vol.26,no.4,november2001,pp.741–768 printedinu.s.a.mansci-web.uai.cl/~thmello/pubs/derivative...

28
MATHEMATICS OF OPERATIONS RESEARCH Vol. 26, No. 4, November 2001, pp. 741–768 Printed in U.S.A. ESTIMATION OF DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES IN REGENERATIVE SYSTEMS TITO HOMEM-DE-MELLO We investigate the problem of estimating derivatives of expected steady-state performance measures in parametric systems. Unlike most of the existing work in the area, we allow those functions to be nonsmooth and study the estimation of directional derivatives. For the class of regenerative Markovian systems we provide conditions under which we can obtain consistent estimators of those directional derivatives. An example illustrates that the conditions imposed must be different from those in the differentiable case. The result also allows us to derive necessary and sufficient conditions for differentiability of the expected steady-state function. We then analyze the process formed by the subdifferentials of the original process, and show that the subdifferential set of the expected steady-state function can be expressed as an average of integrals of multifunctions, which is the approach commonly found in the literature for integrals of sets. The latter result can also be viewed as a limit theorem for more general compact-convex multivalued processes. 1. Introduction. In recent years a great deal of attention has been devoted to the computation of derivatives of performance measures in stochastic systems. The information provided by those quantities is essential to answer the important question: How much will the performance change if some parameters of the system are slightly changed? More for- mally, suppose we have a stochastic process, say X n , depending on a (vector-valued) parameter , and assume that the process converges to a steady-state X . We would like to compute the gradient of the expected value of the process in equilibrium, i.e., ƐX . A typical example is a G/G/1 queue where the distribution of the service times depends on a parameter (e.g., its mean); we may be interested in computing the sensitivity of the expected waiting time ƐW with respect to the parameter . Notice also that the com- putation (or estimation) of derivatives allows one to take an additional step and develop optimization procedures for the underlying performance measure. Such effort brings obvi- ously numerous benefits, and in fact there have been several papers in the literature deal- ing with that issue. See, for instance, Chong and Ramadge (1994), L’Ecuyer and Glynn (1994), Suri and Leung (1989) and references therein for description of methods and further applications. In general, however, closed-form expressions for the steady-state derivatives cannot be obtained, so one must resort to simulation methods like finite differences, perturbation analysis or likelihood ratios in order to estimate gradients (see, e.g., Glasserman 1991, Glynn 1989, L’Ecuyer 1990, Rubinstein and Shapiro 1993, Suri 1989 for discussions on that topic). In addition, it is necessary to show consistency of such estimators, since the steady- state performance measure of the system under scrutiny is a limiting quantity and hence so is its gradient. Extra assumptions that guarantee some type of uniform convergence, such as convexity (see Hu 1992, Shapiro and Wardi 1994, Robinson 1995), are often imposed Received September 1, 1997; revised August 28, 1998; May 24, 1999, and March 15, 2001. MSC 2000 subject classification. Primary: 60Gxx; secondary: 90C30. OR/MS subject classification. Primary: Probability/stochastic model; secondary: programming/nondifferentiable. Key words. Derivative estimation, nonsmooth optimization, regenerative processes, steady-state systems, multi- functions, convex analysis. 741 0364-765X/01/2604/0741/$05.00 1526-5471 electronic ISSN, © 2001, INFORMS

Upload: others

Post on 20-Mar-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

MATHEMATICS OF OPERATIONS RESEARCHVol. 26, No. 4, November 2001, pp. 741–768Printed in U.S.A.

ESTIMATION OF DERIVATIVES OFNONSMOOTH PERFORMANCE MEASURES IN

REGENERATIVE SYSTEMS

TITO HOMEM-DE-MELLO

We investigate the problem of estimating derivatives of expected steady-state performancemeasures in parametric systems. Unlike most of the existing work in the area, we allow thosefunctions to be nonsmooth and study the estimation of directional derivatives. For the class ofregenerative Markovian systems we provide conditions under which we can obtain consistentestimators of those directional derivatives. An example illustrates that the conditions imposed mustbe different from those in the differentiable case. The result also allows us to derive necessary andsufficient conditions for differentiability of the expected steady-state function. We then analyze theprocess formed by the subdifferentials of the original process, and show that the subdifferential setof the expected steady-state function can be expressed as an average of integrals of multifunctions,which is the approach commonly found in the literature for integrals of sets. The latter result canalso be viewed as a limit theorem for more general compact-convex multivalued processes.

1. Introduction. In recent years a great deal of attention has been devoted to thecomputation of derivatives of performance measures in stochastic systems. The informationprovided by those quantities is essential to answer the important question: How much willthe performance change if some parameters of the system are slightly changed? More for-mally, suppose we have a stochastic process, say �Xn����, depending on a (vector-valued)parameter �, and assume that the process converges to a steady-state X����. We would liketo compute the gradient of the expected value of the process in equilibrium, i.e., �ƐX����.A typical example is a G/G/1 queue where the distribution of the service times dependson a parameter � (e.g., its mean); we may be interested in computing the sensitivity of theexpected waiting time ƐW���� with respect to the parameter �. Notice also that the com-putation (or estimation) of derivatives allows one to take an additional step and developoptimization procedures for the underlying performance measure. Such effort brings obvi-ously numerous benefits, and in fact there have been several papers in the literature deal-ing with that issue. See, for instance, Chong and Ramadge (1994), L’Ecuyer and Glynn(1994), Suri and Leung (1989) and references therein for description of methods and furtherapplications.In general, however, closed-form expressions for the steady-state derivatives cannot be

obtained, so one must resort to simulation methods like finite differences, perturbationanalysis or likelihood ratios in order to estimate gradients (see, e.g., Glasserman 1991,Glynn 1989, L’Ecuyer 1990, Rubinstein and Shapiro 1993, Suri 1989 for discussions on thattopic). In addition, it is necessary to show consistency of such estimators, since the steady-state performance measure of the system under scrutiny is a limiting quantity and hence sois its gradient. Extra assumptions that guarantee some type of uniform convergence, suchas convexity (see Hu 1992, Shapiro and Wardi 1994, Robinson 1995), are often imposed

Received September 1, 1997; revised August 28, 1998; May 24, 1999, and March 15, 2001.MSC 2000 subject classification. Primary: 60Gxx; secondary: 90C30.OR/MS subject classification. Primary: Probability/stochastic model; secondary: programming/nondifferentiable.Key words. Derivative estimation, nonsmooth optimization, regenerative processes, steady-state systems, multi-functions, convex analysis.

741

0364-765X/01/2604/0741/$05.001526-5471 electronic ISSN, © 2001, INFORMS

Page 2: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

742 T. HOMEM-DE-MELLO

for that purpose. Further discussion on steady-state derivatives, especially on the topic ofdifferentiability of measures, can be found in Pflug (1996) and Glynn and L’Ecuyer (1995).A particularly neat situation occurs when the underlying process possesses a regenerative

structure; i.e., it “restarts” itself at some points in time (see, for instance, Asmussen 1987).In such cases one can estimate steady-state quantities based on the behavior of the processover regenerative cycles, thus avoiding “warm-up” periods that are typically necessary insimulation (see Bratley, Fox and Schrage 1987). This is expressed in the formula

ƐX����=Ɛ[∑����−1

n=0 Xn���]

Ɛ����

where ���� is the length of a cycle. Notice however that the cycle length often dependson the parameter �, thus making differentiation of those quantities a difficult task. In somecases this problem can be overcome by using the likelihood ratio technique (see, e.g.,Rubinstein and Shapiro 1993, Glasserman and Glynn 1992). The situation is also remediedif the derivative process ��Xn���� regenerates at the same epochs as the original process;we then have that, under some additional assumptions,

�ƐX����=Ɛ[∑����−1

n=0 �Xn���]

Ɛ�����(1.1)

Glasserman (1993) studies conditions under which (1.1) holds for Markov processes (in theone-dimensional case), thus providing a convenient way to estimate the desired derivatives.Most of the work found in the literature treats the case where the expected steady-

state function is differentiable. Indeed, in many situations this is actually the case, forinstance in queueing networks in which the distributions of arrival and service times havedensities (see Suri 1989, Wardi and Hu 1991). This differentiability property, however,does not hold in general; in fact, Shapiro and Wardi (1994) show by a simple examplethat not only can the expected steady-state function be nondifferentiable, but also the setof nondifferentiable points can be a dense subset of the domain. In such cases one isconcerned with estimating subgradients, or more generally, directional derivatives (see §2for definitions). An important contribution in that respect is the work by Robinson (1995):Under the assumption that the functions Xn�· �� are convex w.p.1, Robinson shows thatsubgradients of the expected steady-state function can be consistently estimated; however,no claim can be made regarding the whole subdifferential set, since in that setting onecan only show that limits of subgradients are contained in the subdifferential set of theexpected steady-state function. An interesting application of nonsmooth optimization to themaximization of steady-state throughput in a tandem production line, using the techniquesof Robinson (1995), can be found in Plambeck et al. (1996).In this paper we investigate the problem of estimating directional derivatives of

nonsmooth expected steady-state performance measures of Markov processes. We provideconditions under which we can obtain consistent estimators of those directional derivatives.The basic idea is to show that, under those conditions, the directional derivative functionprocess �X ′

n��0� ·�� regenerates at the same points as the original process �Xn��0�� andhence the directional derivatives of the expected value function can be expressed both as along-run average and as a ratio formula. Here, regeneration plays an essential role: With-out this condition, consistency is unlikely to hold, since typically pointwise convergenceof functions does not imply convergence of directional derivatives. Notice that the situa-tion is more restricted than in the differentiable case, where under proper assumptions—convexity is an example—convergence of derivatives follows from pointwise convergence.An example given in §3 shows that the analysis of regeneration of the directional deriva-tive process is more complicated than in the differentiable case: We exhibit two randomvariables depending on �, say Y1��� and Y2���, such that Y1��� and Y2��� are independentand identically distributed for each �, but their subdifferential sets at � = 0 do not have the

Page 3: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 743

same distribution (the situation can be easily extended to regenerative processes in general).As a consequence, we give a necessary and sufficient condition for the differentiability ofthe expected steady-state function, thus extending a result by Shapiro and Wardi (1994).We then rewrite the results obtained for directional derivatives in terms of subdifferential

sets. Although this reformulation is an immediate consequence of the correspondencebetween sublinear functions and compact convex sets (cf. Hiriart-Urruty and Lemaréchal1993a), the importance of subdifferentials in nonsmooth optimization theory, in terms ofalgorithms and optimality conditions (see Rockafellar 1970, Ioffe and Tihomirov 1979,Hiriart-Urruty and Lemaréchal 1993a,b) justify per se the re-statement of those results.This allows us to draw an important conclusion: The subdifferential of the expectedsteady-state function �ƐX���0� can be expressed both as a limiting average of sets(N−1∑N−1

n=0 �Xn��0�)—where the sum is understood as the Minkowski addition of setsA+ B = �a+ b � a ∈ A b ∈ B�—and as an average of sets over a regenerative cycle(Ɛ∑�−1

n=0 �Xn���/Ɛ�). The latter expression—which involves the integral of a set—can becomputed by using the theory of integration of multifunctions found in the literature (seefor instance Castaing and Valadier 1977, Hiai and Umegaki 1977, Rockafellar 1976). Weshow that this limiting result is valid in general for regenerative compact-convex multivaluedprocesses.The paper is organized as follows: In §2 we define the set-up for the problem and review

concepts on regenerative processes, convex analysis and vector measures. In §3 we illustratewith an example some difficulties of the problem and give conditions to ensure regenerationof the directional derivative functions. In §4 we show that under some additional assump-tions a ratio formula like (1.1) holds for the directional derivatives and then we exhibit con-sistent estimators of the derivatives of the expected steady-state function. As a consequence,we obtain a necessary and sufficient condition for the differentiability of that function. In §5we restrict ourselves to the case of subdifferentiable functions and apply the theory of inte-grals of multifunctions to the results obtained in the previous sections. Section 6 presentstwo examples of application of the results, and in §7 we make some concluding remarks.

