Counting level crossings by a stochastic process by Loren Lutes

download Counting level crossings by a stochastic process  by Loren Lutes

of 8

Transcript of Counting level crossings by a stochastic process by Loren Lutes

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    1/8

    Probabilistic Engineering Mechanics 22 (2007) 293300

    www.elsevier.com/locate/probengmech

    Counting level crossings by a stochastic process

    Loren D. Lutes

    Zachry Department of Civil Engineering, Texas A&M University, 101 Summit Edge Court, Glen Rose, TX 76043, USA

    Received 14 September 2006; received in revised form 16 February 2007; accepted 27 February 2007

    Available online 14 March 2007

    Abstract

    The number of times, N, that a stochastic process has up-crossings of a fixed level within a fixed time interval, T, is investigated. Existingintegral formulas for the moments of N for a stationary Gaussian process are shown also to apply to processes that are neither stationary nor

    Gaussian, and explicit formulas are given for approximating the probability distribution of N from the moment formulas. Particular attention is

    given to simplified results for the limiting situations of very small and very large values of T, and to the behavior of variance, skewness, and

    kurtosis of N. For small T, the number N approaches the well-known Poisson distribution, but the results for large T are significantly different.

    For many stationary processes it is shown that the variance of N tends to grow linearly with T when T is very large, but the large-T growth rate

    is sometimes much smaller than that of the small-T Poisson process. More detailed results and some numerical examples are presented for the

    special case of a stationary Gaussian process crossing its own mean value.c 2007 Elsevier Ltd. All rights reserved.

    Keywords: Crossing rates; Multiple crossings; Number of crossings; Stochastic process

    1. Introduction

    Various models for estimating failure probabilities in

    mechanical or structural systems use information regarding the

    crossings of some level by some stochastic process. In addition

    to the obvious situations involving crossings of a critical stress

    or displacement level by the dynamic response of a system,

    there are also situations involving the estimation of the largest

    value of the response or of the occurrence of peaks and valleys

    within a time history. This latter situation corresponds to zero-

    crossings by the first derivative of the dynamic response.

    This is surely not a new problem, dating back at least

    to the work of Rice in 1944 and 1945 [1]. Rices work, in

    particular, gives the mean number of crossings within a given

    time interval, the mean rate of crossings of a given level, and,

    in some instances, the conditional mean rate of crossings at

    one time given the existence of a crossing at another time.

    Various other investigators have made significant extensions

    to this work, but most of these have restricted their attention

    to the special case of stationary Gaussian processes (e.g. [2

    5]). Particularly pertinent to the current work is the derivation

    Tel.: +1 254 898 2631.E-mail address: [email protected].

    by Cramer and Leadbetter [2] of formulas for the moments of

    the number of crossings of a fixed level within a fixed timeinterval for this special case. Rigorous mathematical analyses

    of various aspects of the crossing problem for stationary

    Gaussian processes are available in the books by Cramer and

    Leadbetter [6] and Leadbetter, Lindgren, and Rootzen [7].

    Some of this work for stationary Gaussian processes can also

    be applied to a category of stationary non-Gaussian processes.

    In particular, crossings by a translation process [8], which can

    be written as a monotonic nonlinear function of a Gaussian

    process, are directly related to crossings by the underlying

    Gaussian process.

    One of the emphases of the current work is the development

    of integral formulas relating to the crossings of a process

    that is neither stationary nor Gaussian, before demonstrating

    how these results simplify as one imposes conditions, first

    of stationarity and finally of Gaussianity. The approach to

    the nonstationary non-Gaussian development is more heuristic

    than rigorous. For example, there is no consideration given to

    measure theoretic demonstrations that uncertain quantities of

    interest are actually random variables.

    Another aspect of the current work is that particular attention

    is given to counting crossings during either very short or very

    0266-8920/$ - see front matter c 2007 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.probengmech.2007.02.003

    http://www.elsevier.com/locate/probengmechmailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.probengmech.2007.02.003http://dx.doi.org/10.1016/j.probengmech.2007.02.003mailto:[email protected]://www.elsevier.com/locate/probengmech
  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    2/8

    294 L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 293300

    long time intervals. Simplified formulas are derived for these

    limiting situations.

