b

DiscoverSearch
About
My stuff
Stochastic Recursive Variance-Reduced Cubic Regularization Methods
2019·arXiv
Abstract
Abstract

Stochastic Variance-Reduced Cubic regularization (SVRC) algorithms have received increasing attention due to its improved gradient/Hessian complexities (i.e., number of queries to stochastic gradient/Hessian oracles) to find local minima for nonconvex finite-sum optimization. However, it is unclear whether existing SVRC algorithms can be further improved. Moreover, the semi-stochastic Hessian estimator adopted in existing SVRC algorithms prevents the use of Hessian-vector product-based fast cubic subproblem solvers, which makes SVRC algorithms computationally intractable for high-dimensional problems. In this paper, we first present a Stochastic Recursive Variance-Reduced Cubic regularization method (SRVRC) using a recursively updated semi-stochastic gradient and Hessian estimators. It enjoys improved gradient and Hessian complexities to find an (ϵ, √ϵ)-approximate local minimum, and outperforms the state-of-the-art SVRC algorithms. Built upon SRVRC, we further propose a Hessian-free SRVRC algorithm, namely SRVRCfree, which only needs �O(nϵ−2 ∧ ϵ−3) stochastic gradient and Hessian-vector product computations, where n is the number of component functions in the finite-sum objective and  ϵis the optimization precision. This outperforms the best-known result �O(ϵ−3.5) achievedby stochastic cubic regularization algorithm proposed in Tripuraneni et al. (2018).

Many machine learning problems can be formulated as empirical risk minimization, which is in the form of finite-sum optimization as follows:

image

where each  fi : Rd → Rcan be a convex or nonconvex function. In this paper, we are particularly interested in nonconvex finite-sum optimization, where each  fiis nonconvex. This is often the case for deep learning (LeCun et al., 2015). In principle, it is hard to find the global minimum of (1.1) because of the NP-hardness of the problem (Hillar and Lim, 2013), thus it is reasonable to resort to finding local minima (a.k.a., second-order stationary points). It has been shown that local minima can be the global minima in certain machine learning problems, such as low-rank matrix factorization (Ge et al., 2016; Bhojanapalli et al., 2016; Zhang et al., 2018b) and training deep linear neural networks (Kawaguchi, 2016; Hardt and Ma, 2016). Therefore, developing algorithms to find local minima is important both in theory and in practice. More specifically, we define an (ϵg, ϵH)-approximate local minimum x of F(x) as follows

image

where  ϵg, ϵH >0 are predefined precision parameters. The most classic algorithm to find the approximate local minimum is cubic-regularized (CR) Newton method, which was originally proposed in the seminal paper by Nesterov and Polyak (2006). Generally speaking, in the k-th iteration, cubic regularization method solves a subproblem, which minimizes a cubic-regularized second-order Taylor expansion at the current iterate  xk. The update rule can be written as follows:

image

where M > 0 is a penalty parameter. Nesterov and Polyak (2006) proved that to find an (ϵ, √ϵ)-approximate local minimum of a nonconvex function F, cubic regularization requires at most  O(ϵ−3/2)iterations. However, when applying cubic regularization to nonconvex finite-sum optimization in (1.1), a major bottleneck of cubic regularization is that it needs to compute n individual gradients ∇fi(xk) and Hessian matrices  ∇2fi(xk) at each iteration, which leads to a total  O(nϵ−3/2) gradientcomplexity (i.e., number of queries to the stochastic gradient oracle  ∇fi(x) for some i and x) andO(nϵ−3/2) Hessian complexity (i.e., number of queries to the stochastic Hessian oracle  ∇2fi(x) forsome i and x). Such computational overhead will be extremely expensive when n is large as is in many large-scale machine learning applications.

To overcome the aforementioned computational burden of cubic regularization, Kohler and Lucchi (2017); Xu et al. (2017) used subsampled gradient and subsampled Hessian, which achieve �O(nϵ−3/2 ∧ϵ−7/2) gradient complexity and  �O(nϵ−3/2 ∧ϵ−5/2) Hessian complexity. Zhou et al. (2018d) proposed a stochastic variance reduced cubic regularization method (SVRC), which uses novel semi-stochastic gradient and semi-stochastic Hessian estimators inspired by variance reduction for first-order finite-sum optimization (Johnson and Zhang, 2013; Reddi et al., 2016; Allen-Zhu and Hazan, 2016), which attains  O(n4/5ϵ−3/2) Second-order Oracle (SO) complexity1. Zhou et al. (2018b); Wang et al. (2018b); Zhang et al. (2018a) used a simpler semi-stochastic gradient compared with (Zhou et al., 2018d), and semi-stochastic Hessian, which achieves a better Hessian complexity, i.e.,  O(n2/3ϵ−3/2). However, it is unclear whether the gradient and Hessian complexities of the aforementioned SVRC algorithms can be further improved. Furthermore, all these algorithms need to use the semi-stochastic Hessian estimator, which is not compatible with Hessian-vector product-based cubic subproblem solvers (Agarwal et al., 2017; Carmon and Duchi, 2016, 2018). Therefore, the cubic subproblem (1.4) in each iteration of existing SVRC algorithms has to be solved by computing the inverse of the Hessian matrix, whose computational complexity is at least  O(dw)2.This makes existing SVRC algorithms not very practical for high-dimensional problems.

In this paper, we first show that the gradient and Hessian complexities of SVRC-type algorithms can be further improved. The core idea is to use novel recursively updated semi-stochastic gradient and Hessian estimators, which are inspired by the stochastic path-integrated differential estimator (SPIDER) (Fang et al., 2018) and the StochAstic Recursive grAdient algoritHm (SARAH) (Nguyen et al., 2017) for first-order optimization. We show that such kind of estimators can be extended to second-order optimization to reduce the Hessian complexity. Nevertheless, our analysis is very different from that in Fang et al. (2018); Nguyen et al. (2017), because we study a fundamentally different optimization problem (i.e., finding local minima against finding first-order stationary points) and a completely different optimization algorithm (i.e., cubic regularization versus gradient method). In addition, in order to reduce the runtime complexity of existing SVRC algorithms, we further propose a Hessian-free SVRC method that can not only use the novel semi-stochastic gradient estimator, but also leverage the Hessian-vector product-based fast cubic subproblem solvers. Experiments on benchmark nonconvex finite-sum optimization problems illustrate the superiority of our newly proposed SVRC algorithms over the state-of-the-art (Due to space limit, we include the experiments in Appendix 6).

In detail, our contributions are summarized as follows:

1. We propose a new SVRC algorithm, namely SRVRC, which can find an (ϵ, √ϵ)-approximate local minimum with �O(nϵ−3/2 ∧ ϵ−3) gradient complexity and �O(n ∧ ϵ−1+  n1/2ϵ−3/2 ∧ ϵ−2) Hessian complexity. Compared with previous work in cubic regularization, the gradient and Hessian complexity of SRVRC is strictly better than the algorithms in Zhou et al. (2018b); Wang et al. (2018b); Zhang et al. (2018a), and better than that in Zhou et al. (2018d); Shen et al. (2019) in a wide regime.

2. We further propose a new algorithm SRVRCfree, which requires �O(ϵ−3 ∧ nϵ−2) stochastic gra-dient and Hessian-vector product computations to find an (ϵ, √ϵ)-approximate local minimum. SRVRCfreeis strictly better than the algorithms in (Agarwal et al., 2017; Carmon and Duchi, 2016; Tripuraneni et al., 2018) when  n ≫1. The runtime complexity of SRVRCfreeis also better than that of SRVRC when the problem dimension d is large.

