This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

11affiliationtext: Cainiao Network, Hangzhou, Zhejiang, China

Online Primal-Dual Algorithms For Stochastic Resource Allocation Problems

Yuwei Chen Equal contribution. Zengde Deng footnotemark: Zaiyi Chen Yinzhi Zhou Yujie Chen Haoyuan Hu
Abstract

This paper studies the online stochastic resource allocation problem (RAP) with chance constraints and conditional expectation constraints. The online RAP is an integer linear programming problem where resource consumption coefficients are revealed column by column along with the corresponding revenue coefficients. When a column is revealed, the corresponding decision variables are determined instantaneously without future information. In online applications, the resource consumption coefficients are often obtained by prediction. An application for such scenario rises from the online order fulfilment task. When the timeliness constraints are considered, the coefficients are generated by the prediction for the transportation time from origin to destination. To model their uncertainties, we take the chance constraints and conditional expectation constraints into the consideration. Assuming that the uncertain variables have known Gaussian distributions, the stochastic RAP can be transformed into a deterministic but nonlinear problem with integer second-order cone constraints. Next, we linearize this nonlinear problem and theoretically analyze the performance of vanilla online primal-dual algorithm for solving the linearized stochastic RAP. Under mild technical assumptions, the optimality gap and constraint violation are both on the order of n\sqrt{n}. Then, to further improve the performance of the algorithm, several modified online primal-dual algorithms with heuristic corrections are proposed. Finally, extensive numerical experiments demonstrate the applicability and effectiveness of our methods.

1 Introduction

1.1 Background

The resource allocation problem (RAP) [2] is to find the best allocation of a fixed amount of resources to various activities, in order to maximize the total revenue. The online RAP has a wide range of applications such as network routing [7], computer resource allocation [19], Internet advertising [24], portfolio selection [14] and so on. This paper studies a multi-dimensional online RAP with uncertainty. There are mm resources and kk resource consumption schemes for each request. The request for the resources arrives one by one. When the ii-th request is revealed, one or none of kk resource consumption schemes is chosen to satisfy this request. If ll-th resource consumption scheme is chosen, the revenue and the consumption of the jj-th resource are ctlc_{tl} and atjla_{tjl} respectively. The decision is irrevocable and has to be decided immediately according to the historical information {(𝒄τ,𝑨τ)}τ=1t\{(\bm{c}_{\tau},\bm{A}_{\tau})\}_{\tau=1}^{t}, without future information. Our aim is to maximize the total revenue with limited resource capacities, given the total number nn of incoming requests and considering the uncertainty of atjla_{tjl}.

This paper models the uncertainty with chance constraints and conditional expectation constraints. On one hand, a number of studies such as [23, 30, 29, 11] have applied the offline chance constrained resource allocation model into engineering fields. Some of these fields are better suited for applying online models. Taking [23] as an example, Lu et al. studied a offline planning model for selecting non-profit operations in a hospital considering the uncertainty of resource consumption. It is an online problem in practice since the decision must be made immediately when a operation request is arrived. On the other hand, the online chance constrained resource allocation model can be applied to the order fulfillment task, which is an important application in the field of the supply chain. In the order fulfillment problem, each order needs to be allocated to a transportation channel which will be delivered from the origin to the destination [27]. The goal is to maximize the total revenue while ensuring the average transportation time does not suppress the given threshold. We refer this average transportation time constraint as the timeliness constraint. A long delivery time highly impairs the customer’s experience. Consequently, timeliness constraint upper bound are often adopted to guarantee the relatively low average transportation time. Due to the complicacy of the transportation process, the transportation time of each channel is attained by prediction and has uncertainty. As far as we know, this is the first work to take both chance constraints and conditional expectation constraints into the consideration for the online RAP.

1.2 Related Works

The deterministic RAP can be modeled as an integer linear programming (ILP) problem. Many recent papers have studied the online ILP problems (see [25, 21, 1, 13, 10, 3, 12, 20] and references therein). Algorithms in [25, 21, 1, 13, 10, 3, 12, 20] are all dual-based which maintain dual prices in iterations and can achieve near-optimal solutions under mild conditions. When a new request arrives, the decision is made immediately based on the dual price vector. Among these studies, researchers [25, 21, 1, 13, 10] construct dual problems by using historical information and solving them to obtain the dual prices. To deal with the disadvantage that solving dual problems may be time-consuming, researchers [3, 12, 20] propose online primal-dual (OPD) algorithms that update the dual prices by utilizing the dual mirror descent or projected stochastic subgradient descent. It is worth to mention that these OPD algorithms are simple and efficient since the dual prices are updated through algebraic computation without solving optimization problems.

However, the optimization models studied in [25, 21, 1, 13, 10, 12, 18, 20, 3] are deterministic and may suffer from poor performance when the resource consumption is uncertain in practice. In the existing articles that study the uncertain online optimization, the uncertainty is modeled by the worst-case scenario value, expectation, regret, or a linear combination of the above (see [5, 4, 16, 22, 17]). These modeling methods are mainly aimed at the uncertainty in the objective function, while almost no chance constraint or conditional expectation constraint is considered in the existing studies.

Chance constrained programming [8] is a widely used stochastic programming technique to model the uncertainty in constraints. In stochastic programming [28], it is assumed that some parameters are uncertain and their distributions are known. If the uncertain parameters in an active inequality constraint are set to the medians, the probability of this constraint not holding is 50%. To avoid this issue in the online RAP, this paper adopts the chance constraints to model the uncertainty. The chance constraint is the constraint on the uncertain parameters whose holding probability is not lower than the prescribed level. The solution methods for chance constrained programming (CCP) have been studied by [8, 15, 9, 26]. If the uncertain parameters have a known multivariate Gaussian distribution, the chance constrained counterparts of linear constraints can be transformed into deterministic second-order cone (SOC) constraints.

The conditional expectation constraint [26] can be served as a supplement to chance constraints, since the chance constraint does not restrict the amount of the violation of inequality constraints directly. In some cases, the expected amount of violation can be large even if the chance constraints are satisfied. To address this concern, the conditional expectation constraint is also taken into consideration in this paper. Similar to chance constraints, the conditional expectation constraints can also be transformed into deterministic SOC constraints under the assumption of multivariate Gaussian distribution [26].

1.3 Main Contributions

This paper studies the online stochastic RAP considering the uncertainty of resource consumption coefficients. The chance constraint and conditional expectation constraint are used to model the uncertainty and both of them can be transformed into the SOC constraints equivalently. The non-linearity and indecomposability of the SOC constraints make the online problem challenging to handle. The main contributions of this paper are as follows.

  1. (1)

    To the best of our knowledge, this is the first time chance constraints and conditional expectation constraints are introduced in the online RAP. A linearization method is presented to transform the SOC constrained problem into a linear form suitable for the online solution.

  2. (2)

    We theoretically analyze the performance of the vanilla OPD algorithm when it is applied to solve the linearized stochastic RAP. Under mild technical assumptions, the expected optimality gap and constraint violation are both O(n)O(\sqrt{n}).

  3. (3)

    We propose modified versions of the OPD approach by leveraging the structure of the SOC constraints to effectively reduce the probability deviation of chance constraints and the constraint violation of conditional expectation constraints in practice.

  4. (4)

    Massive numerical experiments are conducted to demonstrate the applicability and effectiveness of the proposed algorithms.

The rest of this paper is organized as follows. Section 2 formulates the stochastic programming model of RAP and its linear relaxation. Section 3 introduces the proposed heuristic corrections and modified OPD algorithms for solving stochastic RAPs. Section 4 gives the numerical experiment results. Finally in Section 5, conclusions are drawn.

2 Model Description

This section first formulates the deterministic model of the RAP. Then, a stochastic counterpart with chance constraints and conditional expectation constraints is established. Finally, the stochastic programming problem is relaxed into an integer linear problem suitable for online solution.

2.1 Deterministic Problem

Consider the multi-dimensional RAP with nn requests and mm resources. For each request, there are always kk resource consumption schemes that can satisfy it. When a request is revealed, the decision maker chooses one scheme or none. Without loss of generality, a deterministic multi-dimensional RAP can be modeled as follows:

max𝒙\displaystyle\max_{\bm{x}} t=1n𝒄t𝒙t\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t} (1)
s.t.\displaystyle\rm{s.t.} t=1n𝒂tj𝒙tbj,j=1,,m\displaystyle\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}\leq b_{j},\forall j=1,\dots,m
𝟏𝒙t1,𝒙t{0,1}k,t=1,,n\displaystyle\bm{1}^{\top}\bm{x}_{t}\leq 1,\bm{x}_{t}\in\{0,1\}^{k},\forall t=1,\dots,n

where the revenue coefficient vector 𝒄tk\bm{c}_{t}\in\mathbb{R}^{k}, and the resource consumption vector 𝒂tjk\bm{a}_{tj}\in\mathbb{R}^{k}. The decision variables are 𝒙=(𝒙1,,𝒙n)\bm{x}=(\bm{x}_{1},\dots,\bm{x}_{n})^{\top}. xtl=1x_{tl}=1 means that tt-th request is satisfied by resource consumption scheme ll. bjb_{j} is the capacity of resource jj. 𝟏\bm{1} denotes all-one vector. In the online setting of ILP, the input data (𝒄t,𝒂t1,,𝒂tm)(\bm{c}_{t},\bm{a}_{t1},\dots,\bm{a}_{tm}) is revealed one by one and 𝒙t\bm{x}_{t} is determined instantaneously when (𝒄t,𝒂t1,,𝒂tm)(\bm{c}_{t},\bm{a}_{t1},\dots,\bm{a}_{tm}) is revealed.

Throughout this paper, we assume that the input data of deterministic problem and stochastic programming problem is drawn i.i.d. from some distributions. Moreover, nn and 𝒃\bm{b} are assumed to be known and fixed.

2.2 Stochastic Programming Problem

In practice, the value of 𝒂tj\bm{a}_{tj} may be obtained by predicting that yields the uncertainty. Consequently, taken the uncertainty of 𝒂tj\bm{a}_{tj} into consideration, we formulate the following chance constraints:

𝒂tj𝒫a(t=1n𝒂tj𝒙tbj)ηj,j=1,,m,\mathbb{P}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left(\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}\leq b_{j}\right)\geq\eta_{j},\forall j=1,\dots,m, (2)

where \mathbb{P} means probability, ηj\eta_{j} is the given confidence level and 𝒫a\mathcal{P}_{a} is the distribution of 𝒂tj\bm{a}_{tj}.

