmly-nonexpansive [Ryu and Boyd, 2016, and references therein], which, due to Fact 1.3, implies that  P = Id − JF +∂IU is 12-cocoercive.

Finding a point  u ∈ U such that ∥P(u)∥ ≤ ǫis sufficient for approximating monotone inclusion (and (SVI)). This is shown in the following simple proposition, provided here for completeness.

Proposition 2.7. Let  P = Id − JF +∂IU. If ∥P(u)∥ ≤ ǫ, then ¯u = u − P(u) = JF +∂IU(u) satisfies

image

Proof. By the definition of  P and JF +∂IU, u − P(u) = (Id + F + ∂IU)−1(u).Equivalently:

image

As  ∥P(u)∥ ≤ ǫ,the result follows.

If we could compute the resolvent exactly, it would suffice to directly apply the result of Lieder [2017]. However, excluding very special cases, computing the exact resolvent efficiently is generally not possible. However, since F is Lipschitz, the resolvent  JF +∂IU can beapproximated efficiently. This is because it corresponds to solving a VI defined on a closed convex set U with the operator F + Id that is 1-strongly monotone and (L+1)-Lipschitz. Thus, it can be computed by solving a strongly monotone and Lipschitz VI, for which one can use the results of e.g., Nesterov and Scrimali [2011], Mokhtari et al. [2019], Gidel et al. [2019] if L is known, or Stonyakin et al. [2018], if L is not known. For completeness, we provide a simple modification to the Extragradient algorithm of Korpelevich [1977] in Algorithm 4 (Appendix A), for which we prove that it attains the optimal convergence rate without the knowledge of L. The convergence result is summarized in the following lemma, whose proof is provided in Appendix A.

Lemma 2.8. Let  ¯u∗k = JF +IU(uk), where uk ∈ U and F is L-Lipschitz. Then, there exists a parameter- free algorithm that queries  F at most O((L + 1) log(L∥uk−¯u∗k∥ǫ ))times and outputs a point  ¯uk such that∥¯uk − ¯u∗k∥ ≤ ǫ.

To obtain the desired result, we need to prove the convergence of a Halpern iteration with inexact evaluations of the cocoercive operator P. Note that here we do know the cocoercivity parameter of P – it is equal to 1/2. The resulting inexact version of Halpern’s iteration for P is:

image

where ˜P(uk) − P(uk) = JF +∂IU(uk) − ˜JF +∂IU(uk) = ekis the error.

To analyze the convergence of (2.3), we again use the potential function  Ckfrom Eq. (2.1), with P as the operator. For simplicity of exposition, we take the best choice of  λi = 1i+1 that can be obtained from Lemma 2.1 for  Li = L = 2, ∀i.The key result for this setting is provided in the following lemma, whose proof is deferred to the appendix.

Lemma 2.9. Let  Ckbe defined as in Eq. (2.1) with  P as the 12-cocoercive operator, and let  Lk = 2,λk = 1k+1, and Ak = k(k+1)2 , ∀k ≥ 1. If the iterates  ukevolve according to (2.3) for an arbitrary initial point  u0 ∈ U, then:

image

Further, if,  ∀k ≥ 1, ∥ek−1∥ ≤ ǫ4k(k+1), then ∥P(uK)∥ ≤ ǫafter at most  K = 4∥u0−u∗∥ǫiterations.

We are now ready to state the algorithm and prove the main theorem for this subsection.

image

Theorem 2.10. Let F be a monotone and L-Lipschitz operator and let  u0 ∈ Ube an arbitrary initial point. For any  ǫ > 0, Algorithm 3outputs a point with  ∥P(uk)∥ ≤ ǫafter at most 8∥u∗−u0∥ǫiterations, where each iteration can be implemented with  O((L + 1) log((L+1)∥u0−u∗∥ǫ )oracle queries to F. Hence, the total number of oracle queries to  F is: O� (L+1)∥u0−u∗∥ǫ log� (L+1)∥u0−u∗∥ǫ ��.

Proof. Recall that ˜P(uk)−P(uk) = ek and ∥ek∥ = ǫk = ǫ8(k+1)(k+2) < ǫ4.Hence, as Algorithm 3 outputs a point  uk with ∥ ˜P(uk)∥ ≤ 3ǫ4 ,by the triangle inequality,  ∥P(uk)∥ ≤ ǫ.

