Storm 2012

download Storm 2012

of 62

Transcript of Storm 2012

  • 7/30/2019 Storm 2012

    1/62

    Stochastic Operations Research Models

    Hans Reijnierse

    August 2012

    This syllabus deals with the first seven weeks. It is roughly based on topics inA first course in Stochastic Models

    by

    Henk C. Tijms

    Wiley

    ISBN 0-471-49880-7 (Cloth)

    ISBN 0-471-49881-5 (Paper)

  • 7/30/2019 Storm 2012

    2/62

    Contents

    1 Introduction 2

    2 Discrete Time Markov Chains 4

    2.1 Examples of Discrete Time Markov Chains . . . . . . . . . . . . . . . . . 7

    2.2 Multiple Step Transition Probabilities . . . . . . . . . . . . . . . . . . . . 10

    2.3 Mean First Passage Times . . . . . . . . . . . . . . . . . . . . . . . . . . 11

    2.4 The Long Term Behavior of a DTMC . . . . . . . . . . . . . . . . . . . . 12

    2.5 Communication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

    2.6 Cost-Reward Structures on Finite Irreducible DTMCs . . . . . . . . . . 18

    3 Continuous Time Processes 20

    3.1 The Markov Property for Continuous Time Processes . . . . . . . . . . . 20

    3.2 Poisson and Other Counting Processes . . . . . . . . . . . . . . . . . . . 21

    3.3 Continuous Time Markov Chains . . . . . . . . . . . . . . . . . . . . . . 29

    3.3.1 Time Dependent Transition Probabilities and Transition Rates . . 33

    3.3.2 Equilibrium Distributions . . . . . . . . . . . . . . . . . . . . . . 36

    3.3.3 Long Term Behavior . . . . . . . . . . . . . . . . . . . . . . . . . 40

    3.3.4 Cost-Reward Structures . . . . . . . . . . . . . . . . . . . . . . . 42

    3.3.5 The PASTA Theorem and Littles Law . . . . . . . . . . . . . . . 44

    A Appendix; Mathematical Tools to Build Stochastic Models i

    B Index ix

  • 7/30/2019 Storm 2012

    3/62

    1 Introduction

    In order to study Economic situations, it is convenient to rephrase them into a robustmathematical model. Advantages are that ambiguities are (should be) eliminated, one

    can leave side issues out of the model and focus on the main ones, and one can simplify

    the situation if it is too complex to handle as a whole. Of course, too much simpli-

    fication will lead to a poor description of reality; there is always a trade-off between

    effort and accuracy. Another advantage is that mathematical models may reveal general

    phenomena that apply for many situations. The objective of this course can be put in

    the one-liner

    . .... to model and optimize dynamic systems under uncertainty. . .

    The term Operations Research in the name of the course refers to a scientific discipline

    described by Wikipedia as follows:

    ... an interdisciplinary branch of applied mathematics and formal science that uses

    analytical methods to arrive at (near-)optimal solutions to complex decision-making

    problems.

    It is often concerned with determining the maximum (of profit, performance, or yield) or

    minimum (of loss, risk, or cost) of some objective function. This course deals with some

    techniques for modeling and optimizing systems under uncertainty. It aims to increase

    the capability of analyzing managerial problems under uncertainty which occur, for

    example, in

    . .

    scheduling,

    inventory,

    production control,

    telecommunications,

    maintenance,

    insurance.

    . .

    2

  • 7/30/2019 Storm 2012

    4/62

    The emphasis is on providing insight in the theory, on formulating an economic situation

    into a mathematical model and on providing practical examples in which the discussed

    models can be applied. Examples of mathematical models of OR-problems are e.g.

    a time dependent variable counting the number of jobs waiting in a queue,

    a directed graph representing cities and roads, (e.g., to model the Traveling Salesman

    Problem),

    Markov chains; these will form the main subject of the first half of this course.

    . .

    The following exercise is an example of a stochastic optimization problem concerning

    inventory:

    Exercise 1

    A vendor in the lobby of a big office building sells coffee. Suppose that the daily demand

    X for coffee, measured in liters, is exponentially distributed1 with rate > 0. Suppose

    that she must pay e c per liter for coffee, receives e b for each liter sold (assume that

    b > c), but suffers a penalty ofe a per liter in lost goodwill for unfulfilled demand. How

    much coffee should she buy to maximize her expected profit per day?

    Hint: Let I be the amount of coffee the vendor purchases; the decision variable. Then

    E[X] = 1 and show that E[X | X > I] = I +1 . Apply the law of total expectation

    to obtain E[X | X < I].

    Answer: If I is the amount of coffee she buys, it is optimal to choose I subject to

    . .P [X > I] =c

    b + a

    .. .

    Besides uncertainty, time will have a crucial role in the problems that we will consider.

    Real life situations but also mathematical models that involve time and uncertainty

    are called stochastic processes. Examples are the weather, Dow Jones, queues.... The

    1This syllabus contains an appendix with definitions of mathematical notions and an index section

    to find them.

    3

  • 7/30/2019 Storm 2012

    5/62

    following queueing example displays a basic queue and its costs and rewards. The

    questions raised can be answered with the theory provided by this course.

    Example 1 A call center has two servers. Callers who find both servers busy are put

    in a waiting queue. The first come, first served policy is used to handle the services.

    Each caller pays ea, per minute while waiting and each caller pays eb, per minute

    during his own service. Suppose that each idle server costs ec, per hour and each

    busy server costs ed, per hour.

    What would be the net rewards per hour in the long run?

    What will be the average waiting time per client?

    Actually, we will see that a queue like in the previous example is a Markov chain. In

    order to treat the topic of Markov chains (rather) rigourously, we will spend half of

    the course on it. Markov chains can be categorized into two types. Section 2 involves

    discrete time Markov chains and Section 3 treats continuous time Markov chains.

    2 Discrete Time Markov Chains

    Stochastic processes involve time and uncertainty. They describe dynamic systems. A

    random variable X turns into a process when you flag it with a time moment. E.g.,

    X = {X0, X1, X2, X3,...} is a stochastic process, and Y = {Y(t) : t 0} is another one.

    More general, Xn need not be a real number, but can denote the stateof a system. The

    set of all possible states is called the state spaceof the process and is usually denoted by

    I. E.g., ifXn describes the weather type at day n, then I might be I = {sunny, moderate,freezing cold, stormy}. The interpretations ofX7 and Y(3.5) are the probabilistic values

    of X and Y at times 7 and 3.5 respectively. Examples of stochastic processes are the

    weather, the AEX.

    Stochastic processes have a time index set, denoted by T. There are two time index sets

    used in this course:

    Discrete time index set: T = IN0 = {0, 1, 2, 3,...}.

    4

  • 7/30/2019 Storm 2012

    6/62

    The variable n ( IN0) is used for discrete time moments.

    Continuous time index set: T = [0, ).

    The variable t ( IR+; the set of non-negative real numbers) is used for continuous

    time moments.

    A sampleor realization of a stochastic process is a function s : T I, so for each time

    moment, the function tells the state of the system at that moment. IfT = IN0, a sample

    is thereby an infinite sequence of states.

    In general, stochastic processes are extremely complex. In principle, the process is

    described by its so-called fdds (finite dimensional distributions), i.e., allprobabilities of

    the form P[Xt1 x1, . . . , X tn xn] with t1 < < tn T, (x1, . . . , xn) IRn. This is

    of course in general not practical; too complex. Simplifying assumptions are useful, for

    instance the i.i.d. assumption (see the Appendix).

    P[Xt1 x1, . . . , X tn xn] . .(by independency)

    ..= P[Xt1 x1] . . . P[Xtn xn] . .(because of identical distributions)

    ..= F(x1) . . . F(xn).

    The describing probabilities have become a lot more manageable. Actually, all informa-

    tion is given by a single distribution function. A less rigorous assumption is the Markov

    property.

    Sloppily formulated, in the discrete time context, if the process is in some state i at

    time n, then the probability that it moves to state j at time n + 1 depends only on the

    current state i. Let us formalize this:

    Definition 2.1 Let {Xn : n IN0} be a discrete time stochastic process with state

    spaceI. It has the Markov property if for all finite increasing sequences of time moments

    t1 < < tn+1 and all (i1, . . . , in+1) In+1

    P[Xtn+1 = in+1 | Xt1 = i1, . . . , X tn = in] = P[Xtn+1 = in+1 | Xtn = in]. (2.1)

    A process with the Markov property is called a Markov chain.

    An interpretation can be

    5

  • 7/30/2019 Storm 2012

    7/62

    t1, . . . , tn1: (times in the) past,

    tn: present,

    tn+1: moment in the future.

    Past and future are independent, given the present.

    Or, Where you go to doesnt depend on how you got here.

    The one-step transition probabilities are given by P[Xn+1 = j | Xn = i] for all n

    IN0, i , j I. If these do not depend on n, the Markov chain is said to be time-

    homogeneous; i.e., for all n IN0 we have

    pij := P[X1 = j | X0 = i] = P[Xn+1 = j | Xn = i]. (2.2)

    Of course, 0 pij 1 and

    jIpij = 1. They form the (one-step-transition-probability-

    matrix P. In the case I = IN0, it looks like

    . .P =

    p00 p01 p02

    p10 p11 p12

    p20 p21 p22 ...

    ......

    . . .

    ... ... ... . . .

    . .

    P is row-stochastic; each row-sum equals 1.

    If one knows that the distribution of Xn equals , the distribution of the system at the

    next time moment is readily found by right multiplying with P:

    P[Xn+1 = j | Xn ] :=iI

    P[Xn = i | Xn ] P[Xn+1 = j | Xn = i]

    = iI

    ipij

    = (P)j .

    (2.3)

    In principle, a time-homogenous DTMC is determined by its initial distribution (the

    distribution of X0) and its transition probability matrix, i.e., all fdds can be found if

    you know P and FX0.

    6

  • 7/30/2019 Storm 2012

    8/62

    2.1 Examples of Discrete Time Markov Chains

    Example 2 Playing RouletteA gambler starts with 20 fiches and constantly plays red at a roulette table. He bets

    for one fiche at a time and stops either when he is broke or owns 40 fiches. Intuitively,

    this is a finite but unbounded process, the gambler will eventually either reach his goal

    or will end up broke, but there is no bound on the number of turns until this happens.

