11
1 Efficient Selection of a Set of Good Enough Designs with Complexity Preference Shen Yan, Enlu Zhou, Member, IEEE, and Chun-Hung Chen, Senior Member, IEEE Abstract—Many automation or manufacturing systems are large, complex, and stochastic. Since closed-form analytical solutions generally do not exist for such systems, simulation is the only faithful way for performance evaluation. From the practical engineering perspective, the designs (or solution candidates) with low complexity (called simple designs) have many advantages compared with complex designs, such as requiring less computing and memory resources, and easier to interpret and to implement. Therefore, they are usually more desirable than complex designs in the real world if they have good enough performance. Recently, Jia [1] discussed the importance of design simplicity and introduced an adaptive simulation-based sampling algorithm to sequentially screen the designs until one simplest good enough design is found. In this paper, we consider a more generalized problem and introduce two algorithms OCBA-mSG and OCBA-bSG to identify a subset of m simplest and good enough designs among a total of K (K > m) designs. By controlling the simulation allocation intelligently, our approach intends to find those simplest good enough designs using a minimum simulation time. The numerical results show that both OCBA-mSG and OCBA-bSG outperform some other approaches on the test problems. Note to Practitioners—This paper was motivated by two problems from the real world: design of ordering policies in inventory control, and design of node activation rules in sensor networks. In designing a good ordering policy or a good sensor activation rule, simple designs have various advantages in practice, e.g., easy to learn, to implement, to operate, and to maintain. More generally, practitioners want to find a design or a set of designs which not only have good performance but are also simple. Due to the complexity of the systems, simulation is a popular tool in industry to evaluate the performance of different alternative designs before actual implementation. While the advance of new technology has dramatically increased computational power, efficiency is still a big concern. Our proposed approach intelligently controls the simulation of alternative designs, so that the simplest good enough designs can be selected using a minimum computation cost. Our numerical experiments show that the proposed approach is very effective. Index Terms—optimal computing budget allocation, ranking and selection, simulation-based optimization, complexity. Shen Yan was a Master student in the Department of Industrial & Enterprise Systems Engineering, University of Illinois at Urbana-Champaign, IL, 61801 USA ([email protected]). Enlu Zhou is with the faculty in the Department of Industrial & Enterprise Systems Engineering, University of Illinois at Urbana-Champaign, IL, 61801 USA ([email protected]). Chun-Hung Chen (corresponding author) is with the faculty in Department of Electrical Engineering & Institute of Industrial Engineering, National Taiwan University, Taipei, Taiwan ([email protected]). This work was supported by the National Science Foundation under Grants ECCS-0901543 and CMMI-1130273, the Air Force Office of Scien- tific Research under Grant FA-9550-12-1-0250, the Department of Energy under Grant DE-SC0002223, National Institutes of Health under Grant 1R21DK088368-01, and National Science Council of Taiwan under Grant NSC-100-2218-E-002-027-MY3. I. I NTRODUCTION T HE motivation for considering the selection of simplest good enough designs comes from the real world, where simple designs are preferred to the complex ones if the simple designs are good enough to satisfy our requirements. A typical example is to find an ordering policy in inventory control. Consider the problem of ordering a certain amount of products at each period to meet a stochastic demand which follows a probability distribution. In order to minimize the expected cost (including holding cost for excess inventory and shortage cost for unfilled demand), we want to determine the optimal ordering policy in each period. The optimal policy can be found analytically for some models to have the structure of a base-stock policy or an (s; S) policy [2], [3], [4], [5]. The base- stock policy is a threshold function that maps the current stock into the ordering amount. Motivated by this simple structure of the optimal policy for some models, we can approximate the optimal policies for other more complex inventory models by threshold policies. In general, we can approximate the optimal ordering policy better with a greater number of thresholds given the right values of these thresholds. Then the problem is to decide the number and the values of the thresholds. It is clear that with more thresholds in the function we have a more complex ordering policy, making it harder to determine the optimal values of these thresholds and to implement in practice. Conversely, with a small number of thresholds, such as one (base-stock policy) or two, we have a simple ordering policy, which is easier to compute and to implement. If the simple ordering policy can achieve a required cost criterion, it will be more desirable than a complex ordering policy, even though the complex policy may yield a lower cost. Another example is the design of node activation rules in the wireless sensor networks (WSNs), as described in [6], [1]. Each node needs to collaborate with its neighbors in order to have enough power to monitor an area of interest. Similarly, given that the required probability of correct detection can be achieved, we prefer small communication radius of each node (i.e., simple activation rule) to large radius (i.e., complex activation rule). In this paper, we use the word “design” to refer to the object under consideration, such as the ordering policy and the node activation rule in the previous examples. We consider the problem of selecting m (m 1) designs that are simplest (with smallest complexity) and good enough (satisfying a constraint on the performance measure). Selection of multiple designs is sometimes preferred because of at least two reasons. First, a decision maker has to face different objectives and constraints. However, in many cases it is too complicated to include all

Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

1

Efficient Selection of a Set of Good EnoughDesigns with Complexity Preference

Shen Yan, Enlu Zhou, Member, IEEE, and Chun-Hung Chen, Senior Member, IEEE

Abstract—Many automation or manufacturing systems arelarge, complex, and stochastic. Since closed-form analyticalsolutions generally do not exist for such systems, simulationis the only faithful way for performance evaluation. Fromthe practical engineering perspective, the designs (or solutioncandidates) with low complexity (called simple designs) havemany advantages compared with complex designs, such asrequiring less computing and memory resources, and easierto interpret and to implement. Therefore, they are usuallymore desirable than complex designs in the real world if theyhave good enough performance. Recently, Jia [1] discussed theimportance of design simplicity and introduced an adaptivesimulation-based sampling algorithm to sequentially screen thedesigns until one simplest good enough design is found. In thispaper, we consider a more generalized problem and introducetwo algorithms OCBA-mSG and OCBA-bSG to identify asubset of m simplest and good enough designs among a totalof K (K > m) designs. By controlling the simulation allocationintelligently, our approach intends to find those simplest goodenough designs using a minimum simulation time. The numericalresults show that both OCBA-mSG and OCBA-bSG outperformsome other approaches on the test problems.

Note to Practitioners—This paper was motivated by twoproblems from the real world: design of ordering policies ininventory control, and design of node activation rules in sensornetworks. In designing a good ordering policy or a good sensoractivation rule, simple designs have various advantages inpractice, e.g., easy to learn, to implement, to operate, and tomaintain. More generally, practitioners want to find a design ora set of designs which not only have good performance but arealso simple. Due to the complexity of the systems, simulationis a popular tool in industry to evaluate the performanceof different alternative designs before actual implementation.While the advance of new technology has dramatically increasedcomputational power, efficiency is still a big concern. Ourproposed approach intelligently controls the simulation ofalternative designs, so that the simplest good enough designs canbe selected using a minimum computation cost. Our numericalexperiments show that the proposed approach is very effective.

Index Terms—optimal computing budget allocation, rankingand selection, simulation-based optimization, complexity.

Shen Yan was a Master student in the Department of Industrial & EnterpriseSystems Engineering, University of Illinois at Urbana-Champaign, IL, 61801USA ([email protected]).

Enlu Zhou is with the faculty in the Department of Industrial & EnterpriseSystems Engineering, University of Illinois at Urbana-Champaign, IL, 61801USA ([email protected]).

Chun-Hung Chen (corresponding author) is with the faculty in Departmentof Electrical Engineering & Institute of Industrial Engineering, NationalTaiwan University, Taipei, Taiwan ([email protected]).

This work was supported by the National Science Foundation underGrants ECCS-0901543 and CMMI-1130273, the Air Force Office of Scien-tific Research under Grant FA-9550-12-1-0250, the Department of Energyunder Grant DE-SC0002223, National Institutes of Health under Grant1R21DK088368-01, and National Science Council of Taiwan under GrantNSC-100-2218-E-002-027-MY3.

