A New Use of Douglas-Rachford Splitting and ADMM for Classifying Infeasible, Unbounded, and Pathological Conic Programs

Similar documents
A New Use of Douglas-Rachford Splitting and ADMM for Identifying Infeasible, Unbounded, and Pathological Conic Programs

Asynchronous Algorithms for Conic Programs, including Optimal, Infeasible, and Unbounded Ones

Douglas-Rachford Splitting for Pathological Convex Optimization

On the order of the operators in the Douglas Rachford algorithm

Solving conic optimization problems via self-dual embedding and facial reduction: a unified approach

arxiv: v4 [math.oc] 12 Apr 2017

Lecture 5. Theorems of Alternatives and Self-Dual Embedding

Convex Optimization M2

4. Algebra and Duality

On a result of Pazy concerning the asymptotic behaviour of nonexpansive mappings

A STRUCTURAL GEOMETRICAL ANALYSIS OF WEAKLY INFEASIBLE SDPS

THE CYCLIC DOUGLAS RACHFORD METHOD FOR INCONSISTENT FEASIBILITY PROBLEMS

Iterative Convex Optimization Algorithms; Part One: Using the Baillon Haddad Theorem

New stopping criteria for detecting infeasibility in conic optimization

Monotone operators and bigger conjugate functions

5. Duality. Lagrangian

ELEMENTARY REFORMULATION AND SUCCINCT CERTIFICATES IN CONIC LINEAR PROGRAMMING. Minghui Liu

A Simple Derivation of a Facial Reduction Algorithm and Extended Dual Systems

Research overview. Seminar September 4, Lehigh University Department of Industrial & Systems Engineering. Research overview.

Tight Rates and Equivalence Results of Operator Splitting Schemes

Facial Reduction and Geometry on Conic Programming

Convex Optimization Notes

Heinz H. Bauschke and Walaa M. Moursi. December 1, Abstract

Convergence of Fixed-Point Iterations

I.3. LMI DUALITY. Didier HENRION EECI Graduate School on Control Supélec - Spring 2010

Example: feasibility. Interpretation as formal proof. Example: linear inequalities and Farkas lemma

A Dykstra-like algorithm for two monotone operators

Splitting methods for decomposing separable convex programs

Reduction methods in semidefinite and conic optimization

Convex Optimization Boyd & Vandenberghe. 5. Duality

Additional Homework Problems

Extreme Abridgment of Boyd and Vandenberghe s Convex Optimization

Lecture: Duality.

Lecture 8. Strong Duality Results. September 22, 2008

HW1 solutions. 1. α Ef(x) β, where Ef(x) is the expected value of f(x), i.e., Ef(x) = n. i=1 p if(a i ). (The function f : R R is given.

III. Applications in convex optimization

Operator Splitting for Parallel and Distributed Optimization

Approximate Farkas Lemmas in Convex Optimization

Lecture Note 5: Semidefinite Programming for Stability Analysis

On positive duality gaps in semidefinite programming

Convex Optimization and Modeling

SEMIDEFINITE PROGRAM BASICS. Contents

Lecture 3: Semidefinite Programming

6-1 The Positivstellensatz P. Parrilo and S. Lall, ECC

Recent Developments of Alternating Direction Method of Multipliers with Multi-Block Variables

Convex Feasibility Problems

Math 273a: Optimization Subgradients of convex functions

Lecture: Duality of LP, SOCP and SDP

Lecture 5. The Dual Cone and Dual Problem

14. Duality. ˆ Upper and lower bounds. ˆ General duality. ˆ Constraint qualifications. ˆ Counterexample. ˆ Complementary slackness.

Near Equality, Near Convexity, Sums of Maximally Monotone Operators, and Averages of Firmly Nonexpansive Mappings

Optimization for Machine Learning

A convergence result for an Outer Approximation Scheme

Nonlinear Programming 3rd Edition. Theoretical Solutions Manual Chapter 6

Lecture 3: Lagrangian duality and algorithms for the Lagrangian dual problem

INERTIAL ACCELERATED ALGORITHMS FOR SOLVING SPLIT FEASIBILITY PROBLEMS. Yazheng Dang. Jie Sun. Honglei Xu

Enhanced Fritz John Optimality Conditions and Sensitivity Analysis

Conic Linear Optimization and its Dual. yyye

Facial Reduction and Partial Polyhedrality

Coordinate Update Algorithm Short Course Operator Splitting

On Slater s condition and finite convergence of the Douglas Rachford algorithm for solving convex feasibility problems in Euclidean spaces

PARALLEL SUBGRADIENT METHOD FOR NONSMOOTH CONVEX OPTIMIZATION WITH A SIMPLE CONSTRAINT

Primal/Dual Decomposition Methods

Optimization and Optimal Control in Banach Spaces

Some Properties of the Augmented Lagrangian in Cone Constrained Optimization

Math 273a: Optimization Subgradient Methods

Convergence Theorems of Approximate Proximal Point Algorithm for Zeroes of Maximal Monotone Operators in Hilbert Spaces 1

Convex Optimization & Lagrange Duality

Lagrangian-Conic Relaxations, Part I: A Unified Framework and Its Applications to Quadratic Optimization Problems

Math 273a: Optimization Overview of First-Order Optimization Algorithms

Lecture 1: Entropy, convexity, and matrix scaling CSE 599S: Entropy optimality, Winter 2016 Instructor: James R. Lee Last updated: January 24, 2016

A Geometric Framework for Nonconvex Optimization Duality using Augmented Lagrangian Functions

15. Conic optimization

On positive duality gaps in semidefinite programming

Convex Optimization Theory. Chapter 5 Exercises and Solutions: Extended Version

Victoria Martín-Márquez

September Math Course: First Order Derivative

Optimality Conditions for Constrained Optimization

How to generate weakly infeasible semidefinite programs via Lasserre s relaxations for polynomial optimization

arxiv: v1 [math.oc] 15 Apr 2016

M. Marques Alves Marina Geremia. November 30, 2017

A Parametric Simplex Algorithm for Linear Vector Optimization Problems

Lecture 6: Conic Optimization September 8

On a result of Pazy concerning the asymptotic behaviour of nonexpansive mappings

Complexity of the relaxed Peaceman-Rachford splitting method for the sum of two maximal strongly monotone operators