2. Background and basic assumptions. In this section we review some concepts thatwill be used in the sequel. The results presented are known, but some of them are provedhere for the sake of completeness.We start with concepts on regenerative processes. Following Asmussen (1987), we say

that a vector-valued process �Yn� is regenerative if there exists a sequence of iid spacings��j� such that, for each j ≥ 1, the process{

�j+1 �j+2 � � � �Y�j+m m= 0 1 � � � �}

(where �j = �1+· · ·+�j) is independent of �1 � � � �j and its distribution does not depend onj. Observe that this definition does not impose that �Y�j+m m= 0 1 � � � � be independent of�Ym m= 0 � � � �j−1�. Other definitions of regeneration exist; see, for instance, Thorisson(2000) for a detailed discussion.The importance of regenerative processes has both theoretical and practical aspects. On

the theoretical side, it can be shown that if Ɛ�1 <�, then the process �Yn� has a limitingdistribution defined as

F��y�= limN→�

1N

N−1∑n=0

P�Yn ≤ y��(2.1)

Notice that we are following the concept of a limiting distribution in a time-average sense—as defined by Wolff (1989)—rather than in a pointwise sense. Typically, for the latter oneto exist it is necessary to further assume other conditions, such as “�1 does not have a

Page 4: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

744 T. HOMEM-DE-MELLO

lattice distribution.” Since the goal of this paper is to establish conditions for regeneration insome sense, we shall follow the time-average definition in (2.1) for simplicity. Nevertheless,all the results shown in this paper carry out to the pointwise case by imposing the extraassumptions mentioned above. We refer to Wolff (1989) for a comprehensive discussion onthe different concepts of limiting distributions.Let Y� denote a random variable with the limiting distribution defined in (2.1), and let f

be any measurable function. If Ɛ∑�1−1

n=0 �f �Yn�� <�, then we have that

Ɛf �Y��=Ɛ[∑�1−1

n=0 f �Yn�]

Ɛ�1�(2.2)

Furthermore, the same quantity is given by time-averages; that is,

Ɛ f �Y��= limN→�

1N

N−1∑n=0

f �Yn� w.p.1�(2.3)

On the practical side, formula (2.2) provides a way to estimate functions of the process insteady-state, yielding a procedure that is usually called regenerative simulation. Supposewe generate m cycles of sizes �1 � � � �m. Then w = Ɛf �Y�� can be estimated by

w =∑m

i=1

∑�i−1n=�i−1

f �Yn�∑mi=1 �i

The variance of w can be estimated from the same sample; see for instance Shedler (1987)for details.Let us now define formally the underlying processes and make some basic assumptions.

Consider a vector-valued stochastic process X��� �= �Xn���� �= ��X1n��� � � � X

Kn �����,

n = 0 1 � � � , defined on a common probability space �& � P�, and depending on anm-dimensional parameter � belonging to some open set ' ⊂ �m. The choice for vector-valued rather than scalar-valued processes is driven mainly by the fact that many applica-tions fall into that category, like queueing networks for instance. As we shall see later, theanalysis is basically the same for both cases; there are, however, some exceptions, as in thediscussion in the end of §3.Since we are mostly interested in studying the derivatives at some fixed point, we shall

fix from now on some �0 ∈'. Assume the following:Assumption A1. �Xn���� is a Markov process for each � ∈' (possibly with continuous

state space), and the process �Xn��0�� is regenerative with regeneration epochs ��m� m≥ 0,and �0 = 0.Assumption A1 is the basic starting point, since our main goal is to find conditions under

which the derivative process regenerates together with the original process at the points��m�. This assumption will be complemented later, as we shall see in §3. For now, only�Xn��0�� is assumed to be regenerative; the processes �Xn���� for � �= �0 are assumed tobe just Markovian.Before proceeding further, let us review some concepts of convex analysis which will be

used in the sequel (basic references are Rockafellar 1970 and Hiriart-Urruty and Lemaréchal1993a). Let f � ' → � be an arbitrary function. For any � ∈ ', the directional derivativeof f at � in the direction d, when it exists, is given by

f ′���d�= limt↓0

f ��+ td�−f ���

t�(2.4)

Notice that f ′��� ·� is positively homogeneous, i.e., f ′���0�= 0 and f ′���*d�= *f ′���d�for all d ∈ �m and all * > 0.

Page 5: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 745

Definition. A function f � ' ⊂ �m → � is said to be directionally differentiable at�0 ∈ ' if the directional derivative f ′��0�d� defined in (2.4) exists and is finite for alld ∈ �m, and the function f ′��0� ·� is continuous.Definition. A function f � ' ⊂ �m → � is said to be subdifferentiable at �0 ∈ ' if it

is directionally differentiable and, in addition, f ′��0� ·� is convex.Notice that our concept of directional differentiability is stronger than the usual one found

in the literature, since we assume the directional derivative function to be continuous. Whenf is subdifferentiable, we can define the subdifferential of f at � as the set supported bythe function f ′��� ·�, that is,

�f ���= �� ∈ �m � �� d� ≤ f ′���d� for all d ∈ �m��(2.5)

It follows that the subdifferential �f ��� is a convex and compact set. Furthermore, it followsthat in this case the directional derivative f ′���d� is a sublinear function of d (we say thata function g is sublinear if g�*1d1 +*2d2� ≤ *1g�d1�+*2g�d2� for all d1 d2 ∈ �m andall *1 *2 > 0. It can be shown—see Proposition V.1.1.4 in Hiriart-Urruty and Lemaréchal(1993a)—that g is sublinear if and only if g is positively homogeneous and convex). Noticethat when f is locally Lipschitz (as defined below) the concept of subdifferentiabilityincludes that of regularity introduced by Clarke (1990), that is, regular functions are sub-differentiable. Moreover, if f is regular then the subdifferential coincides with the so-calledgeneralized gradient (see Clarke 1990). If f is convex, those two concepts also coincidewith the standard subdifferential of convex analysis (see, e.g., Rockafellar 1970).Relation (2.5) can actually be generalized to arbitrary finite-valued sublinear functions,

resulting in an interesting connection between that class of functions and compact convexsets. Given a finite-valued sublinear function -�·� on �m we can define a compact convexset C- ⊂ �m as

C- = �� ∈ �m � �� d� ≤ -�d� for all d ∈ �m��

Conversely, given a compact convex set C ⊂�m we can construct a finite sublinear function-C�·� on �m as

-C�d�=max��� d� � � ∈ C�

for all d ∈ �m. This correspondence can be shown to be an isometric isomorphism, andmany results can be derived from this equivalence. In §V.3 of Hiriart-Urruty and Lemaréchal(1993a), for instance, one can find a comprehensive discussion of that topic.An important situation occurs when the directional derivative function f ′��0� ·� is linear,

i.e., there exists a vector D�0∈ �m such that f ′��0�d� = �D�0

d� for all d ∈ �m. In thatcase f is said to be Gâteaux-differentiable at �0. A stronger concept is that of Fréchetdifferentiability: f is said to be Fréchet-differentiable at �0 if there exists a vector D�0

∈�m

such thatlimd→0

�f ��0+d�−f ��0�−�D�0 d��/�d� = 0�

It is easy to see that Fréchet-differentiability implies Gâteaux-differentiability. The converseis true if f is locally Lipschitz, i.e., if there exists a positive constant M such that

�f ��1�−f ��2�� ≤M��1−�2�

for all �1 �2 in a neighborhood of �0. If f = �f1 � � � fK� is a mapping from ' to �K we saythat f is directionally (resp. Gâteaux, Fréchet) differentiable at �0 if each fi is directionally(resp. Gâteaux, Fréchet) differentiable at �0, i = 1 � � � K. For a comprehensive discussionon these concepts, including the generalization to infinite-dimensional spaces, see Shapiro(1990) and references therein.

Page 6: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

746 T. HOMEM-DE-MELLO

It is clear from the above discussion that the set of directionally differentiable functionsincludes the Gâteaux-differentiable functions. Another important class included in that cat-egory are the convex continuous functions. More generally, functions that can be writ-ten as a difference of convex continuous functions (sometimes called DC-functions inthe literature) are directionally differentiable. Notice also that composite functions of theform f ��� = g�A����, where g�·� is convex continuous and A � �m → �k is a Gâteaux-differentiable mapping, are directionally differentiable.The ideas discussed so far suggest that a fairly general setting is obtained if we assume

that the underlying functions are directionally differentiable. The goal is then to estimatethe directional derivatives of the expected value function x��·�= ƐX��·� at �0, by usingthe directional derivatives of the process �Xn����. In §5 we specialize our results for thesubdifferentiable case.In studying estimation procedures for the directional derivatives x′

���0�d�, it will beuseful to consider the function x′

���0� ·� rather than specific values of d. We proceed nowin defining the directional derivative process �X ′

n��0� ·�� on an appropriate space. We cannow state the second assumption:Assumption B1. For each n = 0 1 2 � � � and P-almost all � ∈ &, the sample-path

functions Xn�· �� are directionally differentiable at �0.Assumption B1 guarantees, of course, that we can actually take the directional derivatives

pathwise. As pointed out before, a fairly large class of functions satisfies that property.Observe that the state space of each �Xk

n�′��0� ·�, k= 1 � � � K, is the space H+��m� of posi-

tively homogeneous continuous functions defined on �m. This space is clearly isomorphic tothe space C�Sm−1� of continuous functions whose domain is the unit sphere Sm−1 of �m (i.e.,the set �d ∈�m � �d�= 1�) since we can always write f �d�= f ��d�d/�d��=�d�f �d/�d��for d �= 0. Endow C�Sm−1� and H+��m� with the sup-norm

�-� = maxd∈Sm−1

�-�d�� (2.6)

and let �H denote the corresponding collection of Borel sets in H+��m�. Now,�H+��m� �H� is a measurable space and thus, in order to show that each �Xk

n�′��0� ·� is

well defined as a random variable with range in H+��m� we must show that �Xkn�

′��0� ·� is� −�H measurable, i.e., the inverse image of a set in �H belongs to the --field � . Thelemma below shows that this is actually the case.

Lemma 2.1. Let n ∈ �0 1 � � � � and k ∈ �1 � � � K� be arbitrary. Then, underassumption B1, the function 5 � & → H+��m�, defined by 5����·� = �Xk

n�′��0� · ��, is

� −�H measurable.

