b

DiscoverSearch
About
My stuff
Second-order Conditional Gradient Sliding
2020·arXiv
Abstract
Abstract

Constrained second-order convex optimization algorithms are the method of choice when a high accuracy solution to a problem is needed, due to their local quadratic convergence. These algorithms require the solution of a constrained quadratic subproblem at every iteration. We present the Second-Order Conditional Gradient Sliding (SOCGS) algorithm, which uses a projection-free algorithm to solve the constrained quadratic subproblems inexactly and uses inexact Hessian oracles (subject to an accuracy requirement). When the feasible region is a polytope the algorithm converges quadratically in primal gap after a finite number of linearly convergent iterations. Once in the quadratic regime the SOCGS algorithm requires O(log(log 1/𝜀))first-order and inexact Hessian oracle calls and  O(log(1/𝜀) log(log 1/𝜀))linear minimization oracle calls to achieve an  𝜀-optimal solution. This algorithm is useful when the feasible region can only be accessed efficiently through a linear optimization oracle, and computing first-order information of the function, although possible, is costly.

We focus on the optimization problem defined as

image

where X is a polytope and  𝑓 : X → ℝis  𝜇-strongly convex, has  𝐿-Lipschitz continuous gradients and has 𝐿2-Lipschitz continuous Hessian.

An immensely powerful approach to tackle Problem (1.1) is to construct a second-order approximation to  𝑓 (x)at the current iterate using  ∇ 𝑓 (x)and  ∇2 𝑓 (x), and move in the direction that minimizes this approximation, giving rise to a family of methods known as Newton methods, first developed for unconstrained problems (Kantorovich, 1948). Variants of the former converge globally and have a local quadratic convergence rate when minimizing a self-concordant function or a strongly convex function with Lipschitz continuous Hessian (Nesterov & Nemirovskii, 1994; Nesterov, 2013). When the problem at hand is constrained to a convex set, one can use a constrained analog of these methods (Levitin & Polyak, 1966), where a quadratic approximation to the function is minimized over X at each iteration.

However, there are two shortcomings to these methods. First, computing second-order information about 𝑓 (x)is expensive. This has led to the development of Variable-Metric algorithms, which use approximate second-order information. Secondly, in many cases solving the quadratic subproblem to optimality is too costly. This has resulted in numerous Inexact Variable-Metric algorithms, which in many cases inherit many of the favorable properties of Newton methods (Scheinberg & Tang, 2016; Lee et al., 2014).

The Conditional Gradients (CG) algorithm (Levitin & Polyak, 1966) (also known as the Frank-Wolfe algorithm (Frank & Wolfe, 1956)) instead builds a linear approximation to  𝑓 (x)using  ∇ 𝑓 (x), and moves in the direction given by the point that minimizes this linear approximation over X. Instead of solving a constrained quadratic problem at each iteration, it solves a constrained linear problem, which is usually much cheaper. As the algorithm maintains its iterates as convex combinations of extremal points of X obtained from the linear optimization problem it is dubbed projection-free. Conditional Gradients have become the method of choice in many applications where projecting onto X is computationally prohibitive, such as, e.g., in video co-localization (Joulin et al., 2014) or greedy particle optimization in Bayesian inference (Futami et al., 2019).

For constrained problems where the gradient of  𝑓 (x)is relatively hard to compute, using Projected Variable-Metric methods seems counter-intuitive, yet it allows the construction of a quadratic approximation whose gradients are much cheaper to compute. Minimizing a quadratic approximation at each iteration is often costly, but due to the substantial progress it provides per-iteration it can often become competitive in wall-clock time with using first-order algorithms to directly minimize  𝑓 (x) (Schmidt et al., 2009). We consider the case where both the first-order oracle for  𝑓 (x)and the projection oracle onto X are computationally expensive, but linear programming oracles over X are relatively cheap. In this setting, we show how conditional gradient algorithms can be used to compute Inexact Projected Variable-Metric steps, in an approach that is similar in essence to Conditional Gradient Sliding (CGS) (Lan & Zhou, 2016), where the Euclidean projections onto X in  Nesterov’s Accelerated Gradient Descentare computed using the conditional gradient algorithm. We also show how coupling with an independent sequence of conditional gradient steps we can guarantee the global linear convergence in primal gap of the algorithm.

1.1 Contributions and related work

We provide a projection-free Inexact Variable-Metric algorithm, denoted as the Second-order Conditional Gradient Sliding (SOCGS) algorithm which uses inexact second-order information. The algorithm has a stopping criterion that relies on a lower bound on the primal gap, e.g., via smoothness, and achieves global linear convergence and quadratic local convergence when close to the optimum.

The use of a combination of second-order and projection-free methods was first pioneered in Gonçalves & Melo (2017), who proposed an algorithm in which exact unconstrained Newton steps were performed, and were later projected onto X using the Euclidean norm and the CG algorithm. This resulted in a method that showed local linear convergence in distance to the optimum for functions whose derivative satisfied a Hölder-like condition and also for a subclass of analytic functions. This was later extended in Gonçalves & Oliveira (2018) to deal with inexact second-order information, using the inexactness criteria in Morini (1999), and resulting in the same local linear convergence. A variation of the former algorithm, was shown to converge globally (without an explicit convergence rate) using a non-monotone line search strategy (Gonçalves & Oliveira, 2019). Neither of these three algorithms included a complexity analysis on the number of linear minimization oracle calls needed to achieve a certain target accuracy.

An approach that is similar in spirit is the recent Newton Conditional Gradient (NCG) algorithm (Liu et al., 2020) which performs Inexact Newton steps using a conditional gradient algorithm to minimize a self-concordant function over X. This algorithm requires exact second-order information, as opposed to the approximate information used by the SOCGS algorithm, however it does not require the function to be smooth and strongly convex, or the feasible region to be a polytope. After a finite number of damped-steps the NCG algorithm reaches an  𝜀-optimal solution with  O(log 1𝜀 )first order and exact Hessian oracle calls and  O(𝜀−𝜈)linear optimization oracle calls with  𝜈 = 1 + 𝑜(1). Note that Ochs & Malitsky (2019)[Example 4.3] also proposed a conditional gradient-based Variable-Metric algorithm via their Model Function-Based Conditional Gradient algorithm, however their approach is markedly different from ours: the steps performed in their algorithm can be seen as unconstrained Variable-Metric steps which are projected onto X using the Euclidean norm while the SOCGS performs steps which can be interpreted as unconstrained Inexact Variable-Metric steps which are projected onto X using a norm defined by the positive semi-definite matrix that approximates the Hessian. The same can be said regarding the algorithms in (Gonçalves & Oliveira, 2019), moreover, the SOCGS algorithm directly approximately minimizes a quadratic using a CG variant, in an operation that directly represents an Inexact Projected Variable-Metric step, whereas the algorithm in (Gonçalves & Oliveira, 2019) proceeds in a sequential manner, first computing the Newton step, and afterwards projecting onto X using the CG algorithm.

Further CG variants for the minimization of a self-concordant function have been developed in Dvurechensky et al. (2020), in which an algorithm was developed that achieves an  𝜖-optimal solution in primal gap after 𝑂(1/𝜖)first-order, second-order and linear minimization oracle calls. A related second-order CG variant was later developed in Zhao & Freund (2022) for the minimization of the sum of a logarithmically-homogeneous self-concordant barrier function and a non-smooth function with bounded domain. The aforementioned algorithm reaches an  𝜖-optimal solution in primal gap after  𝑂(1/𝜖)first-order, second-order and linear minimization oracle calls.

Later on, other variants in Carderera et al. (2021) and Dvurechensky et al. (2022) were shown to achieve an  𝜖-optimal solution in primal gap when minimizing a generalized self-concordant function after  𝑂(1/𝜖)first-order, domain and linear minimization oracle calls (where the domain oracle call simply checks if a given point is in the domain of the function being minimized). These variants, therefore, do not require second-order information. When the feasible region under consideration is a polytope, a variant presented in Carderera et al. (2021) of the Away-step Conditional Gradient (ACG) algorithm (Wolfe, 1970) with the stepsize of Pedregosa et al. (2020) was shown to achieve an  𝜖-optimal solution in primal gap after  𝑂(log 1/𝜖)first-order, domain and linear minimization oracle calls. Another CG variant in Dvurechensky et al. (2022) was shown to achieve an  𝜖-optimal solution in primal gap after  𝑂(log 1/𝜖)first-order, second-order and linear minimization oracle calls when the feasible region is a polytope.

We denote the unique minimizer of Problem (1.1) by  x∗. Let S𝑛++ and 𝐼𝑛denote the set of symmetric positive definite matrices and the identity matrix in  ℝ𝑛×𝑛. We denote the largest eigenvalue of the matrix  𝐻 ∈ ℝ𝑛×𝑛 as𝜆max (𝐻). Let ∥·∥ and ∥·∥𝐻denote the Euclidean norm and the matrix norm defined by  𝐻 ∈ S𝑛++, respectively. We denote the diameter of the polytope  X as 𝐷 = maxx,y∈X ∥x − y∥, and its vertices by vert (X) ⊆ X. Given anon-empty set  S ⊂ ℝ𝑛 we refer to its  convex hull as conv (S). For any x ∈ Xwe denote by F (x) the minimal face of X that contains x. Lastly, given a matrix  𝐻 ∈ S𝑛++ we denote the  𝐻-scaled projection of y onto X as:

image

2.1 The Conditional Gradients algorithm

We define the linear approximation of the function  𝑓 (x)around the point  x𝑘 as:

image

At each iteration the vanilla Conditional Gradients (CG) algorithm (Levitin & Polyak, 1966; Frank & Wolfe, 1956; Jaggi, 2013) takes steps defined as  x𝑘+1 = x𝑘 + 𝛾𝑘(argminx∈X ˆ𝑙𝑘(x) − x𝑘) with 𝛾𝑘 ∈ (0, 1]. As the iterates are formed as convex combinations of points in X the algorithm is projection-free. A useful quantity that can readily be computed in all steps is  maxv∈X⟨∇ 𝑓 (x𝑘), x𝑘 − v⟩, known as the Frank-Wolfe gap, which provides an upper bound on the primal gap and is often used as a stopping criterion when running the CG algorithm.

However, the vanilla CG algorithm does not converge linearly in primal gap when applied to Problem (1.1) in general. This motivated the development of the Away-step Conditional Gradient (ACG) algorithm (Wolfe, 1970) (shown in Algorithm 4 in Appendix B), which uses Away-steps (shown in Algorithm 1) and converges linearly when coupled with an exact line search (Lacoste-Julien & Jaggi, 2015) or a step size strategy dependent on  𝐿(Pedregosa et al., 2020). The ACG algorithm maintains what is called an  active set S𝑘 ⊆ vert (X)which represents the potentially non-unique set of vertices of X such that  x𝑘 ∈ conv (S𝑘). Associated with this active set  S𝑘we have a set of barycentric coordinates  λ𝑘such that if we denote by  λ𝑘(u) ∈ [0, 1]the element of  λ𝑘associated with  u ∈ S𝑘we have that  x𝑘 = �u∈S𝑘 λ𝑘(u)u, with  �u∈S𝑘 λ𝑘(u) = 1and  λ𝑘(u) ≥ 0for all  u ∈ S𝑘.

2.1.1 Global convergence

The first proof of asymptotic linear convergence of the ACG algorithm relied on the strict complementarity of the problem in Equation (1.1) (shown in Assumption 1), which we will also use in the convergence proof of the SOCGS algorithm. A mild assumption that rules out degeneracy.

Assumption 1 (Strict Complementarity). We have that  ⟨∇ 𝑓 (x∗) , x − x∗⟩ = 0if and only if  x ∈ F (x∗).

If Assumption 1 is satisfied the iterates of the ACG algorithm reach  F (x∗)in a finite number of steps, remaining in  F (x∗)for all subsequent iterations (Guélat & Marcotte, 1986). When inside  F (x∗), the iterates of the ACG algorithm contract the primal gap linearly. This analysis was later significantly extended to provide an explicit global linear convergence rate in primal gap (Theorem 2.1), by making use of the pyramidal width of the polytope X (Lacoste-Julien & Jaggi, 2015). With the pyramidal width one can derive a primal progress guarantee for all steps taken by the ACG algorithm except ‘bad’ away-steps that reduce the

image

1 v ← argminv∈X ⟨∇ 𝑓 (x𝑘) , v⟩, a ← argmaxv∈S𝑘 ⟨∇ 𝑓 (x𝑘) , v⟩2 if ⟨∇ 𝑓 (x𝑘), x𝑘 − v⟩ ≥ ⟨∇ 𝑓 (x𝑘), a − x𝑘⟩ then3 d ← x𝑘 − v, 𝛾max ← 14 else

5 d ← a − x𝑘, 𝛾max ← λ(a)/(1 − λ(a))6 end

7 𝛾𝑘 ← argmin𝛾∈[0,𝛾max] 𝑓 (x𝑘 + 𝛾d)8 x𝑘+1 ← x𝑘 + 𝛾𝑘d9 Update S𝑘 and λ𝑘(see full details in Algorithm 5 in Appendix B)

cardinality of the active set  S𝑘, that is when  ⟨∇ 𝑓 (x𝑘), x𝑘 − v⟩ < ⟨∇ 𝑓 (x𝑘), a − x𝑘⟩and the step size satisfies 𝛾𝑘 = 𝛾maxin Algorithm 1. This cannot happen more than  ⌊𝐾/2⌋times when running the ACG algorithm for  𝐾iterations (as the algorithm cannot drop more vertices with away-steps than it has picked up with Frank-Wolfe steps). This is an important consideration to keep in mind, as it means that the ACG linear primal gap contraction does not hold on a per-iteration basis.

Theorem 2.1 (Primal gap convergence of the ACG algorithm). (Lacoste-Julien & Jaggi, 2015, Theorem 1) Given an initial point  x0 ∈ X, the ACG algorithm applied to Problem (1.1) satisfies after  𝐾 ≥ 0iterations:

image

where  𝐷denotes the diameter of the polytope  X and 𝛿its pyramidal width.

The CG algorithm and its variants make heavy use of the linear approximation ˆ𝑙𝑘(x)in Equation (2.3). What if we consider a quadratic approximation of  𝑓 (x), as opposed to a linear approximation?

2.2 Projected Variable-Metric algorithms

We define the quadratic approximation of the function  𝑓 (x)around the point  x𝑘using a matrix  𝐻𝑘 ∈ S𝑛++, denoted by ˆ𝑓𝑘(x) as:

image

Intuitively, ˆ𝑓𝑘(x)will be a good local approximation to  𝑓 (x)around  x𝑘if  𝐻𝑘is a good approximation to  ∇2 𝑓 (x𝑘). In this case, the quadratic approximation to ˆ𝑓𝑘(x)will contain more information about the local curvature of the function  𝑓 (x)than the linear approximation ˆ𝑙𝑘(x). Methods that minimize quadratic approximations of the function  𝑓 (x)over X to define iterates are commonly known as Projected VariableMetric (PVM) algorithms (Nesterov, 2018; Ben-Tal & Nemirovskii, 2020). These methods could, for example, set  x𝑘+1 = x𝑘 + 𝛾𝑘(argminx∈X ˆ𝑓𝑘(x) − x𝑘), with 𝛾𝑘 ∈ (0, 1].

Minimizing the approximation ˆ𝑓𝑘(x)over X can be interpreted as a scaled projection operation onto X, which is why these methods are considered projection-based, as opposed to the CG algorithm.

Remark 2.2. Minimizing ˆ𝑓𝑘(x)over X can be viewed as the  𝐻𝑘-scaled projection of  x𝑘 − 𝐻−1∇ 𝑓 (x𝑘)onto X, namely:

image

We can recover many well-known algorithms from the PVM formulation, for example, if we set  𝐻𝑘 =∇2 𝑓 (x𝑘)in Equation (2.5) we recover the Projected Newton algorithm. Alternatively, if we use  𝐻𝑘 = 𝐼𝑛we recover the Projected Gradient Descent (PGD) algorithm.

2.2.1 Local convergence

One of the most attractive features of the Projected Newton algorithm with  𝛾𝑘 = 1when applied to Problem (1.1) is its local quadratic convergence in distance to  x∗. This property also extends to PVM algorithms if  𝐻𝑘approximates  ∇2 𝑓 (x𝑘)sufficiently well as  x𝑘approaches  x∗. What do we mean by sufficiently well? As  𝑓 (x)is strongly convex we know that for any  𝐻𝑘 ∈ S𝑛++ and y ∈ X, then for d = y − x𝑘:

image

for  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} and 𝜂𝑘 ≥ 1(see Lemma A.6 in Appendix A.1). The parameter  𝜂𝑘can be used to measure how well  𝐻𝑘approximates  ∇2 𝑓 (x𝑘), and will serve as our accuracy parameter. The chain of inequalities shown in Equation (2.6) is presented as Assumption C in Karimireddy et al. (2018a), where it is used to prove the global convergence of an Inexact Projected Variable-Metric variant. Using  𝐻𝑘 = ∇2 𝑓 (x𝑘)we recover  𝜂𝑘 = 1. We assume that we have access to an oracle  Ω : X → S𝑛++that returns estimates of the Hessian that satisfy:

Assumption 2 (Accuracy of Hessian oracle  Ω).The oracle  Ωqueried with a point x returns a matrix  𝐻with a parameter  𝜂such that:

image

Where  𝜂 = max{𝜆max(𝐻−1∇2 𝑓 (x)), 𝜆max([∇2 𝑓 (x)]−1𝐻)} and 𝜔 ≥ 0denotes a known constant.

Intuitively, the accuracy of the oracle improves as the oracle is queried with points closer to  x∗. If the oracle returns  Ω (x) = ∇2 𝑓 (x) for all x ∈ X then 𝜔 = 0. This assumption allows us to obtain local quadratic convergence in distance to  x∗ for the simplest PVM algorithm, i.e.,  x𝑘+1 = x∗𝑘+1 = argminx∈X ˆ𝑓𝑘(x), as shownin Theorem 2.4 (see Corollary C.12 in Appendix C).

Remark 2.3. Note that finding a matrix  𝐻satisfying Assumption 2 at x, given a fixed  𝜔requires knowledge of a tight lower bound on  ∥x − x∗∥.