Radial Subgradient Descent

The Karush-Kuhn-Tucker (KKT) conditions

ARock: an algorithmic framework for asynchronous parallel coordinate updates

Sparse Optimization Lecture: Dual Certificate in l 1 Minimization

Splitting Techniques in the Face of Huge Problem Sizes: Block-Coordinate and Block-Iterative Approaches

On the convergence rate of a forward-backward type primal-dual splitting algorithm for convex optimization problems

Summer School: Semidefinite Optimization

subject to (x 2)(x 4) u,

arxiv: v1 [math.oc] 12 Mar 2013

A Unified Analysis of Nonconvex Optimization Duality and Penalty Methods with General Augmenting Functions

Lagrangian Duality Theory

Lecture 1. 1 Conic programming. MA 796S: Convex Optimization and Interior Point Methods October 8, Consider the conic program. min.

Finite Dimensional Optimization Part III: Convex Optimization 1

LP Duality: outline. Duality theory for Linear Programming. alternatives. optimization I Idea: polyhedra

Infeasibility detection in the alternating direction method of multipliers for convex optimization

Transcription:

A New Use of Douglas-Rachford Splitting and ADMM for Classifying Infeasible, Unbounded, and Pathological Conic Programs Yanli Liu, Ernest K. Ryu, and Wotao Yin May 23, 2017 Abstract In this paper, we present a method for classifying infeasible, unbounded, and pathological conic programs based on Douglas-Rachford splitting, or equivalently ADMM. When an optimization program is infeasible, unbounded, or pathological, it is well known that the z k iterates of Douglas- Rachford splitting diverge. Somewhat surprisingly, such divergent iterates still provide useful information, which our method uses for classification. As a first-order method, the proposed algorithm relies on simple subroutines, and therefore is simple to implement and has low periteration cost. 1 Introduction Many convex optimization algorithms have strong theoretical guarantees and empirical performance, but they are often limited to non-pathological, feasible problems; under pathologies often the theory breaks down and the empirical performance degrades significantly. In fact, the behavior of convex optimization algorithms under pathologies has been studied much less, and many existing solvers often simply report failure without informing the users of what went wrong upon encountering infeasibility, unboundedness, or pathology. Pathological problem are numerically challenging, but they are not impossible to deal with. As infeasibility, unboundedness, and pathology do arise in practice, designing a robust algorithm that behaves well in all cases is important to the completion of a robust solver. In this paper, we propose a method based on Douglas-Rachford splitting (DRS), or equivalently ADMM, that classifies infeasible, unbounded, and pathological conic programs. First-order methods such as DRS/ADMM are simple and can quickly provide a solution with moderate accuracy. It is well known that the iterates of DRS/ADMM converge to a fixed point if there is one. When there is no fixed point, the iterates diverge unboundedly, but the precise manner in which they do so has been studied much less. Somewhat surprisingly, when iterates of DRS/ADMM diverge, the behavior of the iterates still provide useful information, which we use to classify the conic program. For example, z k+1 z k, the difference between the DRS iterates, provide a separating hyperplane when the conic program is strongly infeasible and an improving direction when there is one. Facial reduction is one approach to handle infeasible or pathological conic programs. Facial reduction reduces an infeasible or pathological problem into a new problem that is strongly feasible, strongly infeasible, or unbounded with an improving direction, which are the easier cases [6, 7, 15, 21]. Many existing methods such as interior point methods or homogeneous self-dual embedding cannot directly handle certain pathologies, such as weakly feasible or weakly infeasible problems, and are forced to use facial reduction [12, 17]. However, facial reduction introduces a new set of computational issues. After completing the facial reduction step, which has its own the computational challenge and cost, the reduced problem must be solved. The reduced problem involves a cone expressed as an 1

intersection of the original cone with an linear subspace, and in general such cones neither are self-dual nor have a simple formula for projection. This makes applying an interior point method or a first-order method difficult, and existing work on facial reduction do not provide an efficient way to address this issue. In contrast, our proposed method directly address infeasibility, unboundedness, and pathology. Being a first-order method, the proposed algorithm relies on simple subroutines; each iteration performs projections onto the cone and the affine space of the conic program and elementary operations such as vector addition. Consequently, the method is simple to implement has a low per-iteration cost. 1.1 Basic definitions Cones. as A set K R n is a cone if K = λk for any λ > 0. We write and define the dual cone of K K = {u R n u T v 0, for all v K}. Throughout this paper, we will focus on nonempty closed convex cones that we can efficiently project onto. In particular, we do not require that the cone be self-dual. Example of such cones include: The positive orthant: R k + = {x R k x i 0, i = 1,..., n} Second order cone: Q k+1 = { } (x 1,..., x k, x k+1 ) R k R + x k+1 x 2 1 + + x2 k. Rotated second order cone: Q k+2 r = { (x 1,..., x k, x k+1, x k+2 ) R k R 2 + 2x k+1 x k+2 x 2 1 + + x 2 k}. Positive semidefinite cone: Exponential cone: S k + = {M = M T R k k x T Mx 0 for any x R k }. K exp = {(x, y, z) R 3 y exp(x/y) z, y > 0} {(x, 0, z) x 0, z 0}. Conic programs. Consider the conic program minimize subject to c T x Ax = b x K, (P) where x R n is the optimization variable, c R n, A R m n, and b R m are problem data, and K R n is a nonempty closed convex cone. We write p to denote the optimal value of (P). For simplicity, we assume m n and A is full rank. The dual problem of (P) is maximize subject to b T y A T y + s = c s K, where y R m and s R n are the optimization variables. We write d to denote the optimal value of (D). (D) 2