The constraints (2) only restrict the probability of violating inequality and do not restrict the amount of violation directly. To address this issue, the conditional expectation constraints (3) are also considered in this paper which can limit the expected violation of these inequality constraints.

𝔼𝒂tj𝒫a[t=1n𝒂tj𝒙tbj|t=1n𝒂tj𝒙tbj>0]γj,j=1,,m,\mathbb{E}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left[\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}\bigg{|}\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}>0\right]\leq\gamma_{j},\forall j=1,\dots,m, (3)

where 𝔼\mathbb{E} denotes expectation and γj\gamma_{j} is the given parameter. This paper assumes that the expectation in (3) always exists.

Replacing the deterministic constraints in (1) by the chance constraints (2) and conditional expectation constraints (3), we formulate the following stochastic programming problem:

max𝒙\displaystyle\max_{\bm{x}} t=1n𝒄t𝒙t\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t} (4)
s.t.\displaystyle\rm{s.t.} 𝒂tj𝒫a(t=1n𝒂tj𝒙tbj)ηj,j=1,,m\displaystyle\mathbb{P}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left(\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}\leq b_{j}\right)\geq\eta_{j},\forall j=1,\dots,m
𝔼𝒂tj𝒫a[t=1n𝒂tj𝒙tbj|t=1n𝒂tj𝒙tbj>0]γj,j=1,,m\displaystyle\mathbb{E}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left[\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}\bigg{|}\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}>0\right]\leq\gamma_{j},\forall j=1,\dots,m
𝟏𝒙t1,𝒙t{0,1}k,t=1,,n.\displaystyle\bm{1}^{\top}\bm{x}_{t}\leq 1,\bm{x}_{t}\in\{0,1\}^{k},\forall t=1,\dots,n.

For convenience, we refer to the problem (4) as a CCP problem throughout this paper, thought (3) are not standard chance constraints.

2.3 Equivalent Transformation

In the online setting of the CCP problem (4), the revenue vector 𝒄t\bm{c}_{t} and the distributions of the resource consumption vectors {𝒂t1,,𝒂tm}\{\bm{a}_{t1},\dots,\bm{a}_{tm}\} are revealed one by one. For any tt and jj, we assume that 𝒂tjN(𝒂¯tj,𝑲tj)\bm{a}_{tj}\sim N(\bar{\bm{a}}_{tj},\bm{K}_{tj}) which means the true value of 𝒂tj\bm{a}_{tj} belongs to a known Gaussian distribution with mean 𝒂¯tj\bar{\bm{a}}_{tj} and covariance matrix 𝑲tj\bm{K}_{tj}. In the following, we will reformulate (4) into a deterministic problem.

First is to reformulate the chance constraints (2). According to the properties of Gaussian distribution, we have that t=1n𝒂tj𝒙tN(t=1n𝒂¯tj𝒙t\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}\sim N(\sum_{t=1}^{n}\bar{\bm{a}}_{tj}^{\top}\bm{x}_{t}, t=1n𝒙t𝑲tj𝒙t)\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}). Thus, the constraints (2) are equivalent to the following constraints [15]:

t=1n𝒂¯tj𝒙t+Φ1(ηj)t=1n𝒙t𝑲tj𝒙tbj,j=1,,m,\sum_{t=1}^{n}\bar{\bm{a}}_{tj}^{\top}\bm{x}_{t}+\Phi^{-1}(\eta_{j})\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\leq b_{j},\forall j=1,\dots,m, (5)

where Φ()\Phi(\cdot) represents the cumulative distribution function of the standard Gaussian distribution.

Next is to reformulate the conditional expectation constraints (3). The value of the conditional expectation 𝔼[t=1n𝒂tj𝒙tbj|t=1n𝒂tj𝒙tbj>0]\mathbb{E}[\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}|\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}>0] is related to the variance t=1n𝒙t𝑲tj𝒙t\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}, which makes the parameter γj\gamma_{j} difficult to quantify. As an alternative, we replace γj\gamma_{j} with the normalized value γ~j\tilde{\gamma}_{j} that equals to γj/t=1n𝒙t𝑲tj𝒙t\gamma_{j}/\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}. Specifically, the constraint (3) is rewritten as

𝔼𝒂tj𝒫a[t=1n𝒂tj𝒙tbjt=1n𝒙t𝑲tj𝒙t|t=1n𝒂tj𝒙tbj>0]γ~j,j=1,,m,\mathbb{E}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left[\frac{\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}}{\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}}\bigg{|}\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}>0\right]\leq\tilde{\gamma}_{j},\forall j=1,\dots,m, (6)

According to [26], the constraints (6) are equivalent to

t=1n𝒂¯tj𝒙t+h1(γj~)t=1n𝒙t𝑲tj𝒙tbj,j=1,,m,\sum_{t=1}^{n}\bar{\bm{a}}_{tj}^{\top}\bm{x}_{t}+h^{-1}(\tilde{\gamma_{j}})\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\leq b_{j},\forall j=1,\dots,m, (7)

where the function h()h(\cdot) has the form111In [26], \sqrt{\cdot} in 2π\sqrt{2\pi} is left out which is a typo.

h(z)=e12z2/2π1Φ(z)z.h(z)=\frac{e^{-\frac{1}{2}z^{2}}/\sqrt{2\pi}}{1-\Phi(z)}-z. (8)

Under the assumption of Gaussian distribution, the constraints (5) and (7) have the same formulations and the only difference lies on the coefficient of t=1n𝒙t𝑲tj𝒙t\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}. Consequently, when both (5) and (7) exist, we can merge them into the following constraint,

t=1n𝒂¯tj𝒙t+ψjt=1n𝒙t𝑲tj𝒙tbj,j=1,,m,\sum_{t=1}^{n}\bar{\bm{a}}_{tj}^{\top}\bm{x}_{t}+\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\leq b_{j},\forall j=1,\dots,m, (9)

where ψj=max{Φ1(ηj),h1(γ~j)}\psi_{j}=\max\{\Phi^{-1}(\eta_{j}),h^{-1}(\tilde{\gamma}_{j})\}. When ηj>50%\eta_{j}>50\% or γ~j<π/2\tilde{\gamma}_{j}<\sqrt{\pi/2}, ψj>0\psi_{j}>0.

Substituting (9) into (4), the CCP problem (4) is equivalent to the following deterministic problem:

max𝒙\displaystyle\max_{\bm{x}} t=1n𝒄t𝒙t\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t} (10)
s.t.\displaystyle\rm{s.t.} t=1n𝒂¯tj𝒙t+ψjt=1n𝒙t𝑲tj𝒙tbj,j=1,,m.\displaystyle\sum_{t=1}^{n}\bar{\bm{a}}_{tj}^{\top}\bm{x}_{t}+\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\leq b_{j},\forall j=1,\dots,m.
𝟏𝒙t1,𝒙t{0,1}k,t=1,,n\displaystyle\bm{1}^{\top}\bm{x}_{t}\leq 1,\bm{x}_{t}\in\{0,1\}^{k},\forall t=1,\dots,n

The problem (10) is an integer SOC programming (ISOCP) problem when ψj>0,j\psi_{j}>0,\forall j. The offline version of (10) can be solved by commercial solvers such as Gurobi. However, in the online setting, it is difficult to solve problem (10) due to its indecomposability: 𝒙t\bm{x}_{t} with different subscripts tt are coupled with each other in t=1n𝒙t𝑲tj𝒙t\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}. In other words, calculating the individual constraint consumption at each time step tt is challenging. Moreover, we make the following assumption throughout this paper.

Assumption 1.

We assume

  1. (a)

    The coefficient set {ctj,𝒂¯tj,𝑲tj}\{c_{tj},\bm{\bar{a}}_{tj},\bm{K}_{tj}\}’s are i.i.d. sampled from an unknown distribution 𝒫\mathcal{P}.

  2. (b)

    The coefficient set {ctj,𝒂¯tj,𝑲tj}\{c_{tj},\bm{\bar{a}}_{tj},\bm{K}_{tj}\}’s are bounded.

  3. (c)

    The right-hand-side 𝒃=n𝒅\bm{b}=n\bm{d}. 𝒅\bm{d} is bounded and its upper and lower bounds are both positive.

2.4 Relaxed Linear Problem

To begin, we provide the following proposition.

Proposition 1.

For arbitrary tt and jj, the following equation holds.

𝒙t𝑲tj𝒙t=𝜸tj𝒙t,𝒙t{𝒙{0,1}k|𝟏𝒙1},\sqrt{\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}=\bm{\bm{\gamma}}^{\top}_{tj}\bm{x}_{t},\forall\bm{x}_{t}\in\{\bm{x}\in\{0,1\}^{k}|\bm{1}^{\top}\bm{x}\leq 1\},\vspace{-0.3cm}

where 𝛄tj\bm{\gamma}_{tj} is formed by concatenating the square roots of the diagonal elements of the matrix 𝐊tj\bm{K}_{tj}.

Proof.

See A. ∎

To address the non-decomposable issue raised by the non-linearity of t=1n𝒙t𝑲tj𝒙t\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}, we linearize this term to decouple different 𝒙t\bm{x}_{t}. Specifically, according to Cauchy-Schwarz inequality nt=1n𝒙t𝑲tj𝒙tt=1n\sqrt{n\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\geq\sum_{t=1}^{n} 𝒙t𝑲tj𝒙t\sqrt{\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}} and Proposition 1, the nonlinear and non-decomposable problem (10) can be approximated by

max𝒙\displaystyle\max_{\bm{x}} t=1n𝒄t𝒙t\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t} (11)
s.t.\displaystyle\rm{s.t.} t=1n(𝒂¯tj+ψjn𝜸tj)𝒙tbj,j=1,,m\displaystyle\sum_{t=1}^{n}\left(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}^{\top}_{tj}\right)\bm{x}_{t}\leq b_{j},\forall j=1,\dots,m
𝟏𝒙t1,𝒙t{0,1}k,t=1,,n.\displaystyle\bm{1}^{\top}\bm{x}_{t}\leq 1,\bm{x}_{t}\in\{0,1\}^{k},\forall t=1,\dots,n.