To bound the number of iterations until  ∥ ˜P(uk)∥ ≤ 3ǫ4 ,note that, again by the triangle inequality, if ∥P(uk)∥ ≤ ǫ/2, then ∥ ˜P(uk)∥ ≤ 3ǫ4 .Applying Lemma 2.9,  ∥P(uk)∥ ≤ ǫ/2after at most  k = 8∥u0−u∗∥ǫiterations, completing the proof of the first part of the theorem.

For the remaining part, using Lemma 2.8, ˜JF +∂IU(uk)can be computed (with target error  ǫk) in O((L+1) log((L+1)∥uk−JF +∂IU (uk)∥ǫk )) = O((L+1) log((L+1)∥P (uk)∥ǫ ))iterations, as  O(log( 1ǫk )) = O(log(1ǫ)) andP(uk) = uk − JF +∂IU(uk),by definition. It remains to use that  ∥P(uk)∥ = O(∥u0 − u∗∥), which can be deduced from, e.g., Eq. (A.7) in the proof of Lemma 2.9.

Similarly as before,  ∥P(uk)∥ ≤ ǫimplies an  ǫ-approximate solution to (MI), by Proposition 2.7. When the diameter D is bounded,  ∥P(uk)∥ ≤ ǫD implies an ǫ-approximate solution to (SVI).

Remark 2.11. In degenerate cases where L << 1, instead of using the resolvent of  F + ∂IU, one could use the resolvent of  F/η + ∂IU for η = O(L),assuming the order of magnitude of L is known (this is typically a mild assumption). Then, each approximate computation of the resolvent would take  O((L/η +1) log((L/η+1)∥u0−u∗∥ǫ )oracle queries to F, and we would need to require that  ∥ ˜P(uk)∥ ≤ 3ǫ4η. Thus, the total number of queries to  F would be O((L + η) log((L+η)∥u0−u∗∥ǫ )).

2.3 Setups with Strongly Monotone and Lipschitz Operators

We now show that by restarting Algorithm 3, we can obtain a parameter-free method with near-optimal oracle complexity. To simplify the exposition, we assume w.l.o.g. that  L = Ω(1).

Theorem 2.12. Given F that is L-Lipschitz and m-strongly monotone, consider running the following algorithm A, starting with  u0 ∈ U:

image

Then,  A outputs uk ∈ U with ∥P(uk)∥ ≤ ǫafter at most  1 + log2(∥u0−u∗∥ǫ )iterations, for any  ǫ ∈ (0, 12].The total number of queries to  F until ∥P(uk)∥ ≤ ǫ is O�(L + Lm) log(∥u0−u∗∥ǫ ) log(L + Lm)�.

Proof. The first part is immediate, as each call to Algorithm 3 ensures, due to Theorem 2.10, that

image

and  ∥P(u0)∥ ≤ 2∥u0 − u∗∥ as Pis 2-Lipschitz (because it is 12-cocoercive) and  P(u∗) = 0.

∥ ˜P(uk)∥ = Θ(∥P(uk)∥), each call to Algorithm 3 takes  O(L∥uk−1−u∗∥∥P (uk−1)∥ log(L∥uk−1−u∗∥∥P (uk−1)∥ )) calls to F. De-note  ¯u∗k−1 = JF +∂IU(uk−1) = uk−1 − P(uk−1).Using Proposition 2.7:

image

On the other hand, as F is m-strongly monotone and  u∗ is an (MVI) solution,

image

Hence:  ∥¯u∗k−1−u∗∥ ≤ 1m∥P(uk−1)∥.It remains to use the triangle inequality and  P(uk−1) = uk−1−¯u∗k−1to obtain:  ∥uk−1 − u∗∥ ≤�1 + 1m�∥P(uk−1)∥.

In this section, we only state the lower bounds, while more details about the oracle model and the proof are deferred to Appendix A.