    We have defined DTMCs on infinite time horizons. Still, this situation can be described

    by a DTMC by a simple modeling trick. Define I := {0, . . . , 40}, P[X0 = 20] = 1, and

    . .pij =

    1 if i = j = 0 or i = j = 40,1937

    if j = i 1,

    1837

    if j = i + 1,

    0 otherwise.

    . .

    States 0 and 40 are called absorbing. When a DTMC reaches an absorbing state, this

    can be interpreted as the end of the process (at least, nothing interesting will happen

    any further).

    Many other games, like the game of checkers (dammen), are Markov chains. Checkershas a huge state space, each possible configuration of men and kings (damstenen en

    dammen), combined with the information who is on turn, defines a state.

    Example 3 Simply counting can be considered to be a DTMC with X1 = 1, X2 =

    2, X3 = 3,... and an infinite state space I = IN.

    Example 4 An insurance problem: Bonus malus in Automobile Insurance

    Suppose that you have your car insured by the following system. There are four pos-sible premium classes with respective premium payments of size Ci. C1 is the highest

    premium, C4 the lowest. Each year without claim leads to a shift to one premium class

    higher next year (if not already in 4). Each claim brings you back to class 1 next year.

    If your car is damaged, should you pay for it yourself, or use your insurance?

    7

  • 7/30/2019 Storm 2012

    9/62

    . .

    E1 1

    '2

    1c1

    E

    1 2

    ' 32

    E

    1 3

    ' 43

    4T1 4

    . .

    Figure 1: A transition probability diagram

    The following policy may be optimal. Determine four numbers a1, . . . , a4, and make a

    claim if the damage is more than eai and your current premium equals Ci.

    Can we model this situation by means of a Markov chain? Since the chosen policy only

    uses information of the current year, (the current premium and damage), the Markov

    property is valid. Let Din be the event that in year n, you face at least once some damage

    to your car that exceeds ai. Assume that P[Din] = P[D

    im] and D

    imD

    in for all m = n.

    Let i := P[Di1].

    First define the state space: I = {1, 2, 3, 4}. The name of the state indicates the level

    of premium you have to pay in a typical year. It might give insights to draw a (one-

    step)-transition-probability-diagram. In such a diagram there is a circle for each state

    and an arrow from one state to another (or itself) if the system can move from the first

    to the latter in one unit of time. It is customary to put the name of the state inside its

    circle and to flag an arrow with the probability that it is used, given that the system is

    in the state at the tail of the arrow. Figure 1 displays the diagram corresponding to this

    Markov chain. It stores the same information as the transition probability matrix P:

    . .P =

    1 1 1 0 0

    2 0 1 2 0

    3 0 0 1 3

    4 0 0 1 4

    . . .

    8

  • 7/30/2019 Storm 2012

    10/62

    Example 5 The s-S-stock inventory system

    Suppose some shop sells some product as a regular stock item. The demand for it is

    stable over time. The demands in the successive weeks are independent of each other.

    Backlogging is not allowed: an empty stock means demand lost and unhappy customers.

    The owner of the shop uses the so-called periodic review s-S control rule for stock

    replenishment of the item. s and S are fixed natural numbers with 0 s < S. At

    Monday mornings, stock is inspected.

    Policy: When at inspection it turns out that the level of stock has decreased under the

    critical level s, a replenishment order of size S minus the current level of stock is

    placed.

    Before the store opens, the stock is back to maximal volume S. Typical question are:

    What is the average ordering frequency?

    What is the average amount of demand that is lost per week?

    These questions can be answered by the theory of Markov chains. Let D1, D2, . . . be

    an i.i.d.-sequence of integer valued random variables. Dn denotes the number of items

    requested in week n. Denote j = P[Dn = j] for all n, j IN0. So is the common

    pmf for the demand of any week. Time is discrete and expressed in weeks: t T =

    {0, 1, 2, . . . }. Assume that at time t = 0 the stock equals S. Let Xn denote the stock at

    the beginning of week n just prior to review.

    We get . .Xn+1 = max(Xn Dn, 0) if Xn s,

    max(S Dn, 0) if Xn < s.. .We get

    X1, . . . , X n1 only implicitly (i.e., via Xn) influence Xn+1. In other words, the process

    {Xn, n IN} possesses the Markov property. It is a discrete time-homogeneous Markov-

    chain with state space I = {0, 1, . . . , S }. How can we find the transition probabilities

    pij?

    If i s: pi0 = P[X1 = 0 | X0 = i] = P[D1 i] =k=i

    k,

    pij = P[D1 = i j] = ij. .(1 j i),

    pij = 0 . .if j > i.

    9

  • 7/30/2019 Storm 2012

    11/62

    If i < s: pi0 = P[D1 S] =k=S

    k,

    pij = P[D1 = Sj] = Sj. .if 0 < j S.

    The probability-matrix is . .

    k=S

    k S1 0

    ......

    ...k=S

    k S1 0

    k=i

    k i1 0 0 0

    ..

    .

    ..

    .

    . ..

    . ..

    ..

    .... i1 0 0k=i

    k i1 0

    i = 011

    ...

    i = s 111

    i = s11

    .

    .....

    i = S11

    . .The probability-matrix is

    . .

    2.2 Multiple Step Transition Probabilities

    If one knows the one step probabilities of a time homogeneous DTMC, it should in

    principle be possible to find out the chances to move in n steps from one state to another.

    The following result shows that these probabilities are, given P, easy to compute (at

    least for a computer). Let p(n)ij := P[Xn = j | X0 = i]. We have p

    (1)ij = pij and, by

    convention, p(0)ij =

    1 if i = j,0 if i = j.

    Theorem 2.2 p(n+m)ij =

    aI

    p(n)ia p

    (m)aj , or, in matrix notation, P

    (n+m) = P(n) P(m).

    In particular, the n-step transition matrix P

    (n)

    equals the n

    th

    power of P, which isdenoted by Pn. Therefore we might as well leave out the brackets again and denote the

    n-step transition probability matrix by Pn. Note that when we would denote p(n)ij simply

    by pnij this would cause ambiguity, so for the entries of Pn we keep the notation p

    (n)ij . To

    prove this theorem, condition upon Xn, given X0 = i:

    10

  • 7/30/2019 Storm 2012

    12/62

    Proof p(n+m)ij = P[Xn+m = j | X0 = i]

    = aI

    P[Xn+m = j | Xn = a, X0 = i] P[Xn = a | X0 = i]

    =aI

    P[Xn+m = j | Xn = a] P[Xn = a | X0 = i]

    =aI

    P[Xm = j | X0 = a] P[Xn = a | X0 = i]

    =aI

    p(m)aj p

    (n)ia .

    The second equality is nothing but the law of total probability. The third is due to the

    Markov property. The fourth is valid because of time homogeneity. . .2.

    2.3 Mean First Passage Times

    Can we find the expected time until a DTMC jumps to some specified state j? Let

    j := min{n 1 : Xn = j}. j is called the first passage time of state j.

    We have . . j IN if the chain ever visits state j after n = 0,

    j = else.. .We have

    The question can be answered by computing for a given state j and each state i the

    conditional mean first passage times ij := E[j | X0 = i] (we call jj the mean returning

    time of state j). This will be done by applying the law of total expectation telling that

    . .E[j] =iI

    ij P[X0 = i]. . .

    So, all we have to find out are, for a given state j, the values of ij (i I).

    Condition upon X1: ij =aI

    P[X1 = a | X0 = i]E[j | X0 = i, X1 = a]

    = P[X1 = j | X0 = i]E[j | X0 = i, X1 = j]

    . .+a=j

    P[X1 = a | X0 = i]E[j | X0 = i, X1 = a]

    = pij 1 +a=j

    piaE[j | X1 = a]

    = pij 1 +a=j

    pia(1 + E[j | X0 = a])

    = 1 +a=j

    piaE[j | X0 = a].

    11

  • 7/30/2019 Storm 2012

    13/62

    The second equality is due to the Markov property. The reason that E[j | X1 = a] =

    1 + E[j | X0 = a] for a = j is as follows. It takes 1 time unit to jump to a, and when

    in a at time moment 1, it takes another E[j | X0 = a] time units to jump to j for the

    first time. This proves the following theorem and corollary:

    Theorem 2.3 Let {Xn : n IN0} be a DTMC with state space I. Let for all i, j I,

    ij be defined by ij = E[minnIN

    : Xn = j| X0 = i]. Then for all i, j I we have

    ij = 1 +a=j

    piaaj. (2.4)

    For a given state j, and i running through the state space, we have |I| unknowns and the

    same amount of equations (one for each state i). Often, the set of equations has a unique

    solution. E.g., if I is finite and j < with probability 1 for all initial distributions.

    Why will be explained later on.

    Corollary 2.4 Let j I and define j to be the first time moment at which to system

    jumps to state j. Then E[j] =

    iIij P[X0 = i], which can be computed if the

    system of equations (2.4) can be solved.

    2.4 The Long Term Behavior of a DTMC

    In general, if we would like to predict the state of a DTMC in the next time unit, it is

    important to know the current state. However, if you would like to give probabilities of

    a state of the system in the far future, the current state might not be relevant at all.

    E.g., if we think of s-S-inventory (Example 5), the stock size at week 52 hardly depends

    on the stock size in week 1. On the other hand, if the gambler in Example 2 would have

    started with 1 fiche, his chances of ending up winning would be much lower than when

    he would have started with 39 fiches. Furthermore, if he starts with an even number of

    fiches, he will remain to have an even number of fiches at every even time moment. So in

    two ways, the initial state does give information about the far future. In this subsection,

    12

  • 7/30/2019 Storm 2012

    14/62

    we will get more insight about these phenomena. First, let us discuss the long term rate

    (frequency) of visits of a specific state j.

    Assume for the time being that X0 = i for some i I.

    The expected number of visits of state j at time 1 equal P[X1 = j] = pij .

    The expected number of visits of state j at time 2 equal P[X2 = j] = p(2)ij .

    ......

    ...

    The expected number of visits of state j at time n equal P[Xn = j] = p(n)ij .

    By aggregating, we find that the expected number of visits of state j in period {1, . . . , n}

    equaln

    k=1p(k)ij and the expected averagenumber of visits of state j in period {1, . . . , n}

    equal 1nn

    k=1p(k)ij . In the long run this leads to a frequency of visits of state j of

    . . limn

    1

    n

    nk=1

    p(k)ij . . .