I. INTRODUCTION

THE motivation for considering the selection of simplestgood enough designs comes from the real world, where

simple designs are preferred to the complex ones if the simpledesigns are good enough to satisfy our requirements. A typicalexample is to find an ordering policy in inventory control.Consider the problem of ordering a certain amount of productsat each period to meet a stochastic demand which followsa probability distribution. In order to minimize the expectedcost (including holding cost for excess inventory and shortagecost for unfilled demand), we want to determine the optimalordering policy in each period. The optimal policy can befound analytically for some models to have the structure of abase-stock policy or an (s,S) policy [2], [3], [4], [5]. The base-stock policy is a threshold function that maps the current stockinto the ordering amount. Motivated by this simple structure ofthe optimal policy for some models, we can approximate theoptimal policies for other more complex inventory models bythreshold policies. In general, we can approximate the optimalordering policy better with a greater number of thresholdsgiven the right values of these thresholds. Then the problemis to decide the number and the values of the thresholds. Itis clear that with more thresholds in the function we have amore complex ordering policy, making it harder to determinethe optimal values of these thresholds and to implement inpractice. Conversely, with a small number of thresholds, suchas one (base-stock policy) or two, we have a simple orderingpolicy, which is easier to compute and to implement. If thesimple ordering policy can achieve a required cost criterion, itwill be more desirable than a complex ordering policy, eventhough the complex policy may yield a lower cost. Anotherexample is the design of node activation rules in the wirelesssensor networks (WSNs), as described in [6], [1]. Each nodeneeds to collaborate with its neighbors in order to have enoughpower to monitor an area of interest. Similarly, given that therequired probability of correct detection can be achieved, weprefer small communication radius of each node (i.e., simpleactivation rule) to large radius (i.e., complex activation rule).

In this paper, we use the word “design” to refer to theobject under consideration, such as the ordering policy and thenode activation rule in the previous examples. We consider theproblem of selecting m (m≥ 1) designs that are simplest (withsmallest complexity) and good enough (satisfying a constrainton the performance measure). Selection of multiple designs issometimes preferred because of at least two reasons. First, adecision maker has to face different objectives and constraints.However, in many cases it is too complicated to include all

Page 2: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

2

objectives and constraints in the simulation model. As a result,the decision maker may not like the best design obtained fromthe model. By offering a set of multiple good designs, thedecision maker can choose the one he/she likes by consideringmore factors. Second, the design space can be extremely largeand the total simulation cost is too expensive. A commonapproach is to first apply a simplified model to screen out somegood alternatives before the full-scale simulation modelinganalysis. Offering a set of multiple good designs is highlyuseful for this purpose.

The complexity of a design is represented by an integernumber, where simpler design has a smaller integer num-ber. The complexity is a deterministic value known beforesimulation. However, the performance of a design is subjectto system noise, and hence, it can only be estimated fromsimulation, which is often computationally expensive. Forexample, it takes a significant amount of computational effortto simulate the inventory system in order to evaluate the costof a particular ordering policy. Hence, our goal is to allocatea given simulation budget efficiently to the designs so as tomaximize the probability of correctly selecting the m simplestgood enough designs out of a total of K designs.

The above problem is closely related with many known re-sults in the literature on ranking and selection (R&S). Severalrecent R&S procedures are discussed and compared by Brankeet al. [7]. Some of the procedures can be further extendedto more generalized simulation optimization problems (e.g.,[8], [9], [10], [11], [12], [13]). For subset selection problem,Gupta [14] proposed the method of selecting a random sizesubset containing the best design with a given probability ofcorrect selection. Later, Santner [15] extended Gupta’s methodby imposing a maximum size m on the subset. Koenig andLaw [16] developed a two-stage procedure for selecting thetop m designs with best performance, following the resultsin Dudewicz and Dalal [17]. Chen et al. [18], [19], [20]developed the optimal computing budget allocation (OCBA)procedure for the selection of one best design, and later Chenet al. [21] extended OCBA to the selection of the m bestdesigns. However, all of this work has focused on optimizinga single-objective performance measure.

Problems of multi-objective optimization and constraintoptimization have also been studied. Lee et al. [22], Teng et al.[23], Chew et al. [24], and Lee et al. [25] extended the OCBAframework to efficiently select designs that optimize multipleperformance measures. Branke and Mattfeld [26] proposed tosearch for solutions that are not only good but also flexiblein dynamic scheduling. Branke and Gamer [27] proposedan efficient sampling procedure in interactive multi-criterionselection. In constraint optimization, Andradottir et al. [28]proposed a two-phase approach which identifies all the feasiblesystems first and then selects the best from them. Szechtmanand Yucesan [29] used large deviation theory to deal withfeasibility determination. Most recently, Pujowidianto et al.[30] developed OCBA further for selecting one single bestdesign under multiple constraints of secondary performancemeasures.

The problem of considering both complexity and perfor-mance evaluation has only been considered recently. It appears

to be a multi-objective optimization or a constraint optimiza-tion problem, but it has its unique problem structure that canbe exploited to design a more efficient sampling procedure.The most relevant problem to ours is probably the selectionof one simplest good design, for which Jia [1] proposed anAdaptive Sampling Algorithm (ASA) to minimize the Type IIerror of the chosen design. The relation between complexityand performance in choosing policies has also been exploredin the context of Markov decision processes [31], [32].

In this paper, we address this problem of selecting multiplesimplest good enough designs by proposing the algorithmOCBA-mSG, abbreviated for optimal computing budget al-location for m simplest good enough designs. Based onOCBA-mSG, we develop another slightly different algorithmcalled OCBA-bSG for selecting the designs with the bestperformance from all the simplest good enough designs, witha slight increase of simulation budget than OCBA-mSG.Numerical results indicate that both OCBA-mSG and OCBA-bSG allocate the simulation budget efficiently to achieve ahigh probability of correct selection.

The rest of the paper is organized as follows. In Section II,we define the two problems of selecting m simplest goodenough designs and selecting the best m simplest good enoughdesigns, respectively. In Section III, we state the main results(proofs are included in the Appendix) and present the algo-rithms. In Section IV, we carry out the numerical experimentson several test problems. Finally, we conclude our paper inSection V.

II. PROBLEM STATEMENT

A. Selecting m Simplest Good Enough Designs

Let θ denote a design, and Θ denote the set of all the K(K > m) designs, i.e.,

Θ = {θ1,θ2, . . . ,θK}.

To simplify notations, we will also use the integers 1,2, . . . ,Kto denote the designs in the following when there is noambiguity. The performance of the design θk is measured by

Jk = E[L(θk,ζ )],

where ζ is a random vector that represents the uncertaintyin the system, and L(θk,ζ ) can only be evaluated throughsimulation of the system. The underlying assumption is thatsuch simulation is expensive. A design is considered better ifits performance measure J is smaller. A good enough designis one that satisfies Jk < J0, where J0 is a given thresholdon the performance. In practice, J0 can be set by the useror chosen based on a few pilot runs which return a roughestimate of the performance of the designs. Please note thatthe definition of “good enough” here is the same as “feasible”,which is different from the definition in the literature onordinal optimization (c.f. [33]). In the rest of the paper we willuse the words “good enough” and “feasible” interchangeably.Hence, the good enough set (or the feasible set) is defined as

F = {k∣Jk < J0,k = 1,2, . . . ,K}.

Page 3: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

3

The complexity of the design θk is represented by thecomplexity C(θk), which is a deterministic value in the set{0,1, . . . ,n},n < K, and is known before simulation. Note thatC(θ) is the result of the user mapping his/her definition ofcomplexity to integer numbers. For example, the user couldmap the number of thresholds of the ordering policy to aninteger; or the user could map a range of communication radiusin the WSN problem to an integer. The complexity set Cicontains all the designs with complexity i, defined as

Ci = {k∣C(θk) = i,k = 1,2, . . . ,K}.