It is worth to mention that the relaxed problem (11) is an integer LP (ILP) problem which is linear and decomposable, and can be solved in the online setting by existing algorithms. The vanilla OPD algorithm for solving this relaxed problem is the basis of our algorithms for solving the CCP problem (10) and we will detail it in the following section.

3 Solution Algorithms

In this section, we introduce several online primal-dual methods to handle the online SOC constrained problem (10). Firstly, we revisit the state-of-the-art OPD algorithm for solving the relaxed problem (11). Then, some heuristic correction methods based on the structure of (10) are proposed to improve the practical performance.

3.1 OPD Algorithm for online ILP

Recall that (11) is an ILP problem and Li et al. [20] have proposed an effective OPD algorithm to solve the online ILP problem. For simplicity, denote 𝒂~tj=𝒂¯tj+ψj𝜸tj/n\tilde{\bm{a}}_{tj}=\bar{\bm{a}}_{tj}+{\psi_{j}}\bm{\gamma}_{tj}{/\sqrt{n}} and we present the OPD method as shown in Algorithm 1.

Input: 𝒅=𝒃/n\bm{d}=\bm{b}/n
Output: 𝒙=(𝒙1,,𝒙n)\bm{x}=(\bm{x}_{1},...,\bm{x}_{n})
1 Initialize: 𝒑1=𝟎\bm{p}_{1}=\bm{0}
2 for t=1,,nt=1,...,n do
3       Set vt=maxl=1,,k(𝒄t𝒑t𝑨~t)𝒆lv_{t}=\max_{l=1,\dots,k}\ (\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\tilde{\bm{A}}_{t})\bm{e}_{l}
4       if vt>0v_{t}>0 then
5             Pick an index ltl_{t} randomly from
{l:vt=(𝒄t𝒑t𝑨~t)𝒆l}\left\{l:v_{t}=(\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\tilde{\bm{A}}_{t})\bm{e}_{l}\right\}
6             Set 𝒙t=𝒆lt\bm{x}_{t}=\bm{e}_{l_{t}}
7      else
8            Set 𝒙t=𝟎\bm{x}_{t}=\bm{0}
9      Compute
𝒑t+1=max{𝒑t+1n(𝑨~t𝒙t𝒅),𝟎}\bm{p}_{t+1}=\max\left\{\bm{p}_{t}+\frac{1}{\sqrt{n}}\left(\tilde{\bm{A}}_{t}\bm{x}_{t}-\bm{d}\right),\mathbf{0}\right\}
Algorithm 1 OPD Algorithm for ILP

In Algorithm 1, we denote 𝑨~t=(𝒂~t1,,𝒂~tm)\tilde{\bm{A}}_{t}=(\tilde{\bm{a}}_{t1}^{\top},\dots,\tilde{\bm{a}}_{tm}^{\top})^{\top}, 𝒃=(b1,,bm)\bm{b}=(b_{1},\dots,b_{m})^{\top} and 𝒙t=(xt1,,xtk)\bm{x}_{t}=(x_{t1},\dots,x_{tk})^{\top}. Algorithm 1 is a dual-based algorithm which maintains a dual vector 𝒑t\bm{p}_{t}. In each round tt, 𝒄t\bm{c}_{t} and 𝑨~t\tilde{\bm{A}}_{t} are revealed. Then, 𝒙t\bm{x}_{t} is determined immediately by choosing ll that maximizes (𝒄t𝒑t𝑨~t)𝒆l(\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\tilde{\bm{A}}_{t})\bm{e}_{l}, where 𝒆lk\bm{e}_{l}\in\mathbb{R}^{k} and 𝒆l\bm{e}_{l} is the unit vector with ll-th coordinate being 11. After determining 𝒙t\bm{x}_{t}, 𝒑t\bm{p}_{t} is updated by a projected stochastic subgradient descent method where (𝒅𝑨~t𝒙t)(\bm{d}-\tilde{\bm{A}}_{t}\bm{x}_{t}) is the subgradient corresponding to 𝒑t\bm{p}_{t}. Li et al. [20] have shown that Algorithm 1 provides a near-optimal solution of the problem (12) that is the linear relaxation of (11), as stated in Theorem 1.

max𝒙\displaystyle\max_{\bm{x}} t=1n𝒄t𝒙t\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t} (12)
s.t.\displaystyle\rm{s.t.} t=1n(𝒂¯tj+ψjn𝜸tj)𝒙tbj,j=1,,m\displaystyle\sum_{t=1}^{n}\left(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}^{\top}_{tj}\right)\bm{x}_{t}\leq b_{j},\forall j=1,\dots,m
𝟏𝒙t1,𝒙t𝟎,t=1,,n.\displaystyle\boldsymbol{1}^{\top}\boldsymbol{x}_{t}\leq 1,\bm{x}_{t}\geq\bm{0},\forall t=1,\dots,n.
Theorem 1 (Theorem 3 in [20]).

Under Assumption 1, the upper bound of the expected optimality gap (13) and the expected constraint violation (14) of Algorithm 1 compared to the optimal solution of the LP problem (12) are on the order of n\sqrt{n}, i.e.,

𝔼ξtj𝒫[R^nLPt=1n𝒄t𝒙t]O(n){\mathbb{E}_{\xi_{tj}\sim\mathcal{P}}}\left[\hat{R}_{n}^{LP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right]\leq O(\sqrt{n}) (13)
𝔼ξtj𝒫[(t=1n𝑨~t𝒙t𝒃)+2]O(n){\mathbb{E}_{\xi_{tj}\sim\mathcal{P}}}\left[\left\|\left(\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}-\bm{b}\right)^{+}\right\|_{2}\right]\leq O(\sqrt{n}) (14)

where R^nLP\hat{R}_{n}^{LP} is the optimal objective value of the LP problem (12) and 𝐱t\bm{x}_{t} is the output of Algorithm 1 and ()+(\cdot)^{+} is the positive part function. ξtj\xi_{tj} denotes the coefficient set {ctj,𝐚¯tj,𝐊tj}\{c_{tj},\bm{\bar{a}}_{tj},\bm{K}_{tj}\} and 𝒫\mathcal{P} is any distribution that satisfies Assumption 1 (b).

Theorem 1 states the upper bound of the expected optimality gap and the expected constraint violation, which are both O(n)O(\sqrt{n}). Then, we provide the lower bound of the expected optimality gap in the following theorem.

Theorem 2.

Under Assumption 1, the lower bound of the expected optimality gap of Algorithm 1 compared to the optimal solution of the LP problem (12) is on the order of n\sqrt{n}, i.e.,

𝔼ξtj𝒫[R^nLPt=1n𝒄t𝒙t]O(n){\mathbb{E}}_{\xi_{tj}\sim\mathcal{P}}\left[\hat{R}_{n}^{LP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right]\geq-O(\sqrt{n}) (15)

where R^nLP\hat{R}_{n}^{LP} is the optimal objective value of the LP problem (12) and 𝐱t\bm{x}_{t} is the output of Algorithm 1.

Proof.

See B. ∎

Next, we will analysis the performance of Algorithm 1 compared to the optimal solution of the ISOCP problem (10). The linear relaxation of the ISOCP problem (10) is

max𝒙\displaystyle\max_{\boldsymbol{x}} t=1n𝒄t𝒙t\displaystyle\sum_{t=1}^{n}\boldsymbol{c}_{t}^{\top}\boldsymbol{x}_{t} (16)
s.t.\displaystyle\rm{s.t.} t=1n𝒂¯tj𝒙t+ψjt=1n𝒙t𝑲tj𝒙tbj,j=1,,m\displaystyle\sum_{t=1}^{n}\bar{\boldsymbol{a}}_{tj}^{\top}\boldsymbol{x}_{t}+\psi_{j}\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}}\leq b_{j},\forall j=1,\dots,m
𝟏𝒙t1,𝒙t𝟎,t=1,,n\displaystyle\boldsymbol{1}^{\top}\boldsymbol{x}_{t}\leq 1,\bm{x}_{t}\geq\bm{0},\forall t=1,\dots,n

where the binary variables are relaxed into continuous variables on [0,1][0,1]. The optimal objective values of the ISOCP problem (10) and SOCP problem (16) are referred to as R^nISOCP\hat{R}_{n}^{ISOCP} and R^nSOCP\hat{R}_{n}^{SOCP}, respectively. Then, we establish the optimality gap between the LP problem (12) and SOCP problem (16), and the constraint violation in the following lemma.

Lemma 1.

Under Assumption 1, we have the following results on the SOCP problem (16) and LP problem (12).

  1. (a)

    If the decision variables (𝒙1,,𝒙n)(\bm{x}_{1}^{\top},\dots,\bm{x}_{n}^{\top})^{\top} of the SOCP problem (16) is set to the optimal solution of the LP problem (12), the optimality gap of (16) satisfies

    O(n)R^nSOCPt=1n𝒄t𝒙^tLP0-O(\sqrt{n})\leq\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}\leq 0 (17)

    where R^nSOCP\hat{R}_{n}^{SOCP} is the optimal objective value of (16) and 𝒙^tLP\hat{\bm{x}}_{t}^{LP} is the optimal solution of (12)).

  2. (b)

    For any 𝒙=(𝒙1,,𝒙n)\bm{x}=(\bm{x}_{1}^{\top},\dots,\bm{x}_{n}^{\top})^{\top} satisfying 𝒙t{𝒙k|𝟏𝒙𝟏,𝒙𝟎},t=1,,n\bm{x}_{t}\in\{\bm{x}\in\mathbb{R}^{k}|\bm{1}^{\top}\bm{x}\leq\bm{1},\bm{x}\geq\bm{0}\},\forall t=1,\dots,n, the gap between the resource consumption of (16) and (12) satisfies

    𝒈(𝒙)t=1n𝑨~t𝒙t2O(n)\left\|\bm{g}\left({\bm{x}}\right)-\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}\right\|_{2}\leq O(\sqrt{n}) (18)

    where gj(𝒙)=t=1n𝒂¯tj𝒙t+ψjt=1n𝒙t𝑲tj𝒙tg_{j}(\bm{x})=\sum_{t=1}^{n}\bar{\boldsymbol{a}}_{tj}^{\top}\bm{x}_{t}+\psi_{j}\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}} is the resource consumption of (16).

Proof.