    Without a proof we postulate that these limits always exist: for every j I, the long

    run frequency of visits of state j is well defined. Furthermore, for states that have a

    finite expected first passage time, these limits have an intuitive value:

    Theorem 2.5 Let j be a state of a DTMC with E[j] < . Then the long run rate of

    visits of a state j equals

    limn

    1

    n

    nk=1

    p(k)ij =

    1

    jj. (2.5)

    Note that even if E[j] < , it might be the case that jj = (can you provide an

    example?). In that case 1jj

    should be interpreted as 0. If we drop the assumption that

    X0 = i, Expression (2.5) turns in to

    . . limn

    1

    n

    nk=1

    iI

    p(k)ij P[X0 = i] =

    1

    jj. . .

    The intuition behind this is straightforward: If a bus station is visited by a bus about

    every half hour, the long run frequency of visits equals to 2 buses per hour. Likewise, if

    it takes jj time units on average between two visits of state j, the chain will visit this

    state 1jj

    times per time unit.

    13

  • 7/30/2019 Storm 2012

    15/62

    One might expect that the probabilities of being in a state will converge in the long run

    to these frequencies. If this is the case, we say that the system reaches equilibrium. The

    corresponding distributions are important enough to be named:

    Definition 2.6 A probability distribution {i : i I} is called an equilibrium or

    stationary distribution for DTMC{Xn, n IN } if

    . .j =kI

    kpkj , or, equivalently, = P. . .

    Such a is called stationary since if X0 , then X1 , and, by induction, Xn

    for all n IN (cf, Equation (2.3)):

    . .P[X1 = j | X0 ] = (P)j = j. . .

    So, if a chain starts in equilibrium, it remains in equilibrium forever. Many chains

    converge to equilibrium however regardless their initial distributions. Can we find suffi-

    cient conditions that guarantee existence of or even convergence to equilibrium? For the

    existence, it suffices to find a so-called regenerative state. A state r is called regenerative

    if ir < for all i. So regardless the initial state of the system, state r will be visited

    in finite time.

    Theorem 2.7 If DTMC {Xn : n IN0} has a regenerative state, then it has a unique

    equilibrium distribution = {j}jI and for any i I:

    . .j = limn

    1

    n

    nk=1

    p(k)ij . . .

    Instead of computing all these limits, it is far more easy to solve the system =

    P,

    iIpi = 1. We will provide the proof for chains with a finite state space, but it isfacultative.

    Proof (facultative) First assume that X0 = r. Define j = limn

    1

    n

    nk=1

    p(k)rj . So j

    is the long run frequency of being in state j. The long run frequency of being somewhere

    is 1, so

    jIj = 1, i.e., is a distribution.2

    2Here we apply that |I| is finite(!)

    14

  • 7/30/2019 Storm 2012

    16/62

    . .

    We have (P)j =I

    limn

    1

    n

    nk=1

    p(k)r

    pj

    = limn

    1n

    nk=1

    I

    p(k)r pj

    = limn

    1

    n

    nk=1

    p(k+1)rj

    = limn

    1

    n

    n+1k=2

    p(k)rj

    = j ,

    . .

    This makes an equilibrium distribution. Now, drop the assumption that X0 = r. Let

    i I be arbitrarily chosen and assume that X0 = i. Since, ir < , we know that, infinite time, state r is visited, say after T time units. Long run frequencies do not change

    if we do not take the first T time units into account:

    limn

    1

    n

    nk=1

    p(k)ij = lim

    n

    1

    n

    Tk=1

    p(k)ij +

    nk=T+1

    p(kT)rj

    limn

    1

    n

    nk=1

    p(k)ij = lim

    n

    1

    n

    Tk=1

    p(k)ij + lim

    n

    1

    n

    nk=T+1

    p(kT)rj

    limn

    1

    n

    n

    k=1

    p(k)ij = 0 + lim

    n

    1

    n

    nT

    k=1

    p(k)rj

    limn

    1

    n

    nk=1

    p(k)ij = lim

    m

    1

    m + T

    mk=1

    p(k)rj

    limn

    1

    n

    nk=1

    p(k)ij = lim

    m+T

    m

    m + T

    1

    m

    mk=1

    p(k)rj

    limn

    1

    n

    nk=1

    p(k)ij = lim

    m

    1

    m

    mk=1

    p(k)rj

    limn

    1

    n

    nk=1

    p(k)ij = j .

    So, regardless which state is the initial one, the long term frequencies form the equilib-

    rium distribution . Why are not there more equilibrium distributions? Suppose y is a

    stationary distribution. We have

    15

  • 7/30/2019 Storm 2012

    17/62

    . .

    y = yP

    y = yP2

    ...

    y = yPn

    yj =1

    n

    nk=1

    (yPk

    )j for all j I.. .

    . .

    Finally, yj =1

    n

    nk=1

    iI

    yip(k)ij

    =iI

    yi1

    n

    nk=1

    p(k)ij

    n

    iI

    yij

    = j.

    . .2

    To apply the previous theorem one needs to find a regenerative state. Therefore we need

    some notions concerning communication. This will be the topic of the next subsection.

    2.5 Communication

    Definition 2.8 State j is accessible from state i if p(n)ij > 0 for somen 0. We denote

    this byi j. If two states are accessible from each other, we say that they communicate,

    notation, i j.

    Properties of the relation are

    1. reflexivity i i for all i I . . (take n = 0: p(0)ii = 1 > 0)

    2. symmetry i j implies j i,

    3. transitivity i j and j k imply i k.

    These three properties make communication an equivalence relation, i.e., communication

    divides I in disjoint equivalence-classes. A Markov chain is called irreducibleif all states

    communicate.

    Example 6 Let I = {0, 1, 2, 3}, P =

    12

    12

    0 0

    12

    12 0 0

    14

    14

    14

    14

    0 0 0 1

    .

    16

  • 7/30/2019 Storm 2012

    18/62

    P does not define a Markov chain alone, an initial distribution, i.e., the distribution of

    X0, is needed as well. However, the state classification does not depend on the initial

    distribution. Draw yourself a transition probability diagram. We have, p33 = 1, so 3 is

    an absorbing state. The communication partition is

    {0, 1}, {2}, {3}

    .

    Definition 2.9 A set of states C I is called closed, if

    . .Xi C = Xi+1 C. . .

    A set is minimal closed if it does not contain a proper closed subset.

    I itself is by this definition always closed. In the previous example, the closed sets are

    I, {0, 1, 3}, {0, 1} and {3}. Without a proof we postulate

    Theorem 2.10 Let {Xn : n IN0} be a DTMC with finite state space I. Then there

    exists a regenerative state if and only if the chain has a unique minimal closed set.

    Corollary 2.11 A DTMC with a finite state space has a unique equilibrium distribution

    if and only if it has a unique minimal closed set.

    The next question that arises is: will this equilibrium always be reached, i.e., is it the

    case that limn

    P[Xn = j | X0 = i] = j for all initial states i? A counter example is

    readily found:

    Example 7 Take I = {0, 1}, P =

    0 1

    1 0

    , X0 = 0. Both states are regenerative,

    and the equilibrium distribution is given by 0 = 1 =12

    . Nevertheless, limn

    P[Xn = 0]

    does not exist since P[Xn = 0] keeps on toggling between 0 and 1.

    To ensure convergence one other notion is needed: periodicity.

    Definition 2.12 Let i be a state with ii < . Theperiod of state i is said to be d if

    d is the greatest common divisor3 of the indicesn 1 for which p(n)ii > 0. A statei with

    period d = 1 is said to be aperiodic.

    3Let A IN. The greatest common divisor, or gcd for short, is the largest natural number such that

    all members ofA are divisible by this number: gcd(A) = max{k IN : ak IN for all a A}.

    17

  • 7/30/2019 Storm 2012

    19/62

    Theorem 2.13 Let {Xn : n IN0} be a DTMC with a finite state space and an ape-

    riodic regenerative state. Then in the long run the chain reaches its unique equilibrium

    , i.e., . . limn

    P[Xn = j | X0 = i] = j for all i, j I. . ., i.e.,

    Resuming, for each DTMC we have

    For every i, j I, the long run frequency of visits of state j given that X0 = i is

    well defined. IfE[j ] < , then limn

    1n

    nk=1

    p(k)ij =

    1

    jj.

    If the process possesses a regenerative state, these limits do not depend on the initial

    state and we can define j

    = limn

    1

    nn

    k=1p

    (k)

    ij. It holds that = P.

    If moreover the state space is finite, then

    jIj = 1 and is the unique equilibrium

    distribution.

    If moreover the regenerative state is aperiodic, the process converges to equilibrium,

    i.e., limn

    P[Xn = j | X0 = i] = j for all i, j I.

    2.6 Cost-Reward Structures on Finite Irreducible DTMCs

    Let r : I IR be a reward function, i.e., every time state i is visited, a reward of

    r(i) is gained. If{X} is finite and irreducible, then we can apply the following intuitive

    theorem.

    Theorem 2.14 Let {Xk : k IN0} be an irreducible DTMC with finite state space I

    and let f : I IR be a cost- (reward-) structure. Let be the unique equilibrium

    distribution of the Markov chain (cf. Theorem 2.7). Then

    limn

    1

    n

    nk=1

    f(Xk) =jI

    jf(j). (2.6)

    The left hand side of Equation (2.6)concerns long run average costs, the right hand

    side concerns the mean costs per time unit at equilibrium. Note that even if the chain

    does not converge to equilibrium (because of periodicity), the theorem is still valid.

    18

  • 7/30/2019 Storm 2012

    20/62

    Often costs (rewards) r(i j) are made at transitions rather than at states. If so, is it

    still possible to apply the theorem above? We can perform the following trick: consider

    the coming transition costs as expected costs of the current state.

    Example 8 A taxi rides between three cities 0,1 and 2, so I = {0, 1, 2}. After each

    ride, it waits in the city where it brings its client to until a new client asks for a ride.

    Let X0 be the city where the taxi starts its working day and let Xn denote the city to

    which the taxi brings its nth client. Suppose the (mean) waiting costs depend on the

    city where the taxi is situated and are expressed by c(i) (i I). The rewards per

    ride depend on both departure and destination and are called r(i j). We can define

    a (net) reward structure by

    . .f(i) = c(i) + E[reward next ride] = c(i) +jI

    pijr(i j) . .