The optimization problem (P) is feasible or infeasible; (P) is feasible if there is an x K {x Ax = b} and infeasible if there is not. When (P) is feasible, it is strongly feasible if there is an x relintk {x Ax = b} and weakly infeasible if there is not, where relint denotes the relative interior. When (P) is infeasible, it is strongly infeasible if there is a non-zero distance between K and {x Ax = b}, i.e., d(k, {x Ax = b}) > 0, and weakly infeasible if d(k, {x Ax = b}) = 0, where d(c 1, C 2 ) = inf { x y x C 1, y C 2 } denotes the Euclidean distance between two nonempty sets C 1 and C 2. When (P) is infeasible we say p = and when feasible p R { }. Likewise, when (D) is infeasible we say d = and when feasible d R { }. As special cases, (P) is called a linear program when K is the positive orthant, a second-order cone program when K is the second-order cone, and a semidefinite program when K is the positive semidefinite cone. 1.2 Classification of conic programs Every conic program of the form (P) falls under exactly one of the following 7 cases. Some of these cases have a corresponding dual characterization, but we skip this discussion as it is not directly relevant to our method. In Section 2, we discuss how to identify these 7 cases. Case (a). p is finite, both (P) and (D) have solutions, and d = p, which is the most common case. For example, the problem minimize x 3 subject to x 1 = 1 x 3 x 2 1 + x2 2 has the solution x = (1, 0, 1) and p = 1. (The inequality constraint corresponds to x Q 3.) The dual problem, after some simplification, is which has the solution y = 1 and d = 1. Case (b). problem maximize y subject to 1 y 2, p is finite, (P) has a solution, but (D) has no solution or d < p. For example, the minimize x 2 subject to x 1 = x 3 = 1 x 3 x 2 1 + x2 2 has the solution x = (1, 0, 1). (The inequality constraint corresponds to x Q 3.) The dual problem, after some simplification, is maximize y 1 1 + y 2 1. By taking y 1 we achieve the dual optimal value d = 0, but no finite y 1 achieves it. As another example, the problem minimize 2x 12 x 11 x 12 x 13 subject to X = x 12 0 x 23 S+, 3 x 13 x 23 x 12 + 1 3

has the solution 0 0 0 X = 0 0 0 0 0 1 and optimal value p = 0. The dual problem, after some simplification, is maximize 2y 2 0 y 2 + 1 0 subject to y 2 + 1 y 1 0 S+, 3 0 0 2y 2 which has the solution y = (0, 1) and optimal value d = 2. Note that case (b) can only happen when (P) is weakly feasible, by standard convex duality [19]. Case (c). p is finite, (P) is feasible, and there is no solution. For example, the problem minimize x 3 subject to x 1 = 2 2x 2 x 3 x 2 1 has an optimal value p = 0, but has no solution as any feasible x satisfies x 3 > 0. (The inequality constraint corresponds to x Q 3 r.) Case (d). p =, (P) is feasible, and there is an improving direction, i.e., there is a u N (A) K satisfying c T u < 0. For example, the problem minimize x 1 subject to x 2 = 0 x 3 x 2 1 + x2 2 has an improving direction u = ( 1, 0, 1). If x is any feasible point, x+tu is feasible, and the objective value goes to as t. (The inequality constraint corresponds to x Q 3.) Case (e). p =, (P) is feasible, and there is no improving direction, i.e., there is no u N (A) K satisfying c T u < 0. For example, the problem For example, consider the problem minimize x 1 subject to x 2 = 1 2x 2 x 3 x 2 1. (The inequality constraint corresponds to x Q 3 r.) Any improving direction u = (u 1, u 2, u 3 ) would satisfy u 2 = 0, and this in turn, with the cone constraint, implies u 1 = 0 and c T u = 0. However, even though there is no improving direction, we can eliminate the variables x 1 and x 2 to verify that p = inf{ 2x 3 x 3 0} =. Case (f). Strongly infeasible, where p = and d(k, {x Ax = b}) > 0. For example, the problem minimize 0 subject to x 3 = 1 x 3 x 2 1 + x2 2 satisfies d(k, {x Ax = b}) = 1. (The inequality constraint corresponds to x Q 3.) 4

Case (g). Weakly infeasible, where p = and d(k, {x Ax = b}) = 0. For example, the problem satisfies d(k, {x Ax = b}) = 0, since minimize 0[ ] [ 0, 1, 1 0 subject to x = 1, 0, 0 1] x 3 x 2 1 + x2 2 d(k, {x Ax = b}) (1, y, y) (1, y, y 2 + 1) 0 as y. (The inequality constraint corresponds to x Q 3.) Remark. In the case of linear programming, i.e., when K in (P) is the positive orthant, only cases (a), (d), and (f) are possible. 1.3 Classification method At a high level, our proposed method for classifying the 7 cases is quite simple. Given an operator T, we call z k+1 = T (z k ) with a starting point z 0 the fixed point iteration of T. Our proposed method runs three similar but distinct fixed-point iterations with the operators T 1 (z) = T (z) + x 0 γdc T 2 (z) = T (z) + x 0 T 3 (z) = T (z) γdc, which are defined and explained in Section 2. We can view T 1 as the DRS operator of (P), T 2 as the DRS operator with c set to 0 in (P), and T 3 as the DRS operator with b set to 0 in (P). We use the information provided by the iterates of the fixed-point iterations to classify the cases, based on the theory of Section 2 and the flowchart shown in Figure 1 as outlined in Section 2.6. 2 Obtaining certificates from Douglas-Rachford Splitting/ADMM The primal problem (P) is equivalent to where minimize f(x) + g(x), (1) f(x) = c T x + δ {x Ax=b} (x) and δ C (x) is the indicator function of a set C defined as { 0 if x C δ C (x) = if x / C. Douglas-Rachford splitting (DRS) applied to (1) is g(x) = δ K (x), (2) x k+1/2 = Prox γg (z k ) x k+1 = Prox γf (2x k+1/2 z k ) (3) z k+1 = z k + x k+1 x k+1/2. 5