Theorem 2.4 (Local quadratic convergence of vanilla PVM algorithm). Given an 𝐿-smooth and  𝜇-stronglyconvex function with  𝐿2-Lipschitz Hessian and a convex set X if Assumption 2 is satisfied, and we set x𝑘+1 = x∗𝑘+1 = argminx∈X ˆ𝑓𝑘(x)we have for all  𝑘 ≥ 0:

image

2.2.2 Global convergence

One of the key questions that remains to be answered in this section is how PVM algorithms behave globally. For Problem (1.1) the vanilla PVM algorithm with unit step size will converge globally, and if we use bounded step sizes, or a exact line search, we can show that the primal gap contracts linearly (Theorem 2.5). The global convergence of these methods can be recast in terms of a notion related to the multiplicative stability of the Hessian, allowing for elegant proofs of convergence (Karimireddy et al., 2018a).

Theorem 2.5 (Primal gap convergence of vanilla PVM algorithm with line search). (Karimireddy et al.,

2018a, Theorem 4) Given an  𝐿-smooth and  𝜇-strongly convex function and a convex set X then the vanilla PVM algorithm with an exact line search or with a step size  𝛾𝑘 = 𝜇𝐿𝜂𝑘 guarantees for all  𝑘 ≥ 0:

image

The discussion of PVM algorithms in Section 2.2 did not address two important concerns:

1. The PVM algorithm requires computing a scaled projection at every iteration. These projections are usually too expensive to compute to optimality. Ideally we would want to solve these scaled projection problems to a certain accuracy, but can we maintain the local quadratic convergence in distance to the optimum shown in Theorem 2.4 when computing approximate scaled projections?

2. The global convergence rate of the PVM algorithm with exact line search and perfect Hessian information (Theorem 2.5 with  𝜂𝑘 = 1) has a worse dependence on the condition number  𝐿/𝜇than the convergence rate of the PGD and the ACG algorithm (see Theorem 2.1 for the latter). Can we couple Inexact PVM steps with ACG steps and improve the global convergence rate in Theorem 2.5?

The Second-order Conditional Gradient Sliding (SOCGS) algorithm (Algorithm 2) is designed with these considerations in mind, providing global linear convergence in primal gap and local quadratic convergence in primal gap and distance to  x∗. The algorithm couples an independent ACG step with line search (Line 4) with an Inexact PVM step with unit step size (Lines 9-12). At the end of each iteration we choose the step that provides the greatest primal progress (Lines 14-18). The ACG steps in Line 4 will ensure global linear convergence in primal gap, and the Inexact PVM steps in Lines 14-18 will provide quadratic convergence. Note that the ACG iterates in Line 4 do not depend on the Inexact PVM steps in Lines 9-12. This is because the ACG steps do not contract the primal gap on a per-iteration basis (see discussion in Section 2.1.1).

We compute the scaled projection in the Inexact PVM step (Lines 9-12) using the ACG algorithm with exact line search, thereby making the SOCGS algorithm (Algorithm 2) projection-free. As the function being minimized in the Inexact PVM steps is quadratic there is a closed-form expression for the optimal step size for ˆ𝑓𝑘 (x)in Line 10. The scaled projection problem is solved to an accuracy  𝜀𝑘such that ˆ𝑓𝑘(˜x𝑘+1) − minx∈X ˆ𝑓𝑘 (x) ≤ 𝜀𝑘, using the Frank-Wolfe gap as a stopping criterion, as in the CGS algorithm (Lan & Zhou, 2016). The accuracy parameter  𝜀𝑘in the SOCGS algorithm depends on a lower bound on the primal gap of Problem 1.1 which we denote by  𝑙𝑏 (x𝑘)that satisfies  𝑙𝑏 (x𝑘) ≤ 𝑓 (x𝑘) − 𝑓 (x∗).

image

Note that this guarantee also holds if we use a line search instead of the step size described above, as the line search is guaranteed to make at least as much progress. Computing the aforementioned quantity comes at no extra cost if  𝐿and  𝐷are known, as the Frank-Wolfe vertex from Line 4 of Algorithm 2 can be reused. Alternatively one could use any CG variant that monotonically decreases the primal gap. It suffices to run an arbitrary number of steps  𝑛of the aforementioned variant to minimize  𝑓 (x)starting from  x𝑘, resulting in  x𝑛𝑘 ∈ X. Simply noting that  𝑓 (x𝑛𝑘) ≥ 𝑓 (x∗)allows us to conclude that  𝑓 (x𝑘) − 𝑓 (x∗) ≥ 𝑓 (x𝑘) − 𝑓 (x𝑛𝑘), and therefore a valid lower bound is  𝑙𝑏 (x𝑘) = 𝑓 (x𝑘) − 𝑓 (x𝑛𝑘). The higher the number of CG steps performed from x𝑘, the tighter the resulting lower bound will be.

Remark 3.3 (Assuming knowledge of a lower bound). In several machine learning applications the value of  𝑓 (x∗)is known a priori, such is the case of the approximate Carathéodory problem (Mirrokni et al., 2017; Combettes & Pokutta, 2019) where  𝑓 (x∗) = 0. In other applications, estimating  𝑓 (x∗)is easier than estimating the strong convexity parameter (see (Barré et al., 2020; Barré & d’Aspremont, 2019; Asi & Duchi, 2019; Hazan & Kakade, 2019) for an in-depth discussion). This allows for tight lower bounds on the primal gap.

image

3.1 Global convergence

The global convergence rate in primal gap of the SOCGS algorithm (Algorithm 2) is driven by the ACG steps in Line 4, as such:

Theorem 3.4. Given x0 ∈ X, then the SOCGS algorithm applied to Problem (1.1) satisfies:

image

where  𝐷denotes the diameter of the polytope  X and 𝛿its pyramidal width.

Proof. As at each step the SOCGS algorithm (Algorithm 2) chooses between the independent ACG step (Line 4) and the Inexact PVM step (Lines 9-12) according to which one provides the greatest primal progress, the primal gap convergence in Theorem 2.1 applies. □

3.2 Local convergence

Despite computing inexact scaled projections in Lines 9-12 of Algorithm 2, the Inexact PVM steps contract the distance to optimum quadratically when close enough to the optimal solution.

Lemma 3.5. Given a 𝜇-strongly convex and  𝐿-smooth function  𝑓 (x) with 𝐿2-Lipschitz Hessian and a convex set X, if Assumption 2 is satisfied then the Inexact PVM steps in Lines 9-12 of Algorithm 2 satisfy for all 𝑘 ≥ 0:

image

where  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} and 𝜔 ≥ 0denotes a constant.

In order to take advantage of the quadratic convergence in distance to the optimum shown in Lemma 3.5, we need to show that at some point the SOCGS algorithm will always choose in Lines 14-19 the Inexact PVM step defined in Lines 9-12. To be more specific, we show that the convergence in primal gap for the Inexact PVM step will also be quadratic. We do this by first showing that there is an iteration  𝐾 ≥ 0such that for all  𝑘 ≥ 𝐾 we have x𝑘 ∈ F (x∗) (Lemma D.8in Appendix D.1).

Lemma 3.6. Given a 𝜇-strongly convex and  𝐿-smooth function  𝑓 (x) with 𝐿2-Lipschitz continuous Hessian and a polytope X, if Assumption 1 and 2 are satisfied, then there is an index  𝐾 ≥ 0such that for  𝑘 ≥ 𝐾we have that  x𝑘 ∈ F (x∗), that is, both the Inexact PVM steps (Lines 9-12 of Algorithm 2) and the ACG step (Line 4 of Algorithm 2) are in  F (x∗).

We can upper bound the right-hand side of Equation (3.2) using strong convexity, and the left-hand side using smoothness, Lemma 3.6 and strict-complementarity (Assumption 1). This allows us to show that there exists an iteration after which the primal progress of the Inexact PMV steps in Lines 9-12 will be quadratic, which ensures the local quadratic convergence of the SOCGS algorithm.

Theorem 3.7 (Quadratic convergence in primal gap of the SOCGS algorithm). Given a 𝜇-strongly convex and  𝐿-smooth function  𝑓 (x) with 𝐿2-Lipschitz Hessian and a polytope X, if Assumption 1 and Assumption 2 are satisfied, then there is a  𝐾 ≥ 0such that for  𝑘 ≥ 𝐾the iterates of the SOCGS algorithm (Algorithm 2) satisfy:

image

where  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} and 𝜔 ≥ 0denotes a constant.

3.3 Complexity analysis

We defer the full details of the complexity analysis to Section D.2 in Appendix D. Throughout this section we make the simplifying assumption that we have at our disposal the tightest possible lower bound  𝑙𝑏(x𝑘) onthe primal gap, that is,  𝑙𝑏(x𝑘) = 𝑓 (x𝑘) − 𝑓 (x∗)(in Remark 3.8 we address a strategy that can be used when the primal gap is not known). Let  𝑟 = min{𝑟ACG, 𝑟PVM} > 0(where  𝑟ACGis described in Theorem D.4 and 𝑟PVMin Corollary D.7),  𝐺 = maxx∈X ∥∇ 𝑓 (x)∥and  𝛽 = max{(2𝐷𝐺)1/4, (2𝐿(1 + 𝜔𝐷2)𝐷3𝐺)1/8}. With these considerations in mind the different oracle complexities are listed in Table 1. As in the classical analysis of PVM algorithms, the SOCGS algorithm shows local quadratic convergence after a number of iterations that is independent of  𝜀(but dependent on  𝑓 (x) and X).

Remark 3.8. Providing a looser lower bound  𝑙𝑏(x𝑘)on the primal gap does not affect the number of first-order or Hessian oracle calls, however it can significantly increase the number of linear optimization oracle calls used to compute the Inexact PVM steps in Lines 9-12. Note that the progress guarantee from a single ACG step that is not an away-step that drops a vertex is  𝑓 (x𝑘) − 𝑓 (x𝑘+1) ≥ 𝜇4𝐿 ( 𝛿𝐷 )2( 𝑓 (𝑥𝑘) − 𝑓 (𝑥∗))(see Theorem 1 in Lacoste-Julien & Jaggi (2015)). If we use as  𝑙𝑏(x𝑘)the progress obtained from such a step (note that  𝑓 (x𝑘) − 𝑓 (x∗) ≥ 𝑓 (x𝑘) − 𝑓 (x𝑘+1)) in the complexity analysis, one can obtain after a finite number of iterations, a  log log 1/𝜀complexity in terms of FO and Hessian oracle calls and  log(1/𝜀) log(log 1/𝜀)LO calls, but with worse constants then the ones in Table 1.

image

Table 1: Complexity to reach an  𝜀-optimal solution to Problem (1.1) for the SOCGS algorithm.

We compare the performance of the SOCGS algorithm with that of other projection-free algorithms, and that of Projected-Gradient Descent (PGD). In all experiments we compare against the vanilla CG algorithm, the ACG algorithm, the Pairwise-Step Conditional Gradients algorithms (PCG) and the Lazy ACG algorithm (Braun et al., 2017) (ACG (L)). In the first experiment we also compare against the Decomposition Invariant Conditional Gradient (DICG) algorithm (Garber & Meshi, 2016), the CGS algorithm (Lan & Zhou, 2016) and the Stochastic Variance-Reduced Conditional Gradients (SVRCG) algorithm (Hazan & Luo, 2016). We were not able to achieve acceptable performance with the CGS algorithm in the second and third experiment and with the SVRFW algorithm in the third experiment. Lastly we also compare against the Newton Conditional Gradients (NCG) algorithm (Liu et al., 2020) which is similar in spirit to the SOCGS algorithm, in the second and third experiment. One of the key features of the NCG algorithm is that it does not require an exact line search strategy, as it provides a specific step size strategy (however it requires selecting five hyperparameters and using an exact Hessian).

In the first problem the Hessian oracle will be inexact, but will satisfy Assumption 2 with  𝜔 = 0.1, moreoverwe will also assume knowledge of the primal gap, by first computing a solution to high accuracy. In the remaining problems the Hessian oracle will be exact, and we will assume that we do not have knowledge of the primal gap, and will use the strategy outlined in Remark 3.8. In the second experiment, in addition to using the exact Hessian, we will also implement SOCGS with an LBFGS Hessian update (SOCGS LBFGS) (note that this does not satisfy Assumption 2). All the line searches that do not have a closed form solution are computed using a golden-section bounded line search between 0 and 1. The full details of the implementation can be found in Appendix E. In the second and third experiment we will also cap the maximum number of inner iterations to 1000 for the SOCGS and NCG algorithms, as is done in the computational experiments of NCG and SVRCG.

image

Remark 4.1 (Hyperparameter search for the NCG algorithm). We tested 27 hyperparameter combinations for the NCG algorithm, and the one that provided the best performance was selected (see Appendix E for the full details).

Sparse coding over the Birkhoff polytope In this example (Figure 1) we minimize the objective function  𝑓 (𝑋) = �𝑚𝑖=1 ∥y𝑖 − 𝑋z𝑖∥2, with  𝑋 ∈ ℝ𝑛×𝑛, over the Birkhoff polytope. This objective function is strongly convex if the vectors  z𝑖, with  𝑚 ∈ [1, 𝑚]form a basis for  ℝ𝑛(See discussion in Appendix E.1). We generate synthetic data by creating a matrix  𝐵 ∈ ℝ𝑛×𝑛with  𝑛 = 80entries sampled from a standard normal distribution, and  𝑚vectors  x ∈ ℝ𝑛(with  𝑚 = 10000in the first experiment and  𝑚 = 100000in the second), with entries sampled from a standard normal distribution, in order to form  𝑍 = {z1, · · · , z𝑚}. For both the experiments we verified numerically that the resulting objective function is strongly convex. The set of vectors  𝑌 = {y1, · · · , y𝑚}is generated by computing  y𝑖 = 𝐵z𝑖for all  𝑖 ∈ ⟦1, 𝑚⟧. The starting point for all the algorithms is  𝐼𝑛. To implement the projection operation used in PGD we use the interior point solver implemented in CVXOPT (Andersen et al., 2011), which we have found to be computationally faster than the Douglas-Rachford approach described in Combettes & Pokutta (2021). Note that the use of this implementation only impacts the performance with respect to time, and not with respect to iteration count.

Structured logistic regression over ℓ1 unit ballIn this last experiment (Figure 2) we minimize a function of the form  𝑓 (𝑥) = 1/𝑚 �𝑚𝑖=1 log�1 + 𝑒−𝑦𝑖⟨x,z𝑖⟩�+ 𝜆/2 ∥x∥2over the  ℓ1unit ball with  𝜆 = 0.05. The labels and samples used are taken from the training set of the gissette (Guyon et al., 2007) and the real-sim (Chang & Lin, 2011) dataset, where  𝑛 = 5000and  𝑚 = 6000and  𝑛 = 72309and  𝑚 = 20958respectively. The starting point for all the algorithms is the vector  (1, 0, · · · , 0).

Inverse covariance estimation over spectrahedron In the second experiment (Figure 3) we minimize the function  𝑓 (𝑋) = − log det(𝑋 + 𝛿𝐼𝑛) + trace (𝑆𝑋) + 𝜆2 ∥𝑋∥2𝐹with  𝑋 ∈ ℝ𝑛×𝑛over the space of positive semidefinite matrices of unit trace, with  𝛿 = 10−5 and 𝜆 = 0.05. This feasible region is not a polytope, and so the guarantees shown in the paper do not apply as they crucially rely on Theorem 2.1, and the pyramidal width of the spectrahedron is zero. However, we include the results to show the promising numerical performance of the method. The matrix  𝑆is generated by computing a random orthonormal basis  B = {v1, · · · , v𝑚}in ℝ𝑚 and computing  𝑆 = �𝑖=1 𝜎𝑖v𝑖v𝑇𝑖 , where 𝜎𝑖is uniformly distributed between  0.5 and 1 for 𝑖 ∈ ⟦1, 𝑚⟧. Thestarting point for all the algorithms is the matrix  1/𝑛𝐼𝑛.

image

Figure 1: Birkhoff polytope: Primal gap comparison for  𝑚 = 10000 (a),(b) and 𝑚 = 100000 (c),(d).

image

Figure 2:  ℓ1-ball: Comparison in terms of primal gap for the gissette (a),(b) and the real-sim (c),(d)

image

Figure 3: Spectrahedron: Comparison in terms of primal gap for  𝑛 = 100 (a),(b) and for 𝑛 = 50 (c),(d).

This paper focuses on the minimization of a smooth and strongly convex function over a polytope in the setting where efficient access to the feasible region is limited to a linear optimization oracle and first-order information about the objective function is expensive to compute. We also assume inexact second-order information subject to an accuracy requirement.

Given these challenges, we present the Second-order Conditional Gradient Sliding (SOCGS) algorithm, which at each iteration computes an Inexact Projected Variable-Metric (PVM) step with unit step size (using the Away-step Conditional Gradient (ACG) algorithm and an accuracy criterion that depends on a lower bound on the primal gap), and an independent ACG step with line search, and chooses the step that provides the greatest primal progress. As the algorithm relies on a linear minimization oracle, as opposed to a projection oracle, it is projection-free. The algorithm can be seen as the second-order analog of the Conditional Gradient Sliding algorithm (Lan & Zhou, 2016), which uses Conditional Gradient steps to compute inexact Euclidean projections in Nesterov’s Accelerated Gradient Descent algorithm. After a finite number (independent of the target accuracy  𝜀) of linearly convergent iterations, the convergence rate of the SOCGS algorithm is quadratic in primal gap. Once inside this phase the SOCGS algorithm reaches an  𝜀-optimal solution after  O (log(log 1/𝜀))Hessian and first-order oracle calls and  O(log(1/𝜀) log(log 1/𝜀))linear minimization oracle calls.

The Newton Conditional Gradient (NCG) (or Newton Frank-Wolfe) algorithm (Liu et al., 2020) uses an approach that is similar in spirit to the one used in the SOCGS algorithm, however with a very different analysis and set of assumptions. The aforementioned algorithm minimizes a self-concordant function over a convex set by performing Inexact Newton steps using a Conditional Gradient algorithm to solve the constrained quadratic subproblems. This algorithm requires exact Hessian information, and after a finite number of iterations (independent of the target accuracy  𝜀), the convergence rate of the NCG algorithm is linear in primal gap. Once inside this phase a  𝜀-optimal solution is reached after  O (log 1/𝜀)exact Hessian and first-order oracle calls and  O(1/𝜀𝜈)linear minimization oracle calls, where  𝜈is a constant greater than one.