    Theorem 2.14 can be applied to find that the average reward per ride will be

    . .

    j{0,1,2}

    jf(j).. .

    19

  • 7/30/2019 Storm 2012

    21/62

    3 Continuous Time Processes

    Like discrete time stochastic processes, continuous time processes have a state space,denoted by I. The time index set of a continuous time process is an interval, usually

    T = [0, ). Where a discrete time process can jump to another state only at integer

    time moments, a continuous time process X = {X(t) : t 0} can jump to another state

    at any moment in T. In this course, we will only consider processes that only switch in

    finite segments of time only finitely many times. Therefore, a sample of such a process is

    a so called step-function. To visualize such a function, put time on the x-axis and state

    space I the y-axis. The graph will be horizontal everywhere but on a sequence of timemoments. A continuous time process starts at time 0 at some state i, i.e., X(0) = i,

    stays there for a while, and then jumps after a random while T1 to another state j, so

    j = i. Then it stays at j for a random period of length T2 before it makes its second

    transition (jump), and so on. Just as in the discrete time case, we will focus on processes

    with the Markov property.

    3.1 The Markov Property for Continuous Time Processes

    The Markov property can be copied from the discrete time context, except the time

    index is flagged differently (X(t) instead of Xn):

    Definition 3.1 A stochastic process {X(t) : t [0, )} with state space I possesses

    the Markov property if for all increasing sequences of time moments t1 < < tn+1 and

    i1, . . . , in+1 in I we have

    P[X(tn+1) = in+1 | X(t1) = i1, . . . , X (tn) = in] = P[X(tn+1) = in+1 | X(tn) = in]. (3.1)

    Such a process is called a Continuous Time Markov Chain or CTMC for short. The

    interpretation of the Markov property remains the same. Given the present, the history

    of a Markov process does not influence its future, or, cryptically,

    . .[X(future) | X(present)] X(past). . .

    20

  • 7/30/2019 Storm 2012

    22/62

    A CTMC is called time-homogeneous if for all u > 0, i, j I,

    P[X(t + u) = j | X(u) = i] = P[X(t) = j | X(0) = i]. (3.2)

    This course, time-homogeneity is assumed.

    Suppose we know the state in which the process starts, let us say X(0) = i. Let T1 be

    the time it takes until the first jump (transition), so X(u) = i for all u [0, T1) and

    X(T1) = i. It is convenient and customary to assume that T1 is a proper random variable

    in the sense that P[T1 = ] = 0. We have

    P[T1 > t + s | T1 > s, X(0) = i]= P[X(u) = i u [s, t + s] | X(u) = i u [0, s]]

    = P[X(u) = i u [s, t + s] | X(s) = i]

    = P[X(u) = i u [0, t] | X(0) = i]

    = P[T1 > t | X(0) = i].

    (3.3)

    The second equality is due to the Markov property and the third one uses time ho-

    mogeneity. We conclude that, given the initial state, the time it takes until the first

    transition has the memoryless property! So, there is a thin line between the Markov

    property and the memoryless property. The first class of a CTMCs we will discuss is

    the class of Poisson processes:

    3.2 Poisson and Other Counting Processes

    A counting process is an idealized stochastic model to count events that occur ran-

    domly in time (generically called renewals or arrivals). Counting processes are used as a

    foundation for building more realistic models. They are often found embedded in other

    stochastic processes, e.g., to describe incoming calls at a service center.

    Definition 3.2 A counting process N = {N(t) : t 0} is a stochastic process with

    states space I = {0, 1, 2, 3, } such that there exists a series of nonnegative random

    variables T1, T2, T3, with

    N(t) = 0 for all t [0, T1),

    21

  • 7/30/2019 Storm 2012

    23/62

    N(t) = 1 for all t [T1, T1 + T2),...

    N(t) = n for all t with T1 + Tn t < T1 + Tn+1.

    Let Sk := T1 + + Tk, so Sk denotes the time moment at which the kth event takes

    place. N(t) denotes the number of events until time t, i.e.,

    . .N(t) := max{k IN | Sk t}. . .

    {N(t) | t 0} or simply N, is called the counting process with interarrival times

    T1, T2, . . . Typical events that are counted in this course are

    customers entering a shop,

    people joining a queue,

    internet users visiting a web page,

    job completions of a server,

    machine breakdowns; replacements.

    In many models it is assumed that the times Tn between the successive arrivals are

    independent and identically distributed. If so, we also speak of a renewal process. If,

    moreover, it is assumed that the interarrival times are exponentially distributed, we

    speak of a Poisson Process.

    Definition 3.3 A counting process {N(t) : t 0} with interarrival times T1, T2,...

    is called a Poisson process (shortly PP or PP()) if the interarrival times form an i.i.d.

    sequence and Ti exp() for some > 0 and all i IN.

    The parameter is the frequency with which the to be counted events occur. It is

    called the arrival-rate, failure-rate, service-rate, depending on what is counted. The

    interpretation of is as follows: it takes on average 1 time units between two events

    that count. So there are events per hour to be expected. So, the main property of a

    Poisson process is that the events occur at a constant rate.

    Why has a Poisson process the Markov property? It is difficult to give a formal proof,

    but an intuitive argument is readily given. Suppose you have been monitoring a Poisson

    22

  • 7/30/2019 Storm 2012

    24/62

    process up to time moment t. You have stored every time moment at which the events

    occur in the list (S1, . . . , S N(t)). What does this tell about the forthcoming part of the

    process? Because of the constant arrival rate, the waiting time until the next arrival

    is exp() distributed, you did not need to observe the process to know that. The only

    relevant information that the current sequence (S1, . . . , S N(t)) gives, is the number of

    arrivals so far, i.e., N(t). Therefore the present situation (N(t)) gives the same relevant

    information about the future of the process as the complete history {N(s) : s [0, t]}.

    The random variable Sn, being the sum ofn independent exponentially distributed ran-

    dom variables, and has a so called Erlang distribution, see e.g., http://en.wikipedia.

    org/wiki/Erlang_distribution . Its distribution function is given by

    P[Sn t] = FSn(t) = 1 n1k=0

    et(t)k

    k!. (3.4)

    This enables us to find the distribution of N(t), the number of arrivals up to time t:

    Proposition 3.4 Let N = {N(t) : t } be a Poisson process with rate . Then for

    each t > 0, N(t) has the Poisson(t) distribution, i.e.,

    P[N(t) = n] = et(t)n

    n!. (3.5)

    Proof The event {N(t) n} can also be described by {Sn t}. Hence,

    . .

    P[N(t) n] = P[Sn t]

    = 1 n1k=0

    et(t)k

    k!.

    . .

    Therefore, P[N(t) = n] = P[N(t) n] P[N(t) n + 1]

    =

    1 n1k=0

    et(t)k

    k!

    1 n

    k=0

    et(t)k

    k!

    = et(t)n

    n!.. .

    23

    http://en.wikipedia.org/wiki/Erlang_distributionhttp://en.wikipedia.org/wiki/Erlang_distributionhttp://en.wikipedia.org/wiki/Erlang_distributionhttp://en.wikipedia.org/wiki/Erlang_distributionhttp://en.wikipedia.org/wiki/Erlang_distribution
  • 7/30/2019 Storm 2012

    25/62

    As a consequence we find that

    E[N(t)] =

    n=1

    n et (t)

    n

    n!

    = ettn=1

    (t)n1

    (n 1)!

    = ettk=0

    (t)k

    k!

    = ett et

    = t.

    (3.6)

    This makes sense; if on average events per time unit take place, then after t time unitsyou can expect t events so far.

    Example 9 Taxis wait for passengers at Central Station in Amsterdam. A so called

    trein-taxi is a cheap alternative for a regular taxi. Several passengers share one taxi.

    Suppose that passengers arrive according to a P P() process with = 0.4, being the

    rate of passengers arriving at the taxi standing per minute. A taxi departs if either 4

    passengers have been collected, or the first passenger has waited for 10 minutes.

    a. What is the probability that the first passenger has to wait for 10 minutes?

    b. What is the probability that the first passenger has to wait for 10 minutes, given

    that a third passenger enters exactly 5 minutes after the first?

    a. Set t = 0 as the moment that the first passenger enters the taxi. He has to wait

    for 10 minutes if less than 3 other passengers arrive at the scene. In terms of a

    Poisson()-process, this is the event {N(10) < 3}. The random variable N(10) has

    a Poisson(10)-distribution, so the answer to question a. equals

    . .P[N(10) < 3] =2

    n=0

    e10(10)n

    n!. . .

    b. Set t = 0 as the moment that the third passenger enters the taxi. Now, the first

    passenger has to wait for 5 more minutes if no passenger shows up in 5 minutes time.

    24

  • 7/30/2019 Storm 2012

    26/62

    This is the event {N(5) = 0}. The answer to question b. equals P[N(5) = 0] =

    e5.. .

    The remainder of this section will be spent on properties of Poisson processes. The first

    four characterize Poisson processes.

    (A) A Poisson process has got Independent increments:

    ..if [a, b] and [c, d] are disjoint time-segments, then N(b) N(a) N(d) N(c). ..

    (B) A Poisson process has got Stationary increments, i.e., the distribution of the number

    of arrivals within a time segment depends on the length of the segment, but not on

    its starting time. In formula,

    . .for all t 0: N(b) N(a) N(b + t) N(a + t). . .

    As a consequence, for all s 0, {N(t + s) N(s), t 0} is another Poisson process.

    (C) If t > 0 is a small increase in time, then the probability of one arrival in an

    interval [a, a + t] is about t. Formally,

    . .P[N(a + t) N(a) = 1] = t + o(t), . .

    in which o(t) denotes an unspecified expression (called the error term) with the

    property that limt0o(t)t = 0.

    The proof of property (C) uses the standard summation ex =

    k=0xk

    k! . We have

    et =1

    0! +(t)

    1! +(t)2

    2! (3.7)

    Hence, P[N(a + t) N(a) = 1] = P[N(t) = 1]

    = et t

    = t[1 t + 12

    (t)2 ]

    = t + o(t).

    25

  • 7/30/2019 Storm 2012

    27/62

    In this elaboration the first equality is due to Property (B), the second uses Proposi-

    tion 3.4, and in the final one all negligible terms (the terms containing a factor t2)

    are put in error term o(t).