Infeasible Thm 7 Alg 2 (f) Strongly infeasible Start Thm 6 Alg 2 (g) Weakly infeasible Feasible Thm 2 Alg 1 (a) There is a primal-dual solution pair with d = p Thm 13 Alg 1 Thm 11,12 Alg 3 (b) There is a primal solution but no dual solution or d < p (c) p is finite but there is no solution Thm 10 Alg 3 (d) Unbounded (p = ) with an improving direction (e) Unbounded (p = ) without an improving direction Figure 1: The flowchart for identifying cases (a) (f). A solid arrow means the cases are always identifiable, a dashed arrow means the cases sometimes identifiable. for any γ > 0, where given a function h Prox γh (x) = arg min z R n { γh(z) + (1/2) z x 2 } denotes the proximal operator with respect to γ. We simplify the DRS iteration (3). Given a nonempty closed convex set C R n, define the projection with respect to C as P C (x) = arg min y x 2 y R n 6

and the reflection with respect to C as R C (x) = 2P C (x) x. Write I to denote both the n n identity matrix and the identity map from R n R n. Write 0 to denote the 0 vector in R n. Define D = I A T (AA T ) 1 A and x 0 = A T (AA T ) 1 b = P {x Ax=b} (0). Write N (A) for the null space of A and R(A T ) for the range of A T. Then P {x Ax=b} (x) = Dx + x 0 and Finally, define P N (A) (x) = Dx. T (z) = 1 2 (I + R N (A)R K )(z). Now we can rewrite the DRS iteration (3) as x k+1/2 = P K (z k ) x k+1 = D(2x k+1/2 z k ) + x 0 γdc (4) z k+1 = z k + x k+1 x k+1/2, or, even more concisely, as Relationship to ADMM. DRS iteration (3) becomes z k+1 = T (z k ) + x 0 γdc. When we define ν k = (1/γ)(z k x k ) and α = 1/γ and reorganize, the x k = arg min {f(x) + x T ν k + α x 2 x xk 1/2 2} x k+1/1 = arg min {g(x) x T ν k + α x 2 x xk 2} ν k+1 = ν k + α(x k x k+1/2 ), which is the alternating direction method of multipliers (ADMM). In a certain sense, DRS and ADMM are equivalent [9, 10], and we can equivalently say that the method of this paper is based on ADMM. Remark. Instead of (2), we could have considered the more general form f(x) = (1 α)c T x + δ {x Ax=b} (x), g(x) = αc T x + δ K (x) with α R. By simplifying the resulting DRS iteration, one can verify that the iterates are equivalent to the α = 0 case. Since the choice of α does not affect the DRS iteration at all, we will only work with the case α = 0. 7

2.1 Convergence of DRS The subdifferential of a function h : R n R { } at x is defined as h(x) = {u R n h(z) h(x) + u T (z x), z R n }. A point x R n is a solution of (1) if and only if 0 (f + g)(x ). DRS, however, converges if and only if there is a point x such that 0 f(x ) + g(x ). In general, f(x) + g(x) (f + g)(x) for all x R, but the two are not necessarily equal. Theorem 1 (Theorem 25.6 of [2] and Theorem 2.1 of [22]). Consider the iteration z k+1 = T (z k ) + x 0 Dc, with any starting point z 0. If there is an x such that 0 f(x ) + g(x ), then z k converges to a limit z, x k+1/2 x = Prox γg (z ), x k+1 x = Prox γg (z ), and 0 f(x ) + g(x ). If there is no x such that then z k diverges in that z k. 0 f(x) + g(x), DRS can fail to find a solution to (P) even when one exists. Slater s constraint qualifications is a sufficient condition that prevents such pathologies: if (P) is strongly feasible, then 0 f(x ) + g(x ) for all solutions x [20, Theorem 23.8]. This fact and Theorem 1 tells us that under Slater s constraint qualifications DRS finds a solution of (P) if one exists. The following theorem, however, provides a stronger, necessary and sufficient characterization of when the DRS iteration converges. Theorem 2. There is an x such that 0 f(x ) + g(x ) if and only if x is a solution to (P), (D) has a solution, and d = p. Based on Theorem 1 and 2 we can determine whether we have case (a) with the iteration with any starting point z 0 and γ > 0. If lim k z k <, we have case (a). z k+1 = T (z k ) + x 0 γdc, 8

If lim k z k =, we do not have case (a). With a finite number of iterations, we test z k+1 z k M for some large M > 0. However, distinguishing the two cases can be numerically difficult as the rate of z k can be very slow. Proof of Theorem 2. This result follows from the exposition of [19], but we provide a proof that matches our notation. The Lagrangian of (P) is Note that and So d = L(x, y, s) = c T x + y T (b Ax) s T x δ K (s). sup L(x, y, s) = c T x + δ {x Ax=b} (x) + δ K (x) y R m,s R m inf x R n L(x, y, s) = bt y δ {(y,s) AT y+s=c}(y, s) δ K (s). sup y R m,s R m x R inf L(x, y, s) inf n x R n gives us weak duality d p. We say (x, y, s ) R n R m R n is a saddle point of L if x arg min x R n L(x, y, s ) (y, s ) arg max y R m,s R n L(x, y, s). sup L(x, y, s) = p y R m,s R n (x, y, s ) is a saddle point of L if and only if x is a solution to (P), (y, s ) is a solution to (D), and p = d. Let us see why. Assume (x, y, s ) is a saddle point of L. This immediately tells us that x is feasible for (P) and (y, s ) is feasible for (D). We also have On the other hand, L(x, y, s ) = inf x R n L(x, y, s ) p = inf x R n sup y R m,s R n inf L(x, y, s) = x R d. n sup L(x, y, s) sup L(x, y, s) = L(x, y, s ). y R m,s R n y R m,s R n Since d p by weak duality, equality holds throughout, and d = p. Since (y, s ) is optimal for (D). Since inf L(x, x R y, s ) = n sup sup L(x, y, s) = inf y R m,s R n y R m,s R n x R x R n inf L(x, y, s) n sup L(x, y, s) y R m,s R n x is optimal for (P). On the other hand, assume x solves (P), (y, s ) solves (D), and d = p. Since (y, s ) is feasible we have x arg min x R n L(x, y, s ). Using the fact that Ax = b and c T x = p, we have L(x, y, s ) = p (s ) T x. 9