Lemma 3.1. For any deterministic algorithm working in the operator oracle model and any L, D > 0, there exists an L-Lipschitz-continuous operator F and a closed convex feasible set U with diameter D such that:

image

Part (c) of Lemma 3.1 certifies that Algorithm 2 is optimal up to a  log(D/ǫ)factor, due to Theorem 2.6. Part (d) certifies that the restarting algorithm from Theorem 2.12 is optimal up to a factor log(D/ǫ) log(L/m) whenever L = Ω(L/m). Note that L = Ω(L/m)can be ensured by a proper scaling of the problem instance, as any such scaling would leave the condition number L/m unaffected and would only impact the target error  ǫ,which only appears under a logarithm.

We showed that variants of Halpern iteration can be used to obtain near-optimal methods for solving different classes of monotone inclusion problems with Lipschitz operators. The results highlight connections between monotone inclusion, variational inequalities, fixed points of nonexpansive maps, and proximal-point-type algorithms. Some interesting questions that merit further investigation remain. In particular, one open question that arises is to close the gap between the upper and lower bounds provided here. We conjecture that the optimal complexity of monotone inclusion is: (i)  Θ(LDǫ )when the operator is either L-Lipschitz or

image

We thank Prof. Ulrich Kohlenbach for useful comments and pointers to the literature. We also thank Howard Heaton for pointing out a typo in the proof of Lemma 2.1 in a previous version of this paper.

Martin Arjovsky and Leon Bottou. Towards principled methods for training generative adversarial networks. In Proc. ICLR’17, 2017.

Martin Arjovsky, Soumith Chintala, and L´eon Bottou. Wasserstein GAN. arXiv preprint arXiv:1701.07875, 2017.

Hilal Asi and John C Duchi. Stochastic (approximate) proximal point methods: Convergence, optimality, and adaptivity. SIAM Journal on Optimization, 29(3):2257–2290, 2019.

Heinz H Bauschke. The approximation of fixed points of compositions of nonexpansive mappings in Hilbert space. Journal of Mathematical Analysis and Applications, 202(1):150–159, 1996.

Heinz H Bauschke and Patrick L Combettes. Convex analysis and monotone operator theory in Hilbert spaces, volume 408. Springer, 2011.

Felix E Browder. Convergence of approximants to fixed points of nonexpansive nonlinear mappings in Banach spaces. Archive for Rational Mechanics and Analysis, 24(1):82–90, 1967.

Cong D Dang and Guanghui Lan. On the convergence properties of non-Euclidean extragradient meth- ods for variational inequalities with generalized monotone operators. Computational Optimization and Applications, 60(2):277–310, 2015.

Damek Davis and Dmitriy Drusvyatskiy. Stochastic model-based minimization of weakly convex functions. SIAM Journal on Optimization, 29(1):207–239, 2019.

Yoel Drori and Marc Teboulle. Performance of first-order methods for smooth convex minimization: a novel approach. Mathematical Programming, 145(1-2):451–482, 2014.

Francisco Facchinei and Jong-Shi Pang. Finite-dimensional variational inequalities and complementarity problems. Springer Science & Business Media, 2003.

Saeed Ghadimi and Guanghui Lan. Accelerated gradient methods for nonconvex nonlinear and stochastic programming. Mathematical Programming, 156(1-2):59–99, 2016.

Gauthier Gidel, Hugo Berard, Ga¨etan Vignoud, Pascal Vincent, and Simon Lacoste-Julien. A variational inequality perspective on generative adversarial networks. In Proc. ICLR’19, 2019.

Ian Goodfellow, Jean Pouget-Abadie, Mehdi Mirza, Bing Xu, David Warde-Farley, Sherjil Ozair, Aaron Courville, and Yoshua Bengio. Generative adversarial nets. In Proc. NIPS’14, 2014.

Benjamin Halpern. Fixed points of nonexpanding maps. Bulletin of the American Mathematical Society, 73 (6):957–961, 1967.

Donghwan Kim. Accelerated proximal point method and forward method for monotone inclusions. arXiv preprint arXiv:1905.05149, 2019.

Donghwan Kim and Jeffrey A Fessler. Optimizing the efficiency of first-order methods for decreasing the gradient of smooth convex functions. arXiv preprint arXiv:1803.06600, 2018.

Ulrich Kohlenbach. Applied proof theory: proof interpretations and their use in mathematics. Springer Science & Business Media, 2008.