    (D) For small t > 0, the probability on two or more arrivals in time segment [a, a + t]

    is negligible. Formally,

    . .P[N(a + t) N(a) 2] = o(t). . .

    The proof of property (D) is facultative:

    . .

    P[N(t) 2] = 1

    P[N(t) = 0]

    P[N(t) = 1]

    = 1 et ett

    = 1 (1 + t)et

    = 1 (1 + t)(1 t + o(t))

    = 2(t)2 + o(t)

    = o(t).

    . .

    The fourth equality is valid because of Lemma A.5.

    Without a proof, we state that the four properties above characterize a Poisson process:

    Theorem 3.5 Any counting process that obeys properties (ABCD) is a Poisson process.

    Corollary In order to show that some counting process is a Poisson process, just

    check properties (ABCD).

    Suppose, traffic is analyzed in order to be able to decrease the congestion at some road.

    The number of vehicles passing a certain road crossing is counted, which turns out to

    produce a Poisson() process {Nv(t) : t 0}. After a while, one realizes that it is

    relevant to count cars and motorcycles separately. Luckily the person who has stored

    so far the arrival moments has also written down the type of the vehicle. This leads to

    two new counting processes {Nc(t) : t 0} and {Nm(t) : t 0} respectively, obeying

    Nv = Nc + Nm. Under which condition will these new processes be Poisson processes as

    26

  • 7/30/2019 Storm 2012

    28/62

    well, and if so, with what rates? The next theorem (the proof is skipped) answers this

    question. A lot of independency has to be assumed, but nothing further.

    Theorem 3.6 [Splitting Theorem for Poisson Processes]

    Let {N(t) | t 0} be PP(). Suppose there are two types of arrivals, 1 and 2. Suppose,

    independent of everything, an arrival is of type i with probability pi (0, 1). Let

    {Ni(t) | t 0} count the arrivals of type i. Then N1 and N2 are two independent

    Poisson processes with arrival rates 1 := p1 and 2 := p2.

    A similar result works the other way round. Given that the processes Nc and Nm are

    independent Poisson processes, the counting of vehicles Nv := Nc + Nm is a Poisson

    process as well:

    Theorem 3.7 [Merging Theorem for Poisson Processes]

    Let {N1(t) : 0} and {N2(t) : t 0} be independent Poisson processes with arrival

    rates 1 and 2. Then {N(t) : t 0} defined by N(t) := N1(t) + N2(t) for all t is a

    Poisson process with arrival rate 1 + 2.

    Let Ik be the type of the kth arrival,

    and Xk the kth interarrival time of the merged process N.

    Then P[Ik = i] = P[Ik = i | Xk = t] =i

    1 + 2(so IkXk).

    The proof of this theorem is beyond the scope of this course as well, but the following

    convenient corollary can be shown relatively easy.

    Corollary 3.8 Let X Poisson(1), Y Poisson(2), XY. Then X + Y

    Poisson(1 + 2).

    Proof Let Ni be a P P(i) process (i {1, 2}) with N1N2. Then X N1(1)

    and Y N2(1). Hence, X + Y [N1 + N2](1). By the merging theorem, N1 + N2 is a

    P P(1 + 2) process. So, [N1 + N2](1) Poisson(1 + 2).. .

    27

  • 7/30/2019 Storm 2012

    29/62

    Finally, we will discuss a relation between Poisson processes and the uniform distribution.

    It is about the conditional distribution of an arrival. Suppose it is known that in a given

    time segment [0, t] exactly one arrival has taken place, i.e., N(t) = 1. What is the

    (conditional) distribution of the time of arrival of this customer? Or, suppose you know

    that a soccer game ended in 1-0. In which minute the goal has been scored? In formula,

    find . .P[S1 x | N(t) = 1]. . .find

    One might have the intuition that time segments of equal length have equal probabilities

    to contain the arrival, so the uniform distribution sounds reasonable. Let U U[0, t] be

    a random variable with the uniform distribution on [0, t].

    Lemma 3.9 For all x [0, t] we have

    P[S1 x | N(t) = 1] = P[U x] =x

    t. (3.8)

    Proof

    . .

    KK

    P[S1 x | N(t) = 1] =P[S1 x, N(t) = 1]

    P[N(t) = 1]KK

    =P[N(x) = 1, N(t) N(x) = 0]

    P[N(t) = 1]KK

    = P[N(x) = 1]P[N(t) N(x) = 0]P[N(t) = 1]K

    K

    =xexe(tx)

    tetKK

    =x

    t.

    . .

    The second equality might be best understood by drawing a sample. Indeed, under the

    knowledge of {N(t) = 1}, the event {S1 x} is also described by telling that N(x) = 1

    and N(t) N(x) = 0. The third equality follows from Properties (B) and (A). The

    fourth equality is valid since all probabilities involve events with Poisson distributions.. .

    As a consequence, S1|{N(t)=1} has distribution U[0, t] and E[S1 | N(t) = 1] =12

    t. The

    same intuition that, given {N(t) = n}, each small interval [a, a + t] [0, t] has the

    same probability to contain an arrival leads to the idea that it is possible to simulate a

    Poisson process conditioned on {N(t) = n} by means of n times applying the uniform

    28

  • 7/30/2019 Storm 2012

    30/62

    distribution on [0, t]. This is expressed by the following theorem, which is a direct

    generalization of Lemma 3.9:

    Theorem 3.10 For each t > 0 and n IN,

    . .P[S1 x1, S2 x2, . . . , S n xn | N(t) = n] . .

    . .equals . .

    . .P[U(1) x1, U(2) x2, . . . , U (n) xn], . .

    in which U1, . . . , U k (i.i.d) and Ui U[0, t], and U(k) is the kth in line if U1, . . . , U n are

    arranged in increasing order.

    In words: Suppose you generate n samples with distribution U(0, t) and arrange the

    outcomes in increasing order, resulting in n outcomes

    . .U(1), . . . , U (n) . .

    with U(1) < < U(n). Then this n-tuple has the same distribution as

    . .S1|{N(t)=n}, . . . , S n|{N(t)=n}. . .

    The theorem itself is facultative, but you might want to apply its consequence:

    Corollary 3.11 For all integers k, n with 1 k n we have

    . .E[Sk | N(t) = n] =kt

    n + 1. . .

    An intuitive argument for the validity of this corollary is as follows. The symmetric

    shape of the uniform density gives E[S1 | N(t) = 1] =t2

    . For notational purposes,

    denote S0 = 0 and Sn+1 = t. The same argument gives that Sk is expected to be

    situated right in the middle of time segment [Sk1, Sk+1] for all k {1, . . . , n}. So, all n

    interarrival times Tk and time segment [Sn, t] have equal expected lengths, i.e., t/(n+1).

    Hence, E[Sk | N(t) = n] = E[T1 + + Tk | N(t) = n] =kt

    n + 1.

    3.3 Continuous Time Markov Chains

    This section returns to general stochastic processes with a continuous time index set

    and the Markov property. Furthermore, we assume time homogeneity and restrict the

    29

  • 7/30/2019 Storm 2012

    31/62

    number of jumps to finitely many in finite time segments. Throughout the section

    X = {X(t) : t 0} is a process with the mentioned properties and state space I.

    Let i I be any state and assume that at some time X(t) = i. Let Ji be the time it

    takes until the first jump (transition) out of state i, so X(u) = i for all u [t, t + Ji) and

    X(t + Ji) = i. Ji is called the sojourn time (verblijfstijd) at i. If the chain never leaves

    state i, then Ji := and state i is called absorbing. To avoid exceptions, we assume

    that there are no absorbing states. In order to determine the distribution function of

    Ji, we might as well take t = 0, because of time-homogeneity. In that case Ji = T1,

    which has been defined to be the time it takes until the first jump of the system (see

    Subsection 3.1). We have seen that T1 has the memoryless property and has therefore

    an exponential distribution. Hence, there exists a positive real number i such that

    Ji exp(i). i is called the transition rate at state i.

    So, given that the system is in state i at a certain time, it stays there for an exponentially

    distributed time. What about the next state, i.e., can we say something about the

    probability that it will be, say, state j? Because of time homogeneity and the Markov

    property, the state to which the system jumps may depend on the current state i, but

    not on any further information about how the system got into state i. Therefore, just

    like in the discrete time setting, we can speak of transition probabilities.

    pij := P[j is the next state | i is the current state]. (3.9)

    Unlike in the discrete time setting, pii > 0 is not possible; the system can only jump

    to an other state since transitions are associated with changes of the state of the sys-

    tem. Whereas we spoke of one step transitions in the previous section, here one jump

    transitions would be more appropriate.

    Just like the class of Poisson processes, the class of CTMCs without absorbing states

    can be characterized by a list of four properties:

    Theorem 3.12 Let X = {X(t) : t 0} be a stochastic process with state space I.

    Then X is a CTMC without absorbing states if and only if it obeys the following four

    30

  • 7/30/2019 Storm 2012

    32/62

    properties:

    a) Each time the process reaches state i, it stays there for a stochastic period with

    length exp(i). . .(i I)

    b) If the chain leaves statei, it jumps to statej with probabilitypij , with

    j=ipij = 1

    and pii = 0.. .(i, j I)

    c) All sojourn times and transitions are independent on each other.

    d) There exists an M IN such that 0 < i M for all i I.

    Property d) might appear out of the blue. This assumption avoids the phenomenon of

    an explosion of jumps, as the following example displays.

    Example 10

    Suppose we have a counting process N = {N(t) : t 0} obeying properties a), b), and

    c) with i = 2i for all i I. Since N is a counting process, pi,i+1 = 1 for each i IN0

    and the chain starts in state N(0) = 0. The expected time until it jumps to state 1

    equals 10

    = 1. When in state 1, the next arrival will take place after an expected time

    of length 11 =12 . Each consecutive jump will happen in expectation twice as fast as the

    previous one. So E[Sk] = 1 +12 + +

    12k1 = 2

    12k . After 2 time units, the expected

    number of jumps is infinite! So after a while one could say that X(t) = , but / I.

    So, the example is not only unrealistic, it is also infeasible. That is why property d) is

    needed.

    Let us turn the attention to a widely applied model:

    Example 11 The M/M/1-queue

    In queueing theory, Kendalls notation is the standard system used to describe and

    classify the queueing model that a queueing system corresponds to. First suggested

    by D. G. Kendall in 1953 as a three factor A/B/C notation system for characterizing

    queues, it has since been extended to include up to six different factors, but we stick to

    the original three. See e.g., http://en.wikipedia.org/wiki/Kendalls_notation if

    you would like to have more information.