See C. ∎

As shown in Lemma 1, both the optimal gap and the constraint violation between the LP problem (12) and SOCP problem (16) are O(n)O(\sqrt{n}). Then, putting Theorem 1, Theorem 2, and Lemma 1 together, we can obtain that the expected optimality gap and constraint violation compared to the optimal solution of the SOCP problem (16) in Theorem 3.

Theorem 3.

Under Assumption 1, the expected optimality gap and constraint violation of Algorithm 1 compared to the optimal solution of the SOCP problem (16) are on the order of n\sqrt{n}, i.e.,

O(n)𝔼ξtj𝒫[R^nSOCPt=1n𝒄t𝒙t]O(n)-O(\sqrt{n})\leq{\mathbb{E}_{\xi_{tj}\sim\mathcal{P}}}\left[\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right]\leq O(\sqrt{n}) (19)
𝔼ξtj𝒫[(𝒈(𝒙)𝒃)+2]O(n){\mathbb{E}_{\xi_{tj}\sim\mathcal{P}}}\left[\left\|\left(\bm{g}\left({{\bm{x}}}\right)-\bm{b}\right)^{+}\right\|_{2}\right]\leq O(\sqrt{n}) (20)

where R^nSOCP\hat{R}_{n}^{SOCP} is the optimal objective value of (16), 𝐱t\bm{x}_{t} is the output of Algorithm 1 and g(𝐱)g(\bm{x}) is the left-hand side of the SOC constraints.

Proof.

See D. ∎

Moreover, inequality R^nISOCPR^nSOCP\hat{R}_{n}^{ISOCP}\leq\hat{R}_{n}^{SOCP} holds due to the fact that the SOCP problem (16) is the linear relaxation of the ISOCP problem (10). Thus, Algorithm 1 achieves O(n)O(\sqrt{n}) regret and constraint violation compared to the optimal solution of the ISOCP problem (10) as stated in Theorem 4.

Theorem 4.

Under Assumption 1, the regret and constraint violation of Algorithm 1 compared to the optimal solution of the ISOCP problem (10) are on the order of n\sqrt{n}, i.e.,

𝔼ξtj𝒫[R^nISOCPt=1n𝒄t𝒙t]O(n){\mathbb{E}_{\xi_{tj}\sim\mathcal{P}}}\left[\hat{R}_{n}^{ISOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right]\leq O(\sqrt{n})\vspace{-0.1cm} (21)
𝔼ξtj𝒫[(𝒈(𝒙)𝒃)+2]O(n){\mathbb{E}_{\xi_{tj}\sim\mathcal{P}}}\left[\left\|\left(\bm{g}\left({{\bm{x}}}\right)-\bm{b}\right)^{+}\right\|_{2}\right]\leq O(\sqrt{n})\vspace{-0.1cm} (22)

where R^nISOCP\hat{R}_{n}^{ISOCP} is the optimal objective value of (10), 𝐱t\bm{x}_{t} is the output of Algorithm 1 and g(𝐱)g(\bm{x}) is the left-hand side of the SOC constraints.

3.2 Modified OPD Algorithms for Online CCP

Although Algorithm 1 has been able to obtain a near-optimal solution of the ISOCP problem (10) according to Theorem 4, its practical performance can be further improved by narrowing the gap between the solutions generated by Algorithm 1 and the offline ISOCP (10). To be specific, this gap mainly comes from the following two points:

  1. (a)

    The error between the offline ILP problem (11) and the offline ISOCP problem (10).

  2. (b)

    The error between the online solution and offline solution of the ILP problem (11).

Input: 𝒅=𝒃/n\bm{d}=\bm{b}/n
Output: 𝒙=(𝒙1,,𝒙n)\bm{x}=(\bm{x}_{1},...,\bm{x}_{n})
1 Initialize: 𝒑1=𝟎\bm{p}_{1}=\bm{0}
2 for t=1,,nt=1,...,n do
3       Compute 𝜷t\bm{\beta}_{t} via equation (23)
4       Set vt=maxl=1,,k(𝒄t𝒑t𝑨^t(𝜷t))𝒆lv_{t}=\max_{l=1,\dots,k}\ (\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\hat{\bm{A}}_{t}(\bm{\beta}_{t}))\bm{e}_{l}
5       if vt>0v_{t}>0 then
6             Pick an index ltl_{t} randomly from
{l:vt=(𝒄t𝒑t𝑨^t(𝜷t))𝒆l}\left\{l:v_{t}=(\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\hat{\bm{A}}_{t}(\bm{\beta}_{t}))\bm{e}_{l}\right\}
7             Set 𝒙t=𝒆lt\bm{x}_{t}=\bm{e}_{l_{t}}
8      else
9            Set 𝒙t=𝟎\bm{x}_{t}=\bm{0}
10      Compute
𝒑t+1=max{𝒑t+1n(𝑨^t(𝜷t)𝒙t𝒅),𝟎}\ \bm{p}_{t+1}=\max\left\{\bm{p}_{t}+\frac{1}{\sqrt{n}}\left(\hat{\bm{A}}_{t}(\bm{\beta}_{t})\bm{x}_{t}-\bm{d}\right),\mathbf{0}\right\}
Algorithm 2 Modified OPD Algorithm for CCP

To address these issues, we propose two modified OPD algorithms (Algorithms 2 and 3) to solve the online CCP problem (10). In Algorithm 2, a heuristic correction is applied to correct the error (a). For the jj-th constraint, we introduce scale factors

βtj={1,t=1 or i=1t1𝜸ij𝒙i=0t1i=1t1𝒙i𝑲ij𝒙ii=1t1𝜸ij𝒙i,t=2,,n and i=1t1𝜸ij𝒙i>0\beta_{tj}=\begin{cases}1,&t=1\text{ or }\sum_{i=1}^{t-1}\bm{\gamma}_{ij}^{\top}\bm{x}_{i}=0\\ \frac{\sqrt{t-1}\sqrt{\sum_{i=1}^{t-1}\bm{x}_{i}^{\top}\bm{K}_{ij}\bm{x}_{i}}}{\sum_{i=1}^{t-1}\bm{\gamma}_{ij}^{\top}\bm{x}_{i}},&t=2,\dots,n\text{ and }\sum_{i=1}^{t-1}\bm{\gamma}_{ij}^{\top}\bm{x}_{i}>0\end{cases}\vspace{-0.15cm} (23)

to reduce the gap between t=1n𝒙t𝑲tj𝒙t\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}} and t=1n𝜸tj𝒙t\sum_{t=1}^{n}{\bm{\gamma}_{tj}^{\top}\bm{x}_{t}} /n/{\sqrt{n}}. Intuitively speaking, recalling that in section 2.4 we utilize Cauchy-Schwarz inequality to linearize the non-decomposable problem (10), equation (23) can be regarded as the empirical correction for the non-decomposable term (t1)i=1t1𝒙i𝑲ij𝒙i\sqrt{(t-1)\sum_{i=1}^{t-1}\bm{x}_{i}^{\top}\bm{K}_{ij}\bm{x}_{i}} with the linear term i=1t1γij𝒙i\sum_{i=1}^{t-1}\gamma_{ij}^{\top}\bm{x}_{i}.

Define

𝒂^tj(βtj)=𝒂¯tj+βtjψjn𝜸tj,\hat{\bm{a}}_{tj}(\beta_{tj})=\bar{\bm{a}}_{tj}+\beta_{tj}\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}, (24)
𝑨^t(𝜷𝒕)=(𝒂^t1(βt1),,𝒂^tm(βtm)).\hat{\bm{A}}_{t}(\bm{\beta_{t}})=(\hat{\bm{a}}_{t1}^{\top}(\beta_{t1}),\dots,\hat{\bm{a}}_{tm}^{\top}(\beta_{tm}))^{\top}. (25)

As shown in Algorithm 2, 𝑨^t(𝜷𝒕)\hat{\bm{A}}_{t}(\bm{\beta_{t}}) is used in place of 𝑨~t\tilde{\bm{A}}_{t}. That is, we use the expression t=1nβtj𝜸tj𝒙t/n\sum_{t=1}^{n}\beta_{tj}\bm{\gamma}_{tj}^{\top}\bm{x}_{t}/\sqrt{n} to approximate t=1n𝒙t𝑲tj𝒙t\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}. In round tt, 𝜷t\bm{\beta}_{t} is calculated according to (23) which is based on the historical decisions and will be used in the next iteration for correction. It is worth noting that 𝜷t=(βt1,,βtm)\bm{\beta}_{t}=(\beta_{t1},\dots,\beta_{tm})^{\top} is calculated in each round and can be computed incrementally with low computational cost.

In the numerical experiments section 4, it is illustrated that Algorithm 2 has better performance than Algorithm 1 in terms of the constraint violation. An intuitive explanation is that Algorithm 2 is more inclined to reject the orders with high uncertainty of resource consumption (i.e., 𝑲tj\bm{K}_{tj}) than Algorithm 1 because 𝜷t𝟏\bm{\beta}_{t}\geq\bm{1}.

Next, we will propose another algorithm to correct the error (b). The error (b) consists of two parts, the optimality gap and constraint violation. In most cases, compared with the optimality gap, the CCP problems have a lower tolerance for the constraint violation, since the constraint violation will cause the probability

𝒂tj𝒫a(t=1n𝒂tj𝒙tbj)\mathbb{P}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left(\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}\leq b_{j}\right) (26)

or the conditional expectations

𝔼𝒂tj𝒫a[t=1n𝒂tj𝒙tbjt=1n𝒙t𝑲tj𝒙t|t=1n𝒂tj𝒙tbj>0]\mathbb{E}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left[\frac{\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}}{\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}}\bigg{|}\sum_{t=1}^{n}\bm{a}_{tj}^{\top}\bm{x}_{t}-b_{j}>0\right] (27)

to deviate from their set values. For instance, if the confidence level of a chance constraint is set to 90% but the actual confidence level of the online solution is 60%, this online solution is unacceptable.

For the online LP problem, Li et al. [20] have proposed a variant of the OPD algorithm that can reduce the amount of the constraint violation in practice. It is a non-stationary algorithm in which the resource consumption is considered while doing the subgradient descent. Specifically, instead of using the static average consumption 𝒅=𝒃/n\bm{d}=\bm{b}/n, the time-varying 𝒅t\bm{d}_{t} calculated by the formula (28) is utilized in Line 9 of Algorithm 1.