In an independent and concurrent work (Shen et al., 2019), two stochastic trust region methods namely STR1 and STR2 were proposed, which are based on the same idea of variance reduction using SPIDER, and are related to our first algorithm SRVRC. Our SRVRC is better than STR1 because it enjoys the same Hessian complexity but a better gradient complexity than STR1. Compared with STR2, our SRVRC has a consistently lower Hessian complexity and lower gradient complexity in a wide regime (i.e.,  ϵ ≫ n−1/2). Since Hessian complexity is the dominating term in cubic regularization method (Zhou et al., 2018b; Wang et al., 2018b), our SRVRC is arguably better than STR2, as verified by our experiments.

For the ease of comparison, we summarize the comparison of methods which need to compute the Hessian explicitly in Table 1, the Hessian-free or Hessian-vector product based methods in Table 2.

Table 1: Comparisons of different methods to find an (ϵ, √ρϵ)-local minimum on gradient and Hessian complexity.

image

In this section, we review additional related work that is not discussed in the introduction section. Cubic Regularization and Trust-Region Methods Since cubic regularization was first proposed by Nesterov and Polyak (2006), there has been a line of followup research. It was extended to adaptive regularized cubic methods (ARC) by Cartis et al. (2011a,b), which enjoy the same iteration complexity as standard cubic regularization while having better empirical performance. The first attempt to make cubic regularization a Hessian-free method was done by Carmon and Duchi (2016), which solves the cubic sub-problem by gradient descent, requiring in total �O(nϵ−2) stochastic gradient and Hessian-vector product computations. Agarwal et al. (2017) solved cubic sub-problem by fast matrix inversion based on accelerated gradient descent, which requires �O(nϵ−3/2+  n3/4ϵ−7/4) stochastic gradient and Hessian-vector product computations. In the pure stochastic optimization setting, Tripuraneni et al. (2018) proposed stochastic cubic regularization method, which uses subsampled gradient and Hessian-vector product-based cubic subproblem solver, and requires �O(ϵ−3.5) stochastic gradient and Hessian-vector product computations. A closely related second-order method to cubic regularization methods are trust-region methods (Conn et al., 2000; Cartis et al., 2009, 2012, 2013). Recent studies (Blanchet et al., 2016; Curtis et al., 2017; Mart´ınez and Raydan, 2017) proved that the trust-region method can achieve the same iteration complexity as the cubic regularization method. Xu et al. (2017) also extended trust-region method to subsampled trust-region method for nonconvex finite-sum optimization.

Local Minima Finding Besides cubic regularization and trust-region type methods, there is

Table 2: Comparisons of different methods to find an (ϵ, √ρϵ)-local minimum both on stochastic gradient and Hessian-vector product computations.

image

another line of research for finding approximate local minima, which is based on first-order optimization. Ge et al. (2015); Jin et al. (2017a) proved that (stochastic) gradient methods with additive noise are able to escape from nondegenerate saddle points and find approximate local minima. Carmon et al. (2018); Royer and Wright (2017); Allen-Zhu (2017); Xu et al. (2018); Allen-Zhu and Li (2018); Jin et al. (2017b); Yu et al. (2017, 2018); Zhou et al. (2018a); Fang et al. (2018) showed that by alternating first-order optimization and Hessian-vector product based negative curvature descent, one can find approximate local minima even more efficiently. Very recently, Fang et al. (2019); Jin et al. (2019) showed that stochastic gradient descent itself can escape from saddle points. Variance Reduction Variance reduction techniques play an important role in our proposed algorithms. Variance reduction techniques were first proposed for convex finite-sum optimization, which use semi-stochastic gradient to reduce the variance of the stochastic gradient and improve the gradient complexity. Representative algorithms include Stochastic Average Gradient (SAG) (Roux et al., 2012), Stochastic Variance Reduced Gradient (SVRG) (Johnson and Zhang, 2013; Xiao and Zhang, 2014), SAGA (Defazio et al., 2014) and SARAH (Nguyen et al., 2017), to mention a few. For nonconvex finite-sum optimization problems, Garber and Hazan (2015); Shalev-Shwartz (2016) studied the case where each individual function is nonconvex, but their sum is still (strongly) convex. Reddi et al. (2016); Allen-Zhu and Hazan (2016) extended SVRG to noncovnex finite-sum optimization, which is able to converge to first-order stationary point with better gradient complexity than vanilla gradient descent. Fang et al. (2018); Zhou et al. (2018c); Wang et al. (2018a); Nguyen et al. (2019) further improved the gradient complexity for nonconvex finite-sum optimization to be (near) optimal.

image

and the minimal function value is bounded.

Assumption 3.1. For any function F(x) and an initial point  x0, there exists a constant 0  < ∆F <∞ such that F(x0) − infx∈Rd F(x) ≤ ∆F .

We also need the following L-gradient Lipschitz and  ρ-Hessian Lipschitz assumption.

Assumption 3.2. For each i, we assume that  fi is L-gradient Lipschitz continuous and  ρ-HessianLipschitz continuous, where we have  ∥∇fi(x) − ∇fi(y)∥2 ≤ L∥x − y∥2 and ∥∇2fi(x) − ∇2fi(y)∥2 ≤ρ∥x − y∥2 for all x, y ∈ Rd.

Note that L-gradient Lipschitz is not required in the original cubic regularization algorithm (Nesterov and Polyak, 2006) and the SVRC algorithm (Zhou et al., 2018d). However, for most other SVRC algorithms (Zhou et al., 2018b; Wang et al., 2018b; Zhang et al., 2018a), they need the L-gradient Lipschitz assumption.

In addition, we need the difference between the stochastic gradient and the full gradient to be bounded.

Assumption 3.3. We assume that F has M-bounded stochastic gradient, where we have  ∥∇fi(x)−∇F(x)∥2 ≤ M, ∀x ∈ Rd, ∀i ∈ [n].

It is worth noting that Assumption 3.3 is weaker than the assumption that each  fiis Lipschitz continuous, which has been made in Kohler and Lucchi (2017); Zhou et al. (2018b); Wang et al. (2018b); Zhang et al. (2018a). We would also like to point out that we can make additional assumptions on the variances of the stochastic gradient and Hessian, such as the ones made in Tripuraneni et al. (2018). Nevertheless, making these additional assumptions does not improve the dependency of the gradient and Hessian complexities or the stochastic gradient and Hessian-vector product computations on  ϵ and n. Therefore we chose not making these additional assumptions on the variances.

In this section, we present SRVRC, a novel algorithm which utilizes new semi-stochastic gradient and Hessian estimators compared with previous SVRC algorithms. We also provide a convergence analysis of the proposed algorithm.

image

4.1 Algorithm Description

In order to reduce the computational complexity for calculating full gradient and full Hessian in (1.3), several ideas such as subsampled/stochastic gradient and Hessian (Kohler and Lucchi, 2017; Xu et al., 2017; Tripuraneni et al., 2018) and variance-reduced semi-stochastic gradient and Hessian (Zhou et al., 2018d; Wang et al., 2018b; Zhang et al., 2018a) have been used in previous work. SRVRC follows this line of work. The key idea is to use a new construction of semi-stochastic gradient and Hessian estimators, which are recursively updated in each iteration, and reset periodically after certain number of iterations (i.e., an epoch). This is inspired by the first-order variance reduction algorithms SPIDER (Fang et al., 2018) and SARAH (Nguyen et al., 2017). SRVRC constructs semi-stochastic gradient and Hessian as in (3.1) and (3.2) respectively. To be more specific, in the t-th iteration when mod(t, S(g)) = 0 or mod(t, S(h)) = 0, where  S(g), S(h)are the epoch lengths of gradient and Hessian, SRVRC will set the semi-stochastic gradient  vtand Hessian  Utto be a subsampled gradient  ∇fJt(xt) and Hessian  ∇2fJt(xt) at point xt, respectively. In the t-th iteration when mod(t, S) ̸= 0 or mod(t, S(h)) ̸= 0, SRVRC constructs semi-stochastic gradient and Hessian vtand  Utbased on previous estimators  vt−1and  Ut−1recursively. With semi-stochastic gradient vt, semi-stochastic Hessian  Utand t-th Cubic penalty parameter  Mt, SRVRC constructs the t-th Cubic subproblem  mtand solves for the solution to  mt as t-th update direction as (3.3). If  ∥ht∥2 isless than a given threshold which we set it as�ϵ/ρ, SRVRC returns  xt+1 = xt + htas its output. Otherwise, SRVRC updates  xt+1 = xt + htand continues the loop.