    31

    http://en.wikipedia.org/wiki/Kendall's_notationhttp://en.wikipedia.org/wiki/Kendall's_notation
  • 7/30/2019 Storm 2012

    33/62

    M: The first M represents exponential interarrival times of the customers that form a

    queue. Usually the arrival rate is called . Let T1, T2, . . . represent the sequence of

    interarrival times. It is assumed that these are i.i.d. Tn exp(), so the related

    process that counts the arrivals of customers is a Poisson() process.

    M: The second M denotes that service times have exponential distributions as well. The

    service completion times SC1, SC2, . . . form another (i.i.d) sequence and the service

    rate is usually denoted by , so SCn exp() for all n IN.

    1: The third factor denotes the number of servers; 1 in this example.

    Let X(t) give the number of customers present at time t, including the one at service (if

    any). Does {X(t) : t 0} obey the Markov-property? Instead of verifying the property

    directly, we might as well check the four characterizing properties a), b), c) and d).

    A transition from one state to another can be caused by two types of events; arrivals

    of customers and service completions. In state i = 0, the server is idle, so J0 exp()

    and 0 = . In any other state i > 0, the transition time Ji until a jump will take

    place can be expressed by Ji = min{T , S C}, in which T stands for a typical interarrival

    time and SC for a typical service completion time. Due to the memoryless property it

    does not matter how long a service is already in progress or how long ago the previous

    customer arrived. Theorem A.2 tells that the minimum of two independent exponentially

    distributed random variables is exponentially distributed itself: Ji = min{T , S C}

    exp( + ). This gives property a): Ji

    exp(), if the server is free,

    exp( + ), if the server is busy.

    b) Theorem A.4 tells that P[min{T,SC} = T] = + , yielding

    . .pij =

    1 if i = 0 and j = 1,

    /( + ) if 0 < i = j + 1,

    /( + ) if 0 < i = j 1,

    0 if |i j| = 1.

    . .

    c) is valid by assumption (or not, but then there is no story).

    d) 0 < i + for all i I.. .

    32

  • 7/30/2019 Storm 2012

    34/62

    3.3.1 Time Dependent Transition Probabilities and Transition Rates

    Let, for all t

    0 and i, j I:

    pij(t) := P[X(j) = t | X(0) = i]. (3.10)

    In words, pij(t) is the probability that when in i, the system is in state j after t time

    units. It is not required that the movement from i to j takes place in a single transition.

    In general, pij(t) is very hard to derive. However, the cases that t is close to zero and

    that t tends to infinity can be manageable. Let t > 0 and think of a number so small

    that t2 is negligible compared to t.

    The probability that two transitions take place in period [t, t + t] is negligible because

    if X(t) = i and ifk is the next state visited after t, then

    P[2 jumps in [t, t + t]] P[Ji < t] P[Jk < t]

    = (1 eit) (1 ekt)

    (1 eMt)2

    = (Mt + o(t))2

    = o(t).

    (3.11)

    The second inequality uses property d) of a CTMC (see Theorem 3.12). The penultimate

    equality is valid by Lemma A.5. Similarly, the probability that three (or any other

    natural number exceeding two) transitions take place in period [t, t + t] is negligible.

    Therefore,

    pij(t) = P[a single jump to j in period [0, t] | X(0) = i] + o(t)

    = P[any jump in [0, t] | X(0) = i] P[first jump is to j | X(0) = i] + o(t)

    = P[Ji < t | X(0) = i] P[first jump is to j | X(0) = i] + o(t)= (1 eit) P[first jump is to j | X(0) = i] + o(t)

    = (it + o(t)) pij + o(t)

    = i pij t + o(t).

    The second equality uses the independence of Ji and the destination of the first jump

    (property c)). For the fifth equality we refer again to Lemma A.5. We conclude that

    . . limt0

    pij(t)

    t= i pij . . .

    33

  • 7/30/2019 Storm 2012

    35/62

    In words, when the chain is in state i, jumps to state j occur at a constant rate ofpij i.

    Hence, we define . .qij = i pij , . .Hence, we define

    and call qij the transition rate from i to j. If it is not possible to jump directly from

    i to j (i.e., pij = 0), this is expressed by a transition rate from i to j equal to 0. The

    notions pij, qij , and i relate to each other by pij = qij

    iif j = i,

    0 if j = i.

    In particular, we havej=i

    qij =j=i

    ipij = i. In words, the sum over all rates out

    of state i with a specified destination equals the rate out of state i without a specified

    destination.

    Instead of drawing a transition probability diagram, it is more natural and more infor-

    mative to draw a transition rate diagram; simply replace the probabilities by the rates

    (the arrows remain the same). E.g., Figure 2 displays the transition rate diagram of

    an M/M/1-queue. Just like a one step transition probability matrix for DTMCs, we

    . .E'

    0 E'

    1 E'

    2 E'

    3 . .

    Figure 2: The transition rate diagram of an M/M/1-queue.

    denote the (one jump) transition probability matrix by P. If the state space I equals

    IN0 it looks like

    P =

    0

    1

    2...

    0 p01 p02 p03 . . .

    p10 0 p12 p13 . . .

    p20 p21 0 p23 . . ....

    . . .. . .

    . . .. . .

    . (3.12)

    Similarly we introduce a transition rate matrix, usually denoted by a capital A and

    entries aij . Here,

    aij := qij if j = i,

    i if j = i,(3.13)

    34

  • 7/30/2019 Storm 2012

    36/62

    so the transition rate matrix for a CTMC with state space IN0 is given by

    A =

    01

    2...

    0 q01 q02 q03 . . .q10 1 q12 q13 . . .

    q20 q21 2 q23 . . ....

    . . . . . . . . . . . .

    . (3.14)

    The diagonal of A contains the unspecified transition rates i with minus signs. There-

    fore, the rows of A sum up to 0. It may seem a rather artificial notation. Why this has

    been done will be clarified when discussing equilibrium distributions.

    Example 12 A single product inventory

    Customers arrive P P(). Each customer buys 1 unit if available, no backlogging is

    allowed. Opportunities to replenish occur P P(). Refill is only possible if the stock

    is empty. Each refill has deterministic size Q. Define X(t) to be the amount of stock at

    time t. Assume that X(0) = Q. Is X = {X(t) : t 0} a time homogeneous CTMC?

    Its state space is I = {0, . . . , Q}. Instead of verifying the definition of a CTMC, we

    apply Theorem 3.12 and check the four characterizing properties. The argumentation

    why properties a) and b) are valid is similar as in the M/M/1-example 11 and therefore

    left out.

    Property a): i =

    if i {1, . . . , Q},

    if i = 0.

    Property b): pij =

    1 if j = i 1 or (i, j) = (0, Q),

    0 else.

    Property c): True by assumption.

    Property d): 0 < i M := max{, }.

    We conclude that X is a CTMC. The transition rate and probability matrices are re-

    spectively

    35

  • 7/30/2019 Storm 2012

    37/62

    . .A =

    0 0

    0 0

    0 . . . ......

    . . . . . . . . . . . ....

    .... . . 0

    0 0

    and P =

    0 0 1

    1 0 0

    0 1 0 ......

    . . . . . . . . ....

    .... . . . . . . . .

    ...

    0 0 1 0

    . . .

    . .

    3.3.2 Equilibrium Distributions

    Intuitively, {X(t) : 0} is in equilibrium if there exists a distribution {pj}jI with

    P[X(t) = j] = pj for all j I, t 0. In words, for all time moments, the probability to

    be in a certain state is the same. If so, we have for all t 0, t > 0,

    P

    system reaches j in [t, t + t]

    = P

    system leaves j in [t, t + t]

    k=j

    Pjump from k to j in [t, t + t]

    + o(t)

    k=j

    Pjump from j to k in [t, t + t]

    + o(t)

    k=j

    P[X(t) = k](qkjt + o(t))k=j

    P[X(t) = j](qjkt + o(t))

    k=j

    pkqkjt + o(t)k=j

    pjqjkt + o(t)

    Divide by t and let t 0 to infer that

    k=j

    pkqkj = pjj. (3.15)

    This gives rise to the following definition of an equilibrium distribution.

    Definition 3.13 A probability distribution {pj, j I} is an equilibrium distribu-

    tion (or stationary distribution) if for all j I. we have

    . .k=j

    pkqkj = pjj . . .

    36

  • 7/30/2019 Storm 2012

    38/62

    The equation can also be found by taking the inner product of p and the jth column

    of A. This is exactly the reason why A has been defined as it has. Apparently, if p is

    an equilibrium distribution, then pA = 0. So, to find an equilibrium distribution, just

    solve the system

    pA = 0,jI

    pj = 1, p 0. (3.16)

    Often it is convenient to search for an x with xA = 0, and then normalize.

    The jth equation of pA = 0 can be interpreted as follows. When the system is in

    equilibrium, events forcing the system to move into state j happen with the same rate

    as events that move the system out of state j:

    rate into state j =k=j

    pkqkj = pjj = rate out ofj. (3.17)

    This is called a balance equation or flow rate equation. More generally,

    . .rate into set A = rate out of set A, . .(A I)

    i.e., . .k/A

    pkjA

    qkj =jA

    pjk/A

    qjk . . .

    Assume that an equilibrium p exists (later on we will discuss when this assumption is

    righteous). Such equations can be visualized by a transition rate diagram in which in

    each circle the equilibrium probability pi of the state i that it represents is put. Let us

    clarify this by discussing the M/M/c-queue.

    Example 13 The M/M/c-queue (the Erlang delay model)

    Two models are named after the Danish Mathematician Agner Krarup Erlang. In the

    Erlang loss model potential customers are not willing to wait when all servers are busy

    and the revenues they generate will be lost. In the Erlang delay model all customers

    patiently wait for their turns. The two Ms in M/M/c denote that both the interarrival

    times between consecutive customers and all service times have exponential distributions.

    Their respective rates are and . c is a natural number and specifies the number of

    servers, all identical. How can we model such a queue by means of a CTMC {X(t) :

    37

    http://nl.wikipedia.org/wiki/Agner_Krarup_Erlanghttp://nl.wikipedia.org/wiki/Agner_Krarup_Erlang
  • 7/30/2019 Storm 2012

    39/62

    t 0}? Let X(t) denote the number of customers in the system at time t, including the

    ones in service. Assume that X(0) = 0.