    The quantities studied here include the variance, skewness,

    and kurtosis of the number of crossings during a given time

    interval, and procedures are also given for approximating the

    probability distribution of that number. The fact that the dual

    crossing rate governing two crossings in an interval can beevaluated in closed form for a stationary Gaussian process

    allows simple computation of the variance of the number of

    crossings and the probability of two crossings within a small

    time interval for this situation. Numerical results are presented

    for selected examples.

    Although the current work is restricted to the counting of up-

    crossings of a given level, one can easily generalize the problem

    presented here to investigate the up-crossing of multiple levels

    or to consider down-crossings at some or all of the times of

    interest.

    2. Problem formulation

    Let {X(t)} be a stochastic process with continuous time

    histories, and let the random variable N be the number of

    up-crossings of a fixed level b within a time interval [0, T].

    Note that {X(t)} is not required to be either stationary or

    Gaussian. One simple way to reach the number N in the limit is

    to divide the time interval into n subintervals by introducing

    points tn j = jn with n = T/n. Let Nn be the number

    of subintervals containing exactly one up-crossing of b. The

    {X(t)} process is now limited to be sufficiently smooth that the

    probability of more than one crossing within any subinterval

    tends to zero more rapidly than the probability of one crossing,

    so that one can consider each subinterval to contain either 0 or

    1 up-crossing for n sufficiently large. Then N is the limit of Nnas n tends to infinity. The indicator function for the event of one

    up-crossing in the j th subinterval, for a given n value, will be

    denoted by In j . That is, In j = 1 if there is an up-crossing in

    subinterval j , and is zero otherwise. One can then say that

    Nn =

    nj =1

    Inj . (1)

    The focus here will be on calculating the moments of the

    random variable N by taking the limits of the moments of the

    random variable Nn . In particular,

    E(Nkn ) =n

    j1=1

    n

    jk=1

    E(In j1In j2 In jk)

    =

    nj1=1

    njk=1

    P(In j1 = 1, . . . , In jk = 1) (2)

    will tend to E(Nk) for n . Presume, now, that there exist

    finite crossing rates of the form used by Rice [ 1]:

    +k (t1, t2, . . . , tk) =

    0

    0

    0

    y1y2 yk

    pX(b, y1, b, y2, . . . b, yk)

    dy1dy2 dyk (3)

    in which X = [X(t1), X(t1), X(t2), X(t2), . . . , X(tk), X(tk)]T.

    Rices result for k = 1 is well known, and he also gave

    some explicit consideration to the dual-crossing rate of(3) with

    k = 2. One can now write the probabilities that are needed in

    (2) using the crossing rate in (3), giving

    P(In j1 = 1, . . . , In jk = 1)

    =

    tnj k+ntnj k

    tnj1 +ntnj 1

    +k (t1, . . . , tk)dt1 dtk

    = +k (tn j1 , . . . , tn jk)kn + O(

    k+1n ) (4)

    provided that the (j1, . . . , jk) arguments in (4) are distinct. This

    restriction of distinct time intervals is obviously required, since

    the order of magnitude in (4) is not correct if some of the

    subintervals are identical. For example, in the extreme situation

    in which all the tn j are identical, one finds that the probability

    on the left-hand side is O(n) rather than O(kn). Also, it may

    be noted that the pX(b, y1, b, y2, . . . b, yk) probability densityfunction in (3) is singular along any surface in (t1, t2, . . . , tk)

    space having any two or more of these arguments equal to

    each other, which gives singularities on these surfaces for the

    multiple crossing rate of(3).

    The moments ofNn can now be evaluated from (2). Note that

    this summation generally includes situations with identical tn jvalues, as well as those with distinct subintervals. Nonetheless,

    (4) does give all the information necessary to compute the

    moments when one counts the ways in which some tn j values

    may be identical and uses the fact that Iln j = In j for any positive

    value ofl. The results will be written out explicitly for the first

    four moments:

    E(Nn) =

    j

    E(In j ) = Mn1 +

    j

    O(2n)

    = Mn1 + O(n1) (5)

    E(N2n ) =

    j

    E(I2n j ) +

    distinct (j1,j2)

    E(In j1In j2 )

    =

    j

    E(In j ) +

    distinct (j1,j2)

    (+2 (tn j1 , tn j2 )

    2n

    + O(3n)) = Mn1 + Mn2 + O(n1) (6)

    E(N3n ) = iE(I3ni ) + 3 distinct (j1,j2)

    E(Ini1I2ni2

    )

    +

    distinct (j1,j2,j3)