The main difference between SRVRC and previous stochastic cubic regularization algorithms (Kohler and Lucchi, 2017; Xu et al., 2017; Zhou et al., 2018d,b; Wang et al., 2018b; Zhang et al., 2018a) is that SRVRC adapts new semi-stochastic gradient and semi-stochastic Hessian estimators, which are defined recursively and have smaller asymptotic variance. The use of such semi-stochastic gradient has been proved to help reduce the gradient complexity in first-order nonconvex finite-sum optimization for finding stationary points (Fang et al., 2018; Wang et al., 2018a; Nguyen et al., 2019). Our work takes one step further to apply it to Hessian, and we will later show that it helps reduce the gradient and Hessian complexities in second-order nonconvex finite-sum optimization for finding local minima.

4.2 Convergence Analysis

In this subsection, we present our theoretical results about SRVRC. While the idea of using variance reduction technique for cubic regularization is hardly new, the new semi-stochastic gradient and Hessian estimators in (3.1) and (3.2) bring new technical challenges in the convergence analysis.

To describe whether a point x is a local minimum, we follow the original cubic regularization work (Nesterov and Polyak, 2006) to use the following criterion  µ(x):

Definition 4.1. For any x, define µ(x) as µ(x) = max{∥∇F(x)∥3/22 , −λ3min�∇2F(x)�/ρ3/2}.

It is easy to note that  µ(x) ≤ ϵ3/2if and only if x is an (ϵ, √ρϵ)-approximate local minimum. Thus, in order to find an (ϵ, √ρϵ)-approximate local minimum, it suffices to find a point x which satisfies  µ(x) ≤ ϵ3/2.

The following theorem provides the convergence guarantee of SRVRC for finding an (ϵ, √ρϵ)-approximate local minimum.

Theorem 4.2. Under Assumptions 3.1, 3.2 and 3.3, set the cubic penalty parameter  Mt= 4ρfor any t and the total iteration number  T ≥40∆F ρ1/2ϵ−3/2. For t such that mod(t, S(g)) ̸= 0 or mod(t, S(h)) ̸= 0, set the gradient sample size  B(g)tand Hessian sample size  B(h)t as

image

For  t such that mod(t, S(g)) = 0 or mod(t, S(h)) = 0, set the gradient sample size  B(g)tand Hessian sample size  B(h)t as

image

Then with probability at least 1  − ξ, SRVRC outputs  xoutsatisfying  µ(xout) ≤600ϵ3/2, i.e., an (ϵ, √ρϵ)-approximate local minimum.

Next corollary spells out the exact gradient complexity and Hessian complexity of SRVRC to find an (ϵ, √ρϵ)-approximate local minimum.

Corollary 4.3. Under the same conditions as Theorem 4.2, if we set  S(g), S(h)as  S(g)= √ρϵ/L ·�n ∧ M2/ϵ2 and S(h) =�n ∧ L/(ρϵ), and set T, {B(g)t }, {B(h)t }as their lower bounds in (4.1)- (4.4), then with probability at least 1  − ξ, SRVRC will output an (ϵ, √ρϵ)-approximate local minimum within

image

stochastic Hessian evaluations and

image

stochastic gradient evaluations.

Remark 4.4. For SRVRC, if we assume  M, L, ρ, ∆Fto be constants, then its gradient complexity is �O(n/ϵ3/2 ∧ √n/ϵ2 ∧ ϵ−3), and its Hessian complexity is �O(n ∧ ϵ−1 + n1/2ϵ−3/2 ∧ ϵ−2). Regarding Hessian complexity, suppose that  ϵ ≪1, then the Hessian complexity of SRVRC can be simplified as �O(n1/2ϵ−3/2 ∧ ϵ−2). Compared with existing SVRC algorithms (Zhou et al., 2018b; Zhang et al., 2018a; Wang et al., 2018b), SRVRC outperforms the best-known Hessian sample complexity by a factor of  n1/6 ∧ n2/3ϵ1/2. In terms of gradient complexity, SRVRC outperforms STR2 (Shen et al., 2019) by a factor of  n3/4ϵ3/2 when ϵ ≫ n−1/2.

Remark 4.5. Note that both Theorem 4.2 and Corollary 4.3 still hold when Assumption 3.3 does not hold. In that case,  M = ∞and SRVRC’s Hessian complexity remains the same, while its gradient complexity can be potentially worse, i.e., �O(n/ϵ3/2 ∧ √n/ϵ2), which degenerates to that of STR1 (Shen et al., 2019).

While SRVRC adapts novel semi-stochastic gradient and Hessian estimators to reduce both the gradient and Hessian complexities, it has three limitations for high-dimensional problems with  d ≫ 1:(1) it needs to compute and store the Hessian matrix, which needs  O(d2) computational time and storage space; (2) it needs to solve cubic subproblem  mtexactly, which requires  O(dw) computational time because it needs to compute the inverse of a Hessian matrix (Nesterov and Polyak, 2006); and (3) it cannot leverage the Hessian-vector product-based cubic subproblem solvers (Agarwal et al., 2017; Carmon and Duchi, 2016, 2018) because of the use of the semi-stochastic Hessian estimator. It is interesting to ask whether we can modify SRVRC to overcome these shortcomings.

5.1 Algorithm Description

We present a Hessian-free algorithm SRVRCfreeto address above limitations of SRVRC for high-dimensional problems, which only requires stochastic gradient and Hessian-vector product computations. SRVRCfreeuses the same semi-stochastic gradient  vtas SRVRC. As opposed to SRVRC which has to construct semi-stochastic Hessian explicitly, SRVRCfreeonly accesses to stochastic

image

Hessian-vector product. In detail, at each iteration t, SRVRCfreesubsamples an index set  Itand defines a stochastic Hessian-vector product function  Ut[·] : Rd → Rd as follows:

image

Note that although the subproblem depends on  Ut, SRVRCfreenever explicitly computes this matrix. Instead, it only provides the subproblem solver access to  Utthrough stochastic Hessian-vector product function  Ut[·]. The subproblem solver performs gradient-based optimization to solve the subproblem  mt(h) as  ∇mt(h) depends on  Utonly via  Ut[h]. In detail, following Tripuraneni et al. (2018), SRVRCfreeuses Cubic-Subsolver (See Algorithms 3 and 4 in Appendix G) and Cubic-Finalsolver from (Carmon and Duchi, 2016), to find an approximate solution  htto the cubic subproblem in (3.3). Both Cubic-Subsolver and Cubic-Finalsolver only need to access gradient  vtand Hessian-vector product function  Ut[·] along with other problem-dependent parameters. With the output  htfrom Cubic-Subsolver, SRVRCfreedecides either to update  xtas  xt+1 ← xt + htor to exit the loop. For the later case, SRVRCfreewill call Cubic-Finalsolver to output  ht, and takes xt+1 = xt + htas its final output.