Proof. Let 5��·� denote the function 5����·�. In order to show that 5 is � −�H

measurable, it is sufficient to show that, given any function -0 ∈ H+��m� and any 6 > 0,the set

A-0= �� ∈& � �5�−-0� ≤ 6�

is � -measurable. Notice that

A-0= �� ∈& � �5��d�−-0�d�� ≤ 6 for all d ∈ Sm−1�

= ⋂d∈Sm−1

�� ∈& � �5��d�−-0�d�� ≤ 6��(2.7)

Observe that, for a fixed d ∈ Sm−1, the set ��∈& � �5��d�−-0�d�� ≤ 6� is � -measurable. Tosee this, notice that 5��d� is defined by the limit of the difference of two measurable func-tions (see the definition (2.4) of the directional derivative) and thus 5��d� is measurable in�. Next, let D be a countable dense subset of Sm−1. Since 5� and -0 are continuous, itfollows that the intersection in (2.7) can be taken over the countable set D instead of Sm−1

and hence we conclude that A-0is � -measurable. �

Page 7: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 747

Let H+K ��

m� be the cartesian product H+K ��

m�=H+��m�×· · ·×H+��m� (K times). Let�HK

denote the corresponding collection of Borel sets of H+K ��

m�. Notice that, because ofthe isomorphism between the metric spaces �C�Sm−1� ��� and �H+��m� ���, it follows thatH+��m� is a separable Banach space since so is C�Sm−1� (see, e.g., Royden 1988) and henceH+

K ��m� is also a separable Banach space. Thus, it makes sense to study the regeneration

(or more generally, Harris recurrence) of �X ′n��0� ·�� �= ���X1

n�′��0� ·� � � � �XK

n �′��0� ·���

as a process on �H+K ��

m� �HK� (see Revuz 1984 or Nummelin 1984 for details on the

latter topic). Note also that �X ′n��0� ·�� is a Markov process with respect to the filtrations

generated by �Xn�·��.Consider now an arbitrary random variable 5 on �H+��m� �H�. As seen above, 5 takes

on values in the separable Banach space H+��m�, so it is clear that the standard Lebesgueintegral cannot be used to compute the expected value Ɛ5. Nevertheless we can resort tothe so-called Bochner integral, which is in a sense an extension of the Lebesgue integralfrom the real line to a Banach space. Concepts and main results about the Bochner integralcan be found for instance in Diestel and Uhl (1977) or Kelley and Srinivasan (1988). Fornow we need the following definitions (cf. Diestel and Uhl 1977, pp. 41, 45):Definition. A function 7 � & → H+��m� is called simple if there exist f1 � � � fN ∈

H+��m� and E1 � � � EN ∈ � such that 7���=∑Ni=1 fi 1Ei

���.Definition. A function 5 � &→ H+��m� is called strongly measurable if there exists

a sequence of simple functions �7n� such that limn→� �5���−7n���� = 0 for P-almostall �.Definition. A strongly measurable function 5 � & → H+��m� is called Bochner

integrable if Ɛ�5� <�.The importance of strongly measurable integrable functions lies in the fact that for this

class of functions the Bochner integral is defined in a natural way. If 7 =∑Ni=1 fi1Ei

isa simple function, its Bochner integral is defined as

∫&7���P�d�� =∑N

i=1 fiP�Ei�. Nowlet 5 be a strongly measurable integrable function and let �7n� be a sequence of simplefunctions converging to 5. Then limn

∫7n exists (see, e.g., p. 45 in Diestel and Uhl 1977)

and hence we can define ∫&5���P�d�� �= lim

n→�

∫&7n���P�d���(2.8)

It is important to note that, because of the separability of H+��m�, � −�H measurablefunctions are strongly measurable; for a proof of this fact, see Lemma 10.18 in Kelley andSrinivasan (1988) (observe that those authors use a different definition for strong measur-ability; the result stated above is translated in terms of the nomenclature adopted here).Hence, by Lemma 2.1 we have that, given �0 ∈', each �Xk

n�′��0� ·� is strongly measurable.

The following assumption will then ensure that Ɛ�Xkn�

′��0� ·� exists:Assumption B2. For each n = 0 1 2 � � � , each k = 1 � � � K, the H+��m�-valued

random variables �Xkn�

′��0� ·� are Bochner integrable.Of course, definition (2.8) does not tell us how to evaluate the resulting function

∫5

pointwise unless we compute the approximating simple functions, which in general cannotbe easily done. The following lemma shows that we can actually evaluate the integralfunction at a point by computing the integrals pointwise.

Lemma 2.2. Let 5 be a Bochner integrable random variable on �H+��m� �H�, and let9 = ∫

&5���P�d�� be the Bochner integral of 5. Then, for each d ∈ �m we have that

9�d�=∫&5����d�P�d�� (2.9)

Page 8: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

748 T. HOMEM-DE-MELLO

where the integral on the right-hand side is the standard Lebesgue integral, and 5����d�denotes the value of 5��� at d.

Proof. Let �H+��m��∗ denote the dual space of H+��m�, i.e., the space of all boundedlinear functionals on H+��m�. Let F ∈ �H+��m��∗. From Lemma 10.21 in Kelley andSrinivasan (1988) we know that F�5� is Lebesgue integrable and

F�9�= F

(∫&5���P�d��

)=∫&F�5����P�d���(2.10)

Next, observe that because of the isomorphism between H+��m� and C�Sm−1� we have that�H+��m��∗ is isomorphic to the space of Borel measures on Sm−1; see for instance p. 358 inRoyden (1988). Hence, given a Borel measure : on Sm−1 there is a unique F ∈ �H+��m��∗

such that

F�;�=∫Sm−1

; d:

for all ; ∈ H+��m�. For each d ∈ Sm−1, let <d denote the atom measure with mass oneat d, and let Fd denote the corresponding linear functional. Clearly, we have that, for any; ∈H+��m�,

Fd�;�=∫Sm−1

; d<d = ;�d�

and by substituting Fd into equation (2.10) it follows that

9�d�= Fd�9�=∫&Fd�5����P�d��=

∫&5����d�P�d��

as stated. �

Notice that the above lemma also implies that if 5 is a Bochner integrable random variableon �H+��m� �H�, then each real-valued random variable 5����d�, d ∈ �m, is (Lebesgue)integrable.

3. Local conditions for regeneration. As remarked in §1, our objective is to studythe regeneration of subdifferentials (or, more generally, directional derivatives) based onthe regeneration of the original process. A much simpler but fundamental question hiddenunder that is: Suppose Y �·� is a random function of a one-dimensional parameter �. Howdoes the distribution of �Y ��0� for some �0 ∈ ' relate to the distribution of the Y ���’s?Let us consider initially the differentiable case. Suppose there is a set A ⊂ & such thatP�A�= 1 and Y �· �� is differentiable at �0 for all � ∈ A. Suppose that Y ��0� is constant,i.e., Y ��0 ��= y0 for some y0 and all �. Then we have that, for � ∈ A, �Y ��0�= �Y ′��0��and

Y ′��0 ��= limh→0

Y ��0+h ��−Y ��0�

h�

Therefore, since convergence w.p.1 implies convergence in distribution, it follows that thedistribution of Y ′��0� is determined by the marginal distribution of the process Y on aneighborhood of �0, that is, by the distribution of each Y ��0 +h� for 0 < h < 6, where 6is any positive fixed number. We must stress here the importance of the assumption thatY ��0� is constant w.p.1 in the above argument; otherwise, different representations of theoriginal process might lead to different derivatives, and in that case one must deal with theso-called process derivatives (see Pflug 1996 for details).In the nondifferentiable case, however, more is needed, as can be verified from the

following simple example. Let &= ��1 �2 �3 �4� with P��i�= 14 , i= 1 � � � 4. Let Y1���

Page 9: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 749

and Y2��� be random variables defined for each � ∈' in the following way:

Y1�� �1�= ��� Y2�� �1�= �+

Y1�� �2�= ��� Y2�� �2�= �−

Y1�� �3�= 0 Y2�� �3�= �+

Y1�� �4�= 0 Y2�� �4�= �−

where �+ = max�� 0�, �− = max�−� 0�. It is clear that Y1��� and Y2��� are independentfor each �. Notice that at �0 �= 0 we have Y1��0� = Y2��0� = 0 w.p.1 and, at � �= 0 anda ∈ � we have that

P�Y1���= a�= P�Y2���= a�={ 1

2 if a= ��� or a= 0

0 otherwise

so Y1��� and Y2��� have the same distribution for any fixed �. Furthermore, at any � �= 0Y1�·� and Y2�·� are differentiable w.p.1 and

P�Y ′1���= a�= P�Y ′

2���= a�={ 1

2 if a= sign��� or a= 0

0 otherwise�

However, at �0 = 0 we have

P��Y1�0�= C� ={ 1

2 if C = −1 1 or C = �0�

0 otherwise,

and

P��Y2�0�= C� ={ 1

2 if C = −1 0 or C = 0 1

0 otherwise

so �Y1�0� and �Y2�0� do not have the same distribution. Note that Y1 is nondifferentiable atzero with probability 1

2 , whereas Y2 is nondifferentiable at zero with probability 1.The above example illustrates the nature of the problem. At a point � where the functions

are differentiable, the subdifferential is just a singleton and hence it is determined by anyof the one-sided derivatives (i.e., the directional derivative at �0 in the direction +1 or −1).At the nondifferentiable point �0 = 0, on the contrary, the subdifferential is given by the setwhose support function is the directional derivative at 0, in this case the convex hull of left-and right-side derivatives, so its distribution depends on the joint distribution of Y ′

i �0�1�and Y ′

i �0�−1�. In summary, knowledge of the distribution of each Y ��� is not sufficient todetermine the distribution of �Y ��0�. On the other hand, of course, the law of �Y ��0� canbe computed if one knows the distribution of the random function Y �·�, that is, all the finite-dimensional distributions �Y ��1� Y ��2� � � � Y ��l��, where l and �1 � � � �l are arbitrary.

The preceding discussion suggests that, in order to obtain conditions for regeneration ofsubdifferentials, one should impose conditions on the random functions Xn�· �� rather thanon each Xn�� ��. This however appears to be quite restrictive, since regeneration of wholefunctions is unlikely to occur in real situations. A more flexible alternative is to study thebehavior of the process �Xn�·�� on (random) neighborhoods of the fixed point �0. The key(although obvious) observation here is that the knowledge of a directionally differentiablefunction f on any neighborhood of �0 is sufficient to compute the directional derivativefunction f ′��0� ·�, as can be verified directly from definition (2.4).The lemma below establishes the connection between the distribution of the directional

derivative and the distribution of the function on a fixed (i.e., independent of �)neighborhood of �0.

Page 10: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

750 T. HOMEM-DE-MELLO

Lemma 3.1. Let V be a neighborhood of �0, and let Y �·� be a real-valued directionallydifferentiable random function on '. Then, the distribution of Y ′��0� ·� is determined by thedistribution of Y �V �·� (the restriction of Y to V ) on cylinder sets, that is, by the elementsof the form

P�Y ��1� ∈ B1 � � � Y ��k� ∈ Bk�

for all k ≥ 1, all �1 � � � �k ∈ V and all B1 � � � Bk Borel sets of the real line.

Proof. Notice first that the distribution of the random function Y ′��0� ·� is given by itsfinite-dimensional distributions(

Y ′��0�d1� Y′��0�d2� � � � Y

′��0�d@�)

for all @ and all d1 � � � d@. Fix then @ d1 � � � d@ and, for each t≥ 0 and each j = 1 � � � @,let Wdj

�t� denote the random variable

Wdj�t�= Y ��0+ tdj�−Y ��0�

t�

By assumption, the joint distribution of Y ��0+ td1� � � � Y ��0+ td@� and Y ��0� is knownfor t small enough. It follows that the joint distribution of(

Wd1�t� � � � Wd@

�t�)

can be computed for t small enough. Moreover, it is clear that Wdj�t�→ Y ′��0�dj� w.p.1

as t → 0 for each j = 1 � � � @, and hence

(Wd1

�t� � � � Wd@�t�) d→ �Y ′��0�d1� � � � Y

′��0�d@��

(whered→ denotes convergence in distribution). The assertion of the lemma follows. �

Clearly, the above lemma implies that, if there exists a neighborhood V of �0 such that therestricted functions Xn�V �·� regenerate, then X ′

n��0� ·� regenerates at the same epochs. Wewould like however to have a less strict condition, since the sample paths Xn�· �� � ∈&,can be quite different from each other—so it is unlikely that they have the same propertiesaround a fixed neighborhood. Similarly, the functions Xn�· �� and Xm�· �� may havedifferent behaviors for n �= m. In other words, we want to allow the neighborhood V todepend on n and �, and still ensure some kind of regeneration. Assumption A2 below is astep in that direction.Recall that, by Assumption A1, ��m� are regeneration points of Xn��0�. We now extend

that assumption and state the main result.Assumption A2. There exist K-dimensional vector-valued directionally differentiable

random functions �Fn�·�� n ≥ 0, which are identically distributed (i.e., have the samefinite-dimensional distributions) and independent of the regeneration epochs ��m�, and, forP-almost all � ∈ &, neighborhoods �Vn��� = �V 1

n ��� � � � V Kn ����� of �0 such that, for

all m n≥ 0, on the event ��m = n� the restricted functions Xn�Vn����· �� and Fn�Vn����· ��coincide.Verification of Assumption A2 must be done on a case-by-case basis. In §6 we shall see

some typical examples of ways to check this condition.