    E(In j1In j2In j3 )

    = Mn1 + 3Mn2 + Mn3 + O(n1) (7)

    E(N4n ) =

    i

    E(I4ni ) + 3

    distinct (i1,i2)

    E(I2ni1I

    2ni2

    )

    + 4

    distinct (i1,i2)

    E(Ini1I

    3ni2

    )

    + 6

    distinct (i1,i2,i3)

    E(Ini1Ini2I

    2ni3

    )

    +

    distinct (j1,j2,j3,j4)

    E(In j1In j2In j3In j4 )

    = Mn1 + 7Mn2 + 6Mn3 + Mn4 + O(n1

    ) (8)

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    3/8

    L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 29 3300 295

    kurtosis(N) =E[(N M1)

    4]

    4N

    =M4 + M3(6 4M1) + M2(7 12M1 + 6M

    21 ) + M1 4M

    21 + 6M

    31 3M

    41

    (M2 + M1 M21 )

    2.

    Box I.

    in which

    Mnk = kn

    distinct (i1,...,ik)

    +nk(tni1 , . . . , tnik). (9)

    Now one can take the limit as n , giving Mnk tending to

    the integral

    Mk

    T0

    T0

    +k (t1, . . . , tk)dt1 dtk (10)

    in which +k (t1, . . . , tk) is the continuous function that is equal

    to the multiple crossing rate +k (t1, . . . , tk) whenever all the

    time arguments are distinct. Since N is the limit of Nn , the

    moments ofN are then given by (5)(8) with each Mnk replaced

    by Mk.The terms Mk in (10) are certainly not new. They were used

    by Cramer and Leadbetter [6] in the context of a stationary

    Gaussian process {X(t)}. For this special case, Cramer and

    Leadbetter gave a rigorous proof that

    Mk = E[N(N 1) (N k + 1)] (11)

    and called this a factorial moment of the number of crossings.

    It is easily shown that the moments obtained as the limits of

    (5)(9) with n are consistent with the factorial moments

    of (11). The intent of the presentation here is not to introduce

    any new idea, but to demonstrate that these factorial moments

    and the usual E(Nk) moments can be applied to more generalprocesses that are nonstationary and/or non-Gaussian.

    The variance of N is, of course, easily obtained by the usual

    method from the limits of (5) and (6):

    2N = M2 + M1 M21 . (12)

    Similarly, the skewness is

    skewness(N) =E[(N M1)

    3]

    3N

    =M3 + 3M2(1 M1) + M1 3M

    21 + 2M

    31

    (M2 + M1 M

    2

    1 )

    3/2(13)

    and the kurtosis is given by the equation given in Box I.

    One may note that E(N) = M1 is of O(T) when T is

    very small and the expressions for 2N, E[(N M1)3], and

    E[(N M1)4] each contain a term M1 that will dominate in

    this situation, since all the other terms are either powers of M1or contain an Mk representing an integral over a higher-order

    cube of dimension T. This behavior is the same as that of the

    well-known Poisson counting process, for which 2N = E(N),

    E[(N M1)3] = E(N), and E[(N M1)

    4] = 3[E(N)]2 +

    E(N) for all values of time.

    One can also use the limits of the moment expressions in

    (6)(8) to learn more about the singularities in the crossing

    rates, resulting from the singularity of the probability density

    function in (4) when times converge. In particular, one can writea moment equation similar to (11) as

    E(Nk) =

    T0

    T0

    +k (t1, . . . , tk)dt1 dtk (14)

    and the moment expressions in (6)(8) are consistent with this

    if +1 (t) = +1 (t) and the higher-order crossing rates have

    singularities as follows:

    +2 (t1, t2) = +2 (t1, t2) +

    +1 (t1) (t1 t2) (15)

    +3 (t1, t2, t3) = +3 (t1, t2, t3)

    + +

    2

    (t1, t2) (t2 t3) + +

    2

    (t2, t3) (t1 t3)

    + +2 (t1, t3) (t1 t2)

    + +1 (t1) (t1 t2) (t1 t3) (16)

    and

    +4 (t1, t2, t3, t4) = +4 (t1, t2, t3, t4) +

    +3 (t1, t3, t4) (t1 t2)

    + +3 (t1, t2, t4) (t1 t3)

    + +3 (t1, t2, t3) (t1 t4)

    + +3 (t1, t2, t4) (t2 t3)