Ulrich Kohlenbach. On quantitative versions of theorems due to fe browder and r. wittmann. Advances in Mathematics, 226(3):2764–2795, 2011.

Daniel K¨ornlein. Quantitative results for halpern iterations of nonexpansive mappings. Journal of Mathematical Analysis and Applications, 428(2):1161–1172, 2015.

GM Korpelevich. Extragradient method for finding saddle points and other problems. Matekon, 13(4): 35–49, 1977.

Laurentiu Leustean. Rates of asymptotic regularity for halpern iterations of nonexpansive mappings. Journal of Universal Computer Science, 13(11):1680–1691, 2007.

Felix Lieder. On the convergence rate of the Halpern-iteration, 2017. http://www.optimization-online.org/DB_FILE/2017/11/6336.pdf.

Hongzhou Lin, Julien Mairal, and Zaid Harchaoui. Catalyst acceleration for first-order convex optimization: From theory to practice. The Journal of Machine Learning Research, 18(1):7854–7907, 2017.

Qihang Lin, Mingrui Liu, Hassan Rafique, and Tianbao Yang. Solving weakly-convex-weakly-concave saddle-point problems as weakly-monotone variational inequality. arXiv preprint arXiv:1810.10207, 2018.

Aleksander Madry, Aleksandar Makelov, Ludwig Schmidt, Dimitris Tsipras, and Adrian Vladu. Towards deep learning models resistant to adversarial attacks. In Proc. ICLR’18, 2018.

Aryan Mokhtari, Asuman Ozdaglar, and Sarath Pattathil. A unified analysis of extra-gradient and optimistic gradient methods for saddle point problems: Proximal point approach. arXiv preprint arXiv:1901.08511, 2019.

Arkadi Nemirovski. Prox-method with rate of convergence O(1/t) for variational inequalities with Lipschitz continuous monotone operators and smooth convex-concave saddle point problems. SIAM Journal on Optimization, 15(1):229–251, 2004.

Yurii Nesterov. Dual extrapolation and its applications to solving variational inequalities and related prob- lems. Mathematical Programming, 109(2-3):319–344, 2007.

Yurii Nesterov. How to make the gradients small. Optima. Mathematical Optimization Society Newsletter, (88):10–11, 2012.

Yurii Nesterov. Lectures on convex optimization, volume 137. Springer, 2018.

Yurii Nesterov and Laura Scrimali. Solving strongly monotone variational and quasi-variational inequalities. Discrete & Continuous Dynamical Systems-A, 31(4):1383–1396, 2011.

Yuyuan Ouyang and Yangyang Xu. Lower complexity bounds of first-order methods for convex-concave bilinear saddle-point problems. Mathematical Programming, Aug 2019.

Ernest K Ryu and Stephen Boyd. Primer on monotone operator methods. Applied and Computational Mathematics, 15(1):3–43, 2016.

Ernest K Ryu, Kun Yuan, and Wotao Yin. ODE analysis of stochastic gradient methods with optimism and anchoring for minimax problems and GANs. arXiv preprint arXiv:1905.10899, 2019.

Fedor Stonyakin, Alexander Gasnikov, Pavel Dvurechensky, Mohammad Alkousa, and Alexander Titov. Generalized mirror prox for monotone variational inequalities: Universality and inexact oracle. arXiv preprint arXiv:1806.05140, 2018.

Rainer Wittmann. Approximation of fixed points of nonexpansive mappings. Archiv der Mathematik, 58 (5):486–491, 1992.

Hong-Kun Xu. Iterative algorithms for nonlinear operators. Journal of the London Mathematical Society, 66(1):240–256, 2002.

A.1 Unconstrained Setting with a Cocoercive Operator

Lemma 2.1. Let  Ckbe defined as in Eq. (2.1) and let  u∗ be the solution to (MI) that minimizes  ∥u0 − u∗∥.Assume further that  ⟨F(u1) − F(u0), u1 − u0⟩ ≥ 1L1∥F(u1) − F(u0)∥2. If Ak+1Ck+1 ≤ AkCk, ∀k ≥ 1,where  {Ai}i≥1is a sequence of positive numbers that satisfies  A1 = 1, then:

image

Proof. The statement holds trivially if  ∥F(uk)∥ = 0,so assume that  ∥F(uk)∥ > 0.Under the assumption of the lemma, we have that  AkCk ≤ C1, ∀k ≥ 1. From (H) and λ1 = 12, u1 = u0 − 1L1 F(u0), and thus:C1 = 1L1 ∥F(u1)∥2 − 1L1 ⟨F(u1), F(u0)⟩ .Let u∗ be an arbitrary solution to (MI) (and thus also to (MVI)). As  ⟨F(u1) − F(u0), u1 − u0⟩ ≥1L1 ∥F(u1) − F(u0)∥2 and u1 = u0 − 1L1F(u0),it follows that  ∥F(u1)∥2 ≤ ⟨F(u0), F(u1)⟩ , and, thusC1 ≤ 0.Further, as  Ak > 0,we also have  Ck ≤ 0,and, hence:

image

where the last line is by  u∗ being a solution to (MVI) and by the Cauchy-Schwarz inequality. The conclusion of the lemma now follows by dividing both sides of  ∥F(uk)∥2 ≤ Lk λk1−λk ∥F(uk)∥·∥u0 −u∗∥ by ∥F(uk)∥and observing that the statement holds for an arbitrary solution  u∗ to (MI), and thus, it also holds for the one that minimizes the distance to  u0.

Lemma 2.2. Let  Ckbe defined as in Eq. (2.1). Let  {Ai}i≥1be defined recursively as  A1 = 1 andAk+1 = Ak λk(1−λk)λk+1 for k ≥ 1.Assume that  {λi}i≥1is chosen so that  λ1 = 12 and for k ≥ 1 : λk+11−2λk+1 ≥

image

Proof. By the assumption of the lemma,

image

which, after expanding the left-hand side, can be equivalently written as:

image

From (H), we have that  uk+1 − uk = λk+11−λk+1(u0 − uk+1) − 2Lk+1F(uk) and uk+1 − uk = λk+1(u0 −

uk) − 2(1−λk+1)Lk+1 F(uk). Hence:

image

Rearranging the last inequality and multiplying both sides by  Ak+1, we have:

image

The left-hand side of the last inequality if precisely  Ak+1Ck+1.The right-hand side is  ≤ AkCk,by the choice of sequences  {Ai}i≥1, {λi}i≥1.

A.2 Operator Mapping

Proposition 2.4. Let  F be an 1L-cocoercive operator and let  Gηbe defined as in Eq. (1.1), where  η ≥ L.Then  Gη is 12η-cocoercive.

image

Hence:

image

As  η ≥ L and F is 1L-cocoercive,  1η ⟨F(u) − F(v), u − v⟩ ≥ 1η2 ∥F(u) − F(v)∥2.It remains to apply Young’s inequality, which implies  ⟨Gη(u) − Gη(v), F(u) − F(v)⟩ ≤ 12∥Gη(u) − Gη(v)∥2 + 12∥F(u) −F(v)∥2.

image

A.3 Approximating the Resolvent

Let us start by proving the convergence of a version of the Extragradient method of Korpelevich [1977] that does not require the knowledge of the Lipschitz constant L (but does require knowledge of the strong monotonicity parameter m; when computing the resolvent we have m = 1). The algorithm is summarized in Algorithm 4. Observe that the update step for  ukfrom Lines 6 and 10 can be written in the form of a projection onto U; we chose to write it in the current form as it is more convenient for the analysis.

We now bound the convergence of Algorithm 4.

Lemma A.1. Let  a0 > 0 and let F be m-strongly monotone and L-Lipschitz. Then, Algorithm 4 outputs a point  uk with ∥uk − u∗∥ ≤ ǫafter at most  k = O� Lm log(L∥u0−u∗∥mǫ �)oracle queries to  F, where u∗

solves (SVI).