𝒅t=1nt(𝒃i=1t𝑨~t𝒙t).\bm{d}_{t}=\frac{1}{n-t}\left(\bm{b}-\sum_{i=1}^{t}\tilde{\bm{A}}_{t}\bm{x}_{t}\right).\vspace{-0.1cm} (28)
Input: 𝒅=𝒃/n\bm{d}=\bm{b}/n
Output: 𝒙=(𝒙1,,𝒙n)\bm{x}=(\bm{x}_{1},...,\bm{x}_{n})
1 Initialize: 𝒑1=𝟎\bm{p}_{1}=\bm{0}, 𝒅1=𝒅\bm{d}_{1}=\bm{d}
2 for t=1,,nt=1,...,n do
3       Compute 𝜷t\bm{\beta}_{t} via equation (23)
4       Set vt=maxl=1,,k(𝒄t𝒑t𝑨^t(𝜷t))𝒆lv_{t}=\max_{l=1,\dots,k}\ (\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\hat{\bm{A}}_{t}(\bm{\beta}_{t}))\bm{e}_{l}
5       if vt>0v_{t}>0 then
6             Pick an index ltl_{t} randomly from
{l:vt=(𝒄t𝒑t𝑨^t(𝜷t))𝒆l}\left\{l:v_{t}=(\bm{c}_{t}^{\top}-\bm{p}_{t}^{\top}\hat{\bm{A}}_{t}(\bm{\beta}_{t}))\bm{e}_{l}\right\}
7             Set 𝒙t=𝒆lt\bm{x}_{t}=\bm{e}_{l_{t}}
8      else
9            Set 𝒙t=𝟎\bm{x}_{t}=\bm{0}
10      Compute 𝒅t\bm{d}_{t} via equation (29)
11       Compute
𝒑t+1=max{𝒑t+1n(𝑨^t(𝜷t)𝒙t𝒅t),𝟎}\ \bm{p}_{t+1}=\max\left\{\bm{p}_{t}+\frac{1}{\sqrt{n}}\left(\hat{\bm{A}}_{t}(\bm{\beta}_{t})\bm{x}_{t}-\bm{d}_{t}\right),\mathbf{0}\right\}
Algorithm 3 Modified Non-stationary OPD Algorithm for CCP

Inspired by the above non-stationary method, we propose the following formula to dynamically adjust the right-hand-side capacity 𝒅\bm{d} in each round:

dtj=\displaystyle d_{tj}= 1nt(bji=1t𝒂¯ij𝒙iψjtni=1t𝒙i𝑲ij𝒙i),j=1,,m,\displaystyle\frac{1}{n-t}\bigg{(}b_{j}-\sum_{i=1}^{t}\bar{\bm{a}}_{ij}^{\top}\bm{x}_{i}-\psi_{j}\sqrt{\frac{t}{n}\sum_{i=1}^{t}\bm{x}_{i}^{\top}\bm{K}_{ij}\bm{x}_{i}}\bigg{)},\forall j=1,\dots,m, (29)

where

i=1t𝒂¯ij𝒙i+ψjtni=1t𝒙i𝑲ij𝒙i=tn(nti=1t𝒂¯ij𝒙i+ψjnti=1t𝒙i𝑲ij𝒙i)\vspace{-0.1cm}\sum_{i=1}^{t}\bar{\bm{a}}_{ij}^{\top}\bm{x}_{i}+\psi_{j}\sqrt{\frac{t}{n}\sum_{i=1}^{t}\bm{x}_{i}^{\top}\bm{K}_{ij}\bm{x}_{i}}=\frac{t}{n}\left(\frac{n}{t}\sum_{i=1}^{t}\bar{\bm{a}}_{ij}^{\top}\bm{x}_{i}+\psi_{j}\sqrt{\frac{n}{t}\sum_{i=1}^{t}\bm{x}_{i}^{\top}\bm{K}_{ij}\bm{x}_{i}}\right)

is an estimation of the resource assumption at period tt.

Then, we obtain Algorithm 3 by merging the correction formula (29) into Algorithm 2. The intuition behind the correction formula (29) is given as follows: if too many resources are spent in the early rounds, the remaining resources will diminish. Then Algorithm 3 will raise the dual price and be more likely to reject an order with high resource consumption as a result. On the other hand, if a large number of orders with high resource consumption are rejected at the start, resulting in an excess of remaining resources, Algorithm 3 will decrease the dual price in order to accept more orders in the future. This correction strategy makes Algorithm 3 perform better than Algorithms 1 and 2 in numerical experiments.

4 Numerical Experiments

In this section, we present extensive numerical experiments to verify our algorithms. Several metrics are used to measure the performance of algorithms. In one trial, the calculation formula of these metrics are given as follows. {𝒙1,𝒙2,,𝒙n}\{\bm{x}_{1},\bm{x}_{2},\dots,\bm{x}_{n}\} is the output of algorithms.

1) Probability deviation:

1mj=1m(ηjΦ(bjt=1n𝒂¯tj𝒙tt=1n𝒙t𝑲tj𝒙t))+\vspace{-0.2cm}\frac{1}{m}\sum_{j=1}^{m}\bigg{(}\eta_{j}-\Phi\bigg{(}\frac{b_{j}-\sum_{t=1}^{n}\bar{\boldsymbol{a}}_{tj}^{\top}\bm{x}_{t}}{\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}}}\bigg{)}\bigg{)}^{+}\vspace{-0.1cm} (30)

where 𝒙t\bm{x}_{t} is the output of the algorithms.

2) Optimality gap:

R^nSOCPt=1n𝒄t𝒙t.\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}.\vspace{-0.1cm} (31)

The exact optimality gap should be calculated with R^nISOCP\hat{R}_{n}^{ISOCP}. For ease of calculation, the offline optimal objective value R^nISOCP\hat{R}_{n}^{ISOCP} is approximated by R^nSOCP\hat{R}_{n}^{SOCP}. Noting that R^nISOCPR^nSOCP\hat{R}_{n}^{ISOCP}\leq\hat{R}_{n}^{SOCP}, the real optimality gap is upper bounded by the value of (31).

3) Competitive ratio:

t=1n𝒄t𝒙t/R^nSOCP×100%.\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}/\hat{R}_{n}^{SOCP}\times 100\%. (32)

4) Normalized constraint violation of conditional expectation constraints:

(𝒗~(𝒙))+2\vspace{-0.1cm}\big{|}\big{|}(\tilde{\bm{v}}(\bm{x}))^{+}\big{|}\big{|}_{2}\vspace{-0.1cm} (33)

where

v~j(𝒙)=𝔼𝒂tj𝒫a[t=1n𝒂tj𝒙tbjt=1n𝒙t𝑲tj𝒙t|t=1n𝒂tj𝒙t>bj]γ~j.\vspace{-0.2cm}\tilde{v}_{j}(\bm{x})=\mathbb{E}_{\bm{a}_{tj}\sim\mathcal{P}_{a}}\left[\frac{\sum_{t=1}^{n}\boldsymbol{a}_{tj}^{\top}\boldsymbol{x}_{t}-b_{j}}{\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}}}\bigg{|}\sum_{t=1}^{n}\boldsymbol{a}_{tj}^{\top}\boldsymbol{x}_{t}>b_{j}\right]-\tilde{\gamma}_{j}. (34)

5) Constraint violation of conditional expectation constraints:

(𝒗(𝒙))+2\vspace{-0.1cm}\big{|}\big{|}(\bm{v}(\bm{x}))^{+}\big{|}\big{|}_{2} (35)

where vj(𝒙)=v~j(𝒙)t=1n𝒙t𝑲tj𝒙t.v_{j}(\bm{x})=\tilde{v}_{j}(\bm{x})\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}}.

4.1 CCP Problem with Chance Constraints (2) Only

In the first subsection, we apply the proposed algorithms to the classic CCP problem which does not contain the conditional expectation constraints (3). Algorithm 1, Algorithm 2, Algorithm 3 and Algorithm 3 without correction (23) are compared in terms of optimality gap and probability deviation. These algorithms are implemented on two different models. Table 1 lists the distributions from which the elements in 𝒄tj\bm{c}_{tj}, 𝒂¯tj\bm{\bar{a}}_{tj} or 𝑲tj\bm{K}_{tj} are i.i.d. sampled. f(χ2(v))f(\chi^{2}(v)) denotes X=f(Y)X=f(Y) and Yχ2(v)Y\sim~{}\chi^{2}(v). The uniform distribution is bounded while the chi-square distribution is unbounded. In other words, the CCP problems in Experiment I satisfy Assumption 1 but the CCP problems in Experiment II do not.

Table 1: models used in the experiments.
Experiment 𝒄tj\bm{c}_{tj} 𝒂¯tj\bm{\bar{a}}_{tj} 𝑲tj\bm{K}_{tj} 𝒅\bm{d}
I U[0, 1] U[0, 4] (U[0,1])2(\text{U}[0,1])^{2} 1
II χ2(3)\chi^{2}(3) 23χ2(4)\frac{2}{3}\chi^{2}(4) (23χ2(2))2(\frac{2}{3}\chi^{2}(2))^{2} 1

4.1.1 Bounded Setting

In Experiment I, we set k=5k=5 and m=4m=4. The confidence levels of chance constraints are set to (0.65, 0.75, 0.85, 0.95). For each value of nn, we run 20 simulation trials. In each trial, coefficients 𝒄tj\bm{c}_{tj}, 𝒂¯tj\bm{\bar{a}}_{tj} and 𝑲tj\bm{K}_{tj} are resampled. All metrics are averaged over all trials.

Refer to caption

(a) optimality gap

Refer to caption

(b) probability deviation

Figure 1: optimality gap and probability deviation with uniform i.i.d. input.

Refer to caption

(a) Algorithm 1

Refer to caption

(b) Algorithm 2

Refer to caption

(c) Algorithm 3 without correction (23)

Refer to caption

(d) Algorithm 3

Figure 2: probability deviation of each chance constraint with uniform i.i.d. input.