Theorem 3.1. Suppose that Assumptions A1, B1 and A2 hold. Then, the process��Xn��0�, X

′n��0� ·��� regenerates at the epochs ��m� m≥ 0.

Page 11: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 751

Proof. Let m≥ 0. By Assumption A2, on the event ��m = n� we have �Xn�Vn�′��0� ·�=

�Fn�Vn�′��0� ·� and thus

X ′n��0� ·�= �Xn�Vn

�′��0� ·�= �Fn�Vn�′��0� ·�= F ′

n��0� ·�

where the above equalities are understood to hold for each � such that �m = n. But fromLemma 3.1 we have that F ′

n��0� ·� d= F ′1��0� ·�, since Fn�·� d= F1�·�. Now, since the functions

�Fn�·�� are independent of the regeneration epochs, it follows that �X�m��0� X

′�m��0� ·��

is independent of �1 � � � �m−1 and independent of m. Moreover, by the MarkovianAssumption A1, we have that X ′

�m+j��0� ·� has the same distribution as X ′j��0� ·�, j =

0 1 � � � . It follows that the process ��Xn��0� X′n��0� ·��� regenerates at �m. �

A particular situation occurs when the original process X��� satisfies a recursion of thetype

Xn+1���= 5�Xn��� Un���� (3.1)

where �Un���� for any fixed � are iid random vectors independent of X���. In those caseswe can often obtain an explicit regenerative structure. In Glasserman (1993), it is assumedthat the original process �Xn��0�� has the following structure: There are an integer r ≥ 1, arecurrent set B⊂�K (“recurrent” here means that B is visited infinitely often w.p.1), subsetsA1 � � � Ar of �

K , and a function h ��rK →�K such that, if Xn��0� ∈ B and Un+i��0� ∈Ai,i = 1 � � � r , then Xn+r ��0� = h�Un��0� � � � Un+r ��0��. This condition can be viewed asa more explicit version of the splitting condition for Harris recurrent chains. Recall that aMarkov process �Zn� with state space �S �S� is a Harris chain if there exists a set A ∈ Ssuch that A is visited infinitely often (say, at times T1 T2 � � � ) and there exist numbersr > 0, p ∈ �0 1 and a probability measure F on �S �S� such that, for any set B ∈�S ,

P�Zr ∈ B �Z0 = x�≥ pF�B� for all x ∈ A�(3.2)

It is possible to show that the above condition implies that the original probability space canbe extended to support 0-1 variables Ik with the following property: P�Ik = 1 �ZTk

= x�= pfor x ∈A and, when the process reaches the set A for the kth time, it regenerates r steps laterif Ik = 1. For details on this construction, see Thorisson (2000). It is easy to see that, in thepresent case, (3.2) holds for �Xn��0�� with p =∏r

i=1 pi, where pi = P�U1��0� ∈ Ai�. Underthis splitting condition, assuming that B and Ai are open and some other differentiabilityassumptions, Glasserman proves that the process ��Xn��0� X

′n��0��� is regenerative. We

refer again to Glasserman (1993) for details.Notice that the assumption that B and Ai are open is actually equivalent to assuming

pointwise regeneration of X��� on neighborhoods of �0, since in that case the splittingcondition can be applied to �Xn���� for � near �0. By strengthening the assumption on theinput functions Un�·�, Theorem 3.6 in Glasserman (1993) can be modified as follows:

Theorem 3.2. Assume A1 and B1, and suppose there is an integer r ≥ 1, a recurrentset B ⊂ �K , subsets A1 � � � Ar of �

K , a Fréchet-differentiable function h � �rK → �K andneighborhoods �Vn���= �V 1

n ��� � � � V kn ����� of �0 such that

if Xn��0 �� ∈ B and Un+i��0 �� ∈ Ai i = 1 � � � r

then Xn+r �· ��= h�Un�· �� � � � Un+r �· ��� on Vn+r ����(3.3)

Suppose further that the input functions �Un�·�� n ≥ 0, are independent and identicallydistributed, and pathwise directionally differentiable. Then, ��Xn��0� X

′n��0� ·��� is

regenerative.

Page 12: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

752 T. HOMEM-DE-MELLO

Proof. Let �-j j ≥ 1� be the regeneration epochs

-j = inf�q > -j−1 � Xq−r ��0� ∈ B and Uq−r+i��0� ∈ Ai i = 1 � � � r��

On �-j = q� we have that Xq�· �� = h�Uq−r �· �� � � � Uq�· ��� on each neighborhoodVq���. Now, let Fq be a random function from ' to RK defined by Fq�·� =h�Uq−r �·� � � � Uq�·��. It can be shown that F is pathwise directionally differentiable, sinceeach Un�·� is directionally differentiable and h is Fréchet-differentiable (see Proposition 3.6in Shapiro 1990). Moreover, the Fq�·� q ≥ 0, are identically distributed. It follows thatAssumption A2 is satisfied and hence the result follows from Theorem 3.1. �

4. Obtaining consistent derivative estimators. One of the most useful applicationsof regenerative processes is the estimation of steady-state quantities by ratio-type formulaslike (2.2). Indeed, from a simulation standpoint, by using the regenerative structure oneavoids “warm-up” periods typically necessary when computing time-averages like (2.3) and,furthermore, variances and consequently confidence intervals can be constructed by standardapplication of the Central Limit Theorem. In many systems, however, the regenerative cyclescan be extremely long, thus making the use of ratio formulas not feasible from a practicalpoint of view. Nevertheless, even in such cases regeneration plays an important role, namelythat of ensuring the existence of a steady-state under mild assumptions. In this sectionwe discuss these issues with respect to the directional derivatives. As we shall see, theassumptions of the previous sections, together with some regularity conditions, guaranteethat the directional derivatives of the expected value function can be expressed both asratio-type and as time-average formulas.Suppose that the processes �Xn���� (for all � in a neighborhood of �0) and �X ′

n��0� ·�� areregenerative with the same regeneration epochs, and suppose that the iid cycle times ��n�are such that Ɛ�1 <�, so a limiting distribution exists. In order to simplify the discussion,we assume throughout this section that K = 1, i.e., Xn��� is a scalar. To ensure the existenceof ratio-type results, we shall also extend Assumption B2 as follows:Assumption B3. For all � in a neighborhood of �0, the random variable

∑�1−1n=0 Xn���

is integrable. Moreover, the H+��m�-valued random variable∑�1−1

n=0 X ′n��0� ·� is Bochner

integrable.Let xn���= ƐXn���, 0≤ n≤�, where X���� denotes the weak limit of Xn���. Recall

that our main goal is to estimate the steady-state directional derivative x′���0� ·� provided it

exists. A direct application of the ratio formula (2.2) gives us

x����= ƐX����=Ɛ[∑�1−1

n=0 Xn���]

Ɛ�1(4.1)

and

ƐJ�0�·�= Ɛ

[∑�1−1n=0 X ′

n��0� ·�]

Ɛ�1 (4.2)

where J�0�·� is the weak limit of X ′

n��0� ·�, and the expected value in (4.2) is understood asthe Bochner integral (see §2). Notice however that neither of the above formulas seems tobe appropriate to compute x′

���0� ·�: Differentiation of (4.1) would require differentiatingƐ�1 (which despite the notation is also a function of �), whereas in (4.2) it does not neces-sarily hold that J�0

�·� = X ′���0� ·�—and even when it does, we still have to be concerned

about interchanging the expectation and the differentiation operators. Indeed, it is not evenclear whether X� is directionally differentiable. Those difficulties do not arise from thenonsmoothness of the functions; they exist in the differentiable case as well, as pointed outby Glasserman (1993).

Page 13: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 753

In order to overcome these problems, we shall impose some stronger conditions. The goalof Assumption A3 below is to ensure that the regeneration epochs are constant in a neigh-borhood of �0 with some positive probability. It has the same spirit of the splitting conditionin Harris chains discussed before Theorem 3.2. Assumption A4, in turn, imposes indepen-dence of the original process between cycles. This assumption is sometimes called “classi-cal regeneration” in the literature. Assumption B4 is used to ensure that the interchange oflimits and expectations is valid.Assumption A3. Assumption A2 holds and there exists an K ∈ �0 1� such that, for any

< ∈ �0 K�, there exists a neighborhood V< of �0 such that

P(X�k

�V<= F�k

�V<

)≥ 1−< for all k = 1 2 � � � �(4.3)

Assumption A4. For each j ≥ 1, the post-�j process �X�j+k��� k = 0 1 � � � � isindependent of the pre-�j process.Assumption B4. Either each sample-path function Xn�· ��, n= 0 1 2 � � � , is convex

for P-almost all � ∈ &, or each Xn�· �� is almost surely locally Lipschitz at �0 withconstant Mn���; that is,

�Xn��1 ��−Xn��2 ��� ≤Mn�����1−�2� for all �1 �2 ∈ Vn���

for P-almost all � ∈&. In the latter case, �Mn� satisfies Ɛ∑�−1

n=0Mn <�.As with Assumption A2, verification of Assumption A3 may involve a thorough study

of the system. However, as we shall see in §6, typically that reduces to showing that thefamily of functions Xn�· �� is equicontinuous for all n and all � in some set of arbitrarilyhigh probability, which can be accomplished by exploiting the structure of the system.Assumption A4, on the other hand, is satisfied in most regenerative systems. AssumptionB4 covers the strong stochastic convexity assumption usually found in the literature (see,e.g., Hu 1992, Robinson 1995) as well as an alternative condition in case convexity is notpresent. Notice that the condition Ɛ

∑�−1n=0Mn < � holds in particular when the Mn’s are

all equal and deterministic, or when �Mn� is independent of �Xn� and ƐMn < �. As weshall see later, however, there are cases where the flexibility provided by Assumption B4 isconvenient.

Theorem 4.1. Suppose that Assumptions B1–B4, A2–A4 hold. Also, assume that the iidcycle times ��n� have finite expectation. Then, the expected value function x��·�= ƐX��·�is directionally differentiable at �0 and

x′���0� ·�= ƐJ�0

�·�= Ɛ[∑�1−1

n=0 X ′n��0� ·�

]Ɛ�1

where, as above, J�0�·� is the weak limit of X ′

n��0� ·�, and the expectation corresponds tothe Bochner integral.

Proof. Fix < ∈ �0 1�, and let V< be the neighborhood of �0 given by Assumption A3.We will show initially that, for any � ∈ V<, �Xn���� regenerates at a subset of theregeneration epochs of �Xn��0��.Consider an extension of the original probability space to support 0-1 random variables

Ik such that P�Ik = 1� = 1−< and X�k�V<

= F�k�V<

on �Ik = 1�. This can be accomplishedby the virtue of Assumption A3 and a construction similar to that in Corollary 3.5.1 inThorisson (2000). Next, let :< be the number of trials between two successive occurrencesof Ik = 1. Clearly, :< has a geometric distribution with success probability q = 1−< > 0.

Page 14: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

754 T. HOMEM-DE-MELLO

Consider now a subset ��kj� of the regeneration points such that Ij = 1, i.e., X�kj

�V<= F�kj

�V<.