The main differences between SRVRC and SRVRCfreeare two-fold. First, SRVRCfreeonly needs to compute stochastic gradient and Hessian-vector product. Since both of these two computations only take O(d) time in many applications in machine learning, SRVRCfreeis suitable for high-dimensional problems. In the sequel, following Agarwal et al. (2017); Carmon et al. (2018); Tripuraneni et al. (2018), we do not distinguish stochastic gradient and Hessian-vector product computations and consider them to have the same runtime complexity. Second, instead of solving cubic subproblem  mtexactly, SRVRCfreeadopts approximate subproblem solver Cubic-Subsolver and Cubic-Finalsolver, both of which only need to access gradient and Hessian-vector product function, and again only take O(d) time. Thus, SRVRCfreeis computational more efficient than SRVRC when  d ≫ 1.

5.2 Convergence Analysis

We now provide the convergence guarantee of SRVRCfree, which ensures that SRVRCfreewill output an (ϵ, √ρϵ)-approximate local minimum.

Theorem 5.1. Under Assumptions 3.1, 3.2, 3.3, suppose  ϵ < L/(4ρ). Set the cubic penalty parameter  Mt= 4ρfor any t and the total iteration number  T ≥25∆F ρ1/2ϵ−3/2. Set the Hessianvector product sample size  B(h)t as

image

For t such that mod(t, S(g)) ̸= 0, set the gradient sample size  B(g)t as

image

For t such that mod(t, S(g)) = 0, set the gradient sample size  B(g)t as

image

Then with probability at least 1  − ξ, SRVRCfree outputs xoutsatisfying  µ(xout) ≤ 1300ϵ3/2, i.e., an(ϵ, √ρϵ)-approximate local minimum.

The following corollary calculates the total amount of stochastic gradient and Hessian-vector product computations of SRVRCfreeto find an (ϵ, √ρϵ)-approximate local minimum.

Corollary 5.2. Under the same conditions as Theorem 5.1, if set  S(g)= √ρϵ/L ·�n ∧ M2/ϵ2and set  T, {B(g)t }, {B(h)t }as their lower bounds in (5.1)-(5.3), then with probability at least 1  − ξ, SRVRCfreewill output an (ϵ, √ρϵ)-approximate local minimum within

image

stochastic gradient and Hessian-vector product computations.

Remark 5.3. For SRVRCfree, if we assume  ρ, L, M, ∆Fare constants, then (5.4) is �O(nϵ−2 ∧ ϵ−3).For stochastic algorithms, the regime  n → ∞is of most interest. In this regime, (5.4) becomes �O(ϵ−3). Compared with other local minimum finding algorithms based on stochastic gradient and Hessian-vector product, SRVRCfreeoutperforms the results achieved by Tripuraneni et al. (2018) and Allen-Zhu (2018) by a factor of  ϵ−1/2. SRVRCfreealso matches the best-known result achieved by a recent first-order algorithm proposed in (Fang et al., 2018). Note that the algorithm proposed by Fang et al. (2018) needs to alternate the first-order finite-sum optimization algorithm SPIDER and negative curvature descent. In sharp contrast, SRVRCfreeis a pure cubic regularization type algorithm and does not need to calculate the negative curvature direction.

Remark 5.4. It is worth noting that both Theorem 5.1 and Corollary 5.2 still hold when Assumption 3.3 does not hold, and SRVRCfree’s runtime complexity remains the same. The only difference is: without Assumption 3.3, we need to use full gradient (i.e.,  B(g)t = n) instead of subsampled gradient at each iteration t.

5.3 Discussions on runtime complexity

We would like to further compare the runtime complexity between SRVRC and SRVRCfree. In specific, SRVRC needs O(d) time to construct semi-stochastic gradient and  O(d2) time to construct semi-stochastic Hessian. SRVRC also needs  O(dw) time to solve cubic subproblem  mtfor each iteration. Thus, with the fact that the total number of iterations is  T = O(ϵ−3/2) by Corollary 4.3, SRVRC needs

image

runtime to find an (ϵ, √ϵ)-approximate local minimum if we regard  M, L, ρ, ∆Fas constants. As we mentioned before, for many machine learning problems, both stochastic gradient and Hessian-vector product computations only need O(d) time, therefore the runtime of SRVRCfree is �O(dnϵ−2 ∧ dϵ−3).We conclude that SRVRCfreeoutperforms SRVRC when d is large, which is in accordance with the fact that Hessian-free methods are superior for high dimension machine learning tasks. On the other hand, a careful calculation can show that the runtime of SRVRC can be less than that of SRVRCfree when dis moderately small. This is also reflected in our experiments in Section 6.

In this section, we present numerical experiments on different nonconvex empirical risk minimization (ERM) problems and on different datasets to validate the advantage of our proposed SRVRC and SRVRCfreealgorithms for finding approximate local minima.

Baselines: We compare our algorithms with the following algorithms: SPIDER+ (Fang et al., 2018), which is the local minimum finding version of SPIDER, stochastic trust region (STR1, STR2) (Shen et al., 2019), subsampled cubic regularization (SCR) (Kohler and Lucchi, 2017), stochastic cubic regularization (STC) (Tripuraneni et al., 2018), stochastic variance-reduced cubic regularization (SVRC) (Zhou et al., 2018d), sample efficient SVRC (Lite-SVRC) (Zhou et al., 2018b; Wang et al., 2018b; Zhang et al., 2018a).

Parameter Settings and Subproblem Solver For each algorithm, we set the cubic penalty parameter  Mtadaptively based on how well the model approximates the real objective as suggested in (Cartis et al., 2011a,b; Kohler and Lucchi, 2017). For SRVRC, we set  S(g)=  S(h)= S for the

image

Figure 1: Plots of logarithmic function value gap with respect to CPU time (in seconds) for nonconvex regularized binary logistic regression on (a) a9a (b) ovtype (c) ijcnn1 and for nonconvex regularized multiclass logistic regression on (d) mnist (e) cifar10 (f) SVHN. Best viewed in color.

simplicity and set gradient and Hessian batch sizes  B(g)t and B(h)tas follows:

image

For SRVRCfree, we set gradient batch sizes  B(g)tthe same as SRVRC and Hessian batch sizes B(h)t = B(h). We tune Sover the grid  {5, 10, 20, 50}, B(g) over the grid {n, n/10, n/20, n/100}, and B(h) over the grid {50, 100, 500, 1000} for the best performance. For SCR, SVRC, Lite-SVRC, and SRVRC, we solve the cubic subproblem using the cubic subproblem solver discussed in (Nesterov and Polyak, 2006). For STR1 and STR2, we solve the trust-region subproblem using the exact trust-region subproblem solver discussed in (Conn et al., 2000). For STC and SRVRCfree, we use Cubic-Subsolver (Algorithm 3 in Appendix G) to approximately solve the cubic subproblem. All algorithms are carefully tuned for a fair comparison.

Datasets and Optimization Problems We use 6 datasets a9a, covtype, ijcnn1 , mnist, cifar10 and SVHN from Chang and Lin (2011) . For binary logistic regression problem with a nonconvex regularizer on a9a, covtype, and ijcnn1, we are given training data  {xi, yi}ni=1, where  xi ∈ Rdand yi ∈ {0, 1}are feature vector and output label corresponding to the i-th training example. The

nonconvex penalized binary logistic regression is formulated as follows

image

where  φ(x) is the sigmoid function and  λ = 10−3. For multiclass logistic regression problem with a nonconvex regularizer on mnist, cifar10 and SVHN, we are given training data  {xi, yi}ni=1, where xi ∈ Rd and yi ∈ Rm are feature vectors and multilabels corresponding to the i-th data points. The nonconvex penalized multiclass logistic regression is formulated as follows

image

where softmax(a) = exp(a)/ �di=1 exp(ai) is the softmax function and  λ = 10−3.We plot the logarithmic function value gap with respect to CPU time in Figure 1. From Figure

1(a) to 1(f), we can see that for the low dimension optimization task on a9a, covtype and ijcnn1, our SRVRC outperforms all the other algorithms with respect to CPU time. We can also observe that the stochastic trust region method STR1 is better than STR2, which is well-aligned with our discussion before. The SPIDER+ does not perform as well as other second-order methods, even though its stochastic gradient and Hessian complexity is comparable to second-order methods in theory. Meanwhile, we also notice that SRVRCfreealways outperforms STC, which suggests that the variance reduction technique is useful. For high dimension optimization task mnist, cifar10 and SVHN, only SPIDER+, STC and SRVRCfreeare able to make notable progress and SRVRCfreeoutperforms the other two. This is again consistent with our theory and discussions in Section 5. Overall, our experiments clearly validate the advantage of SRVRC and SRVRCfree, and corroborate the theory of both algorithms.