Figure 1 shows the average optimality gap and probability deviation over all the simulation trials. From Figure 1, we observe that the optimality gaps of these algorithms are very close, Algorithm 3 has the smallest probability deviation and Algorithm 1 has the largest probability deviation when n5000n\geq 5000. For all nn, Algorithm 3 has all probability deviations less than 1%. Figure 1 (a) also shows that the optimality gaps of these four algorithms are on the order of n\sqrt{n} and verifies Theorem 3. Figure 2 presents the probability deviations of each chance constraint of these four algorithms. For each chance constraint, the probability deviation of Algorithm 3 is the smallest among the four algorithms when n5000n\geq 5000. Figure 1 (b) and Figure 2 both illustrate that the proposed two corrections (23) and (29) can effectively reduce the probability deviation with minor negative effects on the optimality gap.

4.1.2 Unbounded Setting

In Experiment II, kk and mm are still set to 5 and 4. The confidence levels are also the same as those in Experiment I. For each value of nn, we run 20 simulation trials. In each trial, coefficients 𝒄tj\bm{c}_{tj}, 𝒂¯tj\bm{\bar{a}}_{tj} and 𝑲tj\bm{K}_{tj} are i.i.d. sampled from the chi-square distributions which are unbounded.

Refer to caption

(a) optimality gap

Refer to caption

(b) probability deviation

Figure 3: optimality gap and probability deviation with chi-square i.i.d. input.

Figure 3 shows the average optimality gap and probability deviation, and Figure 4 shows the detailed probability deviations of each chance constraint. The results of Experiment II are similar to those of Experiment I: Algorithm 3 has the smallest probability deviation, whereas Algorithm 1 has the largest probability deviation. Experiment II again verifies that corrections (23) and (29) can effectively reduce the probability deviation. Despite the fact that Algorithm 3 produces slightly larger optimality gap, its optimality gap is still approximately on the order of n\sqrt{n}. In this experiment with unbounded input, Algorithm 3 significantly reduces the probability deviations: the probability deviations of the algorithms except Algorithm 3 are larger than 10%, while the probability deviation of Algorithm 3 is less than 1%.

Refer to caption

(a) Algorithm 1

Refer to caption

(b) Algorithm 2

Refer to caption

(c) Algorithm 3 without correction (23)

Refer to caption

(d) Algorithm 3

Figure 4: probability deviation of each chance constraint with chi-square i.i.d. input.
Table 2: competitive ratios of Algorithm 3 in the experiments.
Experiment Number of requests nn
2500 5000 7500 10000 12500 15000
I 95.90% 97.13% 97.69% 97.94% 98.14% 98.38%
II 98.67% 99.09% 99.27% 99.32% 99.42% 99.48%

The competitive ratios of Algorithm 2 in Experiment I and II are larger than 95%, with details provided in Table 2.

4.2 CCP Problem with Conditional Expectation Constraints (3) Only

In this subsection, we apply the proposed algorithms to the CCP problem with only conditional expectation constraints and compare the performance in terms of the normalized constraint violation and the constraint violation. The input data are the same as these in Section 4.1. The parameters kk and mm are also set to 5 and 4, and γ~j\tilde{\gamma}_{j} in four conditional expectation constraints are set to 0.2, 0.3, 0.4 and 0.5, respectively. The scale of γ~j\tilde{\gamma}_{j} matches the standard normal distribution: 0.2, 0.3, 0.4 and 0.5 correspond to 0.2, 0.3, 0.4 and 0.5 times the standard deviation, respectively.

Refer to caption

(a) normalized constraint violation

Refer to caption

(b) constraint violation

Figure 5: normalized constraint violation and constraint violation with uniform i.i.d. input.

Refer to caption

(a) normalized constraint violation

Refer to caption

(b) constraint violation

Figure 6: normalized constraint violation and constraint violation with chi-square i.i.d. input.

Figure 6 and Figure 6 show the normalized constraint violation and constraint violation of four algorithms in the cases with uniform and chi-square inputs. In both scenarios, Algorithm 3 has the smallest normalized constraint violation and constraint violation while Algorithm 1 has the largest violations. It is verified that the proposed corrections can effectively reduce the violation of conditional expectation constraints. Moreover, these experimental results show that the constraint violation of four algorithms in both cases is on the order of n\sqrt{n}. This order is the same as stated in Theorem 3. Note that Theorem 3 only analyzes the performance of Algorithm 1 and its analysis relies on the boundedness assumption of the input. The theoretical analysis of conditional expectation constraints is still an open question.

5 Conclusion

In this paper, we study the online stochastic RAP with chance constraints and conditional expectation constraints under mild assumptions. First, we present a linearization method that decouples the non-linear term in second-order cone constraints and makes the online solution possible. Next, we theoretically analyze the performance of the vanilla OPD algorithm compared with the offline solution of the stochastic RAP. Assuming the input data is bounded and i.i.d. sampled, the expected optimality gap and constraint violation are both O(n)O(\sqrt{n}). After that, we propose modified OPD algorithms with several heuristic corrections to reduce the probability derivation and constraint violation. Numerical results verify the effectiveness of proposed heuristics and reveal that the probability derivation and constraint violation of the proposed Algorithm 3 is very small. Additionally, numerical results confirm that Algorithm 3 can output a near-optimal solution with small regret. In conclusion, the proposed Algorithm 3 is suitable for solving online CCP and can be applied to engineering applications such as online order fulfillment.

References

  • [1] Shipra Agrawal, Zizhuo Wang, and Yinyu Ye. A dynamic near-optimal algorithm for online linear programming. Operations Research, 62(4):876–890, 2014.
  • [2] Arash Asadpour, Xuan Wang, and Jiawei Zhang. Online resource allocation with limited flexibility. Management Science, 66(2):642–666, 2020.
  • [3] Santiago Balseiro, Haihao Lu, and Vahab Mirrokni. The best of many worlds: Dual mirror descent for online allocation problems. arXiv preprint arXiv:2011.10124, 2020.
  • [4] Russell Bent and Pascal Van Hentenryck. Online stochastic optimization without distributions. In Proceedings of the Fifteenth International Conference on International Conference on Automated Planning and Scheduling, pages 171–180. AAAI Press, 2005.
  • [5] Russell Bent and Pascal Van Hentenryck. Online stochastic and robust optimization. In Annual Asian Computing Science Conference, pages 286–300. Springer, 2004.
  • [6] Stephen Boyd and Lieven Vandenberghe. Convex Optimization, pages 237–238. Cambridge University Press, 2004.
  • [7] Niv Buchbinder and Joseph Naor. Online primal-dual algorithms for covering and packing. Mathematics of Operations Research, 34(2):270–286, 2009.
  • [8] Abraham Charnes and William W Cooper. Chance-constrained programming. Management Science, 6(1):73–79, 1959.
  • [9] Abraham Charnes and William W Cooper. Deterministic equivalents for optimizing and satisficing under chance constraints. Operations Research, 11(1):18–39, 1963.
  • [10] Xiao Alison Chen and Zizhuo Wang. A dynamic learning algorithm for online matching problems with concave returns. European Journal of Operational Research, 247(2):379–388, 2015.
  • [11] William W Cooper, Honghui Deng, Zhimin Huang, and Susan X Li. Chance constrained programming approaches to congestion in stochastic data envelopment analysis. European Journal of Operational Research, 155(2):487–501, 2004.
  • [12] Wenzhi Gao, Chunlin Sun, Yuyang Ye, and Yinyu Ye. Boosting method in approximately solving linear programming with fast online algorithm. arXiv preprint arXiv:2107.03570, 2021.
  • [13] Anupam Gupta and Marco Molinaro. How experts can solve lps online. In European Symposium on Algorithms, pages 517–529. Springer, 2014.
  • [14] Masaaki Ida. Portfolio selection problem with interval coefficients. Applied Mathematics Letters, 16(5):709–713, 2003.
  • [15] Raj Jagannathan. Chance-constrained programming with joint constraints. Operations Research, 22(2):358–372, 1974.
  • [16] Jiashuo Jiang, Xiaocheng Li, and Jiawei Zhang. Online stochastic optimization with wasserstein based non-stationarity. arXiv preprint arXiv:2012.06961, 2020.
  • [17] Jiashuo Jiang and Jiawei Zhang. Online resource allocation with stochastic resource consumption. arXiv preprint arXiv:2012.07933, 2020.
  • [18] Thomas Kesselheim, Andreas Tönnis, Klaus Radke, and Berthold Vöcking. Primal beats dual on online packing lps in the random-order model. In Proceedings of the forty-sixth annual ACM Symposium on Theory of Computing, pages 303–312, 2014.
  • [19] James F. Kurose and Rahul Simha. A microeconomic approach to optimal resource allocation in distributed computer systems. IEEE Transactions on Computers, 38(5):705–717, 1989.
  • [20] Xiaocheng Li, Chunlin Sun, and Yinyu Ye. Simple and fast algorithm for binary integer and online linear programming. In Advances in Neural Information Processing Systems, pages 9412–9421, 2020.
  • [21] Xiaocheng Li and Yinyu Ye. Online linear programming: Dual convergence, new algorithms, and regret bounds. Operations Research, 2021.
  • [22] Jialin Liu, Yuantao Gu, and Mengdi Wang. Averaging random projection: A fast online solution for large-scale constrained stochastic optimization. In IEEE ICASSP, pages 3586–3590. IEEE, 2015.
  • [23] Mengshi Lu, Hideaki Nakao, Siqian Shen, and Lin Zhao. Non-profit resource allocation and service scheduling with cross-subsidization and uncertain resource consumptions. Omega, 99:102191, 2021.
  • [24] Aranyak Mehta, Amin Saberi, Umesh Vazirani, and Vijay Vazirani. Adwords and generalized online matching. Journal of the ACM (JACM), 54(5):22–es, 2007.
  • [25] Marco Molinaro and Ramamoorthi Ravi. The geometry of online packing linear programs. Mathematics of Operations Research, 39(1):46–59, 2014.
  • [26] Andras Prekopa. Contributions to the theory of stochastic programming. Mathematical Programming, 4(1):202–221, 1973.
  • [27] Fred Ricker and Ravi Kalakota. Order fulfillment: the hidden key to e-commerce success. Supply chain management review, 11(3):60–70, 1999.
  • [28] Andrzej Ruszczyński and Alexander Shapiro. Stochastic programming models. Handbooks in Operations Research and Management Science, 10:1–64, 2003.
  • [29] Lei Xu and Arumugam Nallanathan. Energy-efficient chance-constrained resource allocation for multicast cognitive ofdm network. IEEE Journal on Selected Areas in Communications, 34(5):1298–1306, 2016.
  • [30] Yujie Zhang and Shaowei Wang. Resource allocation for femtocell networks by using chance-constrained optimization. In IEEE WCNC, pages 1805–1810. IEEE, 2015.