The set of m simplest and good enough designs is defined as

Sm = {m1,m2, . . . ,mm ∣ C(θmi)⩽C(θk),∀k ∈ F ∖Sm},

where F ∖Sm = {k ∈ F ∣k ∕∈ Sm}. Notice that the set Sm may notbe unique, because it is possible that multiple designs in theset F have the same complexity. For example, if there are m(m < m) designs in F with complexity 0 and m (m > m−m)designs in F with complexity 1, then Sm includes all the mdesigns with complexity 0 and any m−m designs of those mdesigns with complexity 1.

Fig. 1 gives a pictorial view of the general case. Supposethat all the designs in the complexity sets C0, C1, . . ., Ct−1(1 ≤ t ≤ n) are not good enough (or infeasible) and the firstfeasible design appears in the set Ct . Moreover, suppose thatthe total number of feasible designs in Ct ,Ct+1, . . . ,Ct ′ is lessthan m until t ′ reaches t+ p (0≤ p≤ n− t). Hence, in generalwe need to consider three types of subsets, which we refer toas: infeasible simplest subsets Sdi , i = 0,1, . . . , t−1; simplestgood enough subsets Ssi , i = 0,1, . . . , p; and infeasible non-simplest subsets Sei , i = 0,1, . . . , p. In particular, the simplestgood enough subsets Ss0 ,Ss1 , . . . ,Ssp satisfy

p−1

∑i=0∣Ssi ∣< m,

p

∑i=0∣Ssi ∣⩾ m,

where ∣ ⋅ ∣ denotes the cardinality of the set. Therefore, ac-cording to the definition of the m simplest and good enoughdesigns, Sm should include all the designs in the subsetsSs0 ,Ss1 , . . . ,Ssp−1 and any (m−∑

p−1i=0 ∣Ssi ∣) designs in the subset

Ssp . Since there are already m designs selected in the lowercomplexity sets C0, . . . ,Ct+p, there is no need to consider thehigher complexity sets.

In simulation, we compute the sample mean J based onthe samples on hand to estimate the performance J for eachdesign, and then order the designs to find the subsets {Ssi , i =0,1, . . . , p}, {Sdi , i = 0,1, . . . , t − 1} and {Sei , i = 0,1, . . . , p}as estimates for Ssi , Sdi and Sei respectively according tothe relationship shown in Fig. 1. Hence, the selected set Smshould include all the designs in Ss0 , Ss1 , . . . , Ssp−1 and any(m−∑

p−1i=0 ∣Ssi ∣) designs in Ssp . To further classify the subsets,

we denote

Ss =p∪

i=0

Ssi , SI ={∪p

i=0Sei

}∪{∪t−1

i=0 Sdi

}.

We define the correct selection CSm as the event that thedesigns in Ss are the true simplest good enough designs, i.e.,

CSm = {Ji < J0 & J j ⩾ J0,∀i ∈ Ss,∀ j ∈ SI},

Fig. 1: Relationship between subsets. Ji j denotes the per-formance of a design whose complexity is i and whoseperformance is the jth smallest in its complexity set Ci.

where Ss includes all the simplest good enough designs, SIincludes all the infeasible (either simplest or non-simplest)designs.

The determination of Ss and SI is based on the estimate(sample mean) of the performance of every design, and theaccuracy of the estimate is determined by the number ofsimulations carried out for that design. Therefore, the decisionvariables that determine the probability of correct selectionP(CSm) are the number of simulations N1,N2, . . . ,NK allocatedfor the designs θ1,θ2, . . . ,θK , respectively. This will be moreclearly shown in the explicit expression (3) for P(CSm) later.Given a fixed total simulation budget T , we want to find theoptimal budget allocation N1,N2, . . . ,NK to the K designs inorder to maximize the probability of correct selection:

maxN1,N2,...,NK

P(CSm)

s.t. N1 +N2 + . . .+NK = T. (1)

B. Selecting the Best m Simplest Good Enough Designs

Consider a simple example that there are two good enoughdesigns with complexity 0, say A and B, so they are bothsimplest good enough designs. If we only need one simplestgood enough design, then we can choose either A or B.However, if A and B have different performance, say JA < JB,then we would prefer A, because it is better than B inperformance and as simple as B. This is what we refer to asthe “best simplest good enough design”. The formal definitionof the best m simplest good enough designs is as follows:

Sb = {b1, . . . ,bm ∈ F ∣ C(θbi)<C(θk) ORJbi < Jk if C(θbi) =C(θk),∀k ∈ F ∖Sb},

where F ∖ Sb = {k ∈ F ∣k ∕∈ Sb}. The key difference from Smis that Sb includes all the feasible designs in the subsetsSs0 ,Ss1 , . . . ,Ssp−1 and the best (m−∑

p−1i=0 ∣Ssi ∣) designs in the

subset Ssp . That implies that the subset Ssp should be furtherdivided into two subsets: optimal subset Sbl , and feasible non-optimal set Sa. Fig. 1 gives a pictorial view of the relationship

Page 4: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

4

between the subsets. Therefore, the optimal set Sb satisfies

Sb ={∪p−1

i=0 Ssi

}∪Sbl , ∣Sb∣= m.

In simulation, we find estimates for these subsets based onthe sample means of the designs, and similarly we denote

Sb− =p−1∪i=0

Ssi , SI ={∪p

i=0Sei

}∪{∪t−1

i=0 Sdi

}.

Then the correct selection CSb is defined as

CSb = {Ji < J0 & J j ⩽ Jk < J0 & Js ⩾ J0,

∀i ∈ Sb− ,∀ j ∈ Sbl ,∀k ∈ Sa,∀s ∈ SI}.

Our goal is to find the optimal budget allocation N1,N2, . . . ,NKto the K designs in order to maximize the probability of correctselection given a fixed total simulation budget T :

maxN1,N2,...,NK

P(CSb)

s.t. N1 +N2 + . . .+NK = T (2)

III. MAIN RESULTS

A. Selecting m Simplest Good Enough Designs

We estimate P(CS) using the same Bayesian model pre-sented in [34] and [35]. Assuming that the performance ofeach design, Ji, has a noninformative normal prior distributionN(0,ν2) with ν2 extremely large, and a sample Ji for Ji is nor-mally distributed as N(Ji,σ

2i ), then the posterior distribution

of Ji has been shown in [34] to be

Ji ∼ N(Ji,σi

2

Ni),

where Ji = 1Ni

∑Nik=1 Ji(k), and Ji(1), Ji(2), . . . , Ji(Ni)

iid∼N(Ji,σ

2). Thus, P(CSm) is as follows:

P(CSm) = P{Ji < J0 & J j ⩾ J0,∀i ∈ Ss,∀ j ∈ SI}= ∏

i∈Ss

P{Ji < J0}∏j∈SI

P{J j ⩾ J0}, (3)

where the second equation is due to the independence betweendesigns. The results are stated in the following theorem, andthe proof is contained in the Appendix.

Theorem 1. P(CSm) is asymptotically (as T →∞) maximizedby the following allocation rule:

Ni

σ2i /(Ji− J0)2 =

N j

σ2j /(J j− J0)2 , (4)

for all i ∈ Ss and j ∈ SI . Nk = 0 for all other k ∈ {1,2, . . . ,K}.