The computational results show that the SOCGS algorithm outperforms other first-order projection-free algorithms and the NCG algorithm in applications where first-order information is costly to compute. The improved performance with respect to other first-order projection-free algorithms is due to the substantial progress per iteration provided by the Inexact PVM steps, which makes up for their higher computational cost, resulting in faster convergence with respect to time. The better performance of the SOCGS algorithm with respect to the NCG algorithm is due to the better global convergence of the SOCGS algorithm, and the use of the Away-step Conditional Gradient algorithm as a subproblem solver in the SOCGS algorithm, as opposed to the vanilla Conditional Gradient algorithm used by the NCG algorithm.

Research reported in this paper was partially supported by NSF CAREER Award CMMI-1452463. We would like to thank Gábor Braun for the helpful discussions, and the anonymous reviewers for their suggestions and comments.

Aharon, M., Elad, M., and Bruckstein, A. K-svd: An algorithm for designing overcomplete dictionaries for sparse representation. IEEE Transactions on Signal Processing, 54(11):4311–4322, 2006.

Andersen, M., Dahl, J., Liu, Z., Vandenberghe, L., Sra, S., Nowozin, S., and Wright, S. Interior-point methods for large-scale cone programming. Optimization for machine learning, 5583, 2011.

Asi, H. and Duchi, J. C. The importance of better models in stochastic optimization. Proceedings of the National Academy of Sciences, 116(46):22924–22930, 2019.

Banerjee, O., Ghaoui, L. E., and d’Aspremont, A. Model selection through sparse maximum likelihood estimation for multivariate gaussian or binary data. Journal of Machine Learning Research, 9(Mar):485–516, 2008.

Barré, M. and d’Aspremont, A. Polyak steps for adaptive fast gradient method. arXiv preprint arXiv:1906.03056, 2019.

Barré, M., Taylor, A., and d’Aspremont, A. Complexity guarantees for Polyak steps with momentum. arXiv preprint arXiv:2002.00915, 2020.

Beck, A. First-order methods in optimization, volume 25. SIAM, 2017.

Ben-Tal, A. and Nemirovskii, A. Optimization III (Spring 2020 Lecture Notes): Convex analysis, nonlinear programming theory and nonlinear programming algorithms. 2020.

Braun, G., Pokutta, S., and Zink, D. Lazifying conditional gradient algorithms. In Proceedings of the 34th International Conference on Machine Learning, pp. 566–575, 2017.

Braun, G., Pokutta, S., Tu, D., and Wright, S. Blended conditonal gradients. In Proceedings of the 36th International Conference on Machine Learning, pp. 735–743, 2019.

Carderera, A., Besançon, M., and Pokutta, S. Simple steps are all you need: Frank-wolfe and generalized self-concordant functions. Advances in Neural Information Processing Systems, 34:5390–5401, 2021.

Chang, C.-C. and Lin, C.-J. Libsvm: A library for support vector machines. ACM transactions on intelligent systems and technology (TIST), 2(3):1–27, 2011.

Combettes, C. W. and Pokutta, S. Revisiting the approximate Carathéodory problem via the frank-wolfe algorithm. arXiv preprint arXiv:1911.04415, 2019.

Combettes, C. W. and Pokutta, S. Complexity of linear minimization and projection on some sets. Operations Research Letters, 49(4):565–571, 2021.

Condat, L. Fast projection onto the simplex and the  ℓ1 ball. Mathematical Programming, 158(1-2):575–585, 2016.

Diakonikolas, J., Carderera, A., and Pokutta, S. Locally accelerated conditional gradients. In International Conference on Artificial Intelligence and Statistics, pp. 1737–1747, 2020.

Dvurechensky, P., Ostroukhov, P., Safin, K., Shtern, S., and Staudigl, M. Self-concordant analysis of frank-wolfe algorithms. In International Conference on Machine Learning, pp. 2814–2824. PMLR, 2020.

Dvurechensky, P., Safin, K., Shtern, S., and Staudigl, M. Generalized self-concordant analysis of frank–wolfe algorithms. Mathematical Programming, pp. 1–69, 2022.

Frank, M. and Wolfe, P. An algorithm for quadratic programming. Naval Research Logistics Quarterly, 3 (1-2):95–110, 1956.

Friedman, J., Hastie, T., and Tibshirani, R. Sparse inverse covariance estimation with the graphical lasso. Biostatistics, 9(3):432–441, 2008.

Futami, F., Cui, Z., Sato, I., and Sugiyama, M. Bayesian posterior approximation via greedy particle optimization. In Proceedings of the AAAI Conference on Artificial Intelligence, volume 33, pp. 3606–3613, 2019.

Garber, D. Revisiting frank-wolfe for polytopes: Strict complementary and sparsity. arXiv preprint arXiv:2006.00558, 2020.

Garber, D. and Hazan, E. A linearly convergent variant of the conditional gradient algorithm under strong convexity, with applications to online and stochastic optimization. SIAM Journal on Optimization, 26(3): 1493–1528, 2016.

Garber, D. and Meshi, O. Linear-memory and decomposition-invariant linearly convergent conditional gradient algorithm for structured polytopes. In Advances in Neural Information Processing Systems 29, pp. 1001–1009, 2016.

Garber, D., Sabach, S., and Kaplan, A. Fast generalized conditional gradient method with applications to matrix recovery problems. arXiv preprint arXiv:1802.05581, 2018.

Ghanbari, H. and Scheinberg, K. Proximal quasi-Newton methods for regularized convex optimization with linear and accelerated sublinear convergence rates. Computational Optimization and Applications, 69(3): 597–627, 2018.

Gonçalves, M. and Oliveira, F. An inexact newton-like conditional gradient method for constrained nonlinear systems. Applied Numerical Mathematics, 132:22–34, 2018.

Gonçalves, M. L. and Melo, J. G. A newton conditional gradient method for constrained nonlinear systems. Journal of Computational and Applied Mathematics, 311:473–483, 2017.

Gonçalves, M. L. and Oliveira, F. On the global convergence of an inexact quasi-newton conditional gradient method for constrained nonlinear systems. Numerical Algorithms, pp. 1–23, 2019.

Guélat, J. and Marcotte, P. Some comments on Wolfe’s ‘away step’. Mathematical Programming, 35(1): 110–119, 1986.

Guyon, I., Li, J., Mader, T., Pletscher, P. A., Schneider, G., and Uhr, M. Competitive baseline methods set new standards for the NIPS 2003 feature selection benchmark. Pattern recognition letters, 28(12): 1438–1444, 2007.

Hazan, E. and Kakade, S. Revisiting the Polyak step size. arXiv preprint arXiv:1905.00313, 2019.

Hazan, E. and Luo, H. Variance-reduced and projection-free stochastic optimization. In Proceedings of the 33th International Conference on Machine Learning, pp. 1263–1271, 2016.

Jaggi, M. Revisiting frank-wolfe: Projection-free sparse convex optimization. In Proceedings of the 30nd International Conference on Machine Learning, number CONF, pp. 427–435, 2013.

Joulin, A., Tang, K., and Fei-Fei, L. Efficient image and video co-localization with Frank-Wolfe algorithm. In European Conference on Computer Vision, pp. 253–268. Springer, 2014.

Kantorovich, L. V. Functional analysis and applied mathematics. Uspekhi Matematicheskikh Nauk, 3(6): 89–185, 1948.

Karimireddy, S. P., Stich, S. U., and Jaggi, M. Global linear convergence of Newton’s method without strong-convexity or Lipschitz gradients. arXiv preprint arXiv:1806.00413, 2018a.

Karimireddy, S. P. R., Stich, S., and Jaggi, M. Adaptive balancing of gradient and update computation times using global geometry and approximate subproblems. In Proceedings of the 35th International Conference on Artificial Intelligence and Statistics, pp. 1204–1213, 2018b.

Lacoste-Julien, S. and Jaggi, M. On the global linear convergence of Frank-Wolfe optimization variants. In Advances in Neural Information Processing Systems 28, pp. 496–504, 2015.

Lan, G. The complexity of large-scale convex programming under a linear optimization oracle. arXiv preprint arXiv:1309.5550, 2013.

Lan, G. and Zhou, Y. Conditional gradient sliding for convex optimization. SIAM Journal on Optimization, 26(2):1379–1409, 2016.

Lee, H., Battle, A., Raina, R., and Ng, A. Y. Efficient sparse coding algorithms. In Advances in Neural Information Processing Systems 20, pp. 801–808, 2007.

Lee, J. D., Sun, Y., and Saunders, M. A. Proximal Newton-type methods for minimizing composite functions. SIAM Journal on Optimization, 24(3):1420–1443, 2014.

Lehoucq, R. B., Sorensen, D. C., and Yang, C. ARPACK users’ guide: solution of large-scale eigenvalue problems with implicitly restarted Arnoldi methods, volume 6. Siam, 1998.

Levitin, E. S. and Polyak, B. T. Constrained minimization methods. USSR Computational Mathematics and Mathematical Physics, 6(5):1–50, 1966.

Liu, D., Cevher, V., and Tran-Dinh, Q. A Newton Frank-Wolfe method for constrained self-concordant minimization. arXiv preprint arXiv:2002.07003, 2020.

Mairal, J., Bach, F., Ponce, J., and Sapiro, G. Online learning for matrix factorization and sparse coding. Journal of Machine Learning Research, 11(Jan):19–60, 2010.

Mirrokni, V., Leme, R. P., Vladu, A., and Wong, S. C.-w. Tight bounds for approximate carathéodory and beyond. In Proceedings of the 34th International Conference on Machine Learning, pp. 2440–2448, 2017.

Morini, B. Convergence behaviour of inexact newton methods. Mathematics of Computation, 68(228): 1605–1613, 1999.

Nemirovski, A. Interior point polynomial time methods in convex programming. Lecture notes, 2004.

Nesterov, Y. Introductory lectures on convex optimization: A basic course, volume 87. Springer Science & Business Media, 2013.

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

Nesterov, Y. and Nemirovskii, A. Interior-point polynomial algorithms in convex programming, volume 13. Siam, 1994.

Nocedal, J. and Wright, S. Numerical optimization. Springer Science & Business Media, 2006.

Ochs, P. and Malitsky, Y. Model function based conditional gradient method with armijo-like line search. In Proceedings of the 36th International Conference on Machine Learning, pp. 4891–4900, 2019.

Pedregosa, F., Negiar, G., Askari, A., and Jaggi, M. Linearly convergent frank-wolfe with backtracking line-search. In International Conference on Artificial Intelligence and Statistics, pp. 1–10. PMLR, 2020.

Rao, N., Shah, P., and Wright, S. Forward–backward greedy algorithms for atomic norm regularization. IEEE Transactions on Signal Processing, 63(21):5798–5811, 2015.

Scheinberg, K. and Tang, X. Practical inexact proximal quasi-Newton method with global complexity analysis. Mathematical Programming, 160(1-2):495–529, 2016.

Schmidt, M., Berg, E., Friedlander, M., and Murphy, K. Optimizing costly functions with simple constraints: A limited-memory projected quasi-Newton algorithm. In Proceedings of the 12th International Conference on Artificial Intelligence and Statistics, pp. 456–463, 2009.

Wolfe, P. Convergence theory in nonlinear programming. Integer and nonlinear programming, pp. 1–36, 1970.

Yuan, M. and Lin, Y. Model selection and estimation in the gaussian graphical model. Biometrika, 94(1): 19–35, 2007.

Zhao, R. and Freund, R. M. Analysis of the frank–wolfe method for convex composite optimization involving a logarithmically-homogeneous barrier. Mathematical Programming, pp. 1–41, 2022.

image

Outline. The appendix of the paper is organized as follows:

• Section A presents the notation and definitions used throughout the appendix, as well as useful material pertaining to the Hessian approximation.

• Section B contains background information about the Conditional Gradients algorithm, pseudocode for the vanilla Conditional Gradients algorithm and the Away-step Conditional Gradients algorithm and theoretical information about the convergence of the Away-step Conditional Gradients algorithm.

• Section C presents information about the vanilla Projected Variable-Metric algorithm, its global linear convergence with exact line search or a bounded stepsize, and its quadratic local convergence in distance to the optimum with unit step size.

• Section D contains the proof of global linear and local quadratic convergence in primal gap of the Secondorder Conditional Gradient Sliding algorithm, as well as an oracle complexity analysis.

• Section E presents a detailed description of the numerical experiments performed.

Appendix A. Notation and Preliminaries

We denote the norm of a vector  v as ∥v∥ =√︁⟨v, v⟩, and the norm of a matrix  𝐴 as ∥𝐴∥ = maxv≠0 ∥𝐴v∥ /∥v∥.Let  S𝑛++ denote the set of symmetric positive definite matrices in  ℝ𝑛×𝑛 and let ∥·∥𝐻 denote the matrix norm defined by  𝐻 ∈ S𝑛++, that is, for a given vector v the norm defined by  𝐻 is ∥v∥𝐻 =√︁⟨v, 𝐻v⟩. We use vmin (𝐻)and  vmax (𝐻)to refer to the eigenvectors of unit norm associated with the minimum and maximum eigenvalues, denoted by  𝜆min (𝐻) and 𝜆max (𝐻)respectively, of the matrix  𝐻 ∈ S𝑛++. Similarly, we use  𝜆𝑖 (𝐻) with 𝑖 ∈ [1, 𝑛]to refer to the  𝑖-th largest eigenvalue of the matrix  𝐻 ∈ S𝑛++. Let 𝜎min(𝐻) and 𝜎max(𝐻)denote the minimum and maximum singular values of the matrix  𝐻. We denote the open ball of radius  𝑟 > 0centered at x as B(x, 𝑟). Let int (X) and rel. int (X) represent the interior and the relative interior of the set X, respectively. Given a function  𝑓 (x) : ℝ𝑛 → ℝ, we say that the function is:

Definition A.1 (𝜇-strongly convex function). The function is  𝜇-strongly convex over X if there exists a 𝜇 > 0such that:

image

for all  x, y ∈ X.

Definition A.2 (𝐿-smooth function). The function is  𝐿-smooth over X if there exists a  𝐿 > 0such that:

image

for all  x, y ∈ X.

A simple schematic representation of the bounds provided by convexity,  𝜇-strong convexity and  𝐿-smoothness can be seen in Figure 4.

Definition A.3 (𝐿2-Lipschitz continuous Hessian). The function has a  𝐿2-Lipschitz continuous Hessian over X if there exists a  𝐿2 > 0such that:

image

for all  x, y ∈ X.

Definition A.4 (Normal cone of X). We define the normal cone of the set X at point x, denoted by  𝑁X (x),as:

image

image

Figure 4: The red line depict the quadratic upper bound from  𝐿-smoothness, the blue line depicts the quadratic lower bound provided by  𝜇-strong convexity, the green line depicts the linear lower bound provided by convexity.

A.1 Hessian Approximation Accuracy

Lemma A.5. Let  𝑃, 𝑄 ∈ S𝑛++. The solution to the fractional quadratic program  maxu∈ℝ𝑛 ∥u∥2𝑄 /∥u∥2𝑃is given by the the largest eigenvalue of the symmetric positive definite matrix  𝑃−1/2𝑄𝑃−1/2, that is, 𝜆max�𝑃−1/2𝑄𝑃−1/2�, which in turn is equal to  𝜆max�𝑃−1𝑄�. Moreover, the solution to the fractional quadratic program  minu∈ℝ𝑛 ∥u∥2𝑄 /∥u∥2𝑃 is given by the smallest eigenvalue of the symmetric positive matrix  𝑃−1/2𝑄𝑃−1/2,that is  𝜆min�𝑃−1/2𝑄𝑃−1/2�, which in turn is equal to  𝜆min�𝑃−1𝑄�.

Proof. Writing out the expression for the quadratic program we have that:

image

Moreover, note that as  𝑃 and 𝑄are positive definite.  𝜆max�𝑃−1/2𝑄𝑃−1/2� = 𝜆max�𝑃−1𝑄�. The second claim follows using a very similar reasoning. □

Lemma A.6. Given two matrices  𝑃, 𝑄 ∈ S𝑛++, then for all  v ∈ ℝ𝑛:

image

with 𝜂 = max�𝜆max�𝑃−1𝑄� , 𝜆max�𝑄−1𝑃��≥ 1.

Proof. Let 𝜆𝑖 (𝑃)denote the  𝑖-th eigenvalue of matrix  𝑃. Note that as  𝑃 and 𝑄are positive definite  𝑃−1 and𝑄−1 are well-defined, furthermore  𝑃−1𝑄 and 𝑄−1𝑃are also positive definite, as the eigenvalues of  𝑃−1𝑄 arethe same as those of the symmetric positive definite matrix  𝑃−1/2𝑄𝑃−1/2, and the eigenvalues of  𝑄−1𝑃 are thesame as those of the symmetric positive definite matrix  𝑄−1/2𝑃𝑄−1/2. In order to show that  𝜂 ≥ 1 note thatif  𝜆max�𝑄−1/2𝑃𝑄−1/2� = 𝜆max�𝑄−1𝑃� ≤ 1, then  𝜆𝑖�𝑄−1/2𝑃𝑄−1/2� = 𝜆𝑖�𝑄−1𝑃� ∈ (0, 1]for all  𝑖 ∈ ⟦1, 𝑛⟧, and therefore the eigenvalues of its inverse satisfy  𝜆𝑖�(𝑄−1𝑃)−1� = 𝜆𝑖�𝑃−1𝑄� ≥ 1for all  𝑖 ∈ ⟦1, 𝑛⟧. Conversely, if 𝜆max�𝑃−1𝑄� ≤ 1, the same reasoning applies, and  𝜆𝑖�𝑄−1𝑃� ≥ 1 for all 𝑖 ∈ ⟦1, 𝑛⟧. Note that the definition of 𝜂together with Lemma A.5 implies that 1𝜂 = min�𝜆min�𝑃−1𝑄� , 𝜆𝑚𝑖𝑛�𝑄−1𝑃��≤ 𝜆min�𝑃−1𝑄� = 𝜆max�𝑄−1𝑃�.Focusing on the first inequality on Equation (A.1) and plugging in the value of  𝜂 leads to:

image

Focusing on the second inequality of Equation (A.1) and noting that  𝜂 = max�𝜆max�𝑃−1𝑄� , 𝜆max�𝑄−1𝑃��≥𝜆max�𝑃−1𝑄�we have that:

image

Which completes the proof. □

Remark A.7. Given two matrices  𝑃, 𝑄 ∈ S𝑛++, then for all  v ∈ ℝ𝑛:

image

If we define the ellipsoid  E𝑃 =�v ∈ ℝ𝑛 | v𝑇𝑃v ≤ 1�for  𝑃 ∈ S𝑛++, we can interpret the value of  𝜂as being the smallest value that ensures that  E𝑃/𝜂 ⊆ E𝑄 ⊆ E𝜂𝑃 for 𝑄 ∈ S𝑛++ (see Figure 5).