In this work we present two faster SVRC algorithms namely SRVRC and SRVRCfreeto find approximate local minima for nonconvex finite-sum optimization problems. SRVRC outperforms existing SVRC algorithms in terms of gradient and Hessian complexities, while SRVRCfree furtheroutperforms the best-known runtime complexity for existing CR based algorithms. Whether our algorithms have achieved the optimal complexity under the current assumptions is still an open problem, and we leave it as a future work.

We define the filtration  Ft = σ(x0, ..., xt) as the  σ-algebra of  x0to  xt. Recall that  vtand  Utare the semi-stochastic gradient and Hessian respectively,  htis the update parameter, and  Mtis the cubic penalty parameter appeared in Algorithm 1 and Algorithm 2. We denote  mt(h) := v⊤h + h⊤Uth/2 + Mt∥h∥32/6 and h∗t = argminh∈Rd mt(h). In this section, we define  δ = ξ/(2T) forthe simplicity.

A.1 Proof of Theorem 4.2

To prove Theorem 4.2, we need the following lemma adapted from Zhou et al. (2018d), which characterizes that  µ(xt + h) can be bounded by  ∥h∥2and the norm of difference between semi-stochastic gradient and Hessian.

Lemma A.1. Suppose that  mt(h) :=  v⊤t h + h⊤Uth/2 +  Mt∥h∥32/6 and  h∗t= argminh∈Rd mt(h). If  Mt/ρ ≥2, then for any  h ∈ Rd, we have

image

Next lemma gives bounds on the inner products  ⟨∇F(xt) − vt, h⟩ and ⟨�∇2F(xt) − Ut�h, h⟩.

Lemma A.2. For any h ∈ Rd, we have

image

We also need the following two lemmas, which show that semi-stochastic gradient and Hessian vt and Utestimators are good approximations to true gradient and Hessian.

Lemma A.3. Suppose that  {B(g)k } satisfies (4.1) and (4.3), then conditioned on  F⌊t/S(g)⌋·S(g), withprobability at least 1  − δ · (t − ⌊t/S(g)⌋ · S(g)), we have that for all  ⌊t/S(g)⌋ · S(g) ≤ k ≤ t,

image

Lemma A.4. Suppose that  {B(h)k } satisfies (4.2) and (4.4), then conditioned on  F⌊t/S(h)⌋·S(h), withprobability at least 1  − δ · (t − ⌊t/S(h)⌋ · S(h)), we have that for all  ⌊t/S(h)⌋ · S(h) ≤ k ≤ t,

image

Given all the above lemmas, we are ready to prove Theorem 4.2.

Proof of Theorem 4.2. Suppose that SRVRC terminates at iteration  T ∗ − 1, then ∥ht∥2 >�ϵ/ρ for

all 0  ≤ t ≤ T ∗ −1. We have

F(xt+1) ≤ F(xt) + ⟨∇F(xt), ht⟩ + 12⟨ht, ∇2F(xt)ht⟩ + ρ6∥ht∥32

image

where the second inequality holds due to the fact that  mt(ht) ≤ mt(0) = 0,  Mt= 4ρand Lemma A.2. By Lemmas A.3 and A.4, with probability at least 1  − 2Tδ, for all 0  ≤ t ≤ T −1, we have that

image

for all 0  ≤ t ≤ T −1. Substituting (A.4) into (A.3), we have

image

Telescoping (A.5) from  t = 0, . . . , T ∗ −1, we have

image

Recall that we have  T ≥ 40∆F √ρ/ϵ3/2 from the condition of Theorem 4.2, then by (A.6), we have T ∗ ≤ T. Thus, we have  ∥hT ∗−1∥2 ≤�ϵ/ρ. Denote �T = T ∗ −1, then we have

image

where the first inequality holds due to Lemma A.1 with  ∇m �T (h �T) = 0 and  ∥h �T ∥2 = ∥h∗�T ∥2. This completes our proof.

A.2 Proof of Corollary 4.3

Proof of Corollary 4.3. Suppose that SRVRC terminates at  T ∗ − 1 ≤ T −1 iteration. Telescoping (A.5) from  t = 0 to T ∗ −1, we have

image

where the last inequality holds since T is set to be 40∆F √ρ/ϵ3/2 as the conditions of Corollary 4.3 suggests. (A.7) implies that �T ∗−1t=0 ∥ht∥32 ≤ 40∆F /ρ. Thus, we have

image

where the first inequality holds due to H¨older’s inequality inequality, and the second inequality is due to  T ∗ ≤ T = 40∆F √ρ/ϵ3/2. We first consider the total gradient sample complexity �T ∗−1t=0 B(g)t ,which can be bounded as

image

where  C1 = 1440 log2(d/δ), the second inequality holds due to (A.8), and the last equality holds due to the choice of  S(g) = √ρϵ/L ·�n ∧ M2/ϵ2. We then consider the total Hessian sample complexity

�T ∗−1t=0 B(h)t, which can be bounded as

image

≤ 40C2

image

where  C2 = 800 log2(d/δ), the second inequality holds due to (A.8), and the last equality holds due to the choice of  S(h) =�n ∧ L/(ρϵ).

In this section, we denote  δ = ξ/(3T) for simplicity.

B.1 Proof of Theorem 5.1

We need the following two lemmas, which bound the variance of semi-stochastic gradient and Hessian estimators.

Lemma B.1. Suppose that  {B(g)k }satisfies (5.2) and (5.3), then conditioned on  F⌊t/S⌋·S, with probability at least 1  − δ · (t − ⌊t/S⌋ · S), we have that for all  ⌊t/S⌋ · S ≤ k ≤ t,

image

Proof of Lemma B.1. The proof is very similar to that of Lemma A.3, hence we omit it.

Lemma B.2. Suppose that  {B(h)k } satisfies (5.1), then conditioned on  Fk, with probability at least 1  − δ, we have that

image

Proof of Lemma B.2. The proof is very similar to that of Lemma A.4, hence we omit it.

We have the following lemma to guarantee that by Algorithm 3 Cubic-Subsolver, the output  htsatisfies that sufficient decrease of function value will be made and the total number of iterations is bounded by  T ′.

image

iterations, where  CS >0 is a constant.

We have the following lemma which provides the guarantee for the dynamic of gradient steps in Cubic-Finalsolver.

Lemma B.4. (Carmon and Duchi, 2016) For  b, A, τ, suppose that  ∥A∥2 ≤ L. We denote that

image

Then for Cubic-Finalsolver, suppose that  η < (4(L+τR))−1, then each iterate ∆ in Cubic-Finalsolver satisfies that  ∥∆∥2 ≤ ∥s∥2, and g(h) is (L + 2τR)-smooth.

With these lemmas, we begin our proof of Theorem 5.1.