image

As F is strongly monotone,  fk ≥ 0, ∀k.By convention, we take  f−1 = 0 and A−1 = 0, so that Akfk −Ak−1fk−1 = ak�⟨F(¯uk), ¯uk − u∗⟩ − m2 ∥¯uk − u∗∥2�.Let us now bound  Akfk − Ak−1fk−1, and observe that  Akfk − Ak−1fk−1 ≥ 0. First, write

image

By the first-order optimality of  uk+1in its definition, we have,  ∀u :

image

and, thus:

image

By the standard three-point identity (which can also be verified directly):

image

Thus, setting  u = u∗ :

image

Observe also that:

image

Thus, we have:

image

By similar arguments:

image

Combining Eq. (A.2)-(A.4):

image

By the condition of the while loop in Line 7 of Algorithm 4, and because  Akfk − Ak−1fk−1 ≥ 0,

image

The condition of the while loop in Line 7 of Algorithm 4 is satisfied for any  ak ≤ 12L, as

image

where we have used the Cauchy-Schwarz inequality, the fact that F is L-Lipschitz, and the Young inequality. Thus, in any iteration,  ak > 14L,and the total number of times the while loop from Line 7 is entered is at most  log2(4L/a0).

image

∥u∗ − uk∥ ≤ δ for k ≥ 16Lm log(∥u∗−u0∥δ ).Consequently, from Eq. (A.5),  ∥¯uk − uk∥ ≤√2δ whenever∥u∗ − uk∥ ≤ δ.In particular, for  δ = akmǫ5√2 ≥ mǫ20√2L, ∥¯uk − uk∥ ≤√2δ = akm5 ǫafter at most k =

16Lm log(20√2L∥u∗−u0∥mǫ(outer loop) iterations.

It remains to show that when  ∥¯uk − uk∥ ≤ δ, ∥uk − u∗∥ ≤ ǫ,and so Algorithm 4 terminates. Observe that  uk − ¯uk = akG1/ak(uk), where G1/akis the operator mapping defined in Eq. (2.2). Thus, using Lemma 2.5 and noting that  ak ≤ 1/Lloc = ∥¯uk−uk∥∥F (¯uk)−F (uk)∥, if ∥¯uk − uk∥ ≤ akm5 ǫ, we have

image

On the other hand, as F is m-strongly monotone, we also have  ⟨F(¯uk), ¯uk − u∗⟩ ≥ m2 ∥¯uk − u∗∥2. Hence,∥¯uk − u∗∥ ≤ 4ǫ5 .Finally, applying the triangle inequality and as  a + k ≤ 1/m :

image

Note that we have already bounded the total number of inner and outer loop iterations. Observing that each inner iteration makes 2 oracle queries to F and each outer iteration makes 2 oracle queries to F outside of the inner iteration, the bound on the total number of oracle queries to F follows.

Lemma 2.8. Let  ¯u∗k = JF +IU(uk), where uk ∈ U and F is L-Lipschitz. Then, there exists a parameter- free algorithm that queries  F at most O((L + 1) log((L+1)∥uk−¯u∗k∥ǫ ))times and outputs a point  ¯uk such that∥¯uk − ¯u∗k∥ ≤ ǫ.

Proof. Observe first that  ¯u∗k solves (SVI) for operator  ¯F(u) = F(u) + u − ukover the set U. This follows from the definition of the resolvent, which implies:

image

A.4 Inexact Halpern Iteration

We start by first proving the following auxiliary result.

Proposition A.2. Given an initial point  u0 ∈ U, let ukevolve according to Eq. (2.3), where  λk = 1k+1.Then,

image

where  u∗ is such that  ∥P(u∗)∥ = 0.

Proof. Let  T = Id−P. Then T(u∗) = u∗. By Fact 1.2, Tis nonexpansive. Observe that we can equivalently write Eq. (2.3) as  uk = λku0 + (1 − λk)T(uk−1) + (1 − λk)ek−1.Thus, using that  u∗ = T(u∗):

image

where we have used the triangle inequality and nonexpansivity of T. The result follows by recursively applying the last inequality and observing that �kj=i(1 − λj) = ik+1.

Using this proposition, we can now prove the following lemma.

Lemma 2.9. Let  Ckbe defined as in Eq. (2.1) with  P as the 12-cocoercive operator, and let  Lk = 2,λk = 1k+1, and Ak = k(k+1)2 , ∀k ≥ 1. If the iterates  ukevolve according to (2.3) for an arbitrary initial point  u0 ∈ U, then:

image

Further, if,  ∀k ≥ 1, ∥ek−1∥ ≤ ǫ4k(k+1), then ∥P(uK)∥ ≤ ǫafter at most  K = 4∥u0−u∗∥ǫiterations.