Remark 1. From (4), we know that the simulation budgetfor each design increases proportionally to its correspondingsample variance. If a design has a larger sample variance,more simulation budget will be allocated to it in order toobtain a more accurate estimate for the performance in thenext iteration. On the other hand, the simulation budget foreach design decreases proportionally to the difference betweenits sample mean and the good enough threshold J0. The designwhose sample mean of the performance is closer to J0 will be

assigned more simulation budget, since it is more sensitive tothe feasibility test. As there are already m designs selectedfrom Ss

∪SI in the lower complexity sets, there is no need to

consider the higher complexity sets, and hence, there is nomore simulation budget allocated to the designs that are notin Ss

∪SI . However, as more simulation is carried out and the

sample means are updated, Ss and SI may become different atthe next iteration and contain some of the higher-complexitysets that are not considered in the previous iteration.

B. Selecting the Best m Simplest Good Enough Designs

It is hard to maximize P(CSb) (problem (2)) analytically,and hence we maximize a lower bound of P(CS)b, which iscalled Approximate Probability of Correct Selection APCSb[18]. APCSb is defined as follows.

P(CS)b = P{Ji < J0 & J j ⩽ Jk < J0 & Js ⩾ J0,

∀i ∈ Sb− ,∀ j ∈ Sbl ,∀k ∈ Sa,∀s ∈ SI}⩾ P{Ji < J0 & J j ⩽ µ & µ ⩽ Jk < J0 & Js ⩾ J0,

∀θi ∈ Sb− ,∀ j ∈ Sbl ,∀k ∈ Sa,∀s ∈ SI}= ∏

i∈Sb−

P{Ji < J0} ∏j∈Sbl

P{J j ⩽ µ}∏k∈Sa

P{µ ⩽ Jk < J0}∏s∈SI

P{Js ⩾ J0} (5)

≜ APCSb,

where the second equation is due to the independence betweenthe designs. It is easy to see that a larger APCSb yields a betterapproximation for P(CSb). Following a similar procedure asin [21], we determine the value of µ as stated in the followingLemma.

Lemma 2. Let θ[r] denote the design with the largest samplemean in the subset Sbl , and θ[r+1] denote the design withthe smallest sample mean in the subset Sa. Then the µ valueintroduced in APCSb is given by

µ =σ[r+1]J[r]+ σ[r]J[r+1]

σ[r]+ σ[r+1], (6)

where σi = σi/√

Ni.

Therefore, instead of solving the maximization problem (2),we consider the following maximization problem.

maxN1,N2,...,NK

APCSb

s.t. N1 +N2 + . . .+NK = T. (7)

The results are given in the following theorem, and the proofis contained in the Appendix.

Theorem 3. APCSb is asymptotically (as T → ∞) maximizedby the following allocation rule:

Case 1: If Sa ∕=∅ (i.e., the total number of feasible designsis greater than m), then

Ni

σ2i /(Ji− J0)2 =

N j

σ2j /(J j−µ)2 =

Ns

σ2s /(Js− J0)2

=Nx

σ2x /(Jx−µ)2 =

Ny

σ2y /(Jy− J0)2 , (8)

Page 5: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

5

for all i ∈ Sb− , j ∈ Sbl , s ∈ SI , x ∈ {k ∈ Sa∣Jk ⩽µ+J0

2 }, y ∈{k ∈ Sa∣Jk >

µ+J02 }. Nk = 0 for all other k ∈ {1,2, . . . ,K}.

Case 2: If Sa =∅ (i.e., the total number of feasible designsis less than or equal to m), then

Ni

σ2i /(Ji− J0)2 =

Ns

σ2s /(Js− J0)2 (9)

for all i ∈ Sb−∪

Sbl and s ∈ SI . Nk = 0 for all other k ∈{1,2, . . . ,K}.

Remark 2. Theorem 2 provides some intuitive results. Wenotice that at the two critical points µ and J0: µ is the thresholdfor the optimality, J0 is the threshold for the feasibility. Thedesigns closer to these two points will be assigned moresimulation budget. For the subsets Sb− and SI , we are onlyinterested in determining wether the designs are good enough,and indeed more simulation budget is assigned to the designsnear J0. Similarly, for the subset Sbl , we are only interestedin comparing the performance of the designs, and moresimulation budget is assigned to the designs around µ . Forthe subset Sa where both µ and J0 are critical points, the lasttwo terms in (8) imply that we should divide the set into twoparts by the midpoint µ+J0

2 : the designs with sample meansin the range µ ⩽ Jx ⩽

µ+J02 will be compared with µ , and the

ones closer to µ will get more simulation budget; the designsfalling into the range µ+J0

2 < Jy ⩽ J0 will be compared withJ0, and be assigned more simulation budget if closer to J0.Please see fig. 2 for a pictorial view of the budget allocationin the complexity set Ct+p, the highest complexity set underconsideration.

Fig. 2: Simulation allocation in the set Ct+p.

Remark 3. Comparing selecting Sb with Sm, the difference isin Ct+p, where the subset Ssp in selecting Sm is divided intotwo subsets Sbl and Sa in selecting Sb. Theorem 3 impliesthat in addition to allocating more simulation budget to thedesigns near J0 in the sets C0,C1 . . . ,Ct+p, we also allocatesimulation budget to designs near µ in the set Ct+p. As aresult, the extra simulation budget assigned to designs nearµ in selecting Sb is approximately 2/(t +2p+2) of the totalsimulation budget in selecting Sm. If t = 0 and p= 0 (i.e., thereare more than m feasible designs in the lowest complexity setC0), then selecting Sb needs approximately double simulationbudget of that in selecting Sm for the same accuracy of thesample means of the design performance. On the other hand,if t + 2p is large, selecting Sb requires little extra simulationbudget, and will be preferred since it yields the best m designsamong all simplest and good enough designs.

C. OCBA-mSG and OCBA-bSG

Based on the above results, we propose the Optimal Com-puting Budget Allocation procedure for selecting m Simplestand Good enough designs (OCBA-mSG) and that for selectingthe Best m Simplest and Good enough designs (OCBA-bSG). Since the two algorithms are similar, we describe themtogether to save space and specify the different steps in thedescription.

OCBA-mSG and OCBA-bSGInput: the total number of the designs K, the number of

designs needed m, the total simulation budget T , the simulationbudget increase at each iteration ∆, the initial simulationbudget for every design n0, the good enough performanceconstraint J0, and the upper bound of the total simulationbudget for one design NU .

Initialize: l = 0.∙ Group the designs according to their complexities to

obtain the complexity sets C0,C1, . . . ,Cn.∙ Perform n0 simulation replications for all designs to

generate samples Xki , k = 1,2, . . . ,n0, i = 1,2, . . . ,K. Set

Nl = Kn0.Loop: while Nl < T , do1) Update:

∙ For each design i, compute the sample mean Ji =1

Nli

∑Nl

ik=1 Xk

i , and the sample standard deviation σi =√∑

Nli

k=1 (Xki − Ji)2/(Nl

i −1). Sort the designs in eachcomplexity set according to their sample means inthe increasing order.

∙ Increase the computing budget Nl+1 = min{Nl +∆,T}.

2) Allocate:OCBA-mSG∙ For each design θi, compute the simulation budget

Nl+1i according to (4).

OCBA-bSG∙ If the total number of feasible designs is greater than

m, compute µ according to (6), and compute thesimulation budget Nl+1

i for each design θi accordingto (8).

∙ Otherwise, compute the simulation budget Nl+1i for

each design θi according to (9).3) Simulate:

∙ If Nl+1i ⩾ NU or Nl+1

i ⩽ Nli , we set Nl+1

i = Nli , and

do not simulate design θi at this iteration.∙ Otherwise, perform (Nl+1

i −Nli ) simulations for de-

sign θi to generate more samples Xki , k = Nl

i +1,Nl

i +2, . . . ,Nl+1i .

4) Update: l → l +1.End of loopOutput: output the feasible designs starting from the lowest

complexity set in the increasing order of their sample meansuntil the total number of such designs reaches m or all thedesigns have been examined.