Appendix A Proof of Proposition 1

First, {𝒙{0,1}k|𝟏𝒙1}\{\bm{x}\in\{0,1\}^{k}|\bm{1}^{\top}\bm{x}\leq 1\} = {𝒆sk|s=1,,k}\{\bm{e}_{s}\in\mathbb{R}^{k}|s=1,\dots,k\}, where 𝒆l\bm{e}_{l} is the unit vector with ll-th coordinate being 1.

Then, for all 𝒆l{𝒆sk|s=1,,k}\bm{e}_{l}\in\{\bm{e}_{s}\in\mathbb{R}^{k}|s=1,\dots,k\}, we have

𝒆l𝑲tj𝒆l=𝑲tjll=𝜸tj𝒆l,\sqrt{\bm{e}_{l}^{\top}\bm{K}_{tj}\bm{e}_{l}}=\sqrt{\bm{K}_{tjll}}=\bm{\gamma}_{tj}^{\top}\bm{e}_{l}, (36)

where 𝑲tjll\bm{K}_{tjll} is ll-th diagonal element of the matrix 𝑲tj\bm{K}_{tj} and 𝜸tj=(𝑲tj11,,\bm{\gamma}_{tj}=(\sqrt{\bm{K}_{tj11}},\dots, ,𝑲tjkk)\dots,\sqrt{\bm{K}_{tjkk}})^{\top}.

Hence, Proposition 1 holds.

Appendix B Proof of Theorem 2

In this section, we denote the output of Algorithm 1 as 𝒙~t\bm{\tilde{x}}_{t} to distinguish it from the decision variables 𝒙t\bm{x}_{t}.

First, we present the following proposition.

Proposition 2.

Under Assumption 1, the problems (12) and (16) have strong duality.

Proof.

Setting 𝒙t=𝟎,t=1,,n\bm{x}_{t}=\bm{0},\forall t=1,\dots,n, we have

t=1n(𝒂¯tj+ψjn𝜸tj)𝒙t=0bj,j=1,,m\displaystyle\sum_{t=1}^{n}\left(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}^{\top}_{tj}\right)\bm{x}_{t}=0\leq b_{j},\forall j=1,\dots,m (37)
t=1n𝒂¯tj𝒙t+ψjt=1n𝒙t𝑲tj𝒙t=0<bj,j=1,,m\displaystyle\sum_{t=1}^{n}\bar{\boldsymbol{a}}_{tj}^{\top}\boldsymbol{x}_{t}+\psi_{j}\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}}=0<b_{j},\forall j=1,\dots,m
𝟏𝒙t1,𝒙t𝟎,t=1,,n\displaystyle\bm{1}^{\top}\bm{x}_{t}\leq 1,\bm{x}_{t}\geq\bm{0},\forall t=1,\dots,n

which means that Slater’s conditions of (12) and (16) are satisfied. Thus, strong duality holds. ∎

Proof of Theorem 2

Based on the strong duality of linear problem (12), we have

R^nLP=\displaystyle\hat{R}_{n}^{LP}= min𝒑0max𝟏𝒙1,𝒙𝟎j=1mpjbj+t=1n(𝒄tj=1mpj(𝒂¯tj+ψjn𝜸tj))𝒙t\displaystyle\min_{\bm{p}\geq 0}\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}p_{j}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top}))\bm{x}_{t}
=\displaystyle= max𝟏𝒙1,𝒙𝟎min𝒑0j=1mpjbj+t=1n(𝒄tj=1mpj(𝒂¯tj+ψjn𝜸tj))𝒙t\displaystyle\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\min_{\bm{p}\geq 0}\sum_{j=1}^{m}p_{j}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top}))\bm{x}_{t}
=\displaystyle= max𝟏𝒙1,𝒙𝟎j=1mp^jLPbj+t=1n(𝒄tj=1mp^jLP(𝒂¯tj+ψjn𝜸tj))𝒙t\displaystyle\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}\hat{p}_{j}^{LP}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}\hat{p}_{j}^{LP}(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top}))\bm{x}_{t} (38)
\displaystyle\geq j=1mp^jLPbj+t=1n(𝒄tj=1mp^jLP(𝒂¯tj+ψjn𝜸tj))𝒙~t\displaystyle\sum_{j=1}^{m}\hat{p}_{j}^{LP}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}\hat{p}_{j}^{LP}(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top}))\tilde{\bm{x}}_{t}
=\displaystyle= t=1n𝒄t𝒙~t+j=1mp^jLP(bj(𝒂¯tj+ψjn𝜸tj)𝒙~t)\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\tilde{\bm{x}}_{t}+\sum_{j=1}^{m}\hat{p}_{j}^{LP}(b_{j}-(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top})\tilde{\bm{x}}_{t})

where p^jLP\hat{p}_{j}^{LP} is the optimal dual value of (12). The second equation holds because of strong duality [6].

Next, since p^jLP\hat{p}_{j}^{LP} is O(1)O(1) [21] and

𝔼[j=1m(bj(𝒂¯tj+ψjnt=1n𝜸tj)𝒙~t)]\displaystyle\mathbb{E}\left[\sum_{j=1}^{m}\left(b_{j}-(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\sum_{t=1}^{n}\bm{\gamma}_{tj}^{\top})\tilde{\bm{x}}_{t}\right)\right] (39)
\displaystyle\geq 𝔼[(t=1n𝑨~t𝒙~t𝒃)+1]𝔼[m(t=1n𝑨~t𝒙~t𝒃)+2]\displaystyle-\mathbb{E}\left[\left\|\left(\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\tilde{\bm{x}}_{t}-\bm{b}\right)^{+}\right\|_{1}\right]\geq-\mathbb{E}\left[\sqrt{m}\left\|\left(\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\tilde{\bm{x}}_{t}-\bm{b}\right)^{+}\right\|_{2}\right]
\displaystyle\geq O(n),\displaystyle-O(\sqrt{n}),

where the last inequality comes from Theorem 1, we get

𝔼[R^nLPt=1n𝒄t𝒙~t]O(n).\mathbb{E}\left[\hat{R}_{n}^{LP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\tilde{\bm{x}}_{t}\right]\geq-O(\sqrt{n}). (40)

Appendix C Proof of Lemma 1

Consider the dual problem of the SOCP problem (16):

min𝒑0max𝟏𝒙1,𝒙𝟎j=1mpjbj+t=1n(𝒄tj=1mpj𝒂¯tj)𝒙tj=1mpjψjt=1n𝒙t𝑲tj𝒙t.\displaystyle\min_{\bm{p}\geq 0}\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}p_{j}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}-\sum_{j=1}^{m}p_{j}\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}. (41)

The optimal value of 𝒑\bm{p} has the following property.

Proposition 3.

Under Assumption 1, the optimal solution of the problem (41) is bounded:

𝒑^SOCP{𝒑m|𝒑𝟎,𝟏𝒑c¯d¯}\hat{\bm{p}}^{SOCP}\in\left\{\bm{p}\in\mathbb{R}^{m}\bigg{|}\bm{p}\geq\bm{0},\bm{1}^{\top}\bm{p}\leq\frac{\bar{c}}{\underline{d}}\right\} (42)

where 𝐩^SOCP\hat{\bm{p}}^{SOCP} is the optimal solution, c¯\bar{c} is the upper bound of ctlc_{tl} and d¯\underline{d} is the lower bound of djd_{j}, which means c¯ctl,t=1,,n,l=1,,k\bar{c}\geq c_{tl},\ \forall t=1,\dots,n,\ l=1,\dots,k and dj=bj/nd¯>0,j=1,,md_{j}=b_{j}/n\geq\underline{d}>0,\ \forall j=1,\dots,m.

Proof.

Use f(𝒑)f(\bm{p}) to denote the value of dual function

f(𝒑):=max𝟏𝒙1,𝒙𝟎j=1mpjbj+t=1n(𝒄tj=1mpj𝒂¯tj)𝒙tj=1mpjψjt=1n𝒙t𝑲tj𝒙t,f(\bm{p}):=\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}p_{j}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}-\sum_{j=1}^{m}p_{j}\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}, (43)

and use L(𝒑,𝒙)L(\bm{p},\bm{x}) to denote the value of Lagrangian function

L(𝒑,𝒙):=j=1mpjbj+t=1n(𝒄tj=1mpj𝒂¯tj)𝒙tj=1mpjψjt=1n𝒙t𝑲tj𝒙t.L(\bm{p},\bm{x}):=\sum_{j=1}^{m}p_{j}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}-\sum_{j=1}^{m}p_{j}\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}. (44)

We only need to prove 𝒅𝒑^SOCPc¯\bm{d}^{\top}\hat{\bm{p}}^{SOCP}\leq\bar{c}. If 𝒅𝒑^SOCP>c¯\bm{d}^{\top}\hat{\bm{p}}^{SOCP}>\bar{c}, then

f(𝒑^SOCP)\displaystyle f(\hat{\bm{p}}^{SOCP}) L(𝒑^SOCP,𝟎)=j=1mp^jSOCPbj\displaystyle\geq L(\hat{\bm{p}}^{SOCP},\bm{0})=\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}b_{j} (45)
=n𝒅𝒑^SOCP>nc¯\displaystyle=n\bm{d}^{\top}\hat{\bm{p}}^{SOCP}>n\bar{c}
max𝟏𝒙1,𝒙𝟎t=1n𝒄t𝒙t=f(𝟎)\displaystyle\geq\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}=f(\bm{0})

which contradicts that 𝒑^SOCP\hat{\bm{p}}^{SOCP} is the optimal solution of the dual problem (41) that is min𝒑𝟎f(𝒑)\min_{\bm{p}\geq\bm{0}}\ f(\bm{p}).

Hence, 𝒅𝒑^SOCPc¯\bm{d}^{\top}\hat{\bm{p}}^{SOCP}\leq\bar{c}. Evidently, 𝟏𝒑^SOCPc¯/d¯\bm{1}^{\top}\hat{\bm{p}}^{SOCP}\leq\bar{c}/\underline{d}. ∎

Proof of Lemma 1

Proof.