Clearly, the length of these new cycles is given by

L �= �:<=

:<∑i=1

�i

(we drop the subscript < from L to simplify the notation). By Assumption A4 we can easilysee that, for any j ≥ 1, the event �:< ≤ j− 1� is independent of the random variable �j .Therefore we can apply Wald’s identity (see, e.g., Chung 1974) to obtain

ƐL= Ɛ:<Ɛ�=Ɛ�

q= Ɛ�

1−<<��

We conclude that, for any � ∈ V<, the process �Xn���� regenerates at a subset of theregeneration epochs of �Xn��0��. Hence, for any sufficiently small t > 0 we have that

limt↓0

x���0+ td�−x���0�t

=

limt↓0

1t

(Ɛ[∑L1−1

n=0 Xn��0+ td�]

ƐL1

− Ɛ[∑L1−1

n=0 Xn��0�]

ƐL1

)

= 1ƐL1

limt↓0

Ɛ

[L1−1∑n=0

Xn��0+ td�−Xn��0�

t

]�

(4.4)

By Assumption B4 (in case the Xn’s are Lipschitz), we have that �1/t��Xn��0 + td�−Xn��0�� ≤ Mn�d�. Furthermore, since Ɛ

∑�−1n=0Mn < � by assumption, it follows that

Ɛ∑L1−1

n=0 Mn <� and thus, by the Dominated Convergence Theorem we have that the limitand the expectation in (4.4) can be interchanged; that is,

x′���0�d� = lim

t↓0x���0+ td�−x���0�

t

= 1ƐL1

Ɛ

[L1−1∑n=0

limt↓0

Xn��0+ td�−Xn��0�

t

]

= Ɛ[∑L1−1

n=0 X ′n��0�d�

]ƐL1

(4.5)

= ƐJ�0�d� (4.6)

the last equality following from the fact that, as seen before, �X ′n��0�d�� also regenerates

at the same points as �Xn��0��. Observe that in case Xn�· �� is convex w.p.1, clearly∑L1−1n=0 Xn is convex as well and hence the interchange of the derivative and expectation

operators follows from the Monotone Convergence Theorem (cf., Proposition 2.1 in Shapiroand Wardi 1994). Next, notice that equations (4.5) and (4.5) hold for all d ∈ �m and henceby Lemma 2.2 we conclude that x′

���0� ·� is the Bochner integral of the correspondingfunctions on the right-hand sides of those equations. Finally, since P�:< = 1�≥ 1−< goesto one as < goes to zero, it follows that P�L= �� → 1 as < → 0. The assertion of thetheorem now follows. �

Remark. A simple situation where Assumption A3 is satisfied occurs when the“regeneration neighborhoods” of Assumption A2 do not depend on n or �. In that case,(4.3) holds with < = 0. Quite often, however, we cannot ensure that such a strong con-dition holds. Suppose for instance that & = �+, �0 ∈ � and ���0� = 1, the regenerationoccurring for all � ∈ �0 − 1/� �0 + 1/�, where � has exponential distribution. Then,

Page 15: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 755

for any positive t, there exists no neighborhood of �0 such that �Xn��0 + t�� regeneratestogether with �Xn��0��. Assumption A3, in turn, covers this case, thus allowing one toapply Theorem 4.1. As we shall see in §6, there are systems, for which we cannot ensureregeneration in a global neighborhood of �0, that satisfy Assumption A3.

Corollary 4.1. Under the assumptions of Theorem 4.1, for any d ∈ �m the randomvariables

1N

N−1∑n=0

X ′n��0�d�(4.7)

and ∑Mm=1

∑�m−1n=�m−1

X ′n��0�d�∑M

m=1��m−�m−1�

(=∑�M−1

n=0 X ′n��0�d�

�M

)(4.8)

are consistent estimators of x′���0�d� (respectively as N and M go to infinity).

Proof. Since X ′n��0� ·� regenerates at the points ��m�, it follows that both above

expressions converge to ƐJ�0�d�. By Theorem 4.1, this quantity is equal to x′

���0�d�. �

Corollary 4.1 provides a way to estimate the directional derivative of x� at �0 in anygiven direction d. Suppose we simulate the system for M regenerative cycles, and recall that�0 = 0 �1 �2 � � � are the regeneration epochs. Then (4.8) gives an estimator of x′

���0�d�,and by using the regenerative property it is possible to estimate the variance of that estimator(for details, see for instance Shedler 1987). Alternatively, one can fix a run length N anduse the estimator (4.7) in that case, variances can be estimated by the batch means method(see, e.g., Bratley et al. 1987).The importance of the above result arises from the fact that, in general, (4.7) and (4.8) are

not consistent estimators of x′���0�d�, because typically pointwise convergence does not

imply convergence of directional derivatives. For instance, the functions fn��� = max��−1/n 0� are all differentiable at 0 and f ′

n�0�= 0; however, fn converges (uniformly) to f ���=max�� 0� and f is not differentiable at 0—we have f ′�0�1� = 1, f ′�0�−1� = 0. In thissense, assumptions of Theorem 4.1 can be viewed as sufficient conditions for convergenceof directional derivatives even in the deterministic case.As another consequence of Theorem 4.1, we can derive necessary and sufficient

conditions for nondifferentiability of the steady-state function in case all Xn are sub-differentiable. This nondifferentiability phenomenon was observed by Shapiro and Wardi(1994). In their paper, they assume that each Xn��� is subdifferentiable and that �Xn���� and��Xn���� are regenerative for each �. They then give a sufficient condition for ƐX���0�not to be differentiable at some fixed point �0, namely, that there exists a nonsingletonconvex compact set C such that

P�C ⊂a �Xk��0� for some k ≤ �1 � > 0

where ⊂a means inclusion up to an additive constant and, as before, �1 is the length of thefirst regenerative cycle; see Proposition 2.2 in Shapiro and Wardi (1994) for details. Thepresent setting allows us to extend that result, as follows. Recall from §2 that �X ′

n��0� ·�� isa process on �H+��m� �H�. Let A⊂H+��m� denote the set

A= �5 ∈H+��m� � 5�·� is not linear��(4.9)

The next lemma shows that P�X ′n��0� ·� ∈ A� is well defined.

Page 16: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

756 T. HOMEM-DE-MELLO

Lemma 4.1. The set A defined above is a �H -Borel set.

Proof. For any r ∈ �m, let Lr�·� ∈ H+��m� denote the linear function �r ·�. Let D ⊂H+��m� denote the set

D = ⋂n≥1

⋃r∈�m

B�Lr�1/n�

where B�Lr�1/n� ⊂ H+��m� is the closed ball B�Lr�1/n� = �5 ∈ H+��m� � �5−Lr� ≤1/n�. We claim that Ac =D. Indeed, let 5 ∈Ac, i.e., 5= Lx for some x ∈�m. Let n≥ 1 bearbitrary. Since �m is dense in �m there exists an rn ∈�m such that �rn−x�2 ≤ 1/n andtherefore we have

�Lx −Lrn� = sup

d∈Sm−1

��x− rn d��

≤ supd∈Sm−1

�x− rn�2�d�2= �x− rn�2 ≤ 1/n

so 5 ∈ B�Lrn�1/n�. Thus, 5 ∈⋃r∈�m B�Lr�1/n� for all n≥ 1 and hence 5 ∈D, so Ac ⊂D.

Conversely, let ; ∈D. Then we know that given any n≥ 1 there exists an rn ∈�m suchthat �;−Lrn

� ≤ 1/n. Fix now an arbitrary N ≥ 1. Let d1 d2 be arbitrary points in Sm−1,and let *1 *2 be any real numbers. Let n = n�*1 *2 N � be any integer greater than orequal to 2N��*1�+ �*2��. Then we have that

�;�*1d1+*2d2�−*1;�d1�−*2;�d2��= �;�*1d1+*2d2�−Lrn

�*1d1+*2d2�

+*1Lrn�d1�+*2Lrn

�d2�−*1;�d1�−*2;�d2��≤ �;�*1d1+*2d2�−Lrn

�*1d1+*2d2��+ �*1��Lrn

�d1�−;�d1��+ �*2��Lrn�d2�−;�d2��

≤ �*1d1+*2d2�2�;−Lrn�+ �*1��Lrn

−;�+ �*2��Lrn−;�

≤ 1n��*1d1+*2d2�2+�*1�+ �*2��

≤ 1n�2�*1�+2�*2��≤

1N

Thus, we have that �;�*1d1 +*2d2�−*1;�d1�−*2;�d2�� ≤ 1/N for any N ≥ 1, so weconclude that ;�*1d1+*2d2�= *1;�d1�+*2;�d2�, i.e., ; is linear. Therefore, D⊂Ac andhence D = Ac. Since D is clearly a �H -Borel set, we conclude that so are Ac and A. �

We can now state the result.

Theorem 4.2. Suppose the assumptions of Theorem 4.1 hold, and suppose furthermorethat Xn�·� is subdifferentiable at �0 ∈' w.p.1 for all n≥ 0. Then, the steady-state functionx��·� �= ƐX��·� is (Gâteaux, Fréchet) differentiable at �0 if and only if

P��� � ∃ j 0 ≤ j ≤ �1��� s.t. Xj�· �� is not differentiable at �0��= 0�(4.10)

For the proof, we shall need the following lemma:

Lemma 4.2. Let 5�·� be a a random variable on �H+��m� �H� such that 5�·� issublinear, i.e., 5�*1d1 + *2d2� ≤ *15�d1�+ *25�d2� w.p.1 for all d1 d2 ∈ �m and all*1 *2 > 0. Then, Ɛ5�·� is linear if and only if P��� � 5�·� is linear��= 0.

Page 17: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 757

Proof. Let A be defined as in (4.9). Suppose first that Ɛ5�·� is linear. Then for anyd1 d2 ∈ �m we have that

Ɛ5�d1+d2�= Ɛ5�d1�+Ɛ5�d2�

(notice the implicit use of Lemma 2.2 here) and hence Ɛ5�d1+d2�−5�d1�−5�d2�= 0.Since the integrand is almost always nonpositive, it follows that 5�d1+d2�=5�d1�+5�d2�w.p.1. Moreover, for any d ∈�m and any *> 0 we have that 0= 5�0�= 5�*d+�−*�d�=*5�d�+5��−*�d� and thus 5��−*�d� = �−*�5�d�. We conclude that 5 is linear w.p.1,i.e., P�5�·� ∈ A�= 0.Conversely, suppose that 5�·� is linear w.p.1. Then, by linearity of the integral we have

that Ɛ5�·� is also linear; that is, Ɛ5�·� � A. �

Proof of Theorem 4.2. Let A be defined as in (4.9). Note initially that Xj�· �� isnondifferentiable at �0 if and only if X ′

j��0� · �� ∈ A. By Lemma 4.1, the set in (4.10) ismeasurable.Let 5 be a random variable on �H+��m� �H� defined as 5�·� =∑�1−1

n=0 X ′n��0� ·�. From

Theorem 4.1 we know that

x′���0� ·�=

Ɛ[∑�1−1

n=0 X ′n��0� ·�

]Ɛ�1

= Ɛ5�·�Ɛ�1

�(4.11)

Notice that the assumption of subdifferentiability of Xn implies that X ′n��0� ·� is convex

w.p.1 (see §2) and hence so is 5�·�. Furthermore, since 5�·� is positively homogeneous, itfollows that 5 is sublinear. Let E denote the set �� � ∃ j ∈ 0 �1��� s.t. X ′

j��0� · �� ∈ A�.We shall prove now that

E = �� � 5�· �� ∈ A��(4.12)

Indeed, let �∈E. Then, there exists some j ≤ �1 such that X′j��0� · ��∈A and hence 5�· ��

is nonlinear; otherwise, 5�· ��−X ′j��0� · �� would be concave. Thus, � ∈ �5�·� ∈ A�.