Page 6: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

6

Remark 4. In the above algorithms, we introduce an upperbound NU on the simulation budget for one single design:if Ni ⩾ NU , we stop allocating new simulation budget tothat design. That is because we obtain the simulation budgetallocation rules under the asymptotic limit T → ∞ but theactual total simulation budget T is finite. When T is infinity,we can assign a large amount of budget to one design atone iteration, and there will always be enough budget left forother designs if needed at future iterations. This is not truewhen T is finite. Thus, we introduce NU and determine itsvalue in the following way. Since the designs near the criticalpoints need more simulation budget, we need to ensure eachof such critical designs will be simulated at least once. Hence,we approximate the upper bound by counting the number ofsubsets related to the critical points after initialization, wherethose subsets are Ssi , Sdi , Sei in OCBA-mSG or Ssi , Sbl , Sa,Sdi , Sei in OCBA-bSG (c.f. Fig. 1).

NU =(T −Kn0)

total number o f sets+n0.

This choice of upper bound works well in our numericalexperiments.

IV. NUMERICAL EXPERIMENTS

In this section, we demonstrate our methods OCBA-mSGand OCBA-bSG on some examples and also compare themwith two other methods - Equal Allocation and Levin Search.

Equal Allocation (EA) allocates the simulation budgetequally among all the designs and do not use any informationsuch as the mean, the variance or the complexity of the design.At iteration l, it allocates ∆ simulation budget according to

Nl+1i −Nl

i = ∆/K, ∀i ∈ {1,2, . . . ,K}.

Levin Search (LS) method [36] allocates simulation budgetto the designs sequentially in the order of the complexity. Itis useful when applied to find one simplest and good enoughdesign [1]. LS first simulates the designs with smallest com-plexity until obtaining a certain accuracy for the estimates ofthe performance, based on which the good enough designs areselected. If only less than m good enough designs are found, itthen continues to simulate the designs in the next complexityset until eventually finding m simplest good enough designseventually. In our implementation, we simulate every designfor n0 times at initialization, and order them according to theirsample means and complexities. Since it is hard to specify agiven accuracy for the estimates in our examples, we evenlyallocate the total remaining budget (T −Kn0)/K to all thedesigns beforehand, but simulate the designs sequentially, i.e.,start simulating the first simplest design for (T−Kn0)/K timesand then move on to the next one to repeat the same procedure.Please note LS often exhibits some “jump” behavior in theP(CS), because the P(CS) stays flat if the design currentlyunder simulation is not good enough and the P(CS) increasesotherwise. If the desirable set of designs is found beforeutilizing all computing budget, LS will terminate and thecorresponding P(CS) curve will level off in the figures. LSis the same as EA when utilizing all the T simulation budget,

but LS often achieves the final P(CS) earlier. In general, LSmethod performs better if the performance deteriorates as thecomplexity increases.

In the numerical experiments, we test three generic exam-ples which mimic different scenarios in real world. In Example1, good designs are also simple designs. In contrast, baddesigns are simple ones in Example 2. In Example 3, weconsider a problem with a larger number of alternative designs.We use P(CS) as the efficiency measurement: for a given totalsimulation budget, the faster the P(CS) converges, the betterthe corresponding method is. Here we estimate P(CS) usingMonte Carlo simulation by computing the ratio of the numberof simulation runs with correct selections to the total numberof simulation runs. In addition, for convenience, we assumethat design θi has complexity ⌊log2 i⌋, so the complexity isnon-decreasing in i.

1) Example 1 (Mean increases as complexity increases):There are 20 designs in total, with the ith design hav-ing L(θi,ζ ) distributed according to the normal distributionN(i,(0.5i)2). We want to find 5 simplest good enough designswith good enough constraint J0 = 6.3. The initial simulationbudget n0 = 20, simulation budget increment ∆ = 200, to-tal simulation budget T = 8000, and total number of sim-ulation runs = 104. The complexity sets are C0 = {θ1},C1 = {θ2,θ3}, C2 = {θ4,θ5,θ6,θ7}, C3 = {θ8, . . . ,θ15} andC4 = {θ16, . . . ,θ20}. The mean increases as the complexityincreases, and the variance increases as the mean increases.

a) OCBA-mSG: The correct selection of the five desir-able designs should include {θ1,θ2,θ3} and any two from{θ4,θ5,θ6}. Fig. 3 shows that OCBA-mSG converges fasterthan EA and LS. EA performs well in this example becauseof the small total number of designs K and the small varianceσ2. LS searches from the simplest sets {θ1}, {θ2,θ3}, . . ., andin this example the correct selection is {θ1,θ2,θ3,θ4,θ5}, soLS terminates in about 7 iterations.

0 1000 2000 3000 4000 5000 6000 7000 80000.7

0.75

0.8

0.85

0.9

0.95

1

Total Simulation Buget

P(C

S)

OCBA−mSGEALS

Fig. 3: Example 1 - selecting 5 simplest good enough designsfrom 20 designs with distribution N(i,(0.5i)2) and J0 = 6.3.

b) OCBA-bSG: The correct selection is{θ1,θ2,θ3,θ4,θ5}. Fig. 4 shows the simulation result.

2) Example 2 (Mean decreases as complexity increases):There are 20 designs, with the ith design having L(θi,ζ )

Page 7: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

7

0 1000 2000 3000 4000 5000 6000 7000 8000

0.4

0.5

0.6

0.7

0.8

0.9

1

Total Simulation Buget

P(C

S)

OCBA−bSGEALS

Fig. 4: Example 1 - selecting the best 5 simplest good enoughdesigns from 20 designs with distribution N(i,(0.5i)2) andJ0 = 6.3.

distributed according to the normal distribution N((21 −i),(0.5i)2). We want to find 5 simplest good enough designswith good enough constraint J0 = 7.3. The initial simulationbudget n0 = 20, simulation budget increment ∆ = 200, totalsimulation budget T = 8000, and total number of simulationruns = 104. The complexity sets are the same as in Example1. The mean decreases as the complexity increases, and thevariance increases as the mean decreases.

a) OCBA-mSG: Correct selection of the five desir-able designs should include {θ14,θ15} and any three from{θ16,θ17,θ18,θ19,θ20}. Fig. 5 shows the simulation result.All three methods converge slower than Example 1, butOCBA-mSG still converges faster than EA and LS. Designswith smaller means have larger variances, and the correctselection is in the set {θ14,θ15,θ16,θ17, θ18,θ19,θ20} whichhave relatively large variances compared to other designs, soOCBA-mSG converges slower than that in Example 1. LS stillsearches from the simplest sets while the correct selection isin the higher complexity sets, so LS method also terminateslater than Example 1.

0 1000 2000 3000 4000 5000 6000 7000 8000

0.4

0.5

0.6

0.7

0.8

0.9

1

Total Simulation Buget

P(C

S)

OCBA−mSGEALS

Fig. 5: Example 2 - selecting 5 simplest good enough designsfrom 20 designs with distribution N((21− i),(0.5i)2) and J0 =7.3.

b) OCBA-bSG: The correct selection is{θ14,θ15,θ18,θ19,θ20}. Fig. 6 shows the simulation result.

0 1000 2000 3000 4000 5000 6000 7000 80000.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Total Simulation Buget

P(C

S)

OCBA−bSGEALS

Fig. 6: Example 2 - selecting the best 5 simplest good enoughdesigns from 20 designs with distribution N((21− i),(0.5i)2)and J0 = 7.3.