    + +3 (t1, t2, t3) (t2 t4)

    + +3 (t1, t2, t3) (t3 t4)

    + +2 (t1, t3) (t1 t2) (t3 t4)

    + +2 (t1, t2) (t1 t3) (t2 t4)

    + +2 (t1, t2) (t1 t4) (t2 t3)

    + +2 (t1, t2) (t1 t3) (t1 t4)

    + +2 (t1, t3) (t1 t2) (t1 t4)

    + +2 (t1, t4) (t1 t2) (t1 t3)

    + +2 (t1, t2) (t2 t3) (t2 t4)

    + +1 (t1) (t1 t2) (t1 t3)(t1 t4) (17)

    in which () denotes the Dirac delta function. Of course, the

    integrals in (14) with these integrands containing Dirac deltafunctions are rigorously defined only in terms of generalized

    functions, but the concept is well known.

    The nature of the singularities can be illustrated by

    considering +4 . Within the four-dimensional space spanned

    by (t1, t2, t3, t4), there is a first-order singularity involving +3

    whenever two of the times are equal and the other two are

    different, a second-order singularity involving +2 whenever the

    times are either two equal pairs or three the same and one

    different, and a third-order singularity involving +1 along the

    line with all four times being the same. Clearly, such an explicit

    listing of the singularities would become tedious for larger

    values ofk in +k . Nonetheless, the pattern is clear.

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    4/8

    296 L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 293300

    In some situations it is convenient to make use of conditional

    rates of crossing given the existence of crossings at other

    specified times. Using the finite subintervals introduced earlier,

    one can define a conditional crossing rate as

    +k (tn ji +1 , . . . , tn jk|upcr. at tn j1 , . . . , tn ji )

    limn

    1

    kin

    P(In j1

    = 1, . . . , In jk

    = 1)

    P(In j1 = 1, . . . , In ji = 1)

    =+k (tn ji +1 , . . . , tn jk)

    +k (tnj i +1 , . . . , tn ji ). (18)

    In particular, +1 (t1|upcr. at t2) = +2 (t1, t2)/

    +1 (t2) is an

    important special case that will be used here. This formula was

    also used by Rice [1], and is easily shown to be consistent

    with the Slepian model that applies for the special case of a

    stationary Gaussian process (e.g. [7]).

    One more type of information that can be obtained from the

    moments relates to the discrete probability distribution of N. In

    particular, let pj P(N = j ). The kth factorial moment from(11) is then

    Mk =

    j=0

    (j [j 1] [j k+ 1])pj =

    j=k

    j !

    (j k)!pj (19)

    in which the first k terms in the summation have been dropped

    in the final form because they are identically zero. The formula

    in (19) is most useful when pj pj +1 for all j 0, as

    would usually be the case when T is very small. In this situation

    one can say that Mk = (k!)pk + O(pk+1) or pk Mk/(k!).

    For T values that are not very small, one anticipates that the

    largest pk value will occur when k is in the neighborhood of

    M1 = +1 T, and that pk will be very small for k M1. Based

    on this, one may choose to ignore the contribution to (19) of

    all pk terms having k > , with M1, by truncating the

    summation at j = . This finite set of simultaneous equations

    with a triangular coefficient matrix is easily solved to give

    pj 1

    j !

    ji =0

    (1)iMj +i

    i !. (20)

    The generalization of the problem presented here to consider

    the up-crossing of multiple levels (b1, b2, . . . , bk) at times

    (t1, t2, . . . , tk) or to consider down-crossings at some or all of

    the times is quite straightforward. The basic modifications are

    in the crossing rates of (3). It should be noted, though, that

    the singularities of the joint probability density function in (3)

    and of the crossing rates, as illustrated in (15)(17), occur only

    when at least one crossing event occurs more than once. That

    is, when there are two or more up-crossings of the same level

    and/or two or more down-crossings of the same level.

    3. Limiting behavior for large time

    Attention will now be limited to stochastic processes for

    which the process and its derivatives at time t1 become

    independent of these same quantities at time t2 when the

    separation between t1 and t2 becomes large. For the special

    case of a Gaussian process, of course, this occurs if the

    autocovariance function of {X(t)} tends to zero for well-

    separated time arguments, but the same behavior also occurs for

    more general processes, including many of practical interest.