Conversely, suppose that � � E. This means that X ′j��0� · �� is linear for all j ≤ �1 and

hence 5�· ��=∑�1���−1n=0 X ′

n��0� · �� is linear, i.e., � � �5�·� ∈ A�. Therefore, (4.12) holds.Finally, from (4.11) we see immediately that x′

���0� ·� is linear (i.e., x��·� is Gâteaux-differentiable at �0) if and only if Ɛ5�·� is linear. Furthermore, as seen in the proofof Theorem 4.1, the properties of Assumption B4 carry out to x��·�; hence Gâteaux-differentiability of x��·� is equivalent to Fréchet differentiability. The assertion of thetheorem follows now from (4.12) and Lemma 4.2. �

5. The subdifferential process. The results presented in the previous sections are basedon the assumption of directional differentiability of the functions Xn�·�. As we have seen,this includes a fairly general class of functions. Suppose now that we restrict ourselvesto subdifferentiable processes, that is, directionally differentiable processes whose direc-tional derivative functions are convex (and hence sublinear). In dealing with optimizationalgorithms, one often encounters the problem of computing the subdifferential set (or atleast obtaining an approximation by using some computed subgradients) in order to checkoptimality conditions or generate a next iterate. In this sense, it is important to formulate theresults discussed in the previous sections in terms of the subdifferentials �Xn��0�, n≥ 0. Thebasic property that makes this possible is the one-to-one correspondence between compactconvex sets and finite sublinear functions mentioned in §2—every finite sublinear functionis the support function of a unique compact convex set. In particular, the subdifferential�Xn��0� (which is compact and convex) corresponds to the directional derivative functionX ′

n��0� ·�.

Page 18: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

758 T. HOMEM-DE-MELLO

Consider the space � of all compact convex subsets of �m, endowed with the Hausdorffmetric OH . It is known (see Debreu 1966) that �� OH� is a complete and separable metricspace, so in principle we can construct the corresponding collection of Borel sets ��

and hence, assuming measurability, we can consider ��Xn��0�� as a process on �� ���.Notice that this measurability assumption follows immediately from Assumption B1 andLemma 2.1. Indeed, the aforementioned correspondence between � and the space � ⊂H+��m� of finite sublinear functions implies that �Xn��0� is � −�� measurable if and onlyif X ′

n��0� ·� is � −�� measurable, where �� are the Borel sets of � . But the sets of ��

are of the form B∩� with B ∈�H . Since X ′n��0� ·� is � −�H measurable by Lemma 2.1

and X ′n��0� ·� ∈� , the conclusion follows.

All results from §3 can be easily reformulated into this new setting. Notice howeverthat, in principle, we cannot apply directly the theory of the Bochner integral to computethe expected value of a random variable on �� ��� as done in §2, since that theory isconstructed for Banach spaces—and � does not have a linear structure. Nevertheless, aspointed out by Hiai and Umegaki (1977) (see also Radström 1952), � can be embedded as aconvex cone in a Banach space, and this suffices to allow the theory of Bochner integrationto be applied to �-valued functions.As with the space �H+��m� �H�, we say that a function P � &→ � is simple if there

exist sets C1 � � � CN ∈� and E1 � � � EN ∈ � such that P���=∑Ni=1Ci1Ei

���. Also, wesay that a function Q � &→ � is strongly measurable if there exists a sequence of simplefunctions �Pn� such that limn→�OH�Q��� Pn����= 0 for P-almost all �. Analogously tothe case of positive homogeneous functions studied in §2, it follows from the separabilityof � that a random variable Q on �� ��� is strongly measurable. Then let P1 P2 � � �(with Pn =

∑Nn

i=1Cni 1En

i) be �-valued simple functions such that, for P-almost all � ∈&,

Q���= limn→�Pn���= lim

n→�

Nn∑i=1

Cni 1En

i��� (5.1)

where the limit is understood to be with respect to the Hausdorff metric. We define theBochner integral of Q as∫

Q dP �= limn→�

∫Pn dP = lim

n→�

Nn∑i=1

Cni P�E

ni ��

It is interesting to notice that the integral defined above for a “random set” Q differsfrom the standard construction of integrals of multifunctions found in the literature (see e.g.,Castaing and Valadier 1977, Hiai and Umegaki 1977, Debreu 1966, Rockafellar 1976). Inthat approach, Q is viewed as a multifunction &→�m (we shall consider here only closed-valued multifunctions); Q is said to be measurable if for each closed subset B of �m, the set

Q−1�B� �= �� ∈& � Q���∩B �= ��

belongs to � . A measurable selection of Q is a measurable function v � &→ �m such thatv��� ∈ Q��� for all � ∈ &. It can be shown that a multifunction Q is measurable if andonly if there is a countable family �vi � i ∈ I� of measurable selections of Q such that

Q���= cl�vi��� � i ∈ I�

for all � ∈&. The integral of Q is defined as the set of integrals of (integrable) measurableselections of Q , that is,∫

Q dP �={∫

v���P�d�� � v ∈ L1�P� v is a measurable selection of Q} (5.2)

Page 19: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 759

where L1�P� is the set of integrable functions with respect to the measure P. Notice thatdefinition (5.2) is general in that no assumptions on compactness or convexity of Q areimposed.A third way to view the integral of a random compact convex set Q is by using the

correspondence between that class of sets and sublinear functions as discussed in §2. Indeed,let ;����·� ∈ H+��m� be the support function of Q��� (so ; is a random variable on�H+��m� �H�), and suppose that ; is Bochner integrable. Let S = ∫

&;���P�d�� be the

Bochner integral of ;. Then, S�·� is finite-valued and sublinear, so there exists a compactconvex set T whose support function is S. We can then take T to be the integral of Q .Observe that, by Lemma 2.2, the computation of the Bochner integral

∫;����·�P�d��

reduces to the computation of the Lebesgue integrals∫;����d�P�d��. A particular advan-

tage of this approach in the present case is that, since the support function of the subdiffer-ential �Xn��0� is the directional derivative function X ′

n��0� ·�, we can easily reinterpret theresults of the previous sections.A natural question of course is: Do the integrals discussed above coincide? As it turns

out, the answer is affirmative. The equivalence between Bochner and multifunction integralsis studied in Theorem 4.5 in Hiai and Umegaki (1977) and §6.5 in Debreu (1966), whereasthe correspondence between multifunction integrals and integrals of support functions canbe found in Theorem V-14 in Castaing and Valadier (1977) and Theorem 2.2 in Hiai andUmegaki (1977). It must be stressed that those equivalences hold for compact-convex-valuedmultifunctions (which is the present case), since otherwise the Bochner integral is not welldefined and the correspondence with support functions does not hold.We now re-state the results of §4 in terms of subdifferentials. First we derive a result

that, although well known (see Rockafellar and Wets 1982, Ioffe and Tihomirov 1969),illustrates an application of the equivalence between integrals discussed above:

Proposition 5.1. Suppose that, for P-almost � ∈&, each Xn�· �� is subdifferentiableat �0. Suppose also that Assumptions B2 and B4 hold. Then, for any finite n≥ 0,

�ƐXn��0�= Ɛ�Xn��0� (5.3)

where the expected value is understood as the multifunction integral defined in (5.2).

Proof. Let xn��� = ƐXn���. By the equivalence of multifunction integrals withintegrals of supporting functions, (5.3) holds if and only if x′

n��0� ·�= ƐX ′n��0� ·�, i.e.,

x′n��0�d�= ƐX ′

n��0�d� for all d ∈ �m�

The above equation holds if we can interchange integrals and limits, which as seen in theproof of Theorem 4.1 follows from Assumption B4. �

Another interesting consequence comes from Lemma 4.2:

Proposition 5.2. Suppose that the assumptions of Proposition 5.1 hold. Then, for anyfinite n ≥ 0, ƐXn�·� is differentiable at �0 if and only if Xn�·� is differentiable at �0 withprobability one.

Proof. First notice that, under the assumptions of the proposition, ƐXn�·� isdifferentiable at �0 if and only if �ƐXn��0� is a singleton, i.e., if the directional deriva-tive of ƐXn�·� at �0 is a linear function. By Lemma 4.2, this occurs if and only if eachX ′

n��0� ·� is linear w.p.1, that is, if and only if Xn�·� is differentiable at �0 w.p.1. �

The following is an immediate consequence of Theorem 4.1 and Corollary 4.1.

Page 20: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

760 T. HOMEM-DE-MELLO

Proposition 5.3. Suppose that the assumptions of Theorem 4.1 hold. Then, the follow-ing equations hold:

�ƐX���0� = limN→�

1N

N−1∑n=0

�Xn��0� w.p.1 (5.4)

�ƐX���0� =Ɛ[∑�1−1

n=0 �Xn��0�]

Ɛ�1 (5.5)

where the sums on the right-hand sides are understood as Minkowski addition of sets, theexpected-values are again understood as the multifunction integrals defined in (5.2), andthe limit in (5.4) is understood in the Hausdorff sense.

As with the directional derivatives, equation (5.4) provides a way to estimate thesubdifferential of ƐX� at �0. Again, suppose we simulate the system for M regenerativecycles, and let �0 = 0 �1 �2 � � � be the regeneration epochs. An estimator for �ƐX���0�is given by

Q�0=∑M

m=1

∑�m−1n=�m−1

�Xn��0�∑Mm=1��m−�m−1�

Clearly, from the above formula we see that the choice of any particular subgradient of eachset �Xn��0� yields an estimator for a subgradient of ƐX� at �0. Furthermore, as remarkedbefore, we can also estimate the variance of the resulting estimators.Observe that Proposition 5.3 can actually be stated in terms of a general �-valued process

�Cn�. Let -n�·� denote the support function of Cn. As seen in §2, -n�·� is a finite-valuedsublinear (i.e., convex and positively homogeneous) function. Moreover, it is easy to see that-n�d� = - ′

n�0�d� for all d ∈ �m, whence we have that Cn = �-n�0�. It follows that if theprocess �-n�·�� satisfies assumptions A3–A4 and B1–B4 discussed in the previous sections(for �0=0), then �Cn� is regenerative and hence we can apply Proposition 5.3. Notice thatAssumptions B2 and B3 are equivalent to assuming Bochner integrability of -n�·� (andtherefore of Cn), and that Assumptions B1 and B4 are automatically satisfied in this casesince -n�·� is convex. Below we summarize this result.

Corollary 5.1. Let �Cn� be a process on �� �C�, and let -n�·� denote the supportfunction of Cn. Suppose that �-n�·�� satisfies Assumptions A3 and A4 for �0=0. Supposealso that the cycle times ��n� have finite expectation and that

∑�1−1n=0 Cn is (Bochner) inte-

grable. Then, �Cn� is regenerative, has a weak limit C�, and the following equations hold:

ƐC� = limN→�

1N

N−1∑n=0

Cn w.p.1 (5.6)

ƐC� =Ɛ[∑�1−1

n=0 Cn

]Ɛ�1

�(5.7)

The above corollary can also be viewed as a result for compact-convex-valuedmultifunction processes. Other types of limiting results have been studied in the literaturefor multivalued processes: in more general settings (i.e., without imposing convexity andcompactness) Hiai (1984) gives various forms of the Strong Law of Large Numbers whenthe Cn are independent, whereas Hiai and Umegaki (1977) study martingales formed bymultivalued processes and provide some convergence theorems. We refer to those papersand references therein for details. To the best of our knowledge, however, there have beenno results on regenerative multivalued processes as given by Corollary 5.1.

Page 21: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 761

6. Examples.(1) A G/D/1 queue. The following system was presented by Shapiro and Wardi (1994) to

illustrate the nondifferentiability of the mean steady-state function. Consider a G/D/1 queuewhere the distribution of the interarrival times An has atoms at two points b and c withb < c. For simplicity, we take P�An = b�= P�An = c�= 1

2 . Suppose that the deterministicservice time is a parameter � ∈'= �b �b+c�/2�. Notice that the assumption � < �b+c�/2guarantees that the queue is stable and hence regenerative.Denote by Tn��� the system time of customer n (i.e., waiting time plus service time).