3) Example 3 (Mid-scale problem): There are 65 designs,with the ith design having L(θi,ζ ) distributed according to thenormal distribution N((66− i),(0.05i)2). We want to find 5simplest good enough designs with good enough constraintJ0 = 6.3. The initial simulation budget n0 = 20, simulationbudget increment ∆ = 200, and total simulation budget T =8000. The complexity sets are C0 = {θ1}, C1 = {θ2,θ3},C2 = {θ4, . . . ,θ7}, C3 = {θ8, . . . ,θ15}, C4 = {θ16, . . . ,θ31},C5 = {θ32, . . . ,θ63} and C6 = {θ64,θ65}.

a) OCBA-mSG: The correct selection of the five de-sirable designs should include {θ60,θ61, θ62,θ63} and anyone from {θ64,θ65}. For this mid-scale problem, OCBA-mSGperforms much better than EA and LS as shown in Fig. 7.Detailed explanation is similar to that for OCBA-bSG in thefollowing.

1000 2000 3000 4000 5000 6000 7000 80000.4

0.5

0.6

0.7

0.8

0.9

1

Total Simulation Buget

P(C

S)

OCBA−mSGEALS

Fig. 7: Example 3 - selecting 5 simplest good enough designsfrom 65 designs with distribution N((66− i),(0.05i)2) andJ0 = 6.3.

b) OCBA-bSG: The correct selection is{θ60,θ61,θ62,θ63,θ65}. For this mid-scale problem, OCBA-bSG performs much better than EA and LS as shown in

Page 8: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

8

Fig. 8. When the total design number K is large, EA convergesslowly since each design is assigned with less simulationbudget at every iteration compared to Examples 1 and 2. ForLS, the first time LS jumps in P(CS) is the time that thetotal simulation budget reaches 4900, which is when it firststarts to simulate designs in the set {θ32, . . . ,θ63} with means{34, . . . ,3}. As we assign the simulation budget according tothe order of the designs in the same complexity set, here wesimulate designs in the order of θ63,θ62, . . .. Since designsθ63,θ62,θ61 and θ60 belong to the correct selection set, LSjumps in P(CS) at this point. The second jump in the P(CS)for LS happens in the end due to the simulation budgetallocation to the design θ65.

1000 2000 3000 4000 5000 6000 7000 80000.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Total Simulation Buget

P(C

S)

OCBA−bSGEALS

Fig. 8: Example 3 - selecting the best 5 simplest good enoughdesigns from 65 designs with distribution N((66− i),(0.05i)2)and J0 = 6.3.

V. CONCLUSION

In this paper, we considered the simulation-based selectionof simplest good enough designs, which is motivated by real-life applications. We proved the optimal simulation budgetallocation rules to asymptotically maximize the probabilityof correct selection (or the approximate probability of correctselection in the case of OCBA-bSG). Based on the asymptoticresults, we proposed the algorithm OCBA-mSG to efficientlyallocate the simulation budget for selecting m simplest goodenough designs out of a total of K designs, and also proposeda slightly different algorithm OCBA-bSG in order to find thebest m simplest good enough designs. Numerical results showthat both methods converge fast on all the test problems, whichindicates OCBA-mSG and OCBA-bSG indeed allocate simu-lation budget efficiently. While our algorithms are motivatedby the asymptotic results, an important future direction is toanalyze the finite-time performance of our algorithms.

VI. APPENDIX

A. Appendix A: Proof of Theorem 1

Since Jk ∼ N(Jk,σk

2

Nk), we have

P(Jk < J0) = Φ

(J0− Jk

σk/√

Nk

),

where Φ is the error function (i.e., the cumulative distribu-tion function (c.d.f.) of the standard normal distribution). ByLagrangian relaxation of P(CSm) and Karush-Kuhn-Tucker(KKT ) condition (c.f. [37]) for the maximization problem (1),we get

F = ∏k∈Ss

P{Jk < J0}∏k∈SI

P{Jk ⩾ J0}−λ (K

∑k=1

Ni−T ).

For i ∈ Ss,

∂F∂Ni

=12 ∏

k∈Ss,k ∕=iP{Jk < J0}∏

k∈SI

P{Jk ⩾ J0} ⋅

φ

(J0− Ji

σi/√

Ni

)J0− Ji

σi√

Ni−λ = 0. (10)

For i ∈ SI ,

∂F∂Ni

= −12 ∏

k∈Ss

P{Jk < J0} ∏k∈SI ,k ∕=i

P{Jk ⩾ J0} ⋅

φ

(J0− Ji

σi/√

Ni

)J0− Ji

σi√

Ni−λ = 0, (11)

where φ denotes the probability density function (p.d.f.) of thestandard normal distribution.

In order to find the relationship between Ni and N j, we needto consider (2

1)+(22) = 3 cases that i, j belong to different sets.

Case 1: i ∈ Ss, j ∈ SI . Equating (10) and (11),

P{J j ⩾ J0}φ(

J0− Ji

σi/√

Ni

)J0− Ji

σi√

Ni

= P{Ji < J0}φ

(J0− J j

σ j/√

N j

)J j− J0

σ j√

N j.

Taking logarithm on both sides, we have

logP{

J j ⩾ J0}− (J0− Ji)

2

2σi2/Ni+ log

(J0− Ji

σi

)− 1

2logNi

= logP{

Ji < J0}−

(J0− J j)2

2σ j2/N j+ log

(J j− J0

σ j

)− 1

2logN j.

Assuming Ni takes continuous values, let Ni =αiT . Taking theasymptotic limit of the above equation as T → ∞, we have

limT→∞

1T{logP(J j ⩾ J0)−

(J0− Ji)2αiT

2σi2+

log(

J0− Ji

σi

)− 1

2log(αiT )}

= limT→∞

1T{logP(Ji < J0)−

(J0− J j)2α jT

2σ j2+

log(

J0− J j

σ j

)− 1

2log(α jT )},

and we obtain

αi

α j=

(J j− J0)2

(Ji− J0)2 ⋅σ2

i

σ2j.

Case 2: i ∈ Ss, j ∈ Ss, i ∕= j. By equating (10) and (10),similarly as above we obtain

αi

α j=

(J j− J0)2

(Ji− J0)2 ⋅σ2

i

σ2j.

Page 9: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

9

Case 3: i ∈ SI , j ∈ SI , i ∕= j. By equating (11) and (11),similarly as above we obtain

αi

α j=

(J j− J0)2

(Ji− J0)2 ⋅σ2

i

σ2j.

Combining all three cases, we prove Theorem 1.

B. Proof of Lemma 2

Our derivation of the value of µ follows the idea and methodin Section 3.3 in [21]. Specifically, if we assume that all thedesigns have equal variances, then we know

P(J[r] ⩽ µ

)⩽ P

(Ji ⩽ µ

), ∀i ∈ Sbl ,

P(J[r+1] ⩾ µ

)⩾ P

(Ji ⩽ µ

), ∀i ∈ Sa.

To maximize APCSb is equivalent to maximizing the productof all the above terms. The smallest terms P

(J[r] ⩽ µ

)and

P(J[r+1] ⩾ µ

)have the most impact on the value of the

product. Hence, to simplify the problem, we consider themaximization of the product of these two terms. A good choiceof µ can be determined by solving the following maximizationproblem

maxN[r],N[r+1] P(J[r] ⩽ µ

)P(µ ⩽ J[r+1]

)s.t. N[r]+N[r+1] = T.

Following the same approach in the proof of Theorem 1, weobtain the asymptotically (as T → ∞) optimal solution (6).

C. Proof for Theorem 3

Since Ji ∼ N(Ji,σi

2

Ni), we have

for i ∈ Sb− , P(Ji < J0) = Φ

(J0− Ji

σi/√

Ni

);

for i ∈ Sbl , P(Ji ⩽ µ) = Φ

(µ− Ji

σi/√

Ni

);

for i ∈ Sa, P(µ ⩽ Ji < J0) = Φ

(J0− Ji

σi/√

Ni

)−Φ

(µ− Ji

σi/√

Ni

);

for i ∈ SI , P(Ji ⩾ J0) = Φ

(Ji− J0

σi/√

Ni

),

where Φ is the error function. By Lagrangian relaxation ofAPCSb and KKT condition, we have

F = ∏k∈Sb−

P{Jk < J0} ∏k∈Sbl

P{Jk ⩽ µ}∏k∈Sa

P{µ ⩽ Jk < J0}

∏k∈SI

P{Jk ⩾ J0}−λ

(K

∑k=1

Nk−T

).