Proof of Theorem 5.1. Suppose that SRVRCfreeterminates at iteration  T ∗ − 1. Then T ∗ ≤ T. Wefirst claim that  T ∗ < T. Otherwise, suppose  T ∗ = T, then we have that for all 0  ≤ t < T ∗,

F(xt+1) ≤ F(xt) + ⟨∇F(xt), ht⟩ + 12⟨ht, ∇2F(xt)ht⟩ + ρ6∥ht∥32

image

where the second inequality holds due to  Mt= 4ρand Lemma A.2. By Lemma B.3 and union bound, we know that with probability at least 1  − Tδ, we have

image

where we use the fact that  Mt= 4ρ. By Lemmas B.1 and B.2, we know that with probability at least 1  − 2Tδ, for all 0  ≤ t ≤ T ∗ −1, we have

image

Substituting (B.2) and (B.3) into (B.1), we have

image

Telescoping (B.4) from  t = 0 to T ∗ −1, we have

image

where the last inequality holds since we assume  T ∗=  T ≥25∆F ρ1/2ϵ−3/2from the condition of Theorem 5.1. (B.5) leads to a contradiction, thus we have  T ∗ < T. Therefore, by union bound, with probability at least 1  − 3Tδ, Cubic-Finalsolver is executed by SRVRCfreeat  T ∗ −1 iteration. We have that  ∥vT ∗−1∥2 < max{MT ∗−1ϵ/(2ρ),�LMT ∗−1/2(ϵ/ρ)3/4} and ∥h∗T ∗−1∥2 <�ϵ/ρ by Lemma

image

The only thing left is to check that we indeed find a second-order stationary point,  xT ∗, by Cubic-Finalsolver. We first need to check that the choice of  η= 1/(16L) satisfies that 1/η >4(L + MtR) by Lemma B.4, where

image

where the first inequality holds due to Lemma A.1, the second inequality holds due to the fact that ∥h �T ∥2 ≤ ∥h∗�T ∥2from Lemma B.4, the last inequality holds due to the facts that  ∥∇m �T(h �T)∥2 ≤ ϵfrom Cubic-Finalsolver and  ∥h∗�T ∥2 ≤�ϵ/ρ by Lemma B.3.

B.2 Proof of Corollary 5.2

We have the following lemma to bound the total number of iterations  T ′′ of Algorithm 4 Cubic-Finalsolver.

Lemma B.5. If  ϵ < 4L2ρ/M 2t, then Cubic-Finalsolver will terminate within  T ′′ = CF L/√ρϵiterations, where  CF >0 is a constant.

Proof of Corollary 5.2. We have that

image

where the first inequality holds due to H¨older’s inequality, the second inequality holds due to the facts that  T ∗ ≤ T= 25∆F ρ1/2/ϵ3/2and ∆F ≥ ρ �T ∗−1t=0 ∥ht∥32/4 by (B.5). We first consider the total stochastic gradient computations, �T ∗−1t=0 B(g)t, which can be bounded as

image

where  C1= 2640 log2(d/δ), the second inequality holds due to (B.6), the last equality holds due to the fact  S(g)= √ρϵ/L ·�n ∧ M2/ϵ2. We now consider the total amount of Hessian-vector product computations T , which includes  T1from Cubic-Subsolver and  T2from Cubic-Finalsolver. By Lemma B.3, we know that at k-th iteration of SRVRCfree, Cubic-Subsolver has  T ′iterations, which needs  B(h)kHessian-vector product computations each time. Thus, we have

image

where  C2 = 1200 log2(d/δ), the first inequality holds due to the fact that  B(h)k = C2n ∧ (L2/ρϵ), thesecond inequality holds due to the fact that  T = 25∆F ρ1/2/ϵ3/2, the last inequality holds due to the fact that  T ′ = CSL/Mt ·�ρ/ϵ = CSL/(4√ρϵ). For T2, we have

image

where the first inequality holds due to the fact that  B(h)T ∗−1 = C2n ∧ (L2/ρϵ), the second inequality holds due to the fact that  T ′′=  CF L/√ρϵby Lemma B.5. Since at each iteration we need  B(h)T ∗−1Hessian-vector computations.

Combining (B.7), (B.8) and (B.9), we know that the total stochastic gradient and Hessian-vector product computations are bounded as

image

C.1 Proof of Lemma A.1

We have the following lemmas from Zhou et al. (2018d)

Lemma C.1. (Zhou et al., 2018d) If  Mt ≥ 2ρ, then we have

image

Lemma C.2. (Zhou et al., 2018d) If  Mt ≥ 2ρ, then we have

image

Proof of Lemma A.1. By Lemma C.1, we have

∥∇F(xt + h)∥3/22 ≤�Mt∥h∥22 +��∇F(xt) − vt��2 + 1Mt

image

where the second inequality holds due to the fact that for any  a, b, c ≥0, we have (a + b + c)3/2 ≤

2(a3/2 + b3/2 + c3/2). By Lemma C.2, we have

image

where the second inequality holds due to the fact that for any  a, b, c ≥0, we have (a + b + c)3 ≤9(a3 + b3 + c3). Thus we have

image

where the inequality holds due to (C.1), (C.2) and the fact that  Mt ≥ 4ρ.

C.2 Proof of Lemma A.2

Proof of Lemma A.2. We have

image

where the first inequality holds due to CauchySchwarz inequality, the second inequality holds due to Young’s inequality. We also have

image

where the first inequality holds due to CauchySchwarz inequality, the second inequality holds due to Young’s inequality.

C.3 Proof of Lemma A.3

We need the following lemma:

Lemma C.3. Conditioned on  Fk, with probability at least 1  − δ , we have

image

We also have

image

Proof of Lemma A.3. First, we have  vt − ∇F(xt) = �tk=⌊t/S(g)⌋·S(g) uk, where

image

Meanwhile, we have  E[uk|Fk−1] = 0. Conditioned on  Fk−1, for mod(k, S(g)) ̸= 0, from Lemma C.3, we have that with probability at least 1  − δthe following inequality holds :

image

where the second inequality holds due to (4.1). For mod(k, S(g)) = 0, with probability at least 1  − δ,we have

image

where the second inequality holds due to (4.3). Conditioned on  F⌊t/S(g)⌋·S(g), by union bound, with probability at least 1  − δ · (t − ⌊t/S(g)⌋ · S(g)) (C.5) or (C.6) holds for all  ⌊t/S(g)⌋ · S(g) ≤ k ≤ t. Then for given k, by vector Azuma-Hoeffding inequality in Lemma F.1, conditioned onFk, with probability at least 1  − δ we have

image

Finally, by union bound, we have that with probability at least 1  − 2δ · (t − ⌊t/S(g)⌋ · S(g)), for all⌊t/S(g)⌋ · S(g) ≤ k ≤ t, we have (C.7) holds.

C.4 Proof of Lemma A.4

We need the following lemma:

Lemma C.4. Conditioned on  Fk, with probability at least 1−δ, we have the following concentration inequality

image

We also have

image

Proof of Lemma A.4. First, we have  Ut − ∇2F(xt) = �tk=⌊t/S(h)⌋·S(h) Vk, where

image

Meanwhile, we have  E[Vk|σ(Vk−1, ..., V0)] = 0. Conditioned on  Fk−1, for mod(k, S(h)) ̸= 0, from Lemma C.4, we have that with probability at least 1  − δ, the following inequality holds :

image

where the second inequality holds due to (4.1). For mod(k, S(h)) = 0, with probability at least 1  − δ,we have

image

where the second inequality holds due to (4.3). Conditioned on  F⌊t/S(h)⌋·S(h), by union bound, with probability at least 1  − δ · (t − ⌊t/S(h)⌋ · S(h)) (C.10) or (C.11)holds for all  ⌊t/S(h)⌋ · S(h) ≤ k ≤ t.Then for given k, by Matrix Azuma inequality Lemma F.2, conditioned onFk, with probability at least 1  − δ we have

image

Finally, by union bound, we have that with probability at least 1  − 2δ · (t − ⌊t/S(h)⌋ · S(h)), for all⌊t/S(h)⌋ · S(h) ≤ k ≤ t, we have (C.12) holds.

image

D.1 Proof of Lemma B.3

We have the following lemma which guarantees the effectiveness of Cubic-Subsolver in Algorithm 3. Lemma D.1. (Carmon and Duchi, 2016) Let  A ∈ Rd×dand  ∥A∥2 ≤ β, b ∈ Rd, τ > 0, ζ > 0, ϵ′ ∈(0, 1), δ′ ∈(0, 1) and  η < 1/(8β+ 2τζ). We denote that g(h) =  b⊤h + h⊤Ah/2 +  τ/6 · ∥h∥32and s = argminh∈Rd g(h). Then with probability at least 1  − δ′, if

image

then x = Cubic-Subsolver(A, b, τ, η, ζ, ϵ′, δ′) satisfies that  g(x) ≤ −(1 − ϵ′)τζ3/12.