On the other hand, using the fact that A T y + s = c and b T y = d, we have Since d = p, we have (s ) T x = 0. So and L(x, y, s ) = d. L(x, y, s) = c T x s T x δ K (s) (y, s ) arg max y R m,s R n L(x, y, s), since s K and (s ) T x = 0. Now assume there is a saddle point (x, y, s ). Define Then since ν = A T y c = s. x arg min x R d {g(x) x T ν }, (δ K (x) x T ν ) = N K (x) + s 0, where we use the fact that s K and (s ) T x = 0 implies s N K (x ). Likewise, since x arg min{x T ν + f(x)}, x R d (x T ν + δ {x Ax=b} (x) + c T x)(x ) = A T y + N {x Ax=b} (x ) 0. This implies ν g(x ) and ν f(x ), and we conclude 0 f(x ) + g(x ). The argument goes the other way as well. Assume there is an x such that 0 f(x ) + g(x ). Then there is a ν such that ν f(x ) and ν g(x ). Since f(x ) = R(A T ) + c, there is a y such that ν = A T y c. Since ν N K (x ), we have s = ν K and (s ) T x = 0. So x minimizes L(x, y, s ) since c A T y s = 0. On the other hand, (y, s ) maximizes since x K and (s ) T x = 0. L(x, y, s) = c T x (s T x + δ K (s)) 2.2 Fixed-point iterations without fixed points We say an operator T : R n R n is nonexpansive if T (x) T (y) 2 x y 2 for all x, y R n. We say T is firmly nonexpansive (FNE) if T (x) T (y) 2 x y 2 (I T )(x) (I T )(y) 2. (FNE operators are nonexpansive.) In particular, the operators defined as T 1 (z) = T (z) + x 0 γdc T 2 (z) = T (z) + x 0 T 3 (z) = T (z) γdc 10

are FNE. It is well known [8] that if a FNE operator T has a fixed point, its fixed-point iteration z k+1 = T (z k ) converges to one with rate z k z k+1 = O(1/ k + 1). Now consider the case where a FNE operator T has no fixed point, which has been studied to a lesser extent. In this case, the fixed-point iteration z k+1 = T (z k ) diverges in that z k [22, Theorem 2.1]. Precisely in what manner z k diverges is characterized by the infimal displacement vector. Given a FNE operator T, we call v = P ran(i T ) (0) the infimal displacement vector of T, which is uniquely defined [6]. To clarify, ran(i T ) denotes the closure of the set ran(i T ) = {(I T )(x) x R n }, and ran denotes the range. We can interpret the infimal displacement vector v as the output of I T corresponding to the best effort to find a fixed point. Lemma 3 (Corollary 2.3 of [1]). Let T be FNE, and consider its fixed-point iteration z k+1 = T (z k ) with any starting point z 0. Then z k z k+1 v = P ran(i T ) (0). When T has a fixed point then v = 0, but v = 0 is possible even when T has no fixed point. In the following sections, we use Lemma 3 to determine the status of a conic program, but, in general, z k z k+1 v has no rate. However, we only need to determine whether lim k (z k+1 z k ) = 0 or lim k (z k+1 z k ) 0, and we do so by checking whether z k+1 z k ε for some tolerance ε > 0. For this purpose, the following rate of approximate convergence is good enough. Theorem 4. Let T be FNE, and consider its fixed point iteration z k+1 = T (z k ), with any starting point z 0. For any ε > 0, there is an M ε > 0 (which depends on T and ε) such that z k z k+1 v + M ε k + 1 + ε/2. Proof of Theorem 4. By Lemma 3, we can pick an x ε such that (I T )T k (x ε ) v ε/2 for k = 0, 1,.... With the triangle inequality, we get z k z k+1 v = (I T )T k (z 0 ) v Since T is FNE, we have (I T )T k (z 0 ) (I T )T k (x ε ) + (I T )T k (x ε ) v (I T )T k (z 0 ) (I T )T k (x ε ) + ε/2. (I T )T k (z 0 ) (I T )T k (x ε ) 2 T k (z 0 ) T k (x ε ) 2 T k+1 (z 0 ) T k+1 (x ε ) 2. Summing this inequality we have k (I T )T j (z 0 ) (I T )T j (x ε ) 2 z 0 x ε 2. (5) j=0 11

Since (I T )T k+1 (z 0 ) (I T )T k+1 (x ε ) 2 = T (I T )T k (z 0 ) T (I T )T k (x ε ) 2 (I T )T k (z 0 ) (I T )T k (x ε ) 2 for k = 0, 1,..., the summand of (5) is monotonically decreasing. So we have (I T )T k (z 0 ) (I T )T k (x ε ) 2 z0 x ε 2. k + 1 By setting M ε = z 0 x ε, we complete the proof. 2.3 Feasibility and infeasibility Sometimes the infimal displacement vector has a simple geometric interpretation: Theorem 5 (Theorem 3.4 of [3]). The operator T 2 defined by T 2 (z) = T (z) + x 0 has the infimal displacement vector v = P K {x Ax=b} (0). In other words, we can interpret v as the best approximation pair between the sets K and {x Ax = b}, and v = d(k, {x Ax = b}). Combining the discussion of Section 2.2 with Theorem 5 gives us Theorem 6 and 7. Theorem 6 (Certificate of feasibility). Consider the iteration z k+1 = T (z k ) + x 0 with any starting point z 0 R n. If (P) is feasible, then z k converges. If (P) is infeasible, then z k is diverges in that z k. Theorem 7 (Certificate of strong infeasibility). Consider the iteration z k+1 = T (z k ) + x 0 with any starting point z 0. If (P) is strongly infeasible, then v 0 and z k z k+1 v. If (P) is weakly infeasible or feasible, then v = 0 and z k z k+1 0. Theorem 8 (Separating hyperplane). Consider the iteration z k+1 = T (z k ) + x 0 with any starting point z 0. If (P) is strongly infeasible, then z k z k+1 v 0, and the hyperplane {x h T x = β}, where h = v K R(A T ) and β = (v T x 0 )/2 > 0, strictly separates K and {x Ax = b}. More precisely, for any y 1 K and y 2 {x Ax = b} we have h T y 1 < β < h T y 2. Based on Theorem 6, 7, and 8, we can determine feasibility, weak infeasiblity, and strong infeasibility and obtain a strictly separating hyperplane if one exists with the iteration with any starting point z 0. If lim k z k <, then (P) is feasible. z k+1 = T (z k ) + x 0, 12