Let φ denote the p.d.f. of the standard normal distribution. Weobtain the following conditions. For i ∈ Sb− ,

0 =∂F∂Ni

=−λ +12 ∏

k∈Sb− ,k ∕=iP{Jk < J0} ∏

k∈Sbl

P{Jk ⩽ µ}

∏k∈Sa

P{µ ⩽ Jk < J0}∏k∈SI

P{Jk ⩾ J0} ⋅φ(

J0− Ji

σi/√

Ni

)⋅ J0− Ji

σi√

Ni. (12)

For i ∈ Sbl ,

0 =∂F∂Ni

=−λ +12 ∏

k∈Sb−

P{Jk < J0} ∏k∈Sbl ,k ∕=i

P{Jk ⩽ µ}

∏k∈Sa

P{µ ⩽ Jk < J0}∏k∈SI

P{Jk ⩾ J0} ⋅φ(

µ− Ji

σi/√

Ni

)⋅ µ− Ji

σi√

Ni. (13)

For i ∈ Sa,

0 =∂F∂Ni

=−λ +12 ∏

k∈Sb−

P{Jk < J0} ∏k∈Sbl

P{Jl ⩽ µ}

∏k∈Sa,k ∕=i

P{µ ⩽ Jk < J0}∏k∈SI

P{Jk ⩾ J0} ⋅[φ

(J0− Ji

σi/√

Ni

)⋅ J0− Ji

σi√

Ni−φ

(µ− Ji

σi/√

Ni

)⋅ µ− Ji

σi√

Ni

].(14)

For i ∈ SI ,

0 =∂F∂Ni

=−λ − 12 ∏

k∈Sb−

P{Jk < J0} ∏k∈Sbl

P{Jk ⩽ µ}

∏k∈Sa

P{µ ⩽ Jk < J0} ∏k∈SI ,k ∕=i

P{Jk ⩾ J0} ⋅

φ

(Ji− J0

σi/√

Ni

)⋅ Ji− J0

σi√

Ni. (15)

In order to find the relationship between Ni and N j, weneed to consider (4

1)+(42) = 10 cases that θi and θ j belong to

different sets.Case 1: θi ∈ Sb− , θ j ∈ Sa. Equating (12) and (14), we have

P{Ji < J0}

(J0− J j

σ j/√

N j

)J0− J j

σ j√

N j−φ

(µ− J j

σ j/√

N j

)µ− J j

σ j√

N j

]

= P{µ ⩽ J j < J0}φ(

J0− Ji

σi/√

Ni

)J0− Ji

σi√

Ni.

Assuming Ni takes continuous values, let Ni =αiT . Taking theasymptotic limit of the above equation as T → ∞, we have

limT→∞

1T

[logP{Ji < J0}+ logA− logσ j−

12

log(α jT )]

= limT→∞

1T{logP{µ ⩽ J j < J0}−

(J0− Ji)2αiT

2σi2+

log(

J0− Ji

σi

)− 1

2log(αiT )}, (16)

where

A = φ

(J0− J j

σ j/√

α jT

)(J0− J j)−φ

(µ− J j

σ j/√

α jT

)(µ− J j).

By L’Hopital’s Rule,

limT→∞

1T

logA

= limT→∞

dA/dTA

= limT→∞

(−(J0−J j)

2

2σ j2/α j− −(µ−J j)

2

2σ j2/α j

)(µ− J j)

exp(−(J0−J j)

2T2σ j2/α j

− −(µ−J j)2T2σ j2/α j

)(J0− J j)− (µ− J j)

+

−(J0− J j)2

2σ j2/α j.

Page 10: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

10

1) If J0− J j ⩾ J j−µ , then exp(−(J0−J j)

2T2σ j2/α j

− −(µ−J j)2T

2σ j2/α j

)→

0 as T → ∞. Hence,

limT→∞

1T

logA =−(J j−µ)2α j

2σ2j

.

2) If J0− J j < J j−µ , then exp(−(J0−J j)

2T2σ j2/α j

− −(µ−J j)2T

2σ j2/α j

)→

∞ as T → ∞. Hence,

limT→∞

1T

logA =−(J0− J j)

2α j

2σ2j

.

By applying the above results of A to equation (16), we get

1) If J j ⩽J0+µ

2 , αiα j

=(µ−J j)

2

(J0−Ji)2 ⋅σ2

iσ2

j.

2) If J j >J0+µ

2 , αiα j

=(J0−J j)

2

(J0−Ji)2 ⋅σ2

iσ2

j.

Similarly, we can obtain the relationship between Ni andN j for the other 9 cases. If Sa = ∅, it reduces to the case ofOCBA-mbG. By combining the results of all the 10 cases, weprove Theorem 3.

ACKNOWLEDGMENT

A preliminary version of the manuscript was presented atthe 2010 Winter Simulation Conference [38].

REFERENCES

[1] Q. S. Jia, “An adaptive sampling algorithm for simulation-based opti-mization with descriptive complexity preference,” IEEE Transactions onAutomation Science and Engineering, vol. 8, no. 4, pp. 720 – 731, 2011.

[2] D. P. Bertsekas, Dynamic Programming and Optimal Control. AthenaScientific, 2005.

[3] H. P. Galliher, P. M. Morse, and M. Simond, “Dynamics of two classesof continuous-review inventory systems,” Operations Research, vol. 7,no. 3, pp. 362–384, 1959.

[4] I. Sahin, “On the objective function behavior in (s, S) inventory models,”Operations Research, vol. 30, no. 4, pp. 709–724, 1992.

[5] A. Federgruen and Y. S. Zheng, “An efficient algorithm for computing anoptimal (r, Q) policy in continuous review stochastic inventory system,”Operations Research, vol. 40, no. 4, pp. 808–813, 1992.

[6] K. Kar, A. Krishnamurthy, and N. Jaggi, “Dynamic node activation innetworks of rechargeable sensors,” IEEE/ACM Transactions on Network-ing, vol. 14, no. 1, pp. 15–26, 2006.

[7] J. Branke, S. E. Chick, and C. Schmidt, “Selecting a selection proce-dure,” Management Science, vol. 53, no. 12, pp. 1916 – 1932, 2007.

[8] S. H. Jacobson and E. Yucesan, “Computational issues for accessibilityin discrete event simulation,” ACM Transactions on Modeling andComputer Simulation, vol. 6, no. 1, pp. 53 – 75, 1996.

[9] ——, “Common issues in discrete-event simulation and discrete opti-mization,” IEEE Transactions on Automatic Control, vol. 47, no. 2, pp.341 – 345, 2002.

[10] M. C. Fu, “Optimization for simulation: Theory vs. practice,” INFORMSJournal on Computing, vol. 14, no. 3, pp. 192 – 215, 2002.

[11] L. Pi, Y. Pan, and L. Shi, “Hybrid nested partitions and mathematicalprogramming approach and its applications,” IEEE Transactions onAutomation Science and Engineering, vol. 4, no. 5, p. 573 586, 2008.

[12] W. Chen, L. Pi, and L. Shi, “An enhanced nested partitions algorithmusing solution value prediction,” IEEE Transactions on AutomationScience and Engineering, vol. 8, no. 2, pp. 412 – 419, 2011.

[13] W. P. Wong, L. H. Lee, and W. Jaruphongsa, “Budget allocation foreffective data collection in predicting an accurate DEA efficiency score,”IEEE Transactions on Automatic Control, vol. 56, no. 6, pp. 1235 –1246, 2011.