    The variance behavior of N for large values of time in this

    situation is better exhibited by using an alternate form given

    by Cramer and Leadbetter [6] for the dual crossing rate. Inparticular, let

    2(t1, t2) = +2 (t1, t2)

    +1 (t1)

    +1 (t2) (21)

    so that (10) and (12) give

    M2 = M21 + H2,

    2N = M1 + H2 (22)

    with H2 defined as

    H2

    T0

    T0

    2(t1, t2)dt1dt2. (23)

    The imposed independence condition implies that 2(t1, t2)

    tends to zero when |t2 t1| is large, and weak restrictions onthe decay of the function imply that H2 is O(T), rather than

    O(T2). In particular, the contribution of2(t1, t2) is significant

    only in the vicinity of the line t1 = t2. Thus, the variance from

    (22) only grows as O(T).

    Similarly, let

    3(t1, t2, t3) = +3 (t1, t2, t3)

    +1 (t1)

    +2 (t2, t3)

    +1 (t2)

    +2 (t1, t3)

    +1 (t3)+2 (t1, t2) + 2

    +1 (t1)

    +1 (t2)

    +1 (t3) (24)

    and

    H3

    T

    0

    T

    0

    T

    03(t1, t2, t3)dt1dt2 dt3 (25)

    so that (10) gives

    M3 = H3 + 3M1M2 2M31 . (26)

    One can divide the large-time domain of the arguments in (25)

    into 5 subsets:

    a: t1, t2, and t3 all well separated,b: t2 and t3 not well separated, but t1 well separated from them,c: t1 and t3 not well separated, but t2well separated from them,d: t1 and t2 not well separated, but t3 well separated from them,

    ande: no separationthe vicinity of the line t1 = t2 = t3.

    Using the independence argument employed with regard to

    (21), one can show that 3(t1, t2, t3) tends to zero in all of these

    subsets except e, so that weak restrictions give H3 as O(T) for

    T .

    One can now use (22) and (26) to rewrite the skewness

    expression. In particular, the numerator in (13) becomes

    E(N M1)3 = H3 + 3H2 + M1 (27)

    which only grows as O(T) for T for appropriate +kfunctions. From (27) and (22)one can see that the skewness in

    (13) tends to zero as O(T

    1/

    2) in this situation.

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    5/8

    L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 29 3300 297

    In the same way, one can show that the function

    4(t1, t2, t3, t4) = +4 (t1, t2, t3, t4)

    +1 (t1)

    +3 (t2, t3, t4)

    +1 (t2)

    +3 (t1, t3, t4)

    +1 (t3)

    +3 (t1, t2, t4)

    +1 (t4)

    +3 (t1, t2, t3)

    +2 (t1, t2)

    +2 (t3, t4)

    +2 (t1, t3)

    +2 (t2, t4)

    +2 (t1, t4)

    +2 (t2, t3)

    + 2+1 (t1)+1 (t2)

    +2 (t3, t4)

    + 2+1 (t1)+1 (t3)

    +2 (t2, t4)

    + 2+1 (t1)+1 (t4)

    +2 (t2, t3)

    + 2+1 (t2)+1 (t3)

    +2 (t1, t4)

    + 2+1 (t2)+1 (t4)

    +2 (t1, t3)

    + 2+1 (t3)+1 (t4)

    +2 (t1, t2)

    6+1 (t1)+1 (t2)

    +1 (t3)

    +1 (t4) (28)

    tends to zero for large-time arguments for all situations except

    the vicinity of the line t1 = t2 = t3 = t4, so that

    H4

    T0

    T0

    T0

    T0

    4(t1, t2, t3, t4)dt1dt2dt3dt4 (29)

    can become O(T) for T . From (2) one now has

    M4 = H4 + 4M1M3 + 3M22 12M

    21M2 + 6M

    41 (30)

    which can be used along with (22) and (26) to rewrite theexpression for kurtosis. The numerator of Box I gives

    E(N M1)4 = 3(M1 + H2)

    2 + H4 + 6H3 + 7H2 + M1

    = 34N + H4 + 6H3 + 7H2 + M1. (31)

    The fact that the denominator in the kurtosis grows as 4N =

    O(T2) then gives the kurtosis as tending to 3 + O(T1) if theHk terms all grow as O(T) for T . The skewness and

    kurtosis results, of course, are consistent with N tending to a

    Gaussian distribution as T tends to infinity. This can also bejustified by the central limit theorem under weak restrictions

    on the {X(t)} process. In particular, it is sufficient to have

    conditions that the autocovariance function of {X(t)} tendsto zero for well-separated time arguments plus a Lyapunov

    condition that the nonstationarity of {X(t)} is not such that the

    magnitude of N for T tending to infinity is dominated by thecontributions during a finite number of finite time intervals [9].