    First, we define the state space I to be {0, 1, } and determine the transition rates.

    An increase of the number of customers is the result of an arrival. Arrivals take place

    by a constant rate of , so

    . .qj,j+1 = for all j I. . .

    A decrease of the number of customers is the result of a service completion. Suppose that

    the system is in a state in which k servers are busy. At what rate service completions

    occur? IfSCj represents the service time of server j, then it is the shortest service timethat forces the state transition, say SC = min{SC1, . . . , S C k}. We know by Theorem

    A.2 already that if k = 2, then SC exp(2). By induction, this statement can be

    easily generalized for general k. Indeed, assume that it holds for k 1, say SC :=

    min{SC1, . . . , S C k1} exp((k 1)), then SC = min{SC,SCk}, so

    . .SC exp((k 1) + ) = exp(k). . .

    In state k there are either k or c servers at work, depending on which of the two numbers

    is the smallest. Hence, we have. .qk,k1 = min{k, c} for all k IN. . .

    This leads to the following transition rate diagram:

    E

    '

    0

    E

    '2

    1

    E

    '3

    2 E

    '(c 1)

    E

    'c

    c 1

    E

    'c

    c

    E

    'c

    c + 1

    How to find an equilibrium probability vector p? Let p be an equilibrium probability

    vector. We are going to describe all coordinates ofp in terms ofpc. Consider the subset

    {0, . . . , c} of I. This subset is visualized by the rectangle in the diagram below. Instead

    of the names of the states, we put the coordinates of p inside the circles:

    38

  • 7/30/2019 Storm 2012

    40/62

    E'

    p0 E'2

    p1 E'3

    p2 E

    '(c 1)

    E'c

    pc1 E'c

    pc E'c

    pc+1

    There is just one arrow pointing from a cell inside the box around set {0, . . . , c} to a cell

    outside the box. In equilibrium it generates an outflow of rate pc qc,c+1. There is just

    one arrow pointing from a cell outside the box to a cell inside the box. In equilibrium it

    generates an inflow of rate pc+1 qc+1,c. In other words, flow out of set {0, . . . , c} = flowinto set {0, . . . , c} gives pc = c pc+1. Hence, pc+1 = pc

    c

    . Let us call this fraction

    ; = c . Now, we consider the set {0, . . . , c + 1}:

    E

    '

    p0

    E

    '2

    p1

    E

    '3

    p2 E

    '(c 1)

    E

    'c

    pc1

    E

    'c

    pc

    E

    'c

    pc+1

    Flow out of set {0, . . . , c+1} = flow into set {0, . . . , c+1} gives pc+1 = c pc+2. Hence,

    pc+2 = pc+1 = pc2. By induction, pc+k = pc

    k for all k IN and

    . .i=c

    pi = pc(1 + + 2 + ) =

    pc1

    if < 1,

    if 1.. .

    This implies that p can only exist if < 1. There is a heuristic argument for this. There

    is a job flow of jobs per hour, each having an expected service length of 1

    hours. So,

    the total number of working hours per hour equals on average. Hence, each individual

    server must work hours per hour on average. This is only possible if 1. Servers

    cannot work all of the time (there are periods in which no customers are in the queue),

    so actually must be strictly smaller than 1. Because of this interpretation of , it is

    called the server utilization.

    39

  • 7/30/2019 Storm 2012

    41/62

    Let us continue with computing p. For i < c, flow out of set {0, . . . , i} = flow into set

    {0, . . . , i} gives pi = (i + 1)pi+1. Hence, pi = pi+1(i+1)

    .

    This describes p in terms of pc. The normalizing equation

    k pk = 1 now fixes p. Let

    us explicitly calculate p in terms of for the case that c = 1 (the M/M/1-queue). We

    have

    p0 = p1/, so p1 = p0 ,

    pk+c = pk+1 = p1k = p0

    k+1,

    Aggregating leads to

    k=0pk = p0

    k=0 k =p01

    . Hence p0 = 1 and pk = k(1 ).

    This comports with the interpretation of being the server utilization; 1 is the

    percentage of the time that the server is free in the long run (or in equilibrium).

    3.3.3 Long Term Behavior

    CTMCs behave similar as DTMCs in the long run. For every CTMC, there is a DTMC

    with almost the same characteristics. They have the same state space, the same minimal

    closed sets, the same regenerative states, and if one of them has a unique equilibrium,

    the other one has a unique equilibrium too. Not the same one however. There is one

    other difference; the phenomenon of periodicity does not occur in the CTMC setting.

    A discrete time indexed process {Xn : n IN0} is called an embedding of a CTMC

    {X(t) : t 0} if there is an increasing sequence of stochastic time moments t0, t1, . . .

    such that Xn = X(tn) for all n IN0. Let, similar to the notation in Section 3.2 about

    Poisson processes, Tk denote the interarrival time between the (k 1)rst and kth state

    transitions of X and let Sk = T1 + + Tk. An important embedding is the one that

    stores the sequence of states that are visited, i.e., X0 := X(0) and Xn = X(Sn) for all

    n IN. It inherits the Markov property and the time homogeneity of the original chain.

    Furthermore, its one step transition probability matrix equals the one jump transition

    probability matrix P of the original process.

    For DTMCs, we have seen theorems that provide sufficient conditions that guarantee

    existence of or convergence to equilibrium. Without a proof, we state that these theorems

    40

  • 7/30/2019 Storm 2012

    42/62

    apply for CTMCs as well. The reason is that the CTMC inherits its long term behavior

    from its embedding. One might wonder why the equilibria of both processes do not

    coincide. The following example might give insight:

    Example 14 Let I = {0, 1, 2} and A =

    200 100 100

    10 20 10

    1 1 2

    . Then P =

    0 1212

    12

    0 12

    12

    12

    0

    .

    Because of the symmetry of P, it is immediate that = (13 ,13 ,

    13) is the equilibrium of

    the embedded DTMC (recall that must satisfy P = ). j gives the long run fraction

    of times that a transition results in a visit of state j. This means that in the long run

    each state is visited equally often. Solving pA = 0 leads to the equilibrium distribution

    p = ( 1111 ,10111 ,

    100111) of the CTMC. The reason of this difference is that in the continuous

    setting the sojourn times in states are taken into account. pj gives the long run fraction

    of time that the system is in state j. Since, 0 = 200 = 10 1 = 100 2, each visit

    of state 0 last on average 10 times shorter than a visit of state 1 and 100 times shorter

    than a visit of state 2. Therefore, the ratios p1 : p2 : p3 relate to13/200 :

    13/20 :

    13/2, i.e.,

    1 : 10 : 100. This explains that p = ( 1111

    , 10111

    , 100111

    ). In general, the way p and relate

    follows from the fact that each visit of state j lasts 1j

    in expectation.

    So, the requirements of having a unique equilibrium distribution are the same as in the

    discrete time setting. The notions accessability, communication, irreducibility, regener-

    ation state, and closedness of a subset ofI can be copied directly from that setting. The

    CTMC has one of the properties above if and only if the embedded DTMC has that

    property. Recall e.g., that a state r is called regenerative if it is visited in finite expected

    time, wherever the chain starts. If such a state exists, there can only be one minimal

    closed set of states and r must be an element of this set (why?).

    If the state space I is finite, having a unique minimal closed set is not only necessary,

    but also sufficient. Each element of it is regenerative.

    If |I| = , more is required. In general, it is very hard to verify whether there is a

    regenerative state. In the case of a queue type of chain, i.e., I = {0, 1, 2, 3, } and

    pij > 0 if and only if |i j| = 1, there must exist a so-called drift to the left, i.e.,

    41

  • 7/30/2019 Storm 2012

    43/62

    pi,i1 > pi,i+1 for all states i in some tail {N, N + 1, N + 2, } of the state space. For

    the M/M/c-queue, this is exactly the case when < 1.

    Theorem 3.14 Let X = {X(t) : t 0} be a CTMC. If the chain has a regenerative

    state, then there exist a unique equilibrium distribution p ofX and a unique equilibrium

    distribution of the embedded chain {Xn : n IN0}, where Xn := X(Sn). They relate

    by . .pj = c jj

    , . .by

    in which c equals the normalizing constant

    kIk/k1

    . Moreover, independent on

    the initial distribution (the distribution of X(0)), the CTMC converges to the equilib-

    rium, i.e., for all i, j I we have

    . .pj = limt

    pij(t). . .

    3.3.4 Cost-Reward Structures

    Similar as in the DTMC setting, cost-reward structures can be defined to model opti-

    mization issues. There, we only defined costs or rewards for states. Here, we follow a

    different approach. Also transitions can lead to revenues or costs. So, a reward structure

    consists of two parts:

    For each state j, a reward rate r(j) denotes the rewards per time unit that are earned

    when the system is in state j. These rewards are called continuous rewards.

    Each transition (j k) can lead to revenues or costs too. The value of such a

    transition is denoted by djk . These rewards are called discrete rewards.

    Let R(t) be the total reward in period [0, t]:

    R(t) =t0

    r(X(u)) du +

    j,k:j=k

    djk

    #{jumps in [0, t] from j to k}

    . (3.18)

    Example 15 A reward structure on an M/M/1-queue

    In an M/M/1-queue, customers pay a starting fee ofe5., pay e0.10 per minute when

    waiting, e10, per hour when being served. Can we find the long run average rewards

    per hour for the server?

    42

  • 7/30/2019 Storm 2012

    44/62

    We can only expect an affirmative answer when the system reaches equilibrium. This is

    the case when = < 1. Set the unit of time to be 1 minute. What is the appropriate

    reward structure? As long as the system stays in state 0, there is no income. We can

    model this by defining r(0) = 0. If there is at least one customer present, then the

    customer being served pays 1060 per time unit, all others pay110 per minute, so define

    r(j) = 1060

    + j110

    for all j 1. Each time somebody joins the queue, 5 is earned. Hence,

    define dj,j+1 = 5. The completion of a service yields no further rewards, so dj,j1 must

    be chosen to be 0.