[14] S. S. Gupta, “On some multiple decision (selection and ranking) rules,”Technometrics, vol. 7, no. 2, pp. 225–245, 1965.

[15] T. J. Santner, “A restricted subset selection approch to ranking andselection problems,” The Annals of Statistics, vol. 3, no. 2, pp. 334–349, 1975.

[16] L. W. Koenig and A. M. Law, “A procedure for selecting a subsetof size m containing the l best of k independent normal populations,”Communications in Statistics - Simulation and Computation, vol. 14, pp.719–734, 1985.

[17] E. J. Dudewicz and S. R. Dalal, “Allocation of observation in rankingand selection with unequal variances,” Sankhya: The Indian Journal ofStatistics, vol. 37, pp. 28–78, 1975.

[18] C.-H. Chen, J. Lin, E. Yucesan, and S. E. Chick, “Simulation budgetallocation for further enhancing the efficiency of ordinal optimization,”Journal of Discrete Event Dynamic Systems: Theory and Applications,vol. 10, pp. 251–270, 2000.

[19] C.-H. Chen, D. He, and M. Fu, “Efficient dynamic simulation allocationin ordinal optimization,” IEEE Transactions on Automatic Control,vol. 51, no. 12, pp. 2005–2009, 2006.

[20] C.-H. Chen, E. Yucesan, L. Dai, and H. Chen, “Efficient computationof optimal budget allocation for discrete event simulation experiment,”IIE Transactions, vol. 42, no. 1, pp. 60 – 70, 2010.

[21] C.-H. Chen, D. H. He, M. Fu, and L. H. Lee, “Efficient simulationbudget allocation for selecting an optimal subset,” INFORMS Journalon Computing, vol. 20, no. 4, pp. 579–595, 2008.

[22] L. H. Lee, E. P. Chew, S. Y. Teng, and D. Goldsman, “Optimalcomputing budget allocation for multi-objective simulation models,” inProceedings of 2004 Winter Simulation Conference, 2004, pp. 586 –594.

[23] S. Y. Teng, L. H. Lee, and E. P. Chew, “Multi-objective ordinaloptimization for simulation optimization problems,” Automatica, vol. 43,no. 11, pp. 1884 – 1895, 2007.

[24] E. P. Chew, L. H. Lee, S. Y. Teng, and C. H. Koh, “Differentiatedservice inventory optimization using nested partitions and MOCBA,”Computers and Operations Research, vol. 36, no. 5, pp. 1703 – 1710,2009.

[25] L. H. Lee, E. P. Chew, S. Y. Teng, and D. Goldsman, “Finding the paretoset for multi-objective simulation models,” IIE Transactions, vol. 42,no. 9, pp. 656 – 674, 2010.

[26] J. Branke and D. Mattfeld, “Anticipation and flexibility in dynamicscheduling,” International Journal of Production Research, vol. 43,no. 15, pp. 3103 – 3129, 2005.

[27] J. Branke and J. Gamer, “Efficient sampling in interactive multi-criteriaselection,” in Proceedings of the 2007 INFORMS Simulation SocietyResearch Workshop, 2007.

[28] S. Andradottir, D. Goldsman, B. W. Schmeiser, L. W. Schruben, andE. Yucesan, “Analysis methodology: Are we done?” in Proceedings ofthe 37th conference on Winter simulation, 2005, pp. 790–796.

[29] R. Szechtman and E. Yucesan, “A new perspective on feasibility deter-mination,” in Proceedings of the 2008 Winter Simulation Conference,2008, pp. 273–280.

[30] N. A. Pujowidianto, L. H. Lee, C.-H. Chen, and C. M. Yap, “Optimalcomputing budget allocation for constrained optimization,” in Proceed-ings of 2009 Winter Simulation Conference, 2009, pp. 584–589.

[31] Q.-S. Jia and Q.-C. Zhao, “Strategy optimization for controlled markovprocess with descriptive complexity constraint,” Science in China SeriesF: Information Sciences, vol. 52, no. 11, pp. 1993 – 2005, 2009.

[32] Q.-S. Jia, “On state aggregation to approximate complex value func-tions in large-scale Markov decision processes,” IEEE Transactions onAutomatic Control, vol. 56, no. 2, pp. 333 – 344, 2011.

[33] Y.-C. Ho, Q.-C. Zhao, and Q.-S. Jia, Ordinal Optimization: Soft Opti-mization for Hard Problems. Springer, 2007.

[34] C.-H. Chen, “A lower bound for the correct subset-selection probabilityand its application to discrete-event system simulations,” IEEE Trans-actions on Automatic Control, vol. 41, no. 8, pp. 1227 – 1231, 1996.

[35] D. He, S. E. Chick, and C. H. Chen, “The opportunity cost and OCBAselection procedures in ordinal optimization,” IEEE Transactions onSystems, Man, and Cybernetics–Part C, vol. 37, no. 5, pp. 951 – 961,2007.

[36] L. A. Levin, “Universal sequential search problems,” Problems ofInformation Transmission, vol. 9, no. 3, pp. 265–266, 1973.

[37] R. C. Walker, Introduction to Mathematical Programming. UpperSaddle River, NJ: Prentice & Hall, 1999.

[38] S. Yan, E. Zhou, and C.-H. Chen, “Efficient simulation budget allo-cation for selecting the best set of simplest good enough designs,” inProceedings of the 2010 Winter Simulation Conference, 2010, pp. 1152– 1159.

Page 11: Efficient Selection of a Set of Good Enough Designs with ...enluzhou.gatech.edu/papers/TASE_May10.pdf · Index Terms—optimal computing budget allocation, ranking and selection,

11

Shen Yan received her Bachelors of Engineeringdegree with highest honors from Chinese Universityof Hong Kong, Hong Kong in 2009, and received theMaster of Science degree in Industrial Engineeringfrom the University of Illinois at Urbana-Champaignin 2011. Her research interest is simulation optimiza-tion.

Enlu Zhou received the B.S. degree with highesthonors in electrical engineering from Zhejiang Uni-versity, China, in 2004, and the Ph.D. degree in elec-trical engineering from the University of Maryland,College Park, in 2009. Since then she has been anAssistant Professor at the Industrial & EnterpriseSystems Engineering Department at the Universityof Illinois Urbana-Champaign. Her research interestsinclude Markov decision processes, stochastic con-trol, and simulation optimization. She is a recipientof the Best Theoretical Paper award at the 2009

Winter Simulation Conference and the 2012 AFOSR Young Investigatoraward.

Chun-Hung Chen received his Ph.D. degree inEngineering Sciences from Harvard University in1994. He is a Professor of Systems Engineering &Operations Research at George Mason Universityand is also affiliated with National Taiwan Uni-versity. Dr. Chen was an Assistant Professor ofSystems Engineering at the University of Pennsylva-nia before joining GMU. Sponsored by NSF, NIH,DOE, NASA, MDA, and FAA, he has worked onthe development of very efficient methodology forstochastic simulation optimization and its applica-

tions to air transportation system, semiconductor manufacturing, healthcare,security network, power grids, and missile defense system. Dr. Chen receivedthe “National Thousand Talents” Award from the central government of Chinain 2011, the Best Automation Paper Award from the 2003 IEEE InternationalConference on Robotics and Automation, 1994 Eliahu I. Jury Award fromHarvard University, and the 1992 MasPar Parallel Computer Challenge Award.Dr. Chen has served as Co-Editor of the Proceedings of the 2002 WinterSimulation Conference and Program Co-Chair for 2007 Informs SimulationSociety Workshop. He has served as a department editor for IIE Transactions,associate editor of IEEE Transactions on Automatic Control, area editor ofJournal of Simulation Modeling Practice and Theory, associate editor ofInternational Journal of Simulation and Process Modeling, and associate editorof IEEE Conference on Automation Science and Engineering.