    4. Simplification for a stationary process

    For a stationary process some of the results can be simplified

    somewhat by rewriting +1 (t) as a time-invariant crossing rate

    +1 and

    +k (t1, . . . , tk) as a function

    +k (1, . . . , k1) ofk 1

    arguments i = tk ti . Similarly, k(t1, . . . , tk) is rewritten

    as k(1, . . . , k1). This allows various multiple integrationexpressions to be reduced by one order. For example, M1 =+1 T, and (10) and (23) become

    M2 = 2

    T0

    (T ) +2 () d,

    H2=

    2T

    0 (T ) 2() d

    (32)

    respectively. Presuming that 2() tends to zero faster than

    2 for large values of , one can then use the approximation

    H2 H20T H21 when T is very large, in which

    H20 = 2

    0

    2()d, H21 = 2

    0

    2()d. (33)

    The formulas in (22) then give

    M2 (+1 T)

    2 + H20T H21,

    2N (+1 + H20)T H21

    (34)

    for large values ofT. Note that (34) gives an explicit expression

    for the limiting linear growth of the variance. The effect of

    the H20 term will be found to be quite significant in some

    situations, so that the rate of growth of the variance may be

    much smaller than for the corresponding Poisson process, for

    which H2 = 0.

    One can generalize (32) to show that

    T0

    T

    0f(t1, . . . , tk)dt1 dtk

    = k

    T0

    T0

    [T max(1, . . . , k1)] f(1, . . . , k1)

    d1 dk1 (35)

    for any function f that has the properties exhibited by +k or kfor a stationary process. In particular, the f(t1, . . . , tk) function

    is unchanged by any permutation of its k arguments and is a

    function only of the k 1 time differences.

    Ifk tends to zero faster than 2 for large values of any of

    the arguments, then

    Hk0 k

    0

    0k(1, . . . , k1) d1 dk1 (36)

    and

    Hk1 k

    0

    0

    max(1, . . . , k1)k(1, . . . , k1)

    d1 dk1 (37)

    are finite and Hk Hk0T Hk1 as T tends to infinity. Using

    (34), (27) and (31)now gives

    skewness(N) (+1 + H30 + 3H20)T H31 3H21

    [(+

    1+ H

    20)T H

    21]3/2

    (+1 + H30 + 3H20)

    (+1 + H20)3/2T1/2

    (38)

    and

    kurtosis(N) 3

    +(+1 + H40 + 6H30 + 7H20)T H41 6H31 7H21

    [(+1 + H20) T H21]2

    3 ++1 + H40 + 6H30 + 7H20

    (+1 + H20)2T

    (39)

    for this large-time situation.

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    6/8

    298 L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 293300

    5. Stationary Gaussian process with known autocorrelation

    function

    Some of the general expressions in the preceding sections

    can be written explicitly in relatively simple forms in the special

    case of a Gaussian {X(t)} process, and some of the integrals in

    the general expressions can be evaluated in closed form in the

    even more special case in which {X(t)} is also stationary and

    the events of interest are crossings of its mean value. Note that

    none of the mathematical results for this stationary Gaussian

    case are fundamentally new, since it is the special case that

    has received considerable study in the past (e.g. [27]). It is

    introduced here, not so much because it is extensively used

    in practical applications, as to provide a simple illustration of

    the general relationships developed in this study. Formulas here

    will be written for the common situation of zero-crossings of a

    mean-zero Gaussian process, but they also apply to any other

    situation with b = X.

    The mean up-crossing rate of zero for a mean-zero stationary

    Gaussian {X(t)} process is well known to be +1 = X/(2 X)

    and the joint probability density functions needed for evaluation

    of higher-order crossing rates from (4) are also well known. For

    the dual-crossing rate the integration can be performed in closed

    form and the result written as

    +2 () =2X

    2X(2 )2(1 20 )

    1/2

    A B (A2 B2) tan1

    B

    A

    (40)

    in which

    A =

    1 2 21

    1 + 0

    1/2,

    B =

    1 + 2

    21

    1 0

    1/2 (41)

    with

    0 =cov[X(t1), X(t2)]

    2X

    , 1 =cov[X(t1), X(t2)]

    XX,

    2 =cov[ X(t1), X(t2)]

    2X

    .