Proof. By the same arguments as in the proof of Lemma 2.1:

image

From (2.3) and the definition of ˜P, we have that

image

Hence:

image

Plugging  λk+1 = 1k+2 in the last inequality and using the definition of  Ckand the choice of  Ak from thestatement of the lemma completes the proof of the first part.

image

Let us now bound each�ei−1, ii+1P(ui−1) − P(ui)�term. Recall that  P(u∗) = 0 and Pis 2-Lipschitz

image

where we have used Proposition A.2 in the last inequality. In particular, if  ∥ei−1∥ ≤ ǫ4i(i+1), then, ∀i ≥ 1:

image

Combining with Eq. (A.6):

image

Observe that if  ∥u0 −u∗∥ ≤ ǫ/2, as Pis 2-Lipschitz and  P(u∗) = 0,we would have  ∥P(u0)∥ ≤ ǫ, and thestatement of the second part of the lemma would hold trivially. Assume from now on that  ∥u0 −u∗∥ > ǫ/2.Suppose that  ∥P(uk)∥ > ǫ and k ≥ 4∥u0−u∗∥ǫ .Then, dividing both sides of Eq. (A.7) by  ∥P(uk)∥/2 andusing that  ∥P(uk)∥ > ǫ and ∥u0 − u∗∥ > ǫ/2, we get:

image

contradicting the assumption that  ∥P(uk)∥ > ǫand completing the proof.

A.5 Strongly Monotone Lipschitz Operators

Theorem 2.12. Given F that is L-Lipschitz and m-strongly monotone, consider running the following algorithm A, starting with  ¯u0 ∈ U:

image

Then, A outputs a point  uk ∈ U with ∥P(uk)∥ ≤ ǫafter at most  log2(∥u0 − u∗∥/ǫ)iterations, where w.l.o.g.  ǫ ≤ 12. The total number of oracle queries to F until this happens is  O�(L + Lm) log(∥u0 −u∗∥/ǫ) log(L + Lm)�.

Proof. The first part of the theorem is immediate, as each call to Algorithm 3 ensures, due to Theorem 2.10, that

image

and  ∥P(u0)∥ ≤ 2∥u0 − u∗∥ as Pis 2-Lipschitz (because it is 12-cocoercive) and  P(u∗) = 0.

∥ ˜P(uk)∥ = Θ(∥P(uk)∥), each call to Algorithm 3 takes  O(L∥uk−1−u∗∥∥P (uk−1)∥ log(L∥uk−1−u∗∥∥P (uk−1)∥ )) calls to F. De-note  ¯u∗k−1 = JF +∂IU(uk−1) = uk−1 − P(uk−1).Using Proposition 2.7:

image

On the other hand, as F is m-strongly monotone and  u∗ is an (MVI) solution,

image

Hence:  ∥¯u∗k−1−u∗∥ ≤ 1m∥P(uk−1)∥.It remains to use the triangle inequality and  P(uk−1) = uk−1−¯u∗k−1to obtain:

image

which completes the proof.

A.6 Lower Bounds

We make use of the lower bound from Ouyang and Xu [2019] and the algorithmic reductions between the problems considered in previous sections to derive (near-tight) lower bounds for all of the problems considered in this paper.

The lower bounds are for deterministic algorithms working in a (first-order) oracle model. For convex-concave saddle-point problems with the objective  Φ(x, y)and closed convex feasible set  X × Y, any suchalgorithm A can be described as follows: in each iteration k, A queries a pair of points  (¯xk, ¯yk) ∈ X × Yto obtain  (∇xΦ(¯xk, ¯yk), ∇yΦ(¯xk, ¯yk)),and outputs a candidate solution pair  (xk, yk) ∈ X × Y. Both thequery points pair  (¯xk, ¯yk)and the candidate solution pair  (xk, yk)can only depend on (i) global problem parameters (such as the Lipschitz constant of  Φ’s gradients or the feasible sets X, Y) and (ii) oracle queries and answers up to iteration k :

image

We start by summarizing the result from [Ouyang and Xu, 2019, Theorem 9].