If lim k z k z k+1 > 0, then (P) is strongly infeasible and Theorem 8 provides a strictly separating hyperplane. If lim k z k = and lim k z k z k+1 = 0, then (P) is weakly infeasible. With a finite number of iterations, we distinguish the three cases by testing z k+1 z k ε and z k M for some small ε > 0 and large M > 0. By Theorem 4, we can distinguish strong infeasibility from weak infeasibility or feasibility at a rate of O(1/ k). However, distinguishing feasibility from weak infeasibility can be numerically difficult as the rate of z k can be very slow when (P) is weakly infeasible. Proof of Theorem 8. Note that v = P K {x Ax=b} (0) = P K+N (A) x0 (0) = P K+N (A) (x 0 ) x 0 = P (K R(A T ))(x 0 ) = P K R(A T )( x 0 ). Since the projection operator is FNE, we have (v 0) T ( x 0 0) P K R(A T )( x 0 ) 2 > 0 and therefore v T x 0 < 0. So for any y 1 K and y 2 {x Ax = b}, we have h T y 1 0 < β < v T x 0 = h T y 2. 2.4 Improving direction We say the dual problem (D) is strongly infeasible if 0 < d(0, K + R(A T ) c) = d({(y, s) A T y + s = c}, {(y, s) s K = c}). Theorem 9 (Certificate of improving direction). Assume (P) is feasible. Exactly one of the following is true: 1. (P) has an improving direction, (D) is strongly infeasible, and P N (A) K ( c) 0 is an improving direction. 2. (P) has no improving direction, (D) is feasible or weakly infeasible, and P N (A) K ( c) = 0. Furthermore, P N (A) K ( c) = P K +R(A T ) c (0). Theorem 10. Assume (P) is feasible, and consider the iteration z k+1 = T (z k ) γdc with any starting point z 0 and γ > 0. If (P) has an improving direction, then is one. If (P) has no improving direction, lim k zk+1 z k = d lim k zk+1 z k = 0. 13

Based on Theorem 9 and 10 we can determine whether there is an improving direction and find one if one exists with the iteration z k+1 = T (z k ) γdc, with any starting point z 0 and γ > 0. If lim k z k+1 z k = 0, there is no improving direction. If lim k z k+1 z k = d 0, then d is an improving direction. With a finite number of iterations, we test z k+1 z k ε for some small ε > 0. By Theorem 4, we can distinguish whether there is an improving direction or not at a rate of O(1/ k). We need the following theorem for Section 2.5. Theorem 11. Consider the iteration z k+1 = T (z k ) γdc with any starting point z 0 and γ > 0. If (D) is feasible, then z k converges. If (D) is infeasible, then z k diverges in that z k. Proof of Theorem 9. This result is known [14], but we provide a proof that matches our notation. (P) has an improving direction if and only if {x R n x N (A) K, c T x < 0} =, which is equivalent to c T x 0 for all N (A) K. This is in turn equivalent to c (N (A) K). So c = P (N (A) K) ( c). if and only if there is no improving direction, which holds if and only if 0 = P N (A) K ( c). Assume there is an improving direction. Since the projection operator is firmly nonexpansive, we have 0 < P N (A) K ( c) 2 (P N (A) K ( c)) T ( c). This simplifies to (P N (A) K ( c)) T c < 0, and we conclude P N (A) K ( c) is an improving direction. Using the fact that (N (A) K) = K + R(A T ), we have P N (A) K ( c) = P N (A) K (c) = (P K +R(A T ) I)(c) = P K +R(A T ) c (0). Proof of Theorem 10 and 11. Using the identities I = P N (A) +P R(AT ), I = P K +P K, and R R(AT ) γc(z) = R R(AT )(z) 2γDc, we have T 3 (z) = T (z) γdc = 1 2 (I + R R(A T ) γcr K )(z). In other words, we can interpret the fixed point iteration z k+1 = T (z k ) γdc 14

as the DRS iteration on minimize 0 subject to x R(A T ) γc x K. This proves Theorem 11. Using Lemma 3, applying Theorem 3.4 of [3] as we did for Theorem 5, and applying Theorem 9, we get z k z k+1 P ran(i T3) (0) = P K R(A T )+γc (0) = γp K +R(A T ) c (0) = γp N (A) K ( c). 2.5 Other cases So far, we have discussed how to identify and certify cases (a), (d), (f), and (g). sufficient but not necessary conditions to certify the remaining cases. The following theorem follows from weak duality. We now discuss Theorem 12 (Certificate of finite p ). If (P) and (D) are feasible, then p is finite. Based on Theorem 11, we can determine whether (D) is feasible with the iteration z k+1 = T (z k ) γdc, with any starting point z 0 and γ > 0. (This is the same iteration as discussed in Section 2.4.) If lim k z k <, (D) is feasible. If lim k z k =, (D) is infeasible. With a finite number of iterations, we test z k M for some large M > 0. However, distinguishing the two cases can be numerically difficult as the rate of z k can be very slow. Theorem 13 (Primal iterate convergence). Consider the DRS iteration as defined in (4) with any starting point z 0. If x k+1/2 x and x k+1 x, then x is primal optimal, even if z k doesn t converge. When running the fixed-point iteration with T 1 (z) = T (z) + x 0 γdc, (the same iteration as discussed in Section 2.1.) if z k but x k+1/2 x and x k+1 x, then we have case (b). Examples for Theorem 12. Consider the following problem in case (c): Its dual problem is minimize x 3 subject to x 1 = 2 2x 2 x 3 x 2 1. maximize 2y subject to y 2 1, which is feasible. Based on diagnostics discussed in the previous sections and the fact that the dual problem is feasible, one can conclude that we have either case (b) or (c) but not case (e). 15