Then, for n≥ 1, Tn��� satisfies the recursion

Tn���= �+ Tn−1���−An+

where x+ =max�x 0� (assume T0�·�≡ 0). Notice that Tn�·� is defined by the maximumof linear functions and therefore it is convex. Given �0 ∈', define random variables �-m�,m= 0 1 2 � � � , with -0 = 0 and

-m = inf�n > -m−1 � Tn−1��0�≤ An��

Observe that �-m�, m = 0 1 2 � � � are the epochs at which an arriving customer findsthe queue empty; hence �Tn��0�� regenerates at those points. Notice also that on theevent �Tn−1��0� ≤ An� we have that Tn��0� = �0. Shapiro and Wardi (1994) show that on�Tn−1��0� = An� the function Tn�·� is not differentiable at �0 and, furthermore, the event�Tn−1��0� = An for some n ≤ -1� has positive probability, which is their basic conditionfor nondifferentiability of the mean steady-state function ƐT��·� at �0.

Consider now the event E = �-m = n�, and observe that we can partition E as E<∪E=,where E< = �Tn−1��0� < An� and E= = �Tn−1��0�=An�. Notice that the continuity of Tn�·�implies that for any � ∈ E< there is a neighborhood Vn��� of �0 such that Tn−1��� < An forall � ∈ Vn��� and hence Tn���= � on that neighborhood. On the other hand, on E= it mighthappen that Tn−1��� > An for � arbitrarily close to �0 and hence Tn���= �+Tn−1���−An.Since the event �Tn−1��0� = An for some n ≥ 1� happens at some -k with probabilityone, it follows that we cannot ensure the existence of neighborhoods �Vn���� satisfyingAssumption A2 and thus we cannot apply Theorem 3.1 directly.We can nevertheless overcome that problem by considering a subset of the regeneration

points as follows. Let 6 > 0 be such that L�An−6≤ Tn−1��0� < An�= 0, which exists sinceAn takes on a finite number of values. Let �Lm�, m= 0 1 2 � � � (with L0 = 0) denote theepochs

Lm = inf�n > Lm−1 � Tn−1��0� < An−6��

As argued before, by continuity of the functions Tn there exist neighborhoods Vn��� of �0,� ∈&, such that on the event F = �Lm = n� we have that Tn���= � on Vn���. Therefore,Assumption A2 is satisfied and hence from Theorem 3.1 it follows that �Tn��0� �Tn��0��regenerates at each Lm.Let us show now that Assumption A3 also holds in this case. Note initially that, since

ƐL1 <�, it follows that for any < ∈ �0 1� there exists an N such that P�L1 > N�< <. Now,on �Lm = n� we can write

Tn−1��0�= �n−1−Lm−1��0−n−1∑

k=Lm−1

Ak < An−6�

Define the setV �=

{� ∈' � ��−�0�<

6

N

}�

Page 22: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

762 T. HOMEM-DE-MELLO

Consider now a point � ∈ V . Since � ∈ ', it follows that the system with service times �is stable and therefore it regenerates at epochs, say, L1 L2 � � � . For each m= 0 1 � � � , letBm denote the event

Bm �= �there exists a Lk such that Lm−1 ≤ Lk ≤ Lm��

It follows that, for any � ∈ V , on �Lm = n�∩Bm we can write

Tn−1���−Tn−1��0�= �n−1���−�0�+Lm−1�0− Lk�+Lk−1∑

j=Lm−1

Aj

= �n−1− Lk���−�0�−[�Lk−Lm−1��0−

Lk−1∑j=Lm−1

Aj

]≤ �n−1− Lk���−�0�

≤ �Lm−Lm−1���−�0�

(6.1)

where the next-to-last inequality follows from the fact that∑@−1

j=Lm−1Aj ≤ �@−Lm−1��0 for

all @ ∈ �Lm−1 Lm. Therefore we have that

�Tn−1���−Tn−1��0�� ≤ �Lm−Lm−1�6

N

and thus, on �Lm = n�∩Bm∩ �Lm−Lm−1 ≤ N� we have

�Tn−1���−Tn−1��0�� ≤ 6�

It follows that there exists some < ∈ �0 1� such that <→ 0 as <→ 0 and

P( ∣∣TLm−1���−TLm−1��0�

∣∣≤ 6)

≥ P( ∣∣TLm−1���−TLm−1��0�

∣∣≤ 6 � Bm Lm−Lm−1 ≤ N)P(Bm Lm−Lm−1 ≤ N

)≥ 1− <

for any � ∈ V . Notice that the condition �TLm−1���−TLm−1��0��< 6 implies that TLm−1��� <Aj , which in turn implies that TLm

��� = �. Therefore, Assumption A3 holds. Incidentally,notice that (6.1) illustrates a case where Assumption B4 holds with variable Lipschitzconstants Mn—in this case, Mn = n—provided that ƐL2

1 <�.Notice that TLm

�·� is differentiable at �0 for all m≥ 0—indeed, T ′Lm��0�= 1—but for the

indices n such that �Tn−1��0�= An� we have that

�Tn��0�= conv(�1�∪ �1�+ �Tn−1��0�

)(6.2)

(where conv�A� denotes the convex hull of A). For all other indices j such that 0< j < L1,we have

�Tj��0�={�1�+ �Tj−1��0� if Tj−1��0� > Aj

�1� if Tj−1��0� < Aj�(6.3)

Observe that with positive probability (i.e., the probability that the event �Tn−1��0� = An�occurs before �Tn−1��0� < An�) there exists n< L1 such that Tn�·� is nondifferentiable at �0.Thus, we see that the sum of the sets �Ti��0� for i within a cycle will be a nonsingletonwith positive probability, and hence by Theorem 4.2 it follows ƐT��·� is not differentiableat �0.

Page 23: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 763

We now check Assumptions B2 and B3. It is not difficult to see from (6.2) and (6.3) that��Tj��0�� ≤ j, for all 0< j < L1. It follows immediately that �Tj��0� is Bochner integrableif ƐL1 <�, and

∑L1−1n=0 �Tn��0� is Bochner integrable if ƐL

21 <�. Assuming these conditions

hold, by applying Proposition 5.3 we obtain

�Ɛ[T���0�

]= Ɛ[∑L1−1

n=0 �Tn��0�]

ƐL1

A consistent estimator of �ƐT���0� can now be computed as follows. Fix some M > 0,and simulate the system for M cycles. Let L0 � � � LM be the regeneration points (withoutloss of generality, assume we start with an empty system, i.e., L0 = 0). Then,∑M

i=1

∑Li−1j=Li−1

�Tj��0�∑Mi=1�Li −Li−1�

is an estimator of �ƐT���0�, where the �Tj��0� are given by (6.2) and (6.3), and �T0��0�=�TLi

��0� = �1�. Another consistent estimator is obtained by simulating the system for Nperiods and computing

1N

N−1∑i=0

�Ti��0��

(2) A tandem queue model. Consider a series of K single-server queues, with generaldistributions for the inter-arrival times An and the service times �Sk

n����, n = 0 1 � � � ,k = 1 � � � K. Suppose that all service time vectors Sn��� �= �S1

n��� � � � SKn ���� are iid,

and that the network is stable. Such system has been studied in the literature in the casethe Sn�·� are differentiable: Some papers deal with optimization issues and consequentlythe computation of derivatives of steady-state quantities, often using perturbation analysistechniques (see, for instance, Wardi and Hu 1991, Hu 1992, Chong and Ramadge 1994),whereas Glasserman (1993) shows that the waiting times Wn��� and the derivative processW ′

n��� regenerate at the same epochs. Here, we do not assume differentiability; instead weassume that Sk

n��� is only a subdifferentiable function of � for each n and each k, andthat the functions Sn�·� �= �S1

n�·� � � � SKn �·�� are iid. We also assume that the service times

satisfy the following property: For any < ∈ �0 1� there exists a constant M< > 0 such that,for any � in a neighborhood of �0,

P(�Sk

n���−Skn��0�� ≤M< ��−�0�

)> 1−<(6.4)

for all n and k. This condition is satisfied for instance if Skn��� has the form max�Y k

n � Zkn�,

where Y kn and Zk

n are random variables with finite expectation.Let T k

n ��� denote the system time of job n upon its completion on server k. It is clearthat one set of regeneration points for the process Tn��� �= �T 1

n ��� � � � TKn ���� consists

of the epochs at which an arriving customer finds the whole network empty. However, aspointed out by Nummelin (1981), those points may never occur, even though the systemis stable. Alternatively, with mild assumptions, Nummelin provides regeneration epochs forthe waiting time process Wk

n ��� and shows that the corresponding regeneration cycles haveexpected finite length. By using an argument similar to Nummelin (1981), it can be shownthat under proper assumptions the system time process �Tn���� is a regenerative Markovchain with regeneration points given by

-j���= inf�n > -j−1��� � Tkn−1−An < bk���−6 T k−1

n > bk���+6 k = 1 � � � K�

where b��� = �b1��� � � � bK���� is a deterministic continuous function of �, and 6 is anarbitrary positive number. Notice that on �-j���= n� the total waiting time of job n is zero

Page 24: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

764 T. HOMEM-DE-MELLO

and hence we have that

T kn ���= S1

n���+· · ·+Skn��� k = 1 � � � K�

For �0 ∈', let B0 ⊂ �K ×� denote the set

B0 = ��x a� � xk−a < bk��0�−6 k = 1 � � � K��

and let D0 ⊂ �K denote the set

D0 = �y � y1+· · ·+yk−1 > bk��0�+6 k = 1 � � � K��

It is easy to see that on ��Tn��0� An+1�∈B0�∩�Sn+1��0�∈D0� we have that Tkn ���−An+1 <

bk���−6, T k−1n+1 > bk���+6 for all k= 1 � � � K and all � in some neighborhood Vn��� of

�0, so it follows that

T kn+1�·�= S1

n+1���+· · ·+Skn+1��� k = 1 � � � K

on Vn���. Furthermore, by writing a recursive expression for T kn ��� it is readily seen

that T kn �·� is subdifferentiable, since so are the service times. Therefore, the conditions of

Theorem 3.2 are satisfied and hence we conclude that �Tn��0� �Tn��0�� regenerates at theepochs �-j��0��.Let us show now that Assumption A3 is satisfied. It is possible to show that, given �∈&,

� ∈ ', k ≤ K and n ≥ 0, there exist sets � and (depending on �, �, n and k) such that� ⊂ �1 � � � K�, ⊂ �-@−1��� � � � -@���− 1� for some @ > 0, -@−1��� ≤ n ≤ -@���− 1and we can write

T kn �� ��=∑

j∈�

∑m∈

Sjm�� ��−

n∑m=r

Am(6.5)

for some r ∈ -@−1��� -@���−1. The sets � and correspond to the solution of a longestpath problem in a graph, as shown in Homem-de-Mello, Shapiro and Spearman (1999);we refer to that paper for details. Let �0, 0, and r0 be the elements corresponding toT kn ��0 ��, i.e.,

T kn ��0 ��= ∑

j∈�0

∑m∈0

Sjm��0 ��−

n∑m=r0

Am�

By the property of longest paths we have that∑j∈�

∑m∈

Sjm�� �� ≥ ∑

j∈�0

∑m∈0

Sjm�� ��

∑j∈�0

∑m∈0

Sjm��0 �� ≥ ∑

j∈�

∑m∈

Sjm��0 ��

and thus ∑j∈�0

∑m∈0

[Sjm�� ��−Sj

m��0 ��] ≤ T k

n �� ��−T kn ��0 ��

≤ ∑j∈�

∑m∈

[Sjm�� ��−Sj

m��0 ��]�

Page 25: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 765

Notice that the term∑n

m=r Am disappears in the above inequality—this follows from anargument similar to the one used in Example 1. It follows that