Theorem A.3. For any deterministic algorithm working in the first-order oracle model described above and any  L, RX , RY > 0, there exists a problem instance with a convex-concave function  Φ(x, y) : X × Y → Rwhose gradients are L-Lipschitz, such that  ∀k = O(d) :

image

where  (xk, yk) ∈ X × Yis the algorithm output after k iterations and  RX , RYdenote the diameters of the feasible sets X, Y, respectively, and where both X, Y, are closed and convex.

The assumption of the theorem that k = O(d) means that the lower bound applies in the high-dimensional regime  d = Ω(L(RX 2+RX RY)ǫ ),which is standard and generally unavoidable.

In the setting of VIs, we consider a related model in which an algorithm has oracle access to F and refer to it as the operator oracle model. Similarly as for the saddle-point problems, we consider deterministic algorithms that on a given problem instance described by (F, U) operate as follows: in each iteration k the algorithm queries a point  ¯uk ∈ U, receives F(¯uk),and outputs a solution candidate  uk ∈ U. Both uk and¯ukcan only depend on (i) global problem parameters (such as the feasible set U and the Lipschitz parameter of F), and (ii) oracle queries and answers up to iteration  k : {¯ui, F(¯ui)}k−1i=0 .Note that all methods described in this paper and most of the commonly used methods for solving VIs, such as, e.g., the mirror-prox method of Nemirovski [2004] and dual extrapolation method of Nesterov [2007], work in this oracle model.

Lemma 3.1. For any deterministic algorithm working in the operator oracle model described above and any L, D > 0, there exists a VI described by an L-Lipschitz-continuous operator F and a closed convex feasible set U with diameter D such that:

image

image

Proof.

Proof of (a): Suppose that this claim was not true. Then we would be able to solve any instance with L-Lipschitz F and U with diameter bounded by D and obtain  uk with maxu∈U ⟨F(uk), uk − u⟩ ≤ ǫ ino(LD2ǫ )iterations, assuming the appropriate high-dimensional regime. In particular, given any fixed convex- concave  Φ(x, y) with L-Lipschitz gradients and feasible sets X, Y whose diameter is  D/2, let u = [xy],F(u) = [ ∇xΦ(x,y)−∇yΦ(x,y)], U = X × Y.Then, it is not hard to verify that F is monotone and L-Lipschitz (see, e.g., Nemirovski [2004], Facchinei and Pang [2003]) and the diameter of U is D. Thus, by assumption, we would be able to construct a point  uk = [xkyk] for which maxu∈U ⟨F(uk), uk − u⟩ ≤ ǫ in o(LD2ǫ )iterations. But then, because  Φis convex-concave, we would also have, for any  x ∈ X, y ∈ Y:

image

In particular, we would get:

image

Because we obtained this bound for an arbitrary L-Lipschitz convex-concave  Φand arbitrary feasible sets X, Y with diameters D/2, Theorem A.3 leads to a contradiction. Proof of (b): If (b) was not true, then we would be able to obtain a point  uk with

image

in  k = LD2ǫiterations. But the same point would satisfy  maxu∈U ⟨F(uk), uk − u⟩ = o(ǫ),which is a contradiction, due to (a). Proof of (c): We prove the claim for L = 2. This is w.l.o.g., due to the standard rescaling argument: if F is

image

then  maxu∈{U∩Buk } ⟨F(uk), uk − u⟩ = Ω(Lǫ).Suppose that the claim was not true for a 12-cocoercive operator F. Then for any M-Lipschitz monotone operator G, we would be able to use the strategy from Section 2.2 to obtain a point  uk with

image

in  k = MDǫiterations. This is a contradiction, due to (b).

Proof of (d): Suppose that the claim was not true, i.e., that there existed an algorithm that, for any m, L > 0, could output  uk with maxu∈{U∩Buk }� ¯F(uk), uk − u�= ǫ/2 in k = o(L/m)iterations, for any m-strongly monotone and L-Lipschitz operator. Then for any L-Lipschitz monotone operator F, we could apply that algorithm to ¯F(·) = F(·) + ǫ2D(· − u0)to obtain a point  uk with maxu∈{U∩Buk }� ¯F(uk), uk − u�= ǫ/2in  k = o(LD/ǫ)iterations. But then we would also have:

image

which is a contradiction, due to (b).


Designed for Accessibility and to further Open Science