Proof of Lemma B.3. We simply set  A = Ut, b = vt, τ = Mt, η= (16L)−1, ζ =�ϵ/ρ, ϵ′= 0.5 and  δ′ = δ. We have ∥Ut∥2 ≤ L, then we set  β = L. With the choice of  Mt where Mt = 4ρ and theassumption that  ϵ < 4L2ρ/M 2t , we can check that  η < 1/(8β + 2τζ). We also have that  s = h∗t and(D.1) holds. Thus, by Lemma D.1, we have

image

By the choice of  T ′ in Cubic-Subsolver, we have

image

D.2 Proof of Lemma B.5

We have the following lemma which provides the guarantee for the function value in Cubic-Finalsolver.

Lemma D.2. (Carmon and Duchi, 2016) We denote that g(h) =  b⊤h + h⊤Ah/2 +  τ/6 · ∥h∥32, s = argminh∈Rd g(h), then g(s) ≥ ∥b∥2∥s∥2/2 − τ∥s∥32/6.

Proof of Lemma B.5. In Cubic-Finalsolver we are focusing on minimizing  mT ∗−1(h). We have that  ∥vt∥2 < max{Mtϵ/(2ρ),�LMt/2(ϵ/ρ)3/4}and  ∥h∗T ∗−1∥2 ≤�ϵ/ρby Lemma B.3. We can check that  η= (16L)−1satisfies that  η <(4(L + τR))−1, where R is defined in Lemma B.4, when ϵ < 4L2ρ/M 2t . From Lemma B.4 we also know that  mT ∗−1 is (L+2MT ∗−1R)-smooth, which satisfies that 1/η > 2(L + 2MT ∗−1R). Thus, by standard gradient descent analysis, to get a point ∆ where ∥∇mT ∗−1(∆)∥2 ≤ ϵ, Cubic-Finalsolver needs to run

image

iterations, where we denote by ∆0the starting point of Cubic-Finalsolver. By directly computing, we have  mT ∗−1(∆0) ≤0. By Lemma D.2, we have

image

Thus, (D.2) can be further bounded as  T ′′ = O(L/√ρϵ).

E.1 Proof of Lemma C.3

Proof of Lemma C.3. We only need to consider the case where  B(g)k = |Jk| < n. For each  i ∈ Jk, let

image

then we have  Eiai = 0, aii.i.d., and

image

where the second inequality holds due to the L-smoothness of  fiand F. Thus by vector Azuma-Hoeffding inequality in Lemma F.1, we have that with probability at least 1  − δ,

image

For each  i ∈ Jk, let

image

then we have  Eibi = 0 and ∥bi∥2 ≤ M. Thus by vector Azuma-Hoeffding inequality in Lemma F.1, we have that with probability at least 1  − δ,

image

E.2 Proof of Lemma C.4

Proof of Lemma C.4. We only need to consider the case where  B(h)k = |Ik| < n. For each i ∈ Ik, let

image

then we have  EiAi = 0, A⊤i = Ai, Aii.i.d. and

image

where the second inequality holds due to  ρ-Hessian Lipschitz continuous of  fiand F. Then by Matrix Azuma inequality Lemma F.2, we have that with probability at least 1  − δ,

image

For each  i ∈ Ik, let

image

then we have  EiBi= 0,  B⊤i=  Bi, and  ∥Bi∥2 ≤ 2L. Then by Matrix Azuma inequality in Lemma F.2, we have that with probability at least 1  − δ,

image

which completes the proof.

We have the following vector Azuma-Hoeffding inequality:

Lemma F.1. (Pinelis, 1994) Consider  {vk}be a vector-valued martingale difference, where E[vk|σ(v1, ..., vk−1)] = 0 and ∥vk∥2 ≤ Ak, then we have that with probability at least 1  − δ,

image

We have the following Matrix Azuma inequality :

Lemma F.2. (Tropp, 2012) Consider a finite adapted sequence  {Xk}of self-adjoint matrices in dimension d, and a fixed sequence  {Ak}of self-adjoint matrices that satisfy

image

Due to space limit, we include the approximate solvers (Carmon and Duchi, 2016) for the cubic subproblem in this section for the purpose of self-containedness.

image

image

Agarwal, N., Allen-Zhu, Z., Bullins, B., Hazan, E. and Ma, T. (2017). Finding approximate local minima faster than gradient descent. In Proceedings of the 49th Annual ACM SIGACT Symposium on Theory of Computing.

Allen-Zhu, Z. (2017). Natasha 2: Faster non-convex optimization than sgd. arXiv preprint arXiv:1708.08694 .

Allen-Zhu, Z. (2018). Natasha 2: Faster non-convex optimization than sgd. In Advances in Neural Information Processing Systems.

Allen-Zhu, Z. and Hazan, E. (2016). Variance reduction for faster non-convex optimization. In International Conference on Machine Learning.

Allen-Zhu, Z. and Li, Y. (2018). Neon2: Finding local minima via first-order oracles. In Advances in Neural Information Processing Systems.

Bhojanapalli, S., Neyshabur, B. and Srebro, N. (2016). Global optimality of local search for low rank matrix recovery. In Advances in Neural Information Processing Systems.

Blanchet, J., Cartis, C., Menickelly, M. and Scheinberg, K. (2016). Convergence rate analysis of a stochastic trust region method for nonconvex optimization. arXiv preprint arXiv:1609.07428 .

Carmon, Y. and Duchi, J. C. (2016). Gradient descent efficiently finds the cubic-regularized non-convex newton step. arXiv preprint arXiv:1612.00547 .

Carmon, Y. and Duchi, J. C. (2018). Analysis of krylov subspace solutions of regularized non-convex quadratic problems. In Advances in Neural Information Processing Systems.

Carmon, Y., Duchi, J. C., Hinder, O. and Sidford, A. (2018). Accelerated methods for nonconvex optimization. SIAM Journal on Optimization 28 1751–1772.

Cartis, C., Gould, N. I. and Toint, P. L. (2009). Trust-region and other regularisations of linear least-squares problems. BIT Numerical Mathematics 49 21–53.

Cartis, C., Gould, N. I. and Toint, P. L. (2011a). Adaptive cubic regularisation methods for unconstrained optimization. part i: motivation, convergence and numerical results. Mathematical Programming 127 245–295.

Cartis, C., Gould, N. I. and Toint, P. L. (2012). Complexity bounds for second-order optimality in unconstrained optimization. Journal of Complexity 28 93–108.

Cartis, C., Gould, N. I. and Toint, P. L. (2013). On the evaluation complexity of cubic regularization methods for potentially rank-deficient nonlinear least-squares problems and its relevance to constrained nonlinear optimization. SIAM Journal on Optimization 23 1553–1574.