The following corollary will allow us to bound the maximum and minimum eigenvalue of the approximation 𝐻𝑘in terms of the maximum and minimum eigenvalue of  ∇2 𝑓 (x𝑘) and 𝜂𝑘 for all 𝑘 ≥ 0, which will be useful in the proofs to follow.

Corollary A.8. Given two matrices  𝑃, 𝑄 ∈ S𝑛++, we have that:

image

where  𝜂 = max�𝜆max�𝑃−1𝑄� , 𝜆max�𝑄−1𝑃�� ≥ 1. This allows us to conclude that 𝜆min(𝑃)𝜂 𝐼𝑛 ⪯ 𝑄 ⪯𝜂𝜆max (𝑃) 𝐼𝑛.

image

Figure 5: Given  𝑃, 𝑄 ∈ S𝑛++, we can always find an  𝜂 such that E𝑃/𝜂 ⊆ E𝑄 ⊆ E𝜂𝑃.

Proof. Let  vmin (𝑄)and  vmax (𝑄)denote the eigenvectors of unit length associated with the minimum and maximum eigenvalue of  𝑄, denoted by  𝜆min (𝑄)and  𝜆max (𝑄)respectively. As  𝑃, 𝑄 ∈ S𝑛++from Lemma A.6 we have that:

image

On the other hand, using similar arguments we have:

image

Moving on to the bound for  𝜆max (𝑄) we have:

image

Similarly, we have that:

image

Combining these bounds completes the proof. □

Particularizing Corollary A.8 with  𝑄 = 𝐻𝑘and  𝑃 = ∇2 𝑓 (x𝑘)allows us to conclude that  𝜇/𝜂𝑘𝐼𝑛 ⪯𝐻𝑘 ⪯ 𝜂𝑘𝐿𝐼𝑛, and so the quadratic approximation ˆ𝑓𝑘(x)in Equation (2.4) will be  𝜇/𝜂𝑘-strongly convex and 𝜂𝑘𝐿-smooth.

Appendix B. The Conditional Gradients algorithm

We define the linear approximation of the function  𝑓 (x)around the point  x𝑘 as:

image

At each iteration the vanilla Conditional Gradients (CG) algorithm (Levitin & Polyak, 1966; Frank & Wolfe, 1956; Jaggi, 2013) (Algorithm 3) takes steps defined as  x𝑘+1 = x𝑘 + 𝛾𝑘(argminx∈X ˆ𝑙𝑘(x) − x𝑘) with 𝛾𝑘 ∈ (0, 1].As the iterates are formed as convex combinations of points in X there is no need for projections onto X, making the algorithm projection-free.

image

A useful quantity that can readily be computed in all CG steps is  ⟨∇ 𝑓 (x𝑘), x𝑘 − v𝑘⟩, known as the Frank-Wolfe gap, which provides an upper bound on the primal gap. If  x∗ ∈ argmin

image

where the last inequality follows from the convexity of  𝑓 (x). This quantity is often used as a stopping criterion when running the CG algorithm. The CG algorithm has seen a renewed interest from the Machine Learning community, as several machine learning problems can be phrased as constrained optimization problems with feasible regions onto which it is hard to project on (Joulin et al., 2014; Futami et al., 2019; Garber et al., 2018).

B.1 Global Convergence

The CG algorithm with exact line search converges linearly in primal gap when applied to Problem (1.1) when  x∗ ∈ int (X)(Guélat & Marcotte, 1986). However, when  x∗ ∈ X \ int (X)the algorithm suffers from a zig-zagging phenomenon - as the iterates get closer to  x∗the directions provided by the algorithm starts to become close to perpendicular to the gradient (Figure 6a). This is remedied by using Away-steps (Algorithm 5), which result in the Away-step Conditional Gradient (ACG) algorithm (Algorithm 4, Figure 6b) (Wolfe, 1970), which converges linearly in primal gap regardless of the location of  x∗when using exact line search (Lacoste-Julien & Jaggi, 2015) or a step size strategy dependent on  𝐿 (Pedregosa et al., 2020).

image

The ACG algorithm maintains what is called an  active set S𝑘 ⊆ vert (X)which represents the potentially non-unique set of vertices of  X such that x𝑘 ∈ conv (S𝑘). Associated with this active set  S𝑘we have a set of barycentric coordinates  λ𝑘such that if we denote by  λ𝑘(u) ∈ [0, 1]the element of  λ𝑘associated with  u ∈ S𝑘we have that  x𝑘 = �u∈S𝑘 λ𝑘(u)u, with �u∈S𝑘 λ𝑘(u) = 1 and λ𝑘(u) ≥ 0 for all u ∈ S𝑘.

image

Figure 6: Qualitative performance comparison of the CG and the ACG algorithm.

image

1 v ← argminv∈X ⟨∇ 𝑓 (x) , v⟩2 a ← argmaxv∈S ⟨∇ 𝑓 (x) , v⟩3 if ⟨∇ 𝑓 (x), x − v⟩ ≥ ⟨∇ 𝑓 (x), a − x⟩ then4 d ← x − v, 𝛾max ← 15 else

6 d ← a − x, 𝛾max ← λ(a)/(1 − λ(a))7 end

8 𝛾 ← argmin𝛾∈[0,𝛾max] 𝑓 (x + 𝛾d)9 x′ ← x + 𝛾d10 if ⟨∇ 𝑓 (x), x − v⟩ ≥ ⟨∇ 𝑓 (x), a − x⟩ then11 if 𝛾 = 1 then

12 S′ ← {v}13 else

14 S′ ← S ∪ {v}15 end

16 λ′(u) ← (1 − 𝛾) λ(u) if u ∈ S \ v17 λ′(v) ← (1 − 𝛾) λ(v) + 𝛾18 else

19 if 𝛾 = 𝛾max then20 S′ ← S \ {a}21 else

22 S′ ← S23 end

24 λ′(u) ← (1 + 𝛾) λ(u) if u ∈ S \ a25 λ′(a) ← (1 + 𝛾) λ(a) − 𝛾26 end

In general one of the easiest ways to maintain the active set is to build a list of previously used vertices and a list of associated barycentric coordinates. If the Frank-Wolfe step adds a new vertex v that is not already in S it is added to the list of vertices and its associated barycentric coordinate is added to the list of barycentric coordinates. If the vertex v is already contained in the list that maintains S, its existing barycentric coordinate is updated in the appropiate list. Note that the barycentric coordinates of the points S \ {v} are also updated at each iteration. The away-steps in Algorithm 5 cannot add new vertices, only remove them from the active set. This type of step also requires updating the barycentric coordinates of the points S \ {a}. For both Frank-Wolfe and away-steps a vertex is removed from the list of vertices and the associated barycentric coordinate removed from the list of coordinates if the value of the barycentric coordinate is zero.

The first proof of asymptotic linear convergence of the ACG algorithm relied on the strict complementarity of the problem in Equation (1.1) (shown in Assumption 1), which we will also use in the convergence proof of the SOCGS algorithm.

Assumption 1 (Strict Complementarity). We have that  ⟨∇ 𝑓 (x∗) , x − x∗⟩ = 0if and only if  x ∈ F (x∗).

Remark B.1. Assumption 1 automatically holds if  x∗ ∈ int (X), that is, if  x∗ is in the strict interior of X. In this case the polytope is fully-dimensional and itself the optimal face, so no off-optimal-face vertices exist.

If Assumption 1 is satisfied the iterates of the ACG algorithm reach  F (x∗)in a finite number of steps, remaining in  F (x∗)for all subsequent iterations (Guélat & Marcotte, 1986). When inside  F (x∗), the iterates of the ACG algorithm contract the primal gap linearly. This analysis was later significantly extended to provide an explicit global linear convergence rate in primal gap (Theorem 2.1), by making use of the pyramidal width of the polytope X (Lacoste-Julien & Jaggi, 2015). With the pyramidal width one can derive a primal progress guarantee for all steps taken by the ACG algorithm except "bad" away-steps that reduce the cardinality of the active set  S𝑘, that is when  ⟨∇ 𝑓 (x𝑘), x𝑘 − v⟩ < ⟨∇ 𝑓 (x𝑘), a − x𝑘⟩and the step size satisfies 𝛾𝑘 = 𝛾maxin Algorithm 5. This cannot happen more than  ⌊𝐾/2⌋times when running the ACG algorithm for  𝐾iterations (as the algorithm cannot drop more vertices with away-steps than it has picked up with Frank-Wolfe steps). This is an important consideration to keep in mind, as it means that the ACG primal gap contraction does not hold on a per-iteration basis.

Theorem B.2 (Primal gap convergence of the ACG algorithm (Algorithm 4)). (Lacoste-Julien & Jaggi,

2015, Theorem 1) Given an  𝐿-smooth and  𝜇-strongly convex function  𝑓 (x), a polytope X and an initial point x0 ∈ X, the ACG algorithm satisfies after  𝐾 ≥ 0iterations:

image

where  𝐷denotes the diameter of the polytope  X and 𝛿its pyramidal width.

See also (Garber & Hazan, 2016; Diakonikolas et al., 2020) for work on linearly convergent CG algorithms, and (Jaggi, 2013; Lan, 2013) for strong lower bounds that limit the linear convergence that can be achieved with algorithms that only access the feasible region through a linear optimization oracle.

Appendix C. Projected Variable-Metric algorithms

In this section we provide theoretical context for the Projected Variable-Metric (PVM) algorithm (Algorithm 6), and we present several well-known results that will be helpful in motivating the SOCGS algorithm.

image

At each iteration the PVM algorithm builds a quadratic approximation of the original function  𝑓 (x), and moves towards the point that minimizes this approximation over X. Formally, we denote the quadratic approximation of  𝑓 (x) at x𝑘 using 𝐻𝑘 ∈ S𝑛++ as:

image

where  𝐻𝑘is an approximation to the Hessian  ∇2 𝑓 (x𝑘). In order to measure how well  𝐻𝑘approximates ∇2 𝑓 (x𝑘)we note that for any  𝐻𝑘 ∈ S𝑛++ and all y ∈ X that:

image

where  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} ≥ 1(see Lemma A.6 in Appendix A.1). We will use the value of  𝜂𝑘to measure the accuracy of how well  𝐻𝑘approximates  ∇2 𝑓 (x𝑘). For example, an  𝜂𝑘 = 1means that  𝐻𝑘 = ∇2 𝑓 (x𝑘). If we were to use  𝐻𝑘 = 𝐼𝑛 we would have that  𝜂𝑘 = max {𝐿, 1/𝜇}.

Just as the steps taken by the Projected Gradient Descent (PGD) algorithm can be interpreted in terms of Euclidean projection operators, the steps taken by the PVM algorithm in Line 2 of Algorithm 6 can be interpreted in terms of scaled projection operators, where the norm of the projection operator is defined by 𝐻𝑘 ∈ S𝑛++. Let  Π𝐻X (x) : ℝ𝑛 → Xdenote the scaled projection of x onto X using the matrix norm  ∥·∥𝐻, more concretely  Π𝐻X (x)def= argminy∈X12 ∥y − x∥2𝐻. We have that:

image

Remark C.1 (First-order optimality condition for PVM subproblems). The solution to the problem in Line 2 of Algorithm 6 (also shown in Equation (C.3)), that is,  ˜x∗𝑘+1 = argminx∈X ˆ𝑓𝑘 (x)satisfies for all  z ∈ X:

image

In both the PGD and the PVM algorithm the only point that is invariant under the steps taken by the algorithms is  x∗. That is, in the case of the PGD algorithm we have that  Π𝐼𝑛X (x − ∇ 𝑓 (x)) = x∗if and only if x = x∗. Similarly, in the case of the PVM algorithm we have that  Π𝐻𝑘

image

Lemma C.2. Given a matrix  𝐻 ∈ S𝑛++, for any  x ∈ Xand  d ∈ 𝑁X (x)(where  𝑁X (x)represents the normal cone of X at x, see Definition A.4) we have that:

image

Proof. From the definition of the normal cone, given a  x ∈ X and d ∈ 𝑁X (x)we know that for all  y ∈ X

image

for all  y ∈ X. This means that the closest point to  x + 𝐻−1dthat is in X, when we measure the distance in the  𝐻norm, is given by x itself, i.e.,  Π𝐻

holds for  𝐻 = 𝐼𝑛. □

Lemma C.3. Given a matrix  𝐻 ∈ S𝑛++, an x ∈ Xsatisfies:

image

if and only if  x = x∗ where x∗ = argminx∈X 𝑓 (x).

Proof. (⇒) Using the first-order optimality conditions for the scaled projection problem, shown in Remark C.1, and particularizing for  ˜x∗𝑘+1 = x𝑘 = xwe have that for all  z ∈ X:

image

which hold true if and only if  x = x∗, as Equation (C.9) represents the first-order optimality conditions for Problem 1.1, of which  x∗ is the unique optimal solution.

(⇐) Assume that  x = x∗, then  −∇ 𝑓 (x∗) ∈ 𝑁X (x∗). By the application of Lemma C.2 we have that for any 𝐻 ∈ S𝑛++ then it holds that  x = Π𝐻X�x − 𝐻−1∇ 𝑓 (x)�. □

Another interesting property of the PVM algorithm is the fact that the direction  ˜x∗𝑘+1 − x𝑘in Line 3 ofAlgorithm 6 is a descent direction regardless of how well  𝐻𝑘 ∈ S𝑛++ approximates the Hessian  ∇2 𝑓 (x𝑘), this isformalized in Lemma C.4. Note that despite this, we cannot guarantee that  𝑓 (˜x∗𝑘+1) ≤ 𝑓 (x𝑘), which is why to ensure primal progress at each iteration a line search or a bounded step size is often used in Line 3 of Algorithm 6.

Lemma C.4 (Descent property of Projected Variable-Metric directions). (Ben-Tal & Nemirovskii, 2020, Section 7.2.1) If  𝐻𝑘 ∈ S𝑛++ and x𝑘 ≠ x∗, then the directions given by  ˜x∗𝑘+1 − x𝑘, where ˜x∗𝑘+1 = argminx∈X ˆ𝑓𝑘(x)are descent directions at point  x𝑘, i.e., they satisfy�−∇ 𝑓 (x𝑘), ˜x∗𝑘+1 − x𝑘�> 0.

Proof. Using the first-order optimality conditions shown in Remark C.1 for the scaled projection subproblem and particularizing for  z = x𝑘:

image

Where the last strict inequality follows from the fact that we have assumed that  x𝑘 ≠ x∗, and consequently ˜x∗𝑘+1 ≠ x𝑘by application of Lemma C.3, and the assumption that  𝐻𝑘 ∈ S𝑛++, thus��˜x∗𝑘+1 − x𝑘��2𝐻𝑘 > 0. □

C.1 Global Convergence

The global primal gap convergence of the PVM algorithm (Algorithm 6) with bounded step sizes is a well-known result that we reproduce here for completeness, as we will compare this global convergence rate with that of other first-order optimization algorithms. In order to prove it, we review Lemma C.5 which will be used in the global convergence proof.

Lemma C.5. (Karimireddy et al., 2018b, Lemma 9) Given a convex domain X and  𝐻𝑘 ∈ S𝑛++then for constants  𝛼 > 0 and 𝜈 > 0 such that 𝛼𝜈 ≥ 1we have that:

image

With the previous Lemma at hand, we can prove the global linear convergence in primal gap of the PVM algorithm with bounded step size when minimizing a  𝜇-strongly convex and  𝐿-smooth function over a convex set X.

Theorem C.6 (Global convergence of Projected Variable-Metric algorithm with bounded step size.). (Karimireddy et al., 2018a, Theorem 4) Given an  𝐿-smooth and  𝜇-strongly convex function and a convex set

X then the Projected Variable-Metric algorithm (Algorithm 6) with a step size  𝛾𝑘 ≤ 𝜇𝐿𝜂𝑘guarantees for all

image

where the parameter  𝜂𝑘measures how well  𝐻𝑘approximates  ∇2 𝑓 (x𝑘).

Proof. The iterate  x𝑘+1can be rewritten as:

image

Using  𝐿-smoothness and the  𝜇-strong convexity of the function  𝑓we can write:

𝑓 (x𝑘+1) − 𝑓 (x𝑘) ≤ ⟨∇ 𝑓 (x𝑘), x𝑘+1 − x𝑘⟩ + 𝐿2𝜇 ∥x𝑘+1 − x𝑘∥2∇2 𝑓 (x𝑘 ) (C.12)≤ ⟨∇  𝑓 (x𝑘), x𝑘+1 − x𝑘⟩ + 𝐿𝜂𝑘2𝜇 ∥x𝑘+1 − x𝑘∥2𝐻𝑘 (C.13)≤ ⟨∇  𝑓 (x𝑘), x𝑘+1 − x𝑘⟩ + 12𝛾𝑘∥x𝑘+1 − x𝑘∥2𝐻𝑘 (C.14)

image

Where the second inequality follows from Equation (2.6) (which in turn is a consequence of  𝐻𝑘 ∈ S𝑛++) and the third inequality follows from the fact that  𝛾𝑘 ≤ 𝜇𝐿𝜂𝑘 . Applying Lemma C.5 to Equation (C.15) and noting that as  𝐻𝑘 ∈ S𝑛++ we can apply Equation (2.6) and transform the minimization problem involving  ∥x − x𝑘∥𝐻𝑘

to one that involves  ∥x − x𝑘∥∇2 𝑓 (x𝑘 ). Continuing with the chain of inequalities:

image

We obtain Equation (C.17) by applying Lemma C.5, and Equation (C.18) from applying Lemma A.6 to the norm term in Equation (C.17), which allows us to use that  1/𝜂𝑘 ∥x − x𝑘∥2𝐻𝑘 ≤ ∥x − x𝑘∥2∇2 𝑓 (x𝑘 ). Equation (C.19)follows from plugging in  x = (1 − 𝛾𝑘)x𝑘 + 𝛾𝑘x∗into Equation (C.18) (as of course  x∗ ∈ X). We obtain Equation (C.20) by considering that  𝛾𝑘 ≤ 1, and Equation (C.21) from the  𝜇-strong convexity and  𝐿-smoothness of the function  𝑓 (x). Reordering the previous expression leads to:

image

As the exact line search strategy makes at least as much progress as choosing any  𝛾𝑘 ≤ 𝜇𝐿𝜂𝑘 , the bound in Theorem C.6 also holds for the Projected Variable-Metric algorithm (Algorithm 6) with exact line search.