Consider the following problem in case (e): Its dual problem is minimize x 1 subject to x 2 = 1 2x 2 x 3 x 2 1 maximize y subject to 1 0, which is infeasible. The diagnostics discussed in the previous sections allows us to conclude that we have case (b), (c), or (e). The fact that the dual problem is infeasible may suggest that we have case (e), there is no such guarantee. Indeed, the dual must be infeasible when we have case (e), but the converse is not necessarily true. Example for Theorem 13 Consider the following problem in case (b): minimize x 2 subject to x 1 = x 3 = 1 x 3 x 2 1 + x2 2. When we run the iteration (4), we can empirically observe that x k+1/2 x and x k+1 x, and conclude that we have case (b). Again, consider the following problem in case (e): minimize x 1 subject to x 2 = 1 2x 2 x 3 x 2 1 When we run the iteration (4), we can empirically observe that x k+1/2 and x k+1 do not converge. The diagnostics discussed in the previous sections allows us to conclude that we have case (b), (c), or (e). The fact that x k+1/2 and x k+1 do not converge may suggest that we have case (c) or (e), but there is no such guarantee. Indeed, x k+1/2 and x k+1 must not converge when we have case (c) or (e), but the converse is not necessarily true. Counterexample for Theorem 12 and 13 The following example shows that the converses of Theorem 12 and 13 are not true. Consider the following problem in case (b): minimize x 1 subject to x 2 x 3 = 0 x 3 x 2 1 + x2 2, which has the solution set {(0, t, t) t R} and optimal value p = 0. Its dual problem is maximize 0 subject to y y 2 + 1, which is infeasible. This immediately tells us that p > is possible even when d =. Furthermore, the x k+1/2 and x k+1 iterates do not converge even though there is a solution. Given z 0 = (z1, 0 z2, 0 0), the iterates z k+1 = (z1 k+1, z2 k+2, z3 k+1 ) are: z k+1 1 = 1 2 zk 1 γ z2 k+1 = 1 2 zk 2 + 1 (z1 k 2 )2 + (z2 k)2 z k+1 3 = 0. 16

So x k+1/2 = P K (z k ) satisfies x k 1 2γ, x k 2 and x k 3, and we can see that x k+1/2 does not converge to the solution set. Proof of Theorem 13. Define Define It s simple to verify that x k+1/2 = Prox γg (z k ) x k+1 = Prox γf (2x k+1/2 z k ) z k+1 = z k + x k+1 x k+1/2 g(x k+1/2 ) = (1/γ)(z k x k+1/2 ) f(x k+1 ) = (1/γ)(2x k+1/2 z k x k+1 ). g(x k+1/2 ) g(x k+1/2 ) f(x k+1/2 ) f(x k+1/2 ) Clearly, g(x k+1/2 ) + f(x k+1 ) = (1/γ)(x k+1/2 x k+1 ). We also have z k+1 = z k γ g(x k+1/2 ) γ f(x k+1 ) = x k+1/2 γ f(x k+1 ) Consider any x K {x Ax = b}. Then g(x k+1/2 ) g(x) + f(x k+1 ) f(x) g(x k+1/2 ) T (x k+1/2 x) + f(x k+1 ) T (x k+1 x) = ( g(x k+1/2 ) + f(x k+1 )) T (x k+1/2 x) + f(x k+1 ) T (x k+1 x k+1/2 ) = (x k+1 x k+1/2 ) T ( f(x k+1 ) (1/γ)(x k+1/2 x)) = (1/γ)(x k+1 x k+1/2 ) T (x z k+1 ) We take the liminf on both sides and use Lemma 14 to get g(x ) + f(x ) g(x) + f(x). Since this holds for any x K {x Ax = b}, x is optimal. Lemma 14. Let 1, 2,... be a sequence in R n. Then Proof. Assume for contradiction that lim inf k ( k ) T lim inf k ( k ) T k i 0. i=1 k i > 2ε for some ε > 0. Since the initial part of the sequence is irrelevant, assume without loss of generality that j ( j ) T i < ε i=1 17 i=1

for j = 1, 2,.... Summing both sides gives us which is a contradiction. 2.6 The algorithm 1 2 k j=1 i=1 k j ( j ) T i < εk j=1 i=1 k ( j ) T i 1{i j} < εk k 2 i + 1 2 i=1 k i 2 < εk, i=1 In this section, we collect the discussed classification results as thee algorithms. The full algorithm is simply running Algorithms 1, 2, and 3, and applying flowchart of Figure 1. Algorithm 1 Finding a solution Parameters: γ, M, ε, z 0 for k = 1,... do x k+1/2 = P K (z k ) x k+1 = D(2x k+1/2 z k ) + x 0 γdc z k+1 = z k + x k+1 x k+1/2 end for if z k < M then Case (a) x k+1/2 and x k+1 solution else if x k+1/2 x and x k+1 x then Case (b) x k+1/2 and x k+1 solution else Case (b), (c), (d), (e), (f), or (g). end if 3 Numerical Experiments We test our algorithm on a library of weakly infeasible SDPs generated by [11]. These semidefinite programs are in the form: minimize C X subject to A i X = b i, i = 1,..., m X S+, n where n = 10 and m = 10 or 20 and A B = n i=1 n j=1 A ijb ij denotes the inner product between two n n matrices A and B. The library provides clean and messy instances. Given a clean instance, a messy instance is 18