    (42)

    Note that 0, 1, and 2 are all functions of t2 t1,

    even though that has not been included explicitly in the

    equations. If one knows the autocovariance function G()

    cov[X(t), X(t+ )] for the process {X(t)}, then the correlation

    coefficients can be written in terms of that function and its first

    two derivatives.

    The expression in (40) confirms the anticipated result that

    +2 () tends to (+1 )

    2 for very large , so that 2() tends to

    zero. In particular, all the correlation coefficients go to zero

    for large , so that A and B both tend to unity. This makes

    the first term in (40) to be unity and the second one to be

    zero, confirming the desired result. No closed-form solution has

    been found for the integrals of 2() in (32). Thus, numerical

    integration is required for evaluating M2, which appears in the

    variance of N, even for this simplified Gaussian problem. The

    integration is simple, though, since it is one-dimensional. Also,

    no closed-form solution has been found for the integrals in (3)

    to evaluate +k () for k 3.

    In order to obtain more detailed information about the dual

    crossing rate and associated properties of the counting processN for small values of time, one can consider a power-series

    expansion for the autocovariance function G():

    G() =

    j=0

    cj ||j . (43)

    Obviously c0 = 2

    X and it is also necessary for the present

    situation to have c1 = 0. In particular, it is necessary that G()

    be twice differentiable in order that finite crossings rates exist.

    This gives 2X= 2c2 and

    +1 () = (2c2/c0)

    1/2/(2 ).

    Three distinct conditions must be considered in evaluating

    the dual crossing rate related to the autocovariance function in

    (43). With no further restriction other than c1 = 0, the formulasin (40) and (41) give the limiting dual crossing rate for small

    as

    +2 (0) =c

    1/20 c3(3

    3/2 )

    6 2(2c2)1/2. (44)

    For small values ofT one can substitute (44) into (32) to obtain

    M2 +2 (0) T

    2. Clearly, this term coming from +2 () makes

    a very small contribution to the variance of N, as given in

    (12), when T is small. Also the probability p2 M2/2 of two

    crossings within a small time interval is very small.

    In some applications, one may choose to use a model with c3

    also equal to zero. In particular, this is a necessary condition forany process with a finite value ofX. For c3 = 0, it is obvious

    from (44) that +2 (0) = 0, which implies that M2 tends to a

    higher-order function ofT. In fact, the limiting result from (40)

    and (41) is

    +2 () 4c0(c5)

    3/25/2

    3 2(c2)1/2(6c0c4 c22)

    1/2. (45)

    From (32) one then finds that M2 = O( T9/2), so that p2 and

    the contribution of M2 to the variance of N are even smaller

    than in the situation represented by (44). One can carry this idea

    one step further by considering c5 also to be zero, as is required

    for a process with a finite value of...X. This condition gives thestill smaller terms of

    +2 () 21/2(5c2c6 2c

    24)

    3/24

    3 2c22(6c4 c22)

    1/2,

    M2 21/2(5c2c6 2c

    24)

    3/2T6

    45 2c22(6c4 c22)

    1/2. (46)

    Note that the limits in (46) depend only on even coefficients

    in the expansion of (43). Thus, the limiting forms of +2 ()

    and M2 in (46) will not be changed by setting more of

    the odd coefficients in (43) to zero. In particular, a process

    with an analytic autocovariance function will have only even

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    7/8

    L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 29 3300 299

    Fig. 1. Covariance function for {X(t)}.

    coefficients in the expansion of(43), but the limiting small-time

    behavior of+2 () and M2 will be as given in (46).

    6. Numerical examples based on dual occurrence rate

    Let {X(t)} be mean-zero and have a spectral density

    SX X() =1

    2 (2 )1/2 (e(+0)

    2

    /(2

    2

    ) + e(0)2

    /(2

    2

    )) (47)

    and autocovariance function

    G() = e22/2 cos(0). (48)

    Note that this has been normalized to give unit variance, so

    the 0 correlation coefficient in (42) is identical to G().