Corollary C.7 (Global convergence of Projected Variable-Metric algorithm with exact line search or 𝛾𝑘 = 𝜇𝐿𝜂𝑘). Given an  𝐿-smooth and  𝜇-strongly convex function and a convex set X then the Projected Variable-Metric algorithm (Algorithm 6) with an exact line search or with a step size  𝛾𝑘 = 𝜇𝐿𝜂𝑘 guarantees for all  𝑘 ≥ 0:

image

where the parameter  𝜂𝑘measures how well  𝐻𝑘approximates  ∇2 𝑓 (x𝑘).

As was mentioned in Lemma C.4 the direction  ˜x∗𝑘+1 − x𝑘in Line 3 of Algorithm 6 is a descent directionregardless of how well  𝐻𝑘 ∈ S𝑛++ approximates the Hessian  ∇2 𝑓 (x𝑘). However, as we can see in Theorem C.6 and Corollary C.7, if we pick a matrix  𝐻𝑘 ∈ S𝑛++that approximates the Hessian  ∇2 𝑓 (x𝑘)well, that is, we have an  𝜂𝑘close to 1, we will be able to guarantee more primal progress per step when using an exact line search or bounded step sizes.

One of the key consequences of Corollary C.7 is that even if we run the PVM algorithm with an exact line search and we use  𝐻𝑘 = ∇2 𝑓 (x𝑘)(which is equivalent to  𝜂𝑘 = 1), we need O(𝐿3/𝜇3 log 1/𝜀)iterations to reach an  𝜀-optimal solution to Problem (1.1). This stands in contrast to the PGD algorithm, which requires O(𝐿/𝜇 log 1/𝜀)iterations, or  Nesterov’s Projected Gradient Descent(NPGD) algorithm, which requires

√︁𝐿/𝜇 log 1/𝜀)iterations to reach an  𝜀-optimal solution. Note that with a small modification of the proof in Theorem C.6 we can recover the same rate for the PGD algorithm and the PVM algorithm with  𝐻𝑘 = 𝐼𝑛.This is expected, as in this case the algorithms are equivalent, except for the bounded step size strategy.

Theorem C.8 (Global convergence of Projected Variable-Metric algorithm with bounded step size and 𝐻𝑘 = 𝐼𝑛). Given an 𝐿-smooth and  𝜇-strongly convex function and a convex set X then the Projected VariableMetric algorithm (Algorithm 6) with a step size  𝛾𝑘 ≤ min{1, 1𝐿 } and 𝐻𝑘 = 𝐼𝑛guarantees for all  𝑘 ≥ 0:

image

Proof. The proof mirrors that of Theorem C.6, and so we only give a brief outline. The iterate  x𝑘+1 can berewritten as:

image

Using  𝐿-smoothness we can write:

𝑓 (x𝑘+1) − 𝑓 (x𝑘) ≤ ⟨∇ 𝑓 (x𝑘), x𝑘+1 − x𝑘⟩ + 𝐿2 ∥x𝑘+1 − x𝑘∥2 (C.24)≤ ⟨∇  𝑓 (x𝑘), x𝑘+1 − x𝑘⟩ + 12𝛾𝑘∥x𝑘+1 − x𝑘∥2 (C.25)

image

Where Equation (C.25) follows from  𝛾𝑘 ≤ min{1, 1𝐿 }and Equation (C.26) follows from Equation (C.23). Applying Lemma C.5 to Equation (C.26) leads to Equation (C.27). Equation (C.28) follows from plugging in  x = (1 − 𝛾𝑘)x𝑘 + 𝛾𝑘x∗ into Equation (C.27) (as of course  x∗ ∈ X) Lastly, in Equation (C.29) we have used 𝜇-strong convexity and the fact that  𝛾𝑘 ≤ min{1, 1𝐿 }. Reordering the terms previous inequality completes the proof. □

C.2 Local Convergence

Despite the lackluster convergence rate in primal gap shown in Theorem C.6, the PVM algorithm can achieve quadratic convergence in distance to the optimum when the iterates are close enough to the optimum and the Hessian approximations are accurate enough. We first review a series of results that will allow us to prove the local quadratic convergence of the PVM algorithm. One of the key properties that is often used in the convergence proof of the PGD algorithm is the non-expansiveness of the Euclidean projection operator onto a convex set X, denoted by  Π𝐼𝑛X . In the local convergence proof of the PVM algorithm we use a generalization of the aforementioned fact, that is, the scaled projection operator onto a convex set X, denoted by  Π𝐻𝑘X where𝐻 ∈ S𝑛++, is also non-expansive (see Lemma C.9).

Lemma C.9. (Beck, 2017)[Theorem 6.42] Given a  𝐻 ∈ S𝑛++and a convex set X, the scaled projection is a contraction mapping (it is firmly-nonexpansive) in the  𝐻-norm:

image

Lemma C.10. (Nesterov, 2018)[Lemma 4.1.1] If a twice differentiable function  𝑓 has 𝐿2-Lipschitz continuous Hessian over X then for all  x, y ∈ X:

image

With the results from Lemma C.9 and Lemma C.10 we can formalize the local convergence of the PVM algorithm.

Lemma C.11 (Local convergence of Projected Variable-Metric algorithm). Given an 𝐿-smooth and  𝜇-stronglyconvex function with  𝐿2-Lipschitz Hessian and a compact convex set X, if  ˜x∗𝑘+1 = argminx∈X ˆ𝑓𝑘(x)then for all 𝑘 ≥ 0:

image

The last inequality is a consequence of the  𝜇-strong convexity of  𝑓which ensures that  ∇2 𝑓 (x𝑘)−1 ⪯ 𝜇−1𝐼𝑛. Using the fact that the Hessian is  𝐿2-Lipschitz and applying Lemma C.10 and using the  𝐿-smoothness of  𝑓leads to:

image

Using Lemma A.6 along with the  𝜇-strong convexity of  𝑓and reordering the expression shown in Equation (C.35) completes the proof. □

As we can see, even if the scaled projection subproblems are solved to optimality we arrive at a convergence rate for��˜x∗𝑘+1 − x∗��that is linear-quadratic in terms of  ∥x𝑘 − x∗∥, and we do not obtain local quadratic convergence without additional assumptions on how well  𝐻𝑘approximates  ∇2 𝑓 (x𝑘), due to  𝜂𝑘 − 1in the second term in Equation (C.35). This can be remedied with Assumption 2:

Corollary C.12. If in addition to the conditions described in Lemma C.11 we also assume that Assumption 2 is satisfied, we have:

image

where  𝜔 ≥ 0is described in Equation (2.7).

Even though��˜x∗𝑘+1 − x∗��may converge quadratically, what we are interested in is in the quadratic convergence of  ∥x𝑘+1 − x∗∥, formed as  x𝑘+1 = x𝑘 + 𝛾𝑘(˜x∗𝑘+1 − x𝑘), that is:

image

We can see from Equation (C.39) that we will only have the desired convergence rate if  (1 − 𝛾𝑘) ≤ 𝛽 ∥x𝑘 − x∗∥for some  𝛽 ≥ 0, that is, we either need to set  𝛾𝑘 = 1, or select a step size strategy that makes  𝛾𝑘converge to 1 fast enough.

Appendix D. Second-order Conditional Gradient Sliding

In Section D.1 we prove that the Inexact PVM steps (Lines 9-12 in Algorithm 2) that the SOCGS algorithm computes contract the distance to the optimum and the primal gap quadratically when close enough to  x∗, by carefully choosing the  𝜀𝑘-parameter at each iteration. First, we review the SOCGS from the main body of the text (shown in Algorithm 7), and then we review a key result in Lemma D.1 that measures the accuracy of the Hessian matrix approximation  𝐻as we approach  x∗, which will be used in the convergence proofs.

image

The algorithm couples an independent ACG step with line search (Line 4) with an Inexact PVM step with unit step size (Lines 9-12). At the end of each iteration we choose the step that provides the greatest primal progress (Lines 14-18). The ACG steps in Line 4 will ensure global linear convergence in primal gap, and the Inexact PVM steps in Lines 14-18 will provide quadratic convergence.

Note that the ACG iterates in Line 4 do not depend on the Inexact PVM steps in Lines Lines 9-12. This is because the ACG steps do not contract the primal gap on a per-iteration basis, and if the active sets of the ACG steps in Line 4 were to be modified using the active set of the PVM steps in Lines 9-12, this would break the proof of linear convergence in Theorem B.2 for the ACG algorithm. The proof in Theorem B.2 crucially relies on the fact that at each iteration of the ACG algorithm we can pick up or drop at most one vertex from the active set, whereas a PVM step may have dropped or picked up multiple vertices from the active set. The line search in the ACG step (Line 4) can be substituted with a step size strategy that requires knowledge of the  𝐿-smoothness parameter of  𝑓 (x) (Pedregosa et al., 2020).

We compute the scaled projection in the Inexact PVM step (Lines 14-18) using the ACG algorithm with exact line search, as the objective function is quadratic, thereby making the SOCGS algorithm (Algorithm 7) projection-free. As the function being minimized in the Inexact PVM steps is quadratic there is a closed-form expression for the optimal step size in Line 10. The scaled projection problem is solved to an accuracy  𝜀𝑘such that ˆ𝑓𝑘(˜x𝑘+1) − minx∈X ˆ𝑓𝑘 (x) ≤ 𝜀𝑘, using the Frank-Wolfe gap as a stopping criterion, as in the CGS algorithm (Lan & Zhou, 2016). The accuracy parameter  𝜀𝑘in the SOCGS algorithm depends on a lower bound on the primal gap of Problem 1.1 which we denote by  𝑙𝑏 (x𝑘)that satisfies  𝑙𝑏 (x𝑘) ≤ 𝑓 (x𝑘) − 𝑓 (x∗).

Lemma D.1. Given a 𝜇-strongly convex and  𝐿-smooth function  𝑓 (x)and a convex set X, then for any  x ∈ Xand any matrix  𝐻 ∈ S𝑛++ that satisfies Assumption 2 at x we have that:

image

Similarly, we also have that:

image

Proof. We can bound the term on the left-hand side of Equation (D.1) as:

image

We obtain Equation (D.4) from the fact that the spectral norm of a matrix is submultiplicative, and both matrices are square. The inequality shown in Equation (D.6) follows from  𝐻 ∈ S𝑛++and Corollary A.8. Proceeding similarly, we can also bound the previous quantity as:

image

Where the inequality in Equation (D.8) follows from fact that  𝜂 ≥ 1. Putting together these bounds, we have that:

image

Each of the terms in the maximization operator in the previous equation can be written as:

image

Where the equality in Equation (D.10) follows from the fact that the maximum singular value of a square matrix is equal to the maximum absolute value of the eigenvalues of the matrix. This allows us to write:

image

image

This means that for all  1 ≤ 𝑖 ≤ 𝑛we have that:

image

Which allows us to write Equation (D.12) as:

image

Which immediately leads to:

image

Where Equation (D.15) follows from the definition of  𝜂and Equation (D.16) follows from Assumption 2. Putting this all together allows us to write:

image

The claim shown in Equation (D.2) follows from a very similar reasoning. With the only difference that:

image

The maximization term on the right-hand side of Equation (D.17) can be bound exactly like in the first claim. □

D.1 Inexact Projected Variable-Metric steps

We first begin by showing that if the PVM steps are computed inexactly using the error criterion shown in the SOCGS algorithm (Line 7 of Algorithm 2) they still achieve local quadratic convergence in distance to the optimum.

Lemma D.2. Given a  𝜇-strongly convex function  𝑓 (x)and a compact convex set X, if  ˜x𝑘+1denotes an 𝜀𝑘-optimal solution to  ˜x∗𝑘+1 = argminx∈X ˆ𝑓𝑘 (x)where  𝜀𝑘 = (𝑙𝑏(x𝑘)/∥∇ 𝑓 (x𝑘)∥)4and  𝑙𝑏(x𝑘)denotes a lower bound on the primal gap such that  𝑙𝑏(x𝑘) ≤ 𝑓 (x𝑘) − 𝑓 (x∗) then:

image

where the parameter  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} ≥ 1measures how well  𝐻𝑘 approx-imates  ∇2 𝑓 (x𝑘).

Proof. By the strong convexity of ˆ𝑓𝑘 (as 𝐻𝑘 ∈ S𝑛++) we have that:

image

The inequality in Equation (D.20) follows from Corollary A.8 and the one in Equation (D.21) from the first-order optimality conditions for the scaled projection problem, of which  ˜x∗𝑘+1 is the exact solution. Rearranging the previous expression allows us to conclude that��˜x𝑘+1 − ˜x∗𝑘+1�� ≤√︁2𝜂𝑘𝜀𝑘/𝜇. If we plug in the value of  𝜀𝑘 inthe previous bound:

image

Where the inequality in Equation (D.24) follows from the fact that  𝑙𝑏 (x𝑘)is a lower bound on the primal gap, the one in Equation (D.25) follows from the convexity of  𝑓 (x)and the last inequality, in Equation (D.26), follows from the Cauchy-Schwarz inequality. □

Using the previous bound along with Corollary C.12 we can show that the iterates will converge quadratically in distance to the optimum (Lemma D.3), despite not solving the problems to optimality.

Lemma D.3 (Quadratic convergence in distance to the optimum of the Inexact Projected-Variable Metric (PMV) steps). Given a 𝜇-strongly convex and  𝐿-smooth function  𝑓 (x) with 𝐿2-Lipschitz Hessian and a compact convex set  X, let ˜𝑥𝑘+1 denote an 𝜀𝑘-optimal solution to  ˜x∗𝑘+1 = argminx∈X ˆ𝑓𝑘 (x) where 𝜀𝑘 = (𝑙𝑏(x𝑘)/∥∇ 𝑓 (x𝑘)∥)4and  𝑙𝑏(x𝑘)denotes a lower bound on the primal gap such that  𝑙𝑏(x𝑘) ≤ 𝑓 (x𝑘) − 𝑓 (x∗), if Assumption 2 is satisfied then:

image

where the parameter  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} ≥ 1measures how well  𝐻𝑘 approx-imates  ∇2 𝑓 (x𝑘) and 𝜔is defined in Assumption 2.

Proof. Using the triangle inequality yields:

image

Where the second inequality follows from using the bounds shown in Corollary C.12 and Lemma D.2. □

The SOCGS algorithm chooses at each iteration between the ACG step and the Inexact PVM step according to which one provides more progress in primal gap (Lines 14-18 of Algorithm 2). Therefore we need to translate the local rate in distance to the optimum of the PVM algorithm in Lemma D.3 to one in primal gap. It is immediate to see that we can upper bound the right-hand side of Equation (D.27) using 𝜇-strong convexity, as:

image

However, when we try to lower bound the norm that appears on the left-hand side of Equation (D.27) using 𝐿-smoothness we arrive at:

image

The only term preventing us from expressing the left-hand side of Equation (D.28) solely in terms of primal gap values is  − ⟨∇ 𝑓 (x∗), ˜x𝑘+1 − x∗⟩. As by Assumption 1 for any  x ∈ F (x∗)we have that  ⟨∇ 𝑓 (x∗) , x − x∗⟩ = 0,if we can show that from some point onward the iterates  ˜x𝑘+1 remain in F (x∗), we will be able conclude that ⟨∇ 𝑓 (x∗), ˜x𝑘+1 − x∗⟩ = 0.

The main tool that we will use for the analysis is based on the idea that for points  x𝑘sufficiently close to x∗, when we minimize ˆ𝑓𝑘(x)over X using the ACG algorithm, the iterates  ˜x𝑘+1of the algorithm will reach F (x∗)in a finite number of iterations, remaining in  F (x∗)for all subsequent iterations, that is, the ACG algorithm "identifies" the optimal face while computing the Inexact PVM steps. This is a variation of the proof originally presented in Guélat & Marcotte (1986), which was used to show for the first time that the ACG algorithm asymptotically converges linearly in primal gap when minimizing a strongly convex and smooth function over a polytope. We reproduce the original proof here, as it will be useful in the technical results to come.

Theorem D.4 (Identification of the optimal face). (Guélat & Marcotte, 1986)[Theorem 5] Given a strongly convex and smooth function  𝑓 (x)and a polytope X, if Assumption 1 is satisfied, then there is a  𝑟ACG > 0such that for  xACG𝑘 ∈ B(x∗, 𝑟ACG) ∩ Xand  xACG𝑘 ∉ F (x∗)then the ACG algorithm (Algorithm 4) with exact line search satisfies that  |SACG𝑘+1 | < |SACG𝑘 | and SACG𝑘 \ SACG𝑘+1 ∉ F (x∗). That is, the ACG algorithm performs an away-step that drops a vertex from  SACG𝑘that is not a vertex of the optimal face  F (x∗). Moreover, there is a  𝐾ACG ≥ 0such that for  𝑘 ≥ 𝐾ACG we have that  xACG𝑘 ∈ F (x∗).