Algorithm 2 Feasibility test Parameters: M, ε, z 0 for k = 1,... do x k+1/2 = P K (z k ) x k+1 = D(2x k+1/2 z k ) + x 0 z k+1 = z k + x k+1 x k+1/2 end for if z k M and z k+1 z k > ε then Case (f) Strictly separating hyperplane defined by (z k+1 z k, ( b T x 0 )/2) else if z k M and z k+1 z k ε then Case (g) else z k < M Case (a), (b), (c), (d), or (e) end if Algorithm 3 Boundedness test Prerequisite: (P) is feasible. Parameters: γ, M, ε, z 0 for k = 1,... do x k+1/2 = P K (z k ) x k+1 = D(2x k+1/2 z k ) γdc z k+1 = z k + x k+1 x k+1/2 end for if z k M and z k+1 z k ε then Case (d) Improving direction z k+1 z k else if z k < M then Case (a), (b), or (c) else Case (a), (b), (c), or (e) end if 19

created with m A i U T ( T ij A j )U for i = 1,..., m b i j=1 m T ij b j for i = 1,..., m, j=1 where T Z m m and U Z n n are random invertible matrices with entries in [ 2, 2]. In [11], four solvers are tested, specifically, Sedumi, SDPT3 and MOSEK from the YALMIP environment, and the preprocessing algorithm of Permenter and Parrilo [18] interfaced with Sedumi. Table 1 reports the numbers of instances determined infeasible out of 100 weakly infeasible instances. The four solvers have varying success in detecting infeasibility of the clean instances, but none of them succeed in the messy instances. m = 10 m = 20 Clean Messy Clean Messy SEDUMI 0 0 1 0 SDPT3 0 0 0 0 MOSEK 0 0 11 0 PP+SEDUMI 100 0 100 0 Table 1: Infeasibility detection by other solvers [11]. Our proposed method performs better. However, it does require many iterations and does fail with some of the messy instances. We run the algorithm with N = 10 7 iterations and label an instance infeasible if 1/ z N 8 10 2 (cf. Theorem 6 and 7). Table 2 reports the numbers of instances determined infeasible out of 100 weakly infeasible instances. m = 10 m = 20 Clean Messy Clean Messy Proposed method 100 21 100 99 Table 2: Infeasibility detection by our algorithm. We would like to note that detecting whether or not a problem is strongly infeasible is easier than detecting whether a problem is infeasible. With N = 5 10 4 and a tolerance of z N z N+1 < 10 3 (c.f Theorem 7) our proposed method correctly determined that all test instances are not strongly infeasible. Table 3 reports the numbers of instances determined not strongly infeasible out of 100 weakly infeasible instances. m = 10 m = 20 Clean Messy Clean Messy Proposed method 100 100 100 100 Table 3: Strong Infeasibility detection by our algorithm. References [1] Bailion, J. B., Bruck, R. E., Reich, S.: On the asymptotic behavior of nonexpansive mappings and semigroups in Banach spaces. (1978). 20

[2] Bauschke, H.H., Combettes, P.L.: Convex Analysis and Monotone Operator Theory in Hilbert Spaces. Springer Science & Business Media (2011). [3] Bauschke, H.H., Combettes, P.L., Luke, D.R.: Finding best approximation pairs relative to two closed convex sets in Hilbert spaces. Journal of Approximation Theory 127.2, 178 192 (2004). [4] Bauschke, H.H., and Moursi, W.M.: On the Douglas Rachford algorithm. Mathematical Programming, 1-22 (2016). [5] Bauschke, H.H., Hare, W.L., Moursi, W.M.: On the range of the Douglas Rachford operator. Mathematics of Operations Research (2016). [6] Borwein, J.M., Wolkowicz, H.: Facial reduction for a cone-convex programming problem. Journal of the Australian Mathematical Society (Series A) 30.03, 369 380 (1981). [7] Borwein, J.M., and Wolkowicz, H.: Regularizing the abstract convex program. Journal of Mathematical Analysis and Applications 83.2, 495 530 (1981). [8] Davis, D., and Yin, W. Faster convergence rates of relaxed Peaceman-Rachford and ADMM under regularity assumptions. Mathematics of Operations Research (2017). [9] Eckstein, J., and Fukushima, M.: Some reformulations and applications of the alternating direction method of multipliers. Large Scale Optimization. Springer US, 115 134 (1994). [10] Eckstein, J.: Splitting Methods for Monotone Operators with Applications to Parallel Optimization, PhD thesis, Massachusetts Institute of Technology (1989). [11] Liu, M., Pataki, G.: Exact duals and short certificates of infeasibility and weak infeasibility in conic linear programming. Mathematical Programming, 1 46 (2017). [12] Lourenço, B.F., Muramatsu, M., Tsuchiya, T. Solving SDP completely with an interior point oracle. arxiv preprint arxiv:1507.08065 (2015). [13] Luo, Z-Q., Sturm, J.F., Zhang, S.: Conic convex programming and self-dual embedding. Optimization Methods and Software 14.3, 169 218 (2000). [14] Luo, Z-Q, Sturm, J.F., Zhang, S.: Duality results for conic convex programming. Econometric Institute, Erasmus University Rotterdam, The Netherlands, Tech. Rep. Report 9719/A (1997). [15] Pataki, G.: A simple derivation of a facial reduction algorithm and extended dual systems. Technical report, Columbia University (2000). [16] Pazy, A.: Asymptotic behavior of contractions in Hilbert space. Israel Journal of Mathematics 9.2, 235 240 (1971). [17] Permenter, F., Friberg, H.A., and Andersen, E.D.:Solving conic optimization problems via selfdual embedding and facial reduction: a unified approach. Optimization Online, September (2015). [18] Permenter, F., and Parrilo, P.: Partial facial reduction: simplified, equivalent SDPs via approximations of the PSD cone. arxiv preprint arxiv:1408.4685 (2014). [19] Rockafellar, R.T.: Conjugate Duality and Optimization. Society for Industrial and Applied Mathematics, 1974. [20] Rockafellar, R.T.: Convex Analysis. Princeton University Press, 1970. [21] Waki, H., Muramatsu, M.: Facial reduction algorithms for conic optimization problems. Journal of Optimization Theory and Applications 158.1, 188 215 (2013). [22] Zhao, J., Yang, Q.: A note on the Krasnoselskii-Mann theorem and its generalizations. Inverse Problems 23.3, 1011 (2007). 21