    Differentiation gives 1 and 2, and +1 and

    +2 () are easily

    evaluated. Since G() is analytic, only even coefficients appear

    in the expansion of(43), and those needed for use in the small-

    time expressions of(46) can be evaluated as

    c2 = 2 + 2

    02

    , c4 =34 + 622

    0+ 4

    024

    ,

    c6 = 156 + 45420 + 15

    240 + 60

    720.

    (49)

    The small-time dual crossing rate then has a limit from (46) of

    +2 () 2(36 + 12420 + 9

    240 + 2 60)

    3/24

    648 2(2 + 20)2(2 + 2 20)

    1/2(50)

    and M2 for small values ofT is similarly obtained.

    Choosing 0 = 0 in (47) gives a simple bell-shaped

    unimodal spectral density and 0 = 5 gives a moderately

    narrowband process. For both of the processes, the values

    of H20 and H21 from (33) are negative. The autocovariancefunctions for the two processes are shown in Fig. 1 and values of

    +2 ()/(+1 )

    2 are then shown in Fig. 2. Note that the normalized

    form used for presenting +2 () is convenient in multiple ways.

    Not only is it dimensionless, but also it can be viewed as a

    normalized form of the conditional rate of up-crossings. In

    particular, (18) gives +1 (t + |upcr. at t)/+1 =

    +2 ()/(

    +1 )

    2.

    Finally, the normalized form approaches unity for large values

    of . Fig. 2 also confirms that +2 () agrees with the O(4)

    behavior of(50) for small values of . The variance of N(T)

    for the two sample situations is shown in Fig. 3. The plot also

    gives the small-T Poisson approximation of+0 T and the large-

    T approximation from (34). Numerical values for M2 are also

    Fig. 2. Dual crossing rate.

    Fig. 3. Variance ofN.

    easily obtained, but are not shown here. For each of the spectral

    densities, the M2 versus T curve tends smoothly from the

    O(T6) limit of(46) for small T to the large-T parabola of(34).

    The transitions of the numerical results in Figs. 2 and 3

    between the small-time and the large-time limits are seento be smooth. It is noted, though, that the transitions are

    quite different in form. In particular, the transitions for 0 =

    0 are monotonic, while those for 0 = 5 approach thelarge-T asymptotes in an oscillatory manner, which loosely

    correlates with the difference between the two autocovariance

    functions. Note, in particular, that 0 = 5 sometimes gives

    +2 () as significantly exceeding (

    +1 )

    2, most notably when approximates one period of the narrowband process. Viewing

    this in terms of the conditional rate of occurrence of crossings

    at time t + , this indicates that the presence of a peak at time

    t significantly increases the likelihood of a crossing one periodlater. This is surely not unexpected for a narrowband process.

    It may also be noted that the large-time rate of increase of

    the variance of N(T) in Fig. 3 is significantly smaller than the

    small-time value for both situations shown, but particularly for0 = 5. For 0 = 0, H20 in (34) is 71% of

    +1 , which

    results in the final rate of growth of the variance being 29% ofthe initial rate, while 0 = 5 gives H20 as 93% of

    +1 so

    that the rate of increase of the variance for large time is only

    about 7% of the initial rate. These decreases in the variance of

    the number of crossings, of course, mean that N(T, 0) has less

    variability than the corresponding Poisson process, in the sense

    that its coefficient of variation is less than (+0 T)1/2.

    7. Summary

    An investigation has been made of the number of times,

    N, that a stochastic process has up-crossings of a fixed level

  • 7/28/2019 Counting level crossings by a stochastic process by Loren Lutes

    8/8

    300 L.D. Lutes / Probabilistic Engineering Mechanics 22 (2007) 293300

    within a fixed time interval, T. Existing integral formulas

    for the moments of N for a stationary Gaussian process

    have been shown also to apply to processes that are neither

    stationary nor Gaussian, and explicit formulas have been given

    for approximating the probability distribution of N from the

    moment formulas. The singularities of the multiple crossing

    rates for equal time arguments have been investigated in somedetail for two to four crossings. Particular attention has been

    given to the limiting situations of very small and very large

    values of T, for which simplified results apply. For small T,

    the number N approaches the well-known Poisson distribution,

    but the results for large T are significantly different. For many

    stationary processes it is shown that the variance ofN does tend

    to grow linearly with T when T is very large, but the large-T

    growth rate is sometimes much smaller than that of the small- T

    Poisson process. The behavior of the skewness and kurtosis of

    N has also been investigated, confirming the convergence of N