    The total rewards in the first t minutes equal R(t). The average rewards per minute in

    time segment [0, t] equal R(t)t

    . Therefore the long run average rewards per minute equal

    limt

    R(t)t

    . Theorem 3.15 below states that the long run average rewards per time unit

    equal the expected reward per time unit in equilibrium. In the example we find that the

    long run average rewards per minute equal

    ..

    jI

    r(j)pj +jI

    pjk=j

    qjkdjk =

    j=1

    1060

    +j 1

    10

    (1 )j +

    j=0

    (1 )j 5

    = 5 +

    j=1

    4 + 6j

    60 (1 )j

    = 5 +k=0

    4 + 6(k + 1)

    60 (1 )k+1

    = 5 +k=0

    10 + 6k

    60 (1 )k

    = 5 +

    6(1 )

    k=0

    k +

    10(1 )

    k=0

    kk

    = 5 +

    6+

    2

    10(1 ).

    ..

    The final equality of the elaboration uses standard summations that can be found in theAppendix (Equations (A.12) and (A.13)).

    Theorem 3.15 [Renewal Reward Theorem]

    Let {X(t) : t 0} be a CTMC with unique equilibrium distribution {pj : j I}. Let

    r and d be the continuous and discrete parts of a cost-reward structure. If the number

    of states is finite or if the chain is irreducible, then the long run average rewards per

    43

  • 7/30/2019 Storm 2012

    45/62

    time unit equal4

    . . limt

    R(t)

    t

    = jI

    r(j)pj +jI

    pjk=j

    qjkdjk . . .

    Corollary 3.16 Let N P P(). Then limt

    N(t)

    t= .

    Proof Let {X(t) : t 0} be an M/M/1-queue and N the Poisson process counting

    arrivals of customers. The discrete reward structure that counts these arrivals is given

    by djk = 1 ifk = j + 1 and djk = 0 ifk = j + 1. Then R(t) = N(t) and

    . . limt

    N(t)

    t=

    jI

    pjk=j

    qjkdjk = jI

    pj = .. .2

    3.3.5 The PASTA Theorem and Littles Law

    This final subsection discusses two generally applicable theorems: the Pasta Theorem

    and Littles Law. Both will be illustrated by an example before being postulated.

    Example 12 A single product inventory (revisited)

    Recall that I = {0, 1, . . . , Q} and that the transition rate and probability matrices in

    this example are respectively

    . .A =

    0 0

    0 0

    0 . . .

    ......

    . . . . . . . . . . . ....

    .... . . 0

    0 0

    and P =

    0 0 1

    1 0 0

    0 1 0...

    .... . . . . . . . .

    ......

    . . . . . . . . ....

    0 0 1 0

    . . .

    The matrix P is the one step transition matrix of the embedded DTMC as well. If

    X(0) = X0 = Q, the embedding is the deterministic and repetitive sequence of states

    . .Q, Q 1, Q 2, . . . , , 1, 0, Q , Q 1, . . . . . .

    4Formally, we should state equal with probability 1.

    44

  • 7/30/2019 Storm 2012

    46/62

    So the embedding does not converge to equilibrium. The reason is that its states are

    periodic. Theorem 3.14 tells however that both the original process as the embedding

    do have a unique equilibrium, called p and respectively. It is readily seen that is the

    uniform distribution on I. p can be found by flow rate equations:

    Let x be a solution of yA = 0. Choose x0 = 1 (later on we normalize anyway). Rate

    into state i = rate out of state i says

    x0 = x1, so x1 = ,

    xj = xj+1, so xj+1 = xj = , . .j {1, . . . , Q 1}

    xQ = x0,

    The last equation follows from the previous ones and can serve as a verification. Hence,

    x = (1, , . . . ,). Find an equilibrium distribution by normalizing:

    p =1

    + Q( , , . . . , ). (3.19)

    Two natural questions that arise are

    (i) What is the long run fraction of time that the store is out of stock?

    (ii) What is the long run fraction of customers that will have to be disappointed?

    In order to answer the first question, define a continuous time cost-reward structure r

    by r(0) = 1 and r(j) = 0 for all j {1, . . . , Q}. We have

    . .

    R(t) =

    t0

    r(X(u)) du

    = the total reward in period [0, t]

    = total time with empty stock in period [0, t].

    . .

    The Renewal Reward Theorem 3.15 claims that

    . . limt

    R(t)

    t=jI

    r(j)pj = p0 =

    + Q. . .

    For Question (ii), we define a discrete cost-reward structure, i.e., r(i) = 0 for all j I.

    Furthermore, let dj,j1 = 1 for all j {1, , . . . , Q} and let d0Q = 0. This reward structure

    counts sold items; R(t) equals the total number of items sold in period [0, t]. So, the

    total number of lost demand equals N(t) R(t) and the total fraction of the demand

    45

  • 7/30/2019 Storm 2012

    47/62

    that is lost equals N(t)R(t)N(t) . The Renewal Reward Theorem tells that in the long run

    the average number of items sold per time unit equals

    . .

    limt

    R(t)

    t=jI

    pjk=j

    qjkdjk =Q

    j=1

    pjqj,j1

    =

    Qj=1

    pj =

    Qj=1

    + Q

    =Q

    + Q.

    . .

    The answer is thereby found as follows:

    . . limt

    N(t) R(t)

    N(t)= 1 lim

    t

    R(t)

    t

    t

    N(t)= 1

    Q

    + Q

    1

    =

    + Q. . .

    The answers coincide! This is no coincidence, but due to a general phenomenon, de-

    scribed by the PASTA theorem. . .

    Before we can postulate the theorem, we need more notations. Let N = {N(t) : t 0}

    be a Poisson process. Let X = {X(t) : t 0} be a process with state space I. It

    need not be a CTMC, but we do assume that it converges to equilibrium; there exists a

    distribution p on I with pi = limt

    P[X(t) = i].

    It is assumed that N is an exogenous factor, i.e., the process of arrivals is independent

    on the past of the chain. In formula, for all t > 0 we have

    . .{N(t + t) N(t) : t > 0} {X(u) : u [0, t)}. . .

    The idea of the theorem is that the arrivals of the Poisson process form a sequence

    of time moments at which process X is observed. In the long run, an observer has a

    probability of pi to find the system in state i. Here, i is the state of the system just

    beforethe arrival of the observer might adjust the state. E.g, in a queue system he finds

    a queue of length i and his arrival enlarges the length of the queue to i + 1.

    Theorem 3.17 [The PASTA Theorem]

    Let X = {X(t) : t 0} be a stochastic process converging to equilibrium p and let

    N = {N(t) : t 0} be an exogenous Poisson process. Let Xn be the state of the system

    46

  • 7/30/2019 Storm 2012

    48/62

    X just prior to the nth arrival of the Poisson process. Then for all j I we have

    limn

    1

    n #

    k {1, . . . , n} : Xk = j

    = pj. (3.20)

    Here #A denotes the number of elements of set A. In words, the PASTA-theorem claims

    that the long run fraction of arrivals finding the system in state j equals the long run

    fraction of time that the system is in j. The name of the theorem refers to the

    abbreviation . .Poisson Arrivals See Time Averages. . .abbreviation

    Are there situations in which PASTA does not hold?

    Example 16

    Consider a U/D/1-queue in which each interarrival time last longer than each service

    time. The U in the notation stands for uniformly distributed interarrival times and the

    D expresses deterministic service times. E.g, Tk U(55, 65) (in minutes) and SCk = 30

    minutes for each service completion. In this situation, each new customer finds an empty

    queue, since the server has ample time. On the other hand, the working load of the server

    equals .5, since on average there will enter one client per hour, costing him 30 minutes of

    time. Hence, the long run fraction of time that the server is busy equals 12

    . The reason

    that PASTA does not apply is that the arrivals do not constitute a Poisson process.

    The following exercise is representative for an exam.

    Exercise 2

    An assembly line for a certain product has two stations in series. Each station has only

    room for a single unit of the product. If the assembly of a unit is completed at station

    1, it is forwarded immediately to station 2 provided station 2 is idle; otherwise the unit

    remains in station 1 until station 2 becomes free. Units for assembly arrive at station 1

    according to a Poisson process with rate , but a newly arriving unit is only accepted by

    station 1 when no other unit is present in station 1. Each unit rejected will be handled

    elsewhere. The assembly times at stations 1 and 2 are exponentially distributed with

    respective rates 1 and 2. Formulate a continuous-time Markov chain to analyze the

    47

  • 7/30/2019 Storm 2012

    49/62

    situation at both stations. Specify the state variable(s) and the transition rate diagram.

    Find the long run fraction of units accepted and the long run average time spent in the

    system by an accepted customer. You dont have to calculate the unique equilibrium

    distribution. Just tell why it exists, call it p and give the answers in terms of p. . .

    Elaboration

    Station 1 can be free (F), busy (B) or waiting (W). Station 2 can be free or busy. It is

    recommended not to speak of the state of a station, but e.g., of the condition, situation,

    or whatever, just to avoid confusion. Save the term state for the complete system. The

    state (W, F) does not occur, so we have five states in total:

    . .I = {(F F), (BF), (F B), (BB), (W B)}. . .

    Let {(X1(t), X2(t)) : t 0} describe the state of the system varying over time. At an

    exam, you only have to verify the Markov property, the time homogeneity property, or

    the four characterizing properties a), b), c), and d) of Theorem 3.12 when explicitly

    asked for. Here, the four characterizing properties are straightforward.

    In order to move e.g., from state (W, B) to state (F, B), a service completion at thesecond station is needed. Given that the system is in state (W, B), these service com-

    pletions occur with a rate of 2. Therefore, q(WB),(FB) = 2. The other six positive

    transition rates can be found in the following transition rate diagram:

    . .

    F, F

    B, F

    F, B

    B, B

    W, B

    dd

    dds

    c c

    '

    '

    2

    1 1

    2

    2

    . .

    48

  • 7/30/2019 Storm 2012

    50/62

    In order to prove the existence of a unique equilibrium p to which the system converges

    (regardless the initial state of the system), it suffices to remark that there are finitely

    many states and that the chain is irreducible and then refer to Theorem 3.14.

    By the PASTA theorem, the fraction pacc of the units that is accepted equals the fraction

    of time that the system is in a state in which the first station is free:

    . .pacc := p(FF) +p(FB). . .

    Let {Nacc(t) : t 0} be the counting process that counts the accepted arrivals. This is

    in general not a Poisson process (the rate of accepted arrivals is not constant), but it is

    if the system is in e