(a) For the optimality gap, R^nSOCPt=1nct𝒙^tLP\hat{R}_{n}^{SOCP}\leq\sum_{t=1}^{n}c_{t}^{\top}\hat{\bm{x}}_{t}^{LP}. Specifically,

R^nSOCP=\displaystyle\hat{R}_{n}^{SOCP}= min𝒑0max𝟏𝒙1,𝒙𝟎j=1mpjbj+\displaystyle\min_{\bm{p}\geq 0}\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}p_{j}b_{j}+
t=1n(𝒄tj=1mpj𝒂¯tj)𝒙tj=1mpjψjt=1n𝒙t𝑲tj𝒙t\displaystyle\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}-\sum_{j=1}^{m}p_{j}\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}
\displaystyle\leq min𝒑0max𝟏𝒙1,𝒙𝟎j=1mpjbj+\displaystyle\min_{\bm{p}\geq 0}\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}p_{j}b_{j}+ (46)
t=1n(𝒄tj=1mpj𝒂¯tj)𝒙tj=1mpjψjnt=1n𝜸tj𝒙t\displaystyle\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}-\sum_{j=1}^{m}p_{j}\frac{\psi_{j}}{\sqrt{n}}\sum_{t=1}^{n}\bm{\gamma}_{tj}^{\top}\bm{x}_{t}
=\displaystyle= t=1n𝒄t𝒙^tLP.\displaystyle\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}.

The first equation is from strong duality of (16) (Proposition 2).

Next is to prove R^nSOCPt=1nct𝒙^tLPO(n)\hat{R}_{n}^{SOCP}\geq\sum_{t=1}^{n}c_{t}^{\top}\hat{\bm{x}}_{t}^{LP}-O(\sqrt{n}). We have

R^nSOCP=\displaystyle\hat{R}_{n}^{SOCP}= max𝟏𝒙1,𝒙𝟎j=1mp^jSOCPbj+t=1n(𝒄tj=1mp^jSOCP𝒂¯tj)𝒙t\displaystyle\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}
j=1mp^jSOCPψjt=1n𝒙t𝑲tj𝒙t\displaystyle-\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}
\displaystyle\geq max𝟏𝒙1,𝒙𝟎j=1mp^jSOCPbj+t=1n(𝒄tj=1mp^jSOCP𝒂¯tj)𝒙t\displaystyle\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}
j=1mp^jSOCP(ψjnt=1n𝒙t𝑲tj𝒙t+ψjt=1n𝒙t𝑲tj𝒙t)\displaystyle-\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}\left(\frac{\psi_{j}}{\sqrt{n}}\sum_{t=1}^{n}\sqrt{\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}+\psi_{j}\sqrt{\sum_{t=1}^{n}\bm{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\right) (47)
\displaystyle\geq max𝟏𝒙1,𝒙𝟎j=1mp^jSOCPbj+t=1n(𝒄tj=1mp^jSOCP𝒂¯tj)𝒙t\displaystyle\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}
j=1mp^jSOCPψjnt=1n𝜸tj𝒙tc¯d¯ψ¯γ¯n\displaystyle-\sum_{j=1}^{m}\hat{p}_{j}^{SOCP}\frac{\psi_{j}}{\sqrt{n}}\sum_{t=1}^{n}\bm{\gamma}_{tj}^{\top}\bm{x}_{t}-\frac{\bar{c}}{\underline{d}}\bar{\psi}\bar{\gamma}\sqrt{n}
\displaystyle\geq max𝟏𝒙1,𝒙𝟎j=1mp^jLPbj+t=1n(𝒄tj=1mp^jLP𝒂¯tj)𝒙t\displaystyle\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}\hat{p}_{j}^{LP}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}\hat{p}_{j}^{LP}\bar{\bm{a}}_{tj}^{\top})\bm{x}_{t}
j=1mp^jLPψjnt=1n𝜸tj𝒙tc¯d¯ψ¯γ¯n\displaystyle-\sum_{j=1}^{m}\hat{p}_{j}^{LP}\frac{\psi_{j}}{\sqrt{n}}\sum_{t=1}^{n}\bm{\gamma}_{tj}^{\top}\bm{x}_{t}-\frac{\bar{c}}{\underline{d}}\bar{\psi}\bar{\gamma}\sqrt{n}
=\displaystyle= t=1nct𝒙^tLPO(n)\displaystyle\sum_{t=1}^{n}c_{t}^{\top}\hat{\bm{x}}_{t}^{LP}-O(\sqrt{n})

where γ¯=K¯\bar{\gamma}=\sqrt{\bar{K}}, K¯\bar{K} is the upper bound of the diagonal elements of 𝑲tj\bm{K}_{tj}, ψ¯=maxjψj\bar{\psi}=\max_{j}{\psi_{j}} and 𝒑^LP\hat{\bm{p}}^{LP} is the optimal solution of the dual problem of (12). In (C), the second inequality comes from Proposition 1, Proposition 3 and the boundedness of 𝑲tj\bm{K}_{tj}. The third inequality is valid because the dual problem

min𝒑𝟎max𝟏𝒙1,𝒙𝟎j=1mpjbj+t=1n(𝒄tj=1mpj(𝒂¯tj+ψjn𝜸tj))𝒙t\min_{\bm{p}\geq\bm{0}}\max_{\bm{1}^{\top}\bm{x}\leq 1,\bm{x}\geq\bm{0}}\sum_{j=1}^{m}p_{j}b_{j}+\sum_{t=1}^{n}(\bm{c}_{t}^{\top}-\sum_{j=1}^{m}p_{j}(\bar{\bm{a}}_{tj}^{\top}+\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top}))\bm{x}_{t} (48)

achieves its minimum value when 𝒑=𝒑^LP\bm{p}=\hat{\bm{p}}^{LP}.

According to (C) and (C),

|R^nSOCPt=1n𝒄t𝒙^tLP|O(n)\bigg{|}\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}\bigg{|}\leq O(\sqrt{n}) (49)

holds for all nn.

(b) For the resource consumption gap, we have

gj(𝒙)t=1n𝒂~tj𝒙t\displaystyle g_{j}(\bm{x})-\sum_{t=1}^{n}\tilde{\bm{a}}_{tj}\bm{x}_{t} =ψjt=1n𝒙t𝑲tj𝒙tt=1nψjn𝜸tj𝒙t\displaystyle=\psi_{j}\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\boldsymbol{x}_{t}}-\sum_{t=1}^{n}\frac{\psi_{j}}{\sqrt{n}}\bm{\gamma}_{tj}^{\top}\bm{x}_{t} (50)
ψjt=1n𝒙t𝑲tj𝒙tψjK¯n.\displaystyle\leq\psi_{j}\sqrt{\sum_{t=1}^{n}\boldsymbol{x}_{t}^{\top}\bm{K}_{tj}\bm{x}_{t}}\leq\psi_{j}\sqrt{\bar{K}}\sqrt{n}.

Hence,

𝒈(𝒙)t=1n𝑨~t𝒙t2O(n)\left\|\bm{g}\left({\bm{x}}\right)-\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}\right\|_{2}\leq O(\sqrt{n}) (51)

holds for all nn. ∎

Appendix D Proof of Theorem 3

Proof.

According to Lemma 1 and Theorem 1, we have

𝔼[R^nSOCPt=1n𝒄t𝒙t]\displaystyle~{}\mathbb{E}\left[\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right] (52)
=\displaystyle= 𝔼[R^nSOCPt=1n𝒄t𝒙^tLP]+𝔼[t=1n𝒄t𝒙^tLPt=1n𝒄t𝒙t]\displaystyle~{}\mathbb{E}\left[\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}\right]+\mathbb{E}\left[\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right]
\displaystyle\leq 0+O(n)=O(n).\displaystyle~{}0+O(\sqrt{n})=O(\sqrt{n}).

According to Lemma 1 and Therorem 2, we have

𝔼[R^nSOCPt=1n𝒄t𝒙t]\displaystyle~{}\mathbb{E}\left[\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right]
=\displaystyle= 𝔼[R^nSOCPt=1n𝒄t𝒙^tLP]+𝔼[t=1n𝒄t𝒙^tLPt=1n𝒄t𝒙t]\displaystyle~{}\mathbb{E}\left[\hat{R}_{n}^{SOCP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}\right]+\mathbb{E}\left[\sum_{t=1}^{n}\bm{c}_{t}^{\top}\hat{\bm{x}}_{t}^{LP}-\sum_{t=1}^{n}\bm{c}_{t}^{\top}\bm{x}_{t}\right] (53)
\displaystyle\geq O(n)O(n)=O(n).\displaystyle~{}-O(\sqrt{n})-O(\sqrt{n})=-O(\sqrt{n}).

For the constraint violation, we have

𝔼[(𝒈(𝒙)𝒃)+2]\displaystyle~{}{\mathbb{E}}\left[\left\|\left(\bm{g}\left({{\bm{x}}}\right)-\bm{b}\right)^{+}\right\|_{2}\right] (54)
\displaystyle\leq 𝔼[(𝒈(𝒙)t=1n𝑨~t𝒙t)+2]+𝔼[(t=1n𝑨~t𝒙t𝒃)+2]\displaystyle~{}{\mathbb{E}}\left[\left\|\left(\bm{g}\left({{\bm{x}}}\right)-\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}\right)^{+}\right\|_{2}\right]+{\mathbb{E}}\left[\left\|\left(\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}-\bm{b}\right)^{+}\right\|_{2}\right]
\displaystyle\leq 𝔼[𝒈(𝒙)t=1n𝑨~t𝒙t2]+𝔼[(t=1n𝑨~t𝒙t𝒃)+2]\displaystyle~{}{\mathbb{E}}\left[\left\|\bm{g}\left({{\bm{x}}}\right)-\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}\right\|_{2}\right]+{\mathbb{E}}\left[\left\|\left(\sum_{t=1}^{n}\tilde{\bm{A}}_{t}\bm{x}_{t}-\bm{b}\right)^{+}\right\|_{2}\right]
\displaystyle\leq O(n)+O(n)=O(n).\displaystyle~{}O(\sqrt{n})+O(\sqrt{n})=O(\sqrt{n}).

Thus we obtain Theorem 3. ∎