Proof. The proof starts by showing that there is an index  𝑇 ≥ 0such that for  𝑘 ≥ 𝑇all the steps taken by the ACG algorithm will be away-steps that reduce the cardinality of the active set if  xACG𝑘 ∉ F (x∗). Let  𝑟𝑖 > 0and  𝑐 > 0be such that:

image

Taking  𝑟ACG = minv𝑖 ∈vert(X) 𝑟𝑖, we know by strong convexity that there is an index  𝑇 ≥ 0such that for  𝑘 ≥ 𝑇we have that  xACG𝑘 ∈ B(x∗, 𝑟ACG) ∩ X. Furthermore, suppose that  xACG𝑘 ∉ F (x∗), then we have that:

image

Where the left-hand side follows from Equation (D.30) and the right-hand side from Equation (D.29). As xACG𝑘 ∉ F (x∗), then SACG𝑘 ∩ vert(X) \ vert(F (x∗)) ≠ ∅, as the active set  SACG𝑘must include vertices that are not in the optimal face  F (x∗)(otherwise we would have  xACG𝑘 ∈ F (x∗)). This means that the ACG algorithm in Line 3 of Algorithm 5 will choose an away-step with a vertex  v𝑖 ∈ SACG𝑘 ∩ vert(X) \ vert(F (x∗)), and not a Frank-Wolfe step with a vertex  v 𝑗 ∈ vert(F (x∗)), for iterations  𝑘 ≥ 𝑇. We denote the vertex chosen in the away-step by  v ∈ SACG𝑘 ∩ vert(X) \ vert(F (x∗)), and we remark that  d = xACG𝑘 − vis a descent direction at xACG𝑘, and so the exact line search will output a step size  𝛾𝑘 ∈ (0, 𝛾max]. The proof proceeds by showing that we must have that  𝛾𝑘 = 𝛾max in Line 8of Algorithm 5 for iterations  𝑘 ≥ 𝑇. Using proof by contradiction, we

assume that  𝛾𝑘 < 𝛾maxand we apply the first-order optimality conditions for the exact line search:

0 =�d, ∇ 𝑓 (xACG𝑘+1 )� (D.31)=�xACG𝑘+1 − v, ∇ 𝑓 (xACG𝑘+1 )�+�xACG𝑘 − xACG𝑘+1 , ∇ 𝑓 (xACG𝑘+1 )� (D.32)=�xACG𝑘+1 − v, ∇ 𝑓 (xACG𝑘+1 )�− 𝛾𝑘�d, ∇ 𝑓 (xACG𝑘+1 )� (D.33)=�xACG𝑘+1 − v, ∇ 𝑓 (xACG𝑘+1 )� (D.34)< −𝑐. (D.35)

Which is the desired contradiction as  𝑐 > 0. The equality in Equation (D.34) is due to�d, ∇ 𝑓 (xACG𝑘+1 )�= 0because of the optimality conditions of the exact line search and the inequality in Equation (D.35) is due to�xACG𝑘+1 − v, ∇ 𝑓 (xACG𝑘+1 )�≤ −𝑐as  v ∈ vert(X) \ vert(F (x∗))and  xACG𝑘+1 ∈ B(x∗, 𝑟ACG)(thus Equation (D.30) holds). This proves that we must have  𝛾𝑘 = 𝛾max and |SACG𝑘 | > |SACG𝑘+1 |. While 𝑘 ≥ 𝑇 and xACG𝑘 ∉ F (x∗) theACG algorithm will drop a vertex  SACG𝑘 ∩vert(X)\vert(F (x∗))using an away-step. As  |SACG𝑘 |is finite, we will have for some  𝐾ACG > 𝑇that  SACG𝐾ACG ∩ vert(X) \ vert(F (x∗)) = ∅, and therefore  SACG𝐾ACG ⊆ vert(F (x∗)). This is equivalent to  xACG𝐾ACG ∈ F (x∗). Lastly, using Equation (D.29) and  SACG𝐾 ∩ vert(X) \ vert(F (x∗)) = ∅ we canshow that the ACG algorithm will not perform any Frank-Wolfe steps with vertices  v ∈ vert(X) \ vert(F (x∗))for  𝑘 ≥ 𝐾ACG, and so x𝑘 ∈ F (x∗). □

The consequence of Theorem D.4 is that after a finite number of iterations  𝐾ACG ≥ 0the iterates of the ACG algorithm applied to Problem (1.1) are "stuck" in the face  F (x∗), that is, we have that  xACG𝑘 ∈ F (x∗)for all  𝑘 ≥ 𝐾ACG. The SOCGS algorithm (Algorithm 2) uses the ACG algorithm to inexactly solve the scaled projection problem of the PVM steps in Lines 14-18 of Algorithm 2. The function being minimized in these steps is not  𝑓 (x), but rather an approximation ˆ𝑓𝑘(x)that changes at each iteration. However for points sufficiently close to  x∗ we show in Theorem D.5 that the ACG steps that solve the scaled projection problem of the PVM steps (in Lines 9-12 of Algorithm 2) will also get "stuck" to  F (x∗), that is, there is a  𝐾 ≥ 0 suchthat we will have that  ˜x𝑘+1 ∈ F (x∗) for all 𝑘 ≥ 𝐾.

Theorem D.5. Let 𝑓 (x)be a strongly convex and smooth function with Lipschitz continuous Hessian and X be a polytope such that Assumption 1 is satisfied. We denote the quadratic approximation of  𝑓 (x)at  x𝑘as ˆ𝑓𝑘(x) = ⟨∇ 𝑓 (x𝑘), x𝑘 − x⟩ + 1/2 ∥x𝑘 − x∥2𝐻𝑘, where  𝐻𝑘satisfies Assumption 2. Assume that we use the ACG algorithm (Algorithm 4) with exact line search to minimize ˆ𝑓𝑘(x)over X, and denote the iterate generated by this algorithm at iteration  𝑡as  ˜x𝑡𝑘+1, then there is a  𝑟 > 0such that if  {x𝑘, ˜x𝑡𝑘+1, ˜x𝑡+1𝑘+1} ⊂ B(x∗, 𝑟) ∩ Xand ˜x𝑡𝑘+1 ∉ F (x∗) then | ˜S𝑡+1𝑘+1| < | ˜S𝑡𝑘+1| and ˜S𝑡𝑘+1 \ ˜S𝑡+1𝑘+1 ∉ F (x∗). That is, at iteration  𝑡the ACG algorithm drops a vertex from the active set ˜S𝑡𝑘+1 that is not a vertex of the optimal face  F (x∗).

Proof. This proof follows relies on the same concepts as the proof in Theorem D.4 from Guélat & Marcotte (1986). Let  𝑟∗𝑖 > 0 and 𝑐∗ > 0be such that:

�v𝑖 − x, ∇ 𝑓 (x∗) + ∇2 𝑓 (x∗)(x − x∗)�≥ −𝑐2 if ∥x − x∗∥ ≤ 𝑟∗𝑖 and v𝑖 ∈ vert(F (x∗)) (D.36)�v𝑖 − x, ∇ 𝑓 (x∗) + ∇2 𝑓 (x∗)(x − x∗)�≥ 𝑐 if ∥x − x∗∥ ≤ 𝑟∗𝑖 and v𝑖 ∈ vert(X) \ vert(F (x∗)). (D.37)

Where  ∇ 𝑓 (x∗) + ∇2 𝑓 (x∗)(x − x∗)is the gradient of the quadratic approximation at  x∗using  ∇2 𝑓 (x∗)(note that the minimizer of this quadratic approximation is  x∗ and that this approximation is strongly convex and smooth). We have that:

image

The term shown in Equation (D.39) can be bounded using the triangle inequality and the fact that the Hessian of  𝑓 (x) is 𝐿2-Lipschitz:

image

The term shown in Equation (D.40), can be bounded using the triangle inequality and Lemma D.1, leading to:

image

where  1 + 𝜔𝐷2 ≥ 𝜂𝑘for all  𝑘 ≥ 0from Assumption 2. Lastly, the term in Equation (D.41) can be bounded using the triangle inequality and the  𝐿2-Lipschitz continuity of the Hessian, which allows us to write:

image

Using these bounds we have:

image

Where we note that ˆ𝑓𝑘(x) = ⟨∇ 𝑓 (x𝑘), x − x𝑘⟩ + 1/2 ∥x − x𝑘∥2𝐻𝑘. Using the bound in Equation (D.50) along with Equations (D.36)-(D.37), and setting  𝐶 = �3𝐿2/2 + 𝐿𝜔𝐷(1 + 𝜔𝐷2)� 𝐷2 we have:

image

Let  𝑟∗ = minv𝑖 ∈vert(X) 𝑟∗𝑖and  𝑟 = min {𝑟∗, 𝑐/(4𝐶)}and assume that  x𝑘 ∈ B (x∗, 𝑟) ∩ X(we know by strong convexity that there is an index  𝑇 ≥ 0such that for  𝑘 ≥ 𝑇the iterates  x𝑘of the SOCGS algorithm (Algorithm 2) will be in the aforementioned ball). If  ˜x𝑡𝑘+1 ∈ B (x∗, 𝑟) ∩ Xthen the bounds in Equations (D.51)-(D.52) hold, as��˜x𝑡𝑘+1 − x∗�� ≤ 𝑟∗, this leads to:

image

Where the inequality in Equation (D.53) follows from Equation (D.52), the inequality in Equation (D.54) from the fact that  ∥x𝑘 − x∗∥ < 𝑟 ≤ 𝑐/(4𝐶)and the last inequality from Equation (D.51). Therefore if ˜x𝑡𝑘+1 ∉ F (x∗)the ACG algorithm will take an away-step with a vertex  v ∈ ˜S𝑡𝑘+1 ∩ vert(X) \ vert(F (x∗))and direction  d = ˜x𝑡𝑘+1 − v (where ˜S𝑡𝑘+1 ∩ vert(X) \ vert(F (x∗)) ≠ ∅ as ˜x𝑡𝑘+1 ∉ F (x∗)). Similarly as in the proof of Theorem D.4, we show that  𝛾𝑘 = 𝛾maxif  ˜x𝑡+1𝑘+1 ∈ B (x∗, 𝑟) ∩ X. We use proof by contradiction, and assume that  𝛾𝑘 < 𝛾max. Using the optimality of the line search:

image

The inequality in Equation (D.60) follows from Equation (D.52), as��˜x𝑡+1𝑘+1 − x∗�� < 𝑟 ≤ 𝑟∗, and the one in Equation (D.61) follows from  ∥x𝑘 − x∗∥ < 𝑟 ≤ 𝑐/(4𝐶). This is the desired contradiction, and we must therefore have that  𝛾𝑘 = 𝛾max. This means that  | ˜S𝑡𝑘+1| > | ˜S𝑡+1𝑘+1|and  ˜S𝑡𝑘+1 \ ˜S𝑡+1𝑘+1 ∉ vert(F (x∗)), or stated equivalently, the ACG algorithm has dropped one of the vertices in its active set ˜S𝑡𝑘+1 that is not present in  F (x∗). □

One of the key requirements in Theorem D.5 is that  {x𝑘, ˜x𝑡𝑘+1, ˜x𝑡+1𝑘+1} ⊂ B(x∗, 𝑟) ∩ X. As the SOCGS algorithm (Algorithm 2) decreases the primal gap of Problem (1.1) at least linearly (Theorem 3.4), we can guarantee by strong convexity that there is an index  𝐾 ≥ 0after which for  𝑘 ≥ 𝐾we have that  x𝑘 ∈ B(x∗, 𝑟)∩X.But in order for Theorem D.5 to apply for all ACG iterations in Line 10, when computing the Inexact PVM step, we also need to ensure that  ˜x𝑡𝑘+1 ∈ B(x∗, 𝑟) ∩ Xfor all  𝑡 ≥ 0. In the next Lemma we show that ��˜x𝑡𝑘+1 − x∗�� ≤ O(∥x𝑘 − x∗∥1/2), allowing us to claim that for any  𝑟 > 0we can ensure that��˜x𝑡𝑘+1 − x∗�� ≤ 𝑟 forsmall enough  ∥x𝑘 − x∗∥.

Lemma D.6. Given a  𝜇-strongly convex and  𝐿-smooth function  𝑓 (x), a polytope X, and a quadratic approximation ˆ𝑓𝑘(x)that satisfies Assumption 2, let  ˜x𝑡𝑘+1 denote the iterate obtained after applying  𝑡 steps ofthe ACG algorithm (Line 10 of Algorithm 2) to minimize ˆ𝑓𝑘(x) over X, starting from  ˜x0𝑘+1 = x𝑘, then for any 𝑡 ≥ 0:

image

Proof. By the triangle inequality we have:

image

The first term in Equation (D.63) can be bounded as follows:

image

Where Equation (D.65) follows from the fact that the ACG algorithm decreases the primal gap at each iteration  𝑡and Equation (D.68) is obtained by applying the Cauchy-Schwarz inequality to the first term in Equation (D.67) and using the fact that  −��˜x∗𝑘+1 − x𝑘��2𝐻𝑘 ≤ 0. Moreover, in Equation (D.69) we have set 𝐺 = maxx∈X ∥∇ 𝑓 (x)∥. Note that the��˜x∗𝑘+1 − x∗��term appearing in Equations (D.63) and (D.70) can be bounded using Corollary C.12, which results in��˜x∗𝑘+1 − x∗�� ≤ O(∥x𝑘 − x∗∥2). Combining the bound shown in Equation (D.70) with the bound in Lemma D.3 allows us to conclude that that:

image

With Lemma D.6 we can guarantee that for any radius  𝑟 > 0, there is a  𝐾 ≥ 0 such that ˜x𝑡𝑘+1 ∈ B(x∗, 𝑟) ∩Xfor all 𝑘 ≥ 𝐾 and all 𝑡 ≥ 0. With this, we can move on to prove that after a finite number of iterations  𝐾 ≥ 0we can guarantee that  x𝑘 ∈ F (x∗) for all 𝑘 ≥ 𝐾.

Corollary D.7. Given a strongly convex and smooth function  𝑓 (x)with Lipschitz continuous Hessian and a polytope X, if Assumptions 1 and 2 are satisfied, then there is a  𝑟PVM > 0such that if  x𝑘 ∈ B(x∗, 𝑟PVM) ∩ Xand for any  𝑡 ≥ 0we have that  ˜x𝑡𝑘+1 ∉ F (x∗) then | ˜S𝑡+1𝑘+1| < | ˜S𝑡𝑘+1| and ˜S𝑡𝑘+1 \ ˜S𝑡+1𝑘+1 ∉ F (x∗).

Proof. Let 𝑟 > 0be the radius in Theorem D.5 such that if  {x𝑘, ˜x𝑡𝑘+1, ˜x𝑡+1𝑘+1} ⊂ B(x∗, 𝑟) ∩ X then | ˜S𝑡+1𝑘+1| < | ˜S𝑡𝑘+1|and ˜S𝑡𝑘+1 \ ˜S𝑡+1𝑘+1 ∉ F (x∗). Since we want this to hold for all  𝑡 ≥ 0for a given  x𝑘, we need to ensure that ˜x𝑡𝑘+1 ∈ B(x∗, 𝑟) ∩ Xfor  𝑡 ≥ 0. This can be accomplished with Lemma D.6, which allows us to ensure that there is a  𝑟PVM > 0such that for any  x𝑘 ∈ B(x∗, 𝑟PVM) ∩ Xwe have that  {x𝑘, ˜x𝑡𝑘+1, ˜x𝑡+1𝑘+1} ⊂ B(x∗, 𝑟) ∩ Xfor all  𝑡 ≥ 0. □

Corollary D.8. Given a strongly convex and smooth function  𝑓 (x)with Lipschitz continuous Hessian and a polytope X, if Assumptions 1 and 2 are satisfied, then there is a  𝐾 > 0such that for all  𝑘 ≥ 𝐾the iterates of the SOCGS algorithm (Algorithm 2) satisfy that  x𝑘 ∈ F (x∗).

Proof. By Theorem D.4 we know that there is a  𝐾ACG ≥ 0such that for  𝑘 ≥ 𝐾ACG we have that  xACG𝑘 ∈ F (x∗).Moreover, from Corollary D.7 we know that there is a radius  𝑟PVM > 0such that if  x𝑘 ∈ B(x∗, 𝑟PVM) ∩ Xthen  {x𝑘, ˜x𝑡𝑘+1, ˜x𝑡+1𝑘+1} ⊂ B(x∗, 𝑟) ∩ Xfor all  𝑡 ≥ 0, where  𝑟 > 0is the radius in Theorem D.5. As the SOCGS algorithm contracts the primal gap at least linearly, there is a  𝐾PVM ≥ 0after which we can guarantee that x𝑘 ∈ B(x∗, 𝑟PVM) ∩ X for all 𝑘 ≥ 𝐾PVM.Assume that  𝐾′ = max{𝐾ACG, 𝐾PVM}and  x𝐾′ ∉ F (x∗). Then for all subsequent iterations  𝑘 ≥ 𝐾′we either choose the ACG step (Line 18 in Algorithm 2) and have that  x𝑘+1 = xACG𝑘+1 ∈ F (x∗)and the claim is true, or we choose the Inexact PVM step (Line 15 in Algorithm 2) and have that  |S𝑘| > |S𝑘+1|and |S𝑘| \ |S𝑘+1| ∈ (vert(X) \ vert(F (x∗)))by Theorem D.5. The latter case can only happen a finite number of times before  x𝐾 ∈ F (x∗)for some  𝐾 > 𝐾′, as  |S𝐾′|is finite. Thereafter we will have that  x𝑘 ∈ F (x∗)for all 𝑘 > 𝐾(as Theorem D.4 and Theorem D.5 will still hold). □

This allows us to conclude in the next theorem that the quadratic convergence in distance to the optimum of the Inexact PVM steps translates into quadratic convergence in the primal gap for the SOCGS algorithm.

Theorem D.9 (Quadratic convergence in primal gap of the SOCGS algorithm). Given a 𝜇-strongly convex and  𝐿-smooth function  𝑓 (x) with 𝐿2-Lipschitz Hessian and a polytope X, if Assumption 1 and Assumption 2 are satisfied, then there is a  𝐾 ≥ 0such that for  𝑘 ≥ 𝐾the iterates of the SOCGS algorithm (Algorithm 2) satisfy:

image

where the parameter  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (x𝑘)), 𝜆max([∇2 𝑓 (x𝑘)]−1𝐻𝑘)} ≥ 1measures how well  𝐻𝑘 approx-imates  ∇2 𝑓 (x𝑘) and 𝜔is defined in Assumption 2.

Proof. From Corollary D.8 we know that there is an index  𝐾 ≥ 0such that for  𝑘 ≥ 𝐾we know that the Inexact PVM iterates and the ACG iterates will be contained in  F (x∗). This allows us to convert the quadratic convergence in distance to the optimum in Lemma D.3 for the Inexact PVM steps to a quadratic convergence in primal gap. Using strong-convexity we can bound bound  ∥x𝑘 − x∗∥2 ≤ 2/𝜇( 𝑓 (x𝑘) − 𝑓 (x∗)). Using  𝐿-smoothness along with the strict-complementary assumption (Assumption 1) and the fact that ˜x𝑘+1 ∈ F (x∗)leads to  ∥˜x𝑘+1 − x∗∥2 ≥ 2/𝐿( 𝑓 (˜x𝑘+1) − 𝑓 (x∗))). Plugging these bounds into the convergence in distance to the optimum from Lemma D.3 results in:

image

As the SOCGS contracts the primal gap at least linearly (see Theorem 3.4), then for small enough  𝑓 (x𝑘)− 𝑓 (x∗)with  𝑘 ≥ 𝐾we know that the quadratic convergence shown in Equation (D.71) for the Inexact PVM steps in Line 9-12 will provide more primal progress than the ACG steps in Line 4. Therefore the Inexact PVM steps will be chosen in Line 14 and we will have that:

image

D.2 Complexity Analysis

Throughout this section we make the simplifying assumption that we have at our disposal the tightest possible lower bound  𝑙𝑏(x𝑘)on the primal gap, that is,  𝑙𝑏(x𝑘) = 𝑓 (x𝑘) − 𝑓 (x∗). Providing a looser lower bound on the primal gap does not affect the number of first-order or Hessian oracle calls, however it can significantly increase the number of linear optimization oracle calls used to compute the Inexact PVM steps. Let  𝑟 = min�𝑟ACG, 𝑟PVM�> 0, where  𝑟ACGis described in Theorem D.4 and  𝑟PVMin Corollary D.7. Note that  𝑟is independent of the target accuracy  𝜀. For ease of exposition we can divide the behaviour of the SOCGS algorithm (Algorithm 2) into three phases:

1.  Phase 1: x𝑘 ∉ B(x∗, 𝑟) ∩ X or xACG𝑘 ∉ B(x∗, 𝑟) ∩ X.In this phase the SOCGS algorithm will contract the primal gap at least linearly, as dictated by Theorem 3.4. Using strong-convexity we can upper bound the number of iterations needed until  {x𝑘, xACG𝑘 } ∈ B(x∗, 𝑟), which marks the end of this first phase.

2.  Phase 2: {x𝑘, xACG𝑘 } ∈ B(x∗, 𝑟) ∩ X and {x𝑘, xACG𝑘 } ∉ F (x∗).The primal gap convergence of the SOCGS algorithm in this phase is also at least linear, and the convergence bound of Theorem 3.4 still holds. However in this phase, the ACG steps in Line 4 and the ACG steps used to compute the Inexact PVM iterates in Lines 9-12 will drop any vertices in their respective active sets that are not in F (x∗). That is, if  xACG𝑘 ∈ B(x∗, 𝑟) ∩ X \ F (x∗) then |SACG𝑘 | > |SACG𝑘+1 | and SACG𝑘 \ SACG𝑘+1 ∉ vert(F (x∗)).Similarly, if  x𝑘 ∈ B(x∗, 𝑟) ∩ X \ F (x∗)then  ˜x𝑘+1in Line 13 in Algorithm 2 satisfies after exiting the while loop in Lines 9-12 that  |S𝑘| > | ˜S𝑘+1|and  S𝑘 \ ˜S𝑘+1 ⊄ vert(F (x∗)). As the cardinality of both active sets is finite, after a finite number of iterations we must have that  {x𝑘, xACG𝑘 } ∈ B(x∗, 𝑟) ∩ F (x∗),which marks the end of this phase.

3.  Phase 3: {x𝑘, xACG𝑘 } ∈ B(x∗, 𝑟) ∩ F (x∗).In this final phase the SOCGS algorithm has a quadratic convergence rate in primal gap, as shown in Theorem D.9. Once  {x𝑘, xACG𝑘 } ∈ B(x∗, 𝑟) ∩ F (x∗)the ACG steps in Line 4 and in Lines 9-12 will not pick up any vertices in  vert(X) \ vert(F (x∗)), and the iterates will remain in  B(x∗, 𝑟) ∩ F (x∗)for all subsequent steps.

As in the classical analysis of PVM and Newton algorithms, the SOCGS algorithm shows local quadratic convergence (in primal gap and distance to the optimum) after a number of iterations that is independent of  𝜀(but dependent on  𝑓 (x)and X). The SOCGS algorithm makes use of three different types of oracle calls, namely, Hessian, first-order and linear optimization oracle calls. The Hessian oracle is called once per iteration (in Line 5), while the first-order oracle is called at most twice (to compute the independent ACG step in Line 4 and to build the quadratic approximation in Line 6). The linear minimization oracle will be called once in Line 4 for the independent ACG step and potentially multiple times in Line 10 while computing the Inexact PVM step.

In order to study the number of linear optimization oracle calls needed to achieve a  𝜀-optimal solution to Problem (1.1) we first review the convergence of the Frank-Wolfe gap of the ACG algorithm, which is used as a stopping criterion in the SOCGS algorithm to compute the Inexact PVM steps (Line 9 in Algorithm 2).

Theorem D.10 (Convergence of the Frank-Wolfe gap of the ACG algorithm). (Lacoste-Julien & Jaggi,

2015, Theorem 2) Given a  𝜇-strongly convex and  𝐿-smooth function  𝑓 (x)and a polytope X, then for any 𝑘 ≥ 0the ACG algorithm satisfies:

image

where  𝐷denotes the diameter of the polytope X.

image

The number of outer iterations needed for  xACG𝑘and  x𝑘to reach  B(x∗, 𝑟) ∩ Xcan be upper bounded using strong convexity. As  𝑓 (x) − 𝑓 (x∗) ≥ 𝜇/2 ∥x − x∗∥2then if  𝑓 (x) − 𝑓 (x∗) ≤ 𝜇/2𝑟2we can conclude that  x ∈ B(x∗, 𝑟) ∩ X. As the iterates  x𝑘and  xACG𝑘have a primal gap convergence that is at least linear (see Theorem 3.4 and Theorem 2.1 respectively) then the number of iterations  𝑇1needed to ensure that {x𝑘, xACG𝑘 } ∈ B(x∗, 𝑟) ∩ X for all 𝑘 ≥ 𝑇1can be upper bounded by:

image

Where we have used the primal gap convergence of Theorem 3.4 and  𝜇-strong convexity. If we denote by  𝑁𝑘,1the number of inner ACG steps in Line 10 that we need to take to satisfy the exit criterion shown in Line 9

of Algorithm 2 at iteration  𝑘during this phase, and we use Theorem D.10 we have that:

image

The inequality follows from the fact that for  x𝑘 ∉ B(x∗, 𝑟) ∩Xwe can bound  𝜇𝑟2/2 ≤ 𝑓 (x𝑘) − 𝑓 (x∗), and thefact that ˆ𝑓𝑘(x𝑘) − ˆ𝑓𝑘(x∗𝑘+1) =�−∇ 𝑓 (x𝑘), x𝑘 − x∗𝑘+1�− 1/2��x𝑘 − x∗𝑘+1��2𝐻𝑘 ≤ ∥∇ 𝑓 (x𝑘)∥��x𝑘 − x∗𝑘+1�� ≤ ∥∇ 𝑓 (x𝑘)∥ 𝐷.If we denote:

image

then, using the fact that  𝜂𝑘 ≤ 1 + 𝜔𝐷2, we can bound the number of inner ACG steps in Line 10 needed for any iteration  𝑘 ≥ 0in the first phase such that  x𝑘 ∉ B(x∗, 𝑟) ∩ X as:

image

As the SOCGS algorithm calls the Hessian oracle once, and the first-order oracle at most twice per iteration we can upper bound the total number of first-order and Hessian oracle calls using the bound shown in Equation (D.72). Combining the aforementioned bound with the bound on the total number of linear minimization oracle calls per iteration in Equation (D.76) we can bound the total number of linear minimization oracle calls. Therefore in this phase we will need:

image

In this phase we can guarantee that if  xACG𝑘 ∈ B(x∗, 𝑟) ∩ X \ F (x∗)then the ACG step in Line 4 will be an away-step that reduces the cardinality of the active set  SACG𝑘, satisfying that  |SACG𝑘 | > |SACG𝑘+1 |and SACG𝑘 \ SACG𝑘+1 ∉ vert(F (x∗)). Similarly, if  x𝑘 ∈ B(x∗, 𝑟) ∩ X \ F (x∗)then the ACG steps in Line 10 will also be away-steps that reduce the cardinality of the active set  S𝑘, that is, after exiting the while loop in Line 12 of Algorithm 2 we have that  |S𝑘| > | ˜S𝑘+1|and  S𝑘 \ ˜S𝑘+1 ⊄ vert(F (x∗)). This behaviour will continue until xACG𝑘 ∈ F (x∗) and ˜x𝑡+1𝑘+1 ∈ F (x∗).

Therefore we need to bound the number of vertices that have to be dropped from both  SACG𝑘and  S𝑘in order for  SACG𝑘 ⊆ vert(F (x∗))and  S𝑘 ⊆ vert(F (x∗)). The ACG algorithm in Line 4 will have picked up at most  𝑇1vertices in the first phase (as each iteration can only add one vertex to  SACGin Line 4), on the other hand, the PVM steps in Lines 9-12 will have picked up at most �𝑇1𝑘=1 𝑁𝑘,1vertices. As once inside the ball all ACG steps (both in Line 4 and Lines 9-12) reduce the cardinality of the active set, and using the bounds in Equation (D.72) and (D.76), we will need:

image

We now need to bound the number of first-order oracle calls needed to drop the aforementioned vertices. The ACG algorithm in Line 4 will need to call the first-order oracle at most  𝑇1times. On the other hand, we need to bound the number of vertices that the PVM steps will drop per first-order oracle call in Lines 9-12, for which we will use the following Lemma:

Lemma D.11. If 𝑓 (x𝑘) − 𝑓 (x∗) ≤ 4𝜇2 then the Inexact PVM steps in Lines 9-12 of Algorithm 2 will perform at least one ACG step in Line 10.

Proof. We use proof by contradiction, and we assume that to compute the Inexact PVM step to the necessary accuracy we did not perform any ACG steps in Line 10, that is:

image

Where the last inequality follows from convexity. Using the previous chain of inequalities along with 𝑓 (x𝑘) − 𝑓 (x∗) ≤ ∥∇ 𝑓 (x𝑘)∥2 /2𝜇from  𝜇-strong convexity we have that  𝑓 (x𝑘) − 𝑓 (x∗) > 4𝜇2, which is the desired contradiction. □We assume that  𝑟 < √8𝜇, which allows us to claim that the primal gap for any point  x𝑘 ∈ B(x∗, 𝑟) satisfies𝑓 (x𝑘) − 𝑓 (x∗) ≤ 4𝜇2 (otherwise it simply takes a constant number of iterations to achieve this once in  B(x∗, 𝑟),as the primal gap contracts at least linearly). Therefore in this phase we will need:

image

Let  𝑇denote the first iteration of the final phase, where  {x𝑇, xACG𝑇 } ∈ B(x∗, 𝑟) ∩ F (x∗)and the quadratic rate dominates over the linear rate. Using the quadratic convergence in primal gap shown in Theorem D.9 we have that:

image

Where we have used the fact that by Assumption 2 we have that  𝜂𝑘 ≤ 1 + 𝜔 ∥x𝑘 − x∗∥2 ≤ 1 + 𝜔𝑟2. Therefore in order to reach a  𝜀-optimal solution starting from this phase we need:

image

Where we have only included the dependence on  𝜀for notational convenience. If we denote by  𝑁𝑘,3the number of inner ACG steps in Line 10 that we need to take to satisfy the exit criterion shown in Line 9 of Algorithm 2 at iteration  𝑘during this last phase and we use the fact that  𝑓 (x𝑘) − 𝑓 (x∗) ≥ 𝜀for all suboptimal iterates, resulting in:

image

Therefore combining the bound on the total number of iterations in this phase with the bound on the number of linear minimization oracle calls per iteration we need:

image

The results for all these phases can be seen in Table 2.

image

Table 2: Oracle complexity to reach an  𝜀-optimal solution to Problem 1.1 for the SOCGS algorithm (Algorithm 2).

Remark D.12. The constant  𝑟is an invariant of the function and feasible region under consideration and has been used in a similar fashion in (Wolfe, 1970; Guélat & Marcotte, 1986) and more recently in (Garber, 2020), and although unknown, still makes the convergence analysis and complexity estimate conceptually useful, as it adds at most a constant number of iterations independent of  𝜀.

Remark D.13. Note that for simplicity we are implicitly assuming in the complexity analysis that the last iterate of the SOCGS algorithm at the end of Phase 2 satisfies  𝑓 (x𝑘) − 𝑓 (x∗) ≤ [𝐿𝜂𝑘/(2𝜇4)(√8𝜇(1 +√𝐿𝜔) +√𝜂𝑘𝐿2)]−2, as otherwise the convergence guarantee in Theorem D.9 does not provide a contraction. If this is not the case at the end of Phase 2, then after an additional finite number of linearly convergent iterations in primal gap, the iterates will indeed satisfy  𝑓 (x𝑘) − 𝑓 (x∗) ≤ [𝐿𝜂𝑘/(2𝜇4)(√8𝜇(1 +√𝐿𝜔) + √𝜂𝑘𝐿2)]−2, after which the complexity analysis from Phase 3 will apply.

Appendix E. Computational Results

In this section we compare the performance of the SOCGS algorithm with that of other first-order projection-free algorithms for several problems of interest. In the first problem the Hessian oracle will be inexact, but will satisfy Assumption 2 with  𝜔 = 0.1, moreover we will also assume knowledge of the primal gap, by first computing a solution to high accuracy. In the remaining problems the Hessian oracle will be exact, and we will assume that we do not have knowledge of the primal gap, and will use the strategy outlined in Remark 3.8. In the second experiment, in addition to using the exact Hessian, we will also implement SOCGS with an LBFGS Hessian update (SOCGS LBFGS) (note that this does not satisfy Assumption 2). In the second and third experiment we will also cap the maximum number of inner iterations for the SOCGS and NCG algorithms, as is done in the computational experiments of NCG and SVRCG.

In all three experiments we compare the performance of the SOCGS algorithm with the vanilla Conditional Gradients algorithm (denoted by CG), the Away-Step and Pairwise-Step Conditional Gradients algorithms (ACG and PCG), the Lazy Away-Step Conditional Gradients algorithm (Braun et al., 2017) (ACG (L)). In the first problem the Hessian oracle will be inexact, but will satisfy Assumption 2. In the remaining problems the Hessian oracle will be exact.

In the first experiment we also compare the performance of the algorithm with the Decomposition Invariant Conditional Gradient (DICG) algorithm (Garber & Meshi, 2016), as the feasible region is a  0 − 1 polytope.

We also compare against the Conditional Gradient Sliding (CGS) algorithm (Lan & Zhou, 2016) in the first experiment. This algorithm was also used in the second and third experiment, however the results were not competitive with the ones obtained for the other algorithms, both in terms of iteration count and wall-clock time, and so the CGS results are not included in the images for the second and third experiment.

Additionally, in the first experiment we also compare against the Stochastic Variance-Reduced Conditional Gradients (SVRCG) algorithm (Hazan & Luo, 2016), as we can take stochastic first-order oracles of the objective function in question. The third experiment has an objective function that is also amenable to stochastic first-order oracle calls, however the results obtained were not competitive with the other algorithms, both in terms of iteration count and wall-clock time, and so the results for this algorithm were not included in the images for the third experiment.

In the second and third experiments, which use an exact second-order oracle, we also compare the performance against the Newton Conditional Gradients (NCG) algorithm in Liu et al. (2020) which is similar in spirit to the SOCGS algorithm. One of the key features of this algorithm is that it does not require an exact line search strategy, as it provides a specific step size strategy (however it requires selecting five hyperparameters), and it does not require estimating an upper bound on the primal gap.

Remark E.1 (Hyperparameter search for the NCG algorithm). We tested 27 hyperparameters for the NCG algorithm, and the one that provided the best performance was selected. The parameters used (see (Liu et al., 2020) for their meaning) were combinations of  𝐶1 ∈ {0.1, 0, 25, 0.4}, 𝛿 ∈ {0.01, 0, 5, 0.99} and 𝐶 = {1.1, 1.5, 2}.The two remaining hyperparemeters were chosen as  𝛽 = 12 (1 − 12−1/𝐶 )and  𝜎 = 1𝐶(1−𝛽) + 𝛽(1−2𝛽) (1−𝛽)2so as to satisfy the requirements in Theorem 4.2 in (Liu et al., 2020). The hyperparameters that gave the best performance were  𝜎 = 0.96, 𝛽 = 1/6.0, 𝐶 = 2.0, 𝐶1 = 0.25 and 𝛿 = 0.99.

One of the key challenges that we found when implementing the NCG algorithm is the management of the active set. Starting from a given point  x𝑘the algorithm builds a quadratic approximation and performs a series of CG variant steps until the algorithm reaches a certain Frank-Wolfe gap (like in the SOCGS algorithm), which we denote by  ˜xNCG𝑘. At that point the algorithm either takes a step with  𝛾𝑘 = 1 (what iscalled a full step), or it takes a step size  𝛾𝑘 ≠ 1(which is called a damped step). In the former case the active set and the barycentric coordinates used for  x𝑘+1are simply those of  ˜xNCG𝑘, which is the point returned by the CG variant steps. In the latter case, however, we set  x𝑘+1 = x𝑘 + 𝛾𝑘(˜xNCG𝑘 − x𝑘) with 𝛾𝑘 ≠ 1, and we need to combine the active sets and barycentric coordinates of the points  x𝑘and  ˜xNCG𝑘to form  x𝑘+1. This is a computationally expensive task in general, as the CG variant can drop and pick-up an arbitrary number of vertices going from  x𝑘to  ˜xNCG𝑘, and we need to reconcile the two active sets and barycentric coordinates. This process involves checking if each vertex in the active set of  ˜xNCG𝑘is in the active set of  x𝑘, and vice-versa. When the dimensionality of the problem and the cardinality of the active set is high this can become too costly. That is why in general this algorithm is easiest to implement with CG variants that do not maintain an active set, like the vanilla CG algorithm or the DICG algorithm. We have chosen to use the vanilla CG algorithm in out implementation, as it gave good performance. Note however that there are simple feasible regions where updating the active set and the barycentric coordinates is trivial, like in the probability simplex.

The experiments were run on a laptop with Windows 10, an Intel Core i7 2.4GHz CPU and 6GB RAM.

E.1 Sparse Coding over the Birkhoff Polytope

Given a set of  𝑚input data points  𝑌 = [y1, · · · , y𝑚] with y𝑖 ∈ ℝ𝑑, sparse dictionary learning attempts to find a dictionary  𝑋 ∈ ℝ𝑑×𝑛 and a sparse representation  𝑍 = [z1, · · · , z𝑚] with z𝑖 ∈ ℝ𝑛 that minimizes:

image

Where  C = {𝑋 ∈ ℝ𝑑×𝑛 | �𝑛𝑗=1 𝑋2𝑗,𝑖 ≤ 1, ∀𝑖 ∈ [1, 𝑑]}is the set of matrices with columns with  ℓ2norm less than one. This problem is of interest as many signal processing tasks see performance boosts when given a learned dictionary  𝑋that is able to give a sparse representation (Mairal et al., 2010), as opposed to a predefined dictionary obtained from Fourier or wavelet transforms. The elements in this learned dictionary are not required to be orthogonal, and they can form an undercomplete or an overcomplete dictionary.

image

The gradient of  𝑓 (𝑋)amounts to computing  ∇ 𝑓 (𝑋) = �𝑚𝑖=1 −2(y𝑖 −𝑋z𝑖)z𝑇𝑖 and the Hessian is given by the block diagonal matrix  ∇2 𝑓 (𝑋) ∈ ℝ𝑛2×𝑛2 with ∇2 𝑓 (𝑋) = diag [𝐵, · · · , 𝐵] where 𝐵 ∈ ℝ𝑛×𝑛 has the form  𝐵 = �𝑚𝑖=1 z𝑖z𝑇𝑖 .Therefore  𝐵will be positive definite as long as we can form a basis for  ℝ𝑛 with the vectors  z𝑖, with 𝑚 ∈ [1, 𝑚].This is verified numerically. As the eigenvalues of a block-diagonal matrix are the eigenvalues of the blocks that form the diagonal, and as we verify that  𝐵is positive definite, the function  𝑓 (𝑋)is  𝜇-strongly convex and  𝐿-smooth. The complexity of the gradient computation scales as  O(𝑚𝑛2).

Remark E.2 (On the complexity of linear oracles for the Birkhoff polytope). Solving an LP exactly over the Birkhoff polytope using the Hungarian algorithm (from combinatorial optimization) has complexity  O(𝑛3). Thus it is more expensive to compute the gradient  ∇ 𝑓 (𝑋)than it is to solve an LP over the Birkhoff polytope if  𝑚 is large.

Remark E.3 (On the complexity of projection oracles for the Birkhoff polytope). There are no known algorithms to compute exact projections onto the Birkhoff polytope, and as such projections onto this feasible region have to be computed approximately. For example, if we use an interior-point method to compute a projection onto the Birkhoff polytope, the projection is computed to a certain accuracy (say  ˆ𝜀), and as such the complexity will depends on a  log 1/ˆ𝜀term. Moreover, to represent the constraints of the Birkhoff polytope we need  𝑛2 linear inequality constraints and  2𝑛 − 1linear equality constraints. We can get rid of the equality constraints by adding  2(2𝑛 − 1)inequality constraints. We can transform the projection problem with a quadratic objective function and linear inequality constraints into a problem with a linear objective function and quadratic/linear inequality constraints using standard optimization techniques. This means that we have  O(𝑛2)inequality constraints, and the dimensionality of our problem is  𝑛2. If we use a path following interior-point method, and we use the complexity guarantee from Equation 10.12 in (Nemirovski, 2004) the resulting complexity to reach an  ˆ𝜀-optimal solution is  O(𝑛7 log 1/ˆ𝜀). Note that in the complexity guarantee in the reference, the ambient dimension is  𝑛, whereas in our case it is  𝑛2, and the number of constraints is  𝑛2as opposed to  𝑚. The cost of these projection oracles justifies the use of conditional gradient algorithms to minimize convex functions over the Birkhoff polytope.

We generate synthetic data by creating a matrix  𝐵 ∈ ℝ𝑛×𝑛 with 𝑛 = 80and entries sampled from a standard normal distribution, and  𝑚vectors  x ∈ ℝ𝑛, with entries sampled from a standard normal distribution, in order to form  𝑍 = {z1, · · · , z𝑚}. The set of vectors  𝑌 = {y1, · · · , y𝑚}is generated by computing  y𝑖 = 𝐵z𝑖for all  𝑖 ∈ ⟦1, 𝑚⟧.

Let us denote the Frobenius norm by  ∥·∥2𝐹, and the uniform distribution between  𝑎and  𝑏as  U(𝑎, 𝑏). In this problem the Hessian oracle will return a matrix  𝐻𝑘 = ∇2 𝑓 (𝑋𝑘) + 𝛽𝑘𝜔 ∥𝑋𝑘 − 𝑋∗∥2𝐹 𝐼𝑛, where  𝛽𝑘 ∈U(−𝜆max(∇2 𝑓 (𝑋𝑘))/(𝜔 ∥𝑋𝑘 − 𝑋∗∥2𝐹 + 1), 𝜆min(∇2 𝑓 (𝑋𝑘))).

Remark E.4. The approximate matrix  𝐻𝑘 = ∇2 𝑓 (𝑋𝑘) + 𝛽𝑘𝜔 ∥𝑋𝑘 − 𝑋∗∥2𝐹 𝐼𝑛 with:

image

satisfies Assumption 2.

Proof. To see this note that  𝜂𝑘 = max{𝜆max(𝐻−1𝑘 ∇2 𝑓 (𝑋𝑘)), 𝜆max([∇2 𝑓 (𝑋𝑘)]−1𝐻𝑘)}and if we plug in the approximation for the Hessian we have that:

𝜆max([∇2 𝑓 (𝑋𝑘)]−1𝐻𝑘) = 𝜆max([∇2 𝑓 (𝑋𝑘)]−1(∇2 𝑓 (𝑋𝑘) + 𝛽𝑘𝜔 ∥𝑋𝑘 − 𝑋∗∥2𝐹 𝐼𝑛)) (E.3)= 1 + 𝛽𝑘𝜔 ∥𝑋𝑘 − 𝑋∗∥2𝐹 𝜆max([∇2 𝑓 (𝑋𝑘)]−1) (E.4)= 1 + 𝛽𝑘𝜔 ∥𝑋𝑘 − 𝑋∗∥2𝐹 /𝜆min(∇2 𝑓 (𝑋𝑘)). (E.5)

On the other hand:

image

The results for  𝑚 = 10000and  𝑚 = 100000can be seen in Figure 7 and Figure 8 respectively. In both cases, the initial point used for all the algorithms is the identity matrix  𝐼𝑛×𝑛. We can see that the SOCGS algorithm (with the DICG algorithm as a subproblem solver for the PVM steps) outperforms all the other algorithms being considered for both moderate to high values of  𝑚. The performance of the SVRCG algorithm improves relative to the other algorithms as we increase the value of  𝑚, as expected. We use the original implementation of the CGS algorithm for strongly-convex and smooth functions shown in Lan & Zhou (2016), which uses CG to solve the Euclidean projection subproblems that arise in Nesterov’s Accelerated Gradient Descent. The poor performance of the CGS algorithm can be explained with the fact that the CG algorithm does not contract the Frank-Wolfe gap linearly in general, and the accuracy to which the subproblems are solved increases with each iteration, and so at some point the subproblems become very computationally expensive to solve.

image

Given a binary classification task with  𝑚labels  𝑌 = {y1, · · · , y𝑚}and  𝑚samples  𝑍 = {z1, · · · , z𝑚}with 𝑦𝑖 ∈ {−1, 1} and z𝑖 ∈ ℝ𝑛 for all 𝑖 ∈ [1, 𝑚], we wish to solve:

image

where X is the  ℓ1unit ball centered at the origin and  𝜆 = 1/𝑚. Although projecting into the  ℓ1ball has complexity  O(𝑛) (Condat, 2016), and so projections are cheap, this feasible region is often used to compare the performance of projection-free algorithms between each other (see Lacoste-Julien & Jaggi (2015); Rao et al. (2015); Braun et al. (2019)). Solving a linear program over the  ℓ1ball also has complexity  O(𝑛). Thisexperiment was also considered in Ghanbari & Scheinberg (2018) and Scheinberg & Tang (2016) to compare the performance of several Proximal Quasi-Newton methods in the context of minimization with a projection oracle. The gradient of the objective function has the form given by:

image

The Hessian of the objective function can be written as:

image

Note that the  ∇2 𝑓 (x) ∈ ℝ𝑛×𝑛 in Equation (E.10), and so for large  𝑛even storing the Hessian might become problematic. However, the quadratic approximation does not need to store the matrix, as the function ˆ𝑓𝑘(x)

can be written as:

image

Which means that the gradient of ˆ𝑓𝑘(x)is given by:

image

When computing the Inexact PVM steps we compute  ∇ 𝑓 (x𝑘)and  1/((1 + 𝑒−𝑦𝑖⟨x𝑘,z𝑖⟩)(1 + 𝑒𝑦𝑖⟨x𝑘,z𝑖⟩))for each 𝑖 ∈ [1, 𝑚]at the beginning of the iteration, as these quantities do not change for a fixed  𝑘. This significantly decreases the time it takes to compute an ACG step with  ∇ ˆ𝑓𝑘(x) in Line 4of Algorithm 7, as we only perform operations with transcendental operations once at the beginning of the PVM step. Moreover, as in the previous numerical experiments, we can find a closed-form expression for the line search, that is:

image

Where we only need to compute a series of inner products with quantities that in many cases we have already pre-computed in previous operations and stored. This makes line searches with ˆ𝑓𝑘(x)significantly cheaper than line searches with  𝑓 (x).

The labels and samples used are taken from the training set of the gissette (Guyon et al., 2007) (Figure 9) and the real-sim (Chang & Lin, 2011) (Figure 10) dataset, where  𝑛 = 5000 and 𝑚 = 6000 and 𝑛 = 72309 and𝑚 = 20958, respectively. Figure 2 shows the performance of Algorithm 2 with the Lazy Away-Step Conditional Gradient algorithm (Braun et al., 2019). We also limit the maximum number of inner iterations that the SOCGS algorithm and the NCG algorithm perform at each outer iteration to 1000. In this last example we substituted the step size strategy of the NCG algorithm with a line search, as otherwise we were not getting comparable performance to the other algorithms using the step size strategy defined in Liu et al. (2020). We use a golden-section bounded line search for all the line searches for which we cannot find a closed-form solution.

The results for this experiment can be seen in Figure 9 and 10. The initial point used for all the algorithms is the vector  x0 = (1, 0, · · · , 0). We can see that the SOCGS algorithm (with the AFW algorithm as a subproblem solver for the PVM steps) and the NCG algorithm outperform all the other algorithms, with the SOCGS performing better than the NCG algorithm. The quadratic approximation in this example is easier to evaluate than the original function, as we only need to perform operations with transcendental functions once when we build the approximation, reusing these quantities for all remaining inner iterations. Like in the previous two examples, the SOCGS algorithm and the NCG algorithm benefit from the fact that there is a closed-form solution to the step size at each inner iteration when computing the PVM steps, and so avoid a potentially expensive golden section line search.

E.3 Inverse covariance estimation over spectrahedron

In many applications the relationships between variables can be modeled with the use of undirected graphical models, such is the case for example in gene expression problems, where the goal is to find out which groups of genes are responsible for producing a certain outcome, given a gene dataset. When the underlying distribution of these variables is Gaussian, the problem of determining the relationship between variables boils down to finding patterns of zeros in the inverse covariance matrix  Σ−1of the distribution. A common approach to solving this problem relies on finding a  ℓ1-regularized maximum likelihood estimator of  Σ−1, so as to encourage sparsity, over the positive definite cone (Banerjee et al., 2008; Friedman et al., 2008), this is often called the Graphical Lasso.

Several optimization algorithms have been used to tackle this problem, such as interior point methods (Yuan & Lin, 2007), block coordinate descent or accelerated first-order algorithms (Banerjee et al., 2008), coordinate descent algorithms (Friedman et al., 2008) and even projected limited-memory quasi-Newton algorithms (Schmidt et al., 2009). We solve a variation of the Graphical Lasso problem over the space of positive semidefinite matrices of unit trace, that is:

image

Where  𝛿 > 0is a small constant that we add to make to problem smooth,  𝑆 = �𝑁𝑖=1(z𝑖 − 𝜇)(z𝑖 − 𝜇)𝑇is the empirical covariance matrix of a set of datapoints  𝑍 = {z1, · · · , z𝑁 }drawn from a Gaussian distribution with z𝑖 ∈ ℝ𝑚and  𝜆 > 0is a regularization parameter. This feasible region (known as the spectrahedron) is not a polytope, and so the guarantees shown in the paper do not apply as they crucially rely on Theorem 2.1. However, we include the results to show the promising numerical performance of the method. Evaluating 𝑓 (𝑋)has complexity  O(𝑛3)if we compute the determinant with a LU decomposition, and evaluating the gradient  ∇ 𝑓 (𝑋) = −(𝑋 + 𝛿𝐼𝑛)−1 + 𝑆 + 𝜆𝑋has complexity  O(𝑛3), dominated by the matrix inversion. Solving the linear program  min𝑌 ∈X�𝑛𝑖, 𝑗=1(∇ 𝑓 (𝑋) ⊗ 𝑌)𝑖, 𝑗, where ⊗denotes the Hadamard product, amounts to finding the largest eigenvector of  −∇ 𝑓 (𝑋). We do this approximately by using the Implicitly Restarted Lanczos algorithm (Lehoucq et al., 1998) (implemented in eigsh in the scipy.sparse.linalg library).

The quadratic approximation ˆ𝑓𝑘(𝑋)of  𝑓 (𝑋)that the PVM steps in Line 10 of Algorithm 7 uses can be written as:

ˆ𝑓𝑘(𝑋) = trace ��−(𝑋𝑘 + 𝛿𝐼𝑛)−1 + 𝑆 + 𝜆𝑋𝑘� (𝑋 − 𝑋𝑘)� (E.12)+ 12

image

This allows us to write the gradient  ∇ ˆ𝑓𝑘(𝑋)of the quadratic approximation as:

image

The complexity of evaluating the gradient of ˆ𝑓𝑘(𝑋) is also O(𝑛3), dominated by the matrix inversion and the matrix multiplication operations. In practice, we only invert the matrix  (𝑋𝑘 + 𝛿𝐼𝑛)−1 once per iteration when we form the quadratic approximation in Line 6 of Algorithm 7. Nevertheless, this means that the complexity of computing  ∇ 𝑓 (𝑋) and ∇ ˆ𝑓𝑘(𝑋)is the same, so in this respect there is no advantage to using the quadratic approximation. However, for the quadratic approximation ˆ𝑓𝑘(𝑋)we can find a closed-form expression for the optimal step size when moving along a direction  𝐷. It suffices to take the derivative of ˆ𝑓𝑘(𝑋 + 𝛾𝐷) withrespect to  𝛾using the expression shown in Equation (E.15) and set the derivative to zero. This leads to:

image

If we use a golden section search to perform a line search over the original function  𝑓 (𝑋)to compute the optimal step size we will potentially need to evaluate  𝑓 (𝑋)multiple times, and each evaluation has complexity O(𝑛3). On the other hand, to compute the exact line search for ˆ𝑓𝑘(𝑋)we only need to evaluate the expression in Equation (E.16) once, with complexity  O(𝑛3). This makes the line search operation with ˆ𝑓𝑘(𝑋)significantly cheaper than the line search with  𝑓 (𝑋), and makes the ACG iterations in Line 10 of Algorithm 7 significantly cheaper than the iterations in Line 18 of Algorithm 7.

The matrix  𝑆is generated by computing a random orthonormal basis  B = {v1, · · · , v𝑚}in  ℝ𝑚and computing  𝑆 = �𝑖=1 𝜎𝑖v1v𝑇1, where  𝜎𝑖is uniformly distributed between 0.5 and 1 for  𝑖 ∈ [1, 𝑚]. We use 𝜆 = 0.05and  𝛿 = 10−5in the experiments. We also limit the maximum number of inner iterations that the SOCGS algorithm and the NCG algorithm perform at each outer iteration to 1000. We use a golden-section bounded line search for all the line searches for which we cannot find a closed-form solution.

We also implemented an LBFGS algorithm to build an approximate Hessian from first order information from previous iterations. This is specially useful if we cannot find an analytical expression to the exact Hessian, or its matrix-vector products. Note however that the matrix outputted by the LBFGS algorithm does not satisfy Assumption 2, and so the best we can hope for is for the linear-quadratic convergence in primal gap of the SOCGS algorithm. The implementation used stores the Hessian approximation in outer-product form, and so does not explicitly store the full Hessian matrix, as that could be computationally prohibitive (see Section 7.2 in Nocedal & Wright (2006)).

The results for this experiment can be seen in Figures 11 and 12. The initial point for all the algorithms is the matrix  1/𝑛𝐼𝑛. We can see that the SOCGS (with the PCG algorithm as a subproblem solver for the PVM steps) and the NCG algorithm outperform all the other algorithms, with the SOCGS performing better than the NCG algorithm. Note that the in this case the main advantage that the SOCGS and the NCG algorithms have over all the other algorithms is the fact that there is a closed-form solution to the step size at each inner iteration when computing the PVM steps. As discussed earlier, the complexity of evaluating the original function  𝑓 (𝑋)is the same as that of evaluating ˆ𝑓𝑘(𝑋). The SOCGS algorithm that uses the LBFGS algorithm to build up an approximate Hessian also performs well in terms of iterations and in terms of time, despite Assumption 2 not holding in this case.

image

Figure 7: Sparse Coding over the Birkhoff polytope: Algorithm comparison for  𝑚 = 10, 000 (mediumsize) samples in terms of primal gap (a),(b), Frank-Wolfe gap (c),(d) and distance to the optimum (e),(f).

image

Figure 8: Sparse Coding over the Birkhoff polytope: Algorithm comparison for  𝑚 = 100, 000(large size) samples in terms of primal gap (a),(b), Frank-Wolfe gap (c),(d) and distance to the optimum (e),(f).

image

Figure 9:  Structured Logistic Regression over ℓ1 unit ball:Algorithm comparison in terms of primal gap (a),(b), Frank-Wolfe gap (c),(d) and distance to the optimum (e),(f) for the gissette (Guyon et al., 2007) dataset, where  𝑛 = 5000 and 𝑚 = 6000.

image

Figure 10:  Structured Logistic Regression over ℓ1 unit ball:Algorithm comparison in terms of primal gap (a),(b), Frank-Wolfe gap (c),(d) and distance to the optimum (e),(f) for the real-sim (Chang & Lin, 2011) dataset, where  𝑛 = 72309 and 𝑚 = 20958.

image

Figure 11: Inverse covariance estimation over spectrahedron: Algorithm comparison for  𝑛 = 100 interms of primal gap (a),(b), Frank-Wolfe gap (c),(d) and distance to the optimum (e),(f).

image

Figure 12: Inverse covariance estimation over spectrahedron: Algorithm comparison for  𝑛 = 50in terms of primal gap (a),(b), Frank-Wolfe gap (c),(d) and distance to the optimum (e),(f).


Designed for Accessibility and to further Open Science