Cartis, C., Gould, N. I. M. and Toint, P. L. (2011b). Adaptive cubic regularisation methods for unconstrained optimization. Part II: worst-case function- and derivative-evaluation complexity. Springer-Verlag New York, Inc.

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

Conn, A. R., Gould, N. I. and Toint, P. L. (2000). Trust region methods. SIAM.

Curtis, F. E., Robinson, D. P. and Samadi, M. (2017). A trust region algorithm with a worst-case iteration complexity of  o(ϵ−3/2) for nonconvex optimization. Mathematical Programming 162 1–32.

Defazio, A., Bach, F. and Lacoste-Julien, S. (2014). Saga: A fast incremental gradient method with support for non-strongly convex composite objectives. In Advances in Neural Information Processing Systems.

Fang, C., Li, C. J., Lin, Z. and Zhang, T. (2018). Spider: Near-optimal non-convex optimization via stochastic path integrated differential estimator. arXiv preprint arXiv:1807.01695 .

Fang, C., Lin, Z. and Zhang, T. (2019). Sharp analysis for nonconvex sgd escaping from saddle points. arXiv preprint arXiv:1902.00247 .

Garber, D. and Hazan, E. (2015). Fast and simple pca via convex optimization. arXiv preprint arXiv:1509.05647 .

Ge, R., Huang, F., Jin, C. and Yuan, Y. (2015). Escaping from saddle pointsonline stochastic gradient for tensor decomposition. In Conference on Learning Theory.

Ge, R., Lee, J. D. and Ma, T. (2016). Matrix completion has no spurious local minimum. In Advances in Neural Information Processing Systems.

Golub, G. H. and Van Loan, C. F. (1996). Matrix Computations (3rd Ed.). Johns Hopkins University Press, Baltimore, MD, USA.

Hardt, M. and Ma, T. (2016). Identity matters in deep learning. arXiv preprint arXiv:1611.04231 .

Hillar, C. J. and Lim, L.-H. (2013). Most tensor problems are np-hard. Journal of the ACM (JACM) 60 45.

Jin, C., Ge, R., Netrapalli, P., Kakade, S. M. and Jordan, M. I. (2017a). How to escape saddle points efficiently. arXiv preprint arXiv:1703.00887 .

Jin, C., Netrapalli, P., Ge, R., Kakade, S. M. and Jordan, M. I. (2019). Stochastic gradient descent escapes saddle points efficiently. arXiv preprint arXiv:1902.04811 .

Jin, C., Netrapalli, P. and Jordan, M. I. (2017b). Accelerated gradient descent escapes saddle points faster than gradient descent. arXiv preprint arXiv:1711.10456 .

Johnson, R. and Zhang, T. (2013). Accelerating stochastic gradient descent using predictive variance reduction. In Advances in neural information processing systems.

Kawaguchi, K. (2016). Deep learning without poor local minima. In Advances in Neural Information Processing Systems.

Kohler, J. M. and Lucchi, A. (2017). Sub-sampled cubic regularization for non-convex optimization. arXiv preprint arXiv:1705.05933 .

LeCun, Y., Bengio, Y. and Hinton, G. (2015). Deep learning. Nature 521 436–444.

Mart´ınez, J. M.and Raydan, M. (2017). Cubic-regularization counterpart of a variable-norm trust-region method for unconstrained minimization. Journal of Global Optimization 68 367–385.

Nesterov, Y. and Polyak, B. T. (2006). Cubic regularization of newton method and its global performance. Mathematical Programming 108 177–205.

Nguyen, L. M., Liu, J., Scheinberg, K. and  Tak´aˇc, M.(2017). Sarah: A novel method for machine learning problems using stochastic recursive gradient. In International Conference on Machine Learning.

Nguyen, L. M., van Dijk, M., Phan, D. T., Nguyen, P. H., Weng, T.-W. and Kalagnanam, J. R. (2019). Optimal finite-sum smooth non-convex optimization with sarah. arXiv preprint arXiv:1901.07648 .

Pinelis, I. (1994). Optimum bounds for the distributions of martingales in banach spaces. The Annals of Probability 1679–1706.

Reddi, S. J., Hefny, A., Sra, S., Poczos, B. and Smola, A. (2016). Stochastic variance reduction for nonconvex optimization 314–323.

Roux, N. L., Schmidt, M. and Bach, F. R. (2012). A stochastic gradient method with an exponential convergence rate for finite training sets. In Advances in Neural Information Processing Systems.

Royer, C. W. and Wright, S. J. (2017). Complexity analysis of second-order line-search algorithms for smooth nonconvex optimization. arXiv preprint arXiv:1706.03131 .

Shalev-Shwartz, S. (2016). Sdca without duality, regularization, and individual convexity. In International Conference on Machine Learning.

Shen, Z., Zhou, P., Fang, C. and Ribeiro, A. (2019). A stochastic trust region method for non-convex minimization. arXiv preprint arXiv:1903.01540 .

Tripuraneni, N., Stern, M., Jin, C., Regier, J. and Jordan, M. I. (2018). Stochastic cubic regularization for fast nonconvex optimization. In Advances in Neural Information Processing Systems.

Tropp, J. A. (2012). User-friendly tail bounds for sums of random matrices. Foundations of computational mathematics 12 389–434.

Wang, Z., Ji, K., Zhou, Y., Liang, Y. and Tarokh, V. (2018a). Spiderboost: A class of faster variance-reduced algorithms for nonconvex optimization. arXiv preprint arXiv:1810.10690 .

Wang, Z., Zhou, Y., Liang, Y. and Lan, G. (2018b). Sample complexity of stochastic variance-reduced cubic regularization for nonconvex optimization. arXiv preprint arXiv:1802.07372 .

Xiao, L. and Zhang, T. (2014). A proximal stochastic gradient method with progressive variance reduction. SIAM Journal on Optimization 24 2057–2075.

Xu, P., Roosta-Khorasani, F. and Mahoney, M. W. (2017). Newton-type methods for non-convex optimization under inexact hessian information. arXiv preprint arXiv:1708.07164 .

Xu, Y., Rong, J. and Yang, T. (2018). First-order stochastic algorithms for escaping from saddle points in almost linear time. In Advances in Neural Information Processing Systems.

Yu, Y., Xu, P. and Gu, Q. (2018). Third-order smoothness helps: faster stochastic optimization algorithms for finding local minima. In Advances in Neural Information Processing Systems.

Yu, Y., Zou, D. and Gu, Q. (2017). Saving gradient and negative curvature computations: Finding local minima more efficiently. arXiv preprint arXiv:1712.03950 .

Zhang, J., Xiao, L. and Zhang, S. (2018a). Adaptive stochastic variance reduction for subsampled newton method with cubic regularization. arXiv preprint arXiv:1811.11637 .

Zhang, X., Wang, L., Yu, Y. and Gu, Q. (2018b). A primal-dual analysis of global optimality in nonconvex low-rank matrix recovery. In International conference on machine learning.

Zhou, D., Xu, P. and Gu, Q. (2018a). Finding local minima via stochastic nested variance reduction. arXiv preprint arXiv:1806.08782 .

Zhou, D., Xu, P. and Gu, Q. (2018b). Sample efficient stochastic variance-reduced cubic regularization method. arXiv preprint arXiv:1811.11989 .

Zhou, D., Xu, P. and Gu, Q. (2018c). Stochastic nested variance reduced gradient descent for nonconvex optimization. In Advances in Neural Information Processing Systems.

Zhou, D., Xu, P. and Gu, Q. (2018d). Stochastic variance-reduced cubic regularized Newton methods. In Proceedings of the 35th International Conference on Machine Learning. PMLR, 5990–5999.

Zhou, D., Xu, P. and Gu, Q. (2019). Stochastic variance-reduced cubic regularization methods. Journal of Machine Learning Research 20 1–47.


Designed for Accessibility and to further Open Science