�T kn �� ��−T k

n ��0 ���≤ max(∣∣∣∣ ∑

j∈�0

∑m∈0

Sjm�� ��−Sj

m��0 ��

∣∣∣∣ ∣∣∣∣∑j∈�

∑m∈

Sjm�� ��−Sj

m��0 ��

∣∣∣∣)�(6.6)

Next, by the assumption (6.4) on service times it follows from (6.6) that, for any < ∈ �0 1�,there exists an event Bk

n with positive probability such that, on Bkn, we have

�T kn ���−T k

n ��0�� ≤∑j∈�1

∑m∈1

M<��−�0�

where �1 and 1 correspond to the maximizer of the right-hand side of (6.6). Let: = card��1� card�1�. Then, we have that P�Bk

n �: = r� ≥ �1− <�r . Moreover, since�1 ⊂ �1 � � � K� and 1 ⊂ �-@−1 � � � -@� for some @ > 0, it follows that card��1�≤K andcard�1�≤ �@, where �@ is the length of the @th cycle (notice that here - and � have argu-ments � or �0, according to the maximizer of the right-hand side of (6.6)). Since Ɛ�@ <�,there exists a constant Q< > 0 such that

P(card�1�≤Q<

)> 1−<

which implies thatP(: ≤ KQ<

)> 1−<�

It follows that on Bkn ∩ �: ≤ KQ<� we can write

�T kn ���−T k

n ��0�� ≤ :M<��−�0� ≤ KQ<M<��−�0��(6.7)

Let V denote the neighborhood V = �� � ��− �0� < 6/KQ<M<�. Then, from (6.7) we seethat, on Bk

n ∩ �: ≤ KQ<�, we have, for all � ∈ V , �T kn ���−T k

n ��0�� ≤ 6. Finally, since

P(Bk

n ∩ �: ≤ KQ<�)= KQ<∑

r=1

P(Bk

n � : = r)P�: = r�≥

KQ<∑r=1

�1−<�rP�: = r�

it follows that there exists some < ∈ �0 1� such that <→ 0 as <→ 0 and

P(�T k

n ���−T kn ��0�� ≤ 6

)> 1− <

for all n and k, and all � ∈ V . We conclude that

P(T k-m���∣∣V= S1

-m���+· · ·+Sk

-m���)> 1− <

for all m and k. Therefore, Assumption A3 holds.Notice also that from (6.5) we can see that if �Sk

n��� is integrable for all n and k—which happens, for example, if Sk

n��� has the form Y kn �, where Y k

n is a random vari-able with finite expectation—then Assumptions B2 and B3 will follow, as in the previousexample, from finiteness of first and second moments of the cycle lengths. Thus, by applyingProposition 5.3 we obtain

�ƐT k���0�=

Ɛ[∑-1−1

n=0 �T kn ��0�

]Ɛ-1

Page 26: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

766 T. HOMEM-DE-MELLO

We can now estimate an element of �ƐT k���0�, as follows. Fix some M > 0, and

simulate the system for M cycles. Let -0 � � � -M be the regeneration points (without lossof generality, assume we start with an empty system, i.e., -0 = 0). Then,∑M

i=1

∑-i−1j=-i−1

�T kj ��0�∑M

i=1�-i −-i−1�

is an estimator of �ƐT k���0�. Notice that for any j ≥ 0 we have

T k-j��0�= S1

-j��0�+· · ·+Sk

-j��0� k = 1 � � � K

and hence�Tk

-j��0�= �S1

-j��0�+· · ·+�Sk

-j��0� k = 1 � � � K

where �f denotes an arbitrary subgradient of f . For the indices n such that -j−1 < n < -j

a subgradient of T k-j

at �0 can be computed by using the corresponding solution of thelongest-path problem mentioned above, together with subgradients of the service times; werefer again to Homem-de-Mello et al. (1999) for details. Another consistent estimator isobtained by simulating the system for N periods and computing

1N

N−1∑i=0

�Tki ��0��

7. Concluding remarks. In this paper we provided results that can potentially enlargethe scope of applications of sensitivity analysis and optimization in stochastic systems toinclude “nonsmooth” processes. In particular, we have shown that, under some assump-tions, one can consistently estimate directional derivatives (and consequently subdifferentialsets and subgradients) of expected steady-state functions both by ratio-type and long-runaverage formulas. Those formulas are convenient in that they allow the computation of thederivatives to be done simultaneously with the original process during the simulation. Wehave given some examples showing potential applications of these results.From the more theoretical viewpoint, our contribution is twofold: First, we extended the

result in Shapiro and Wardi (1994) by exhibiting a necessary and sufficient condition for thedifferentiability of the expected steady-state function. Second, we showed a limit theoremfor compact-convex-valued multifunctions by proving that, under proper assumptions, theaverage of a sequence of regenerative random multifunctions converge with probability one.

We hope this work will allow many results from the well-studied fields of nonsmoothanalysis and multifunctions to be incorporated into the area of stochastic processes.

Acknowledgments. The author is grateful to Alexander Shapiro for some enlighteningdiscussions during the preparation of this work. The author also thanks three anonymousreferees for some detailed comments and suggestions that greatly helped to improve thequality of the paper, and David McDonald and A. Ioffe for discussions on the Bochnerintegral. This work was supported, in part, by CNPq, Brasília, Brazil, through a DoctoralFellowship under Grant 200595/93-8. It was also supported in part by grant Grant DMI-9713878 from the National Science Foundation.

References

Asmussen, S. 1987. Applied Probability and Queues. Wiley, New York.Bratley, P., B. L. Fox, L. E. Schrage. 1987. A Guide to Simulation, 2nd. Edition. Springer-Verlag, New York.Castaing, C., M. Valadier. 1977. Convex Analysis and Measurable Multifunctions. Springer-Verlag, Berlin,

Germany.

Page 27: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

DERIVATIVES OF NONSMOOTH PERFORMANCE MEASURES 767

Chong, E. K. P., P. J. Ramadge. 1994. Stochastic optimization of regenerative systems using infinitesimalperturbation analysis. IEEE Trans. Automat. Control 39 (7) 1400–1410.

Chung, K. L. 1974. A Course in Probability Theory, 2nd Edition. Academic Press, San Diego, CA.Clarke, F. H. 1990. Optimization and Nonsmooth Analysis. Reprint, SIAM (originally published by Wiley, New

York, 1983).Debreu, G. 1966. Integration of correspondences. Proc. Fifth Berkeley Sympos. Math. Statist. Probab. Vol. II,

Part I, 351–372.Diestel, J., J. J. Uhl, Jr. 1977. Vector Measures. American Mathematical Society, Providence, RI.Glasserman, P. 1991. Gradient Estimation via Perturbation Analysis. Kluwer, Norwell, MA.

. 1993. Regenerative derivatives of regenerative sequences. Adv. Appl. Probab. 25 116–139., P. W. Glynn. 1992. Gradient estimation for regenerative processes. Proc. 1992 Winter Simulation Conf.IEEE Press, Piscataway, NJ, 280–288.

Glynn, P. W. 1989. Optimization of stochastic systems via simulation. Proc. 1989 Winter Simulation Conf. IEEEPress, Piscataway, NJ, 90–105., P. L’Ecuyer. 1995. Likelihood ratio gradient estimation for stochastic recursions. Adv. Appl. Probab. 271019–1053.

Hiai, F. 1984. Strong laws of large numbers for multivalued random variables. A. Dold and B. Eckmann, Eds.Multifunctions and Integrands. Springer-Verlag, Berlin, Germany., H. Umegaki. 1977. Integrals, conditional expectations and martingales of multivalued functions.J. Multivariate Anal. 7 149–182.

Hiriart-Urruty, J.-B., C. Lemaréchal. 1993. Convex Analysis and Minimization Algorithms I. Springer-Verlag,Berlin, Germany., . 1993a. Convex Analysis and Minimization Algorithms II. Springer-Verlag, Berlin, Germany.

Homem-de-Mello, T., A. Shapiro, M. L. Spearman. 1999. Finding optimal material release times using simulationbased optimization. Management Sci. 45 86–102.

Hu, J. Q. 1992. Convexity of sample path performance and strong consistency of infinitesimal perturbation analysis.IEEE Trans. Automat. Control 37 (2) 258–262.

Ioffe, A. D., V. M. Tihomirov. 1969. On the minimization of integral functionals. Funct. Anal. Appl. 3 218–227., . 1979. Theory of Extremal Problems. North-Holland, Amsterdam, The Netherlands.

Kelley, J. L., T. P. Srinivasan. 1988. Measure and Integral. Springer-Verlag, New York.L’Ecuyer, P. 1990. A unified view of the IPA, SF and LR gradient estimation techniques. Management Sci. 36

1364–1383., P. W. Glynn. 1994. Stochastic optimization by simulation: Convergence proofs for the GI/G/1 queue insteady-state. Management Sci. 40 1562–1578.

Nummelin, E. 1981. Regeneration in tandem queues. Adv. Appl. Probab. 13, 221–230.. 1984. General Irreducible Markov Chains and Non-negative Operators. Cambridge University Press,Cambridge, MA.

Pflug, G. Ch. 1996. Optimization of Stochastic Models. Kluwer, Norwell, MA.Plambeck, E. L., B. R. Fu, S. M. Robinson, R. Suri. 1996. Sample-path optimization of convex stochastic

performance functions. Math. Programming Ser. B 75 137–176.Radström, H. 1952. An embedding theorem for spaces of convex sets. Proc. Amer. Math. Soc. 3 165–169.Revuz, D. 1984. Markov Chains. North-Holland, Amsterdam, The Netherlands.Robinson, S. M. 1995. Convergence of subdifferentials under strong stochastic convexity. Management Sci. 41

1397–1401.Rockafellar, R. T. 1970. Convex Analysis. Princeton University Press, Princeton, NJ.

. 1976. Integral functionals, normal integrands and measurable selections. J.P. Gossez, E.J. Lami Dozo,J. Mawhin and L. Waelbroeck, Eds. Nonlinear Operators and the Calculus of Variations. Springer-Verlag,Berlin, Germany., R. J.-B. Wets. 1982. On the interchange of subdifferentiation and conditional expectation for convexfunctionals. Stochastics 7 173–182.

Royden, H. 1988. Real Analysis. Macmillan, New York.Rubinstein, R. Y., A. Shapiro. 1993. Discrete Event Systems: Sensitivity Analysis and Stochastic Optimization by

the Score Function Method. John Wiley & Sons, New York.Shapiro, A. 1990. On concepts of directional differentiability. J. Optim. Theory Appl. 66 (3) 477–487.

, Y. Wardi. 1994. Nondifferentiability of the steady-state function in discrete event dynamical systems.IEEE Trans. Automat. Control 39 (8) 1707–1711.

Shedler, G. S. 1987. Regeneration and Networks of Queues. Springer-Verlag, New York.Suri, R. 1989. Perturbation analysis: The state of the art and research issues explained via the GI/G/1 queue. Proc.

IEEE 77 114–137., Y. T. Leung. 1989. Single run optimization of discrete event simulations—an empirical study using theM/M/1 queue. IIE Trans. 21 35–49.

Thorisson, H. 2000. Coupling, Stationarity and Regeneration. Springer-Verlag, New York.

Page 28: Vol.26,No.4,November2001,pp.741–768 PrintedinU.S.A.mansci-web.uai.cl/~thmello/pubs/derivative estimation MOR... · 2007-11-03 · which the derivative process regenerates together

768 T. HOMEM-DE-MELLO

Wardi, Y., J. Q. Hu. 1991. Strong consistency of infinitesimal perturbation analysis for tandem queueing networks.J. Discrete Event Dynamic Systems: Theory and Applications 1 37–59.

Wolff, R. W. 1989. Stochastic Modeling and the Theory of Queues. Prentice-Hall, NJ.

T. Homem-de-Mello: Department of Industrial, Welding & Systems Engineering, The Ohio State University,1971 Neil Avenue, Columbus, Ohio 43210; e-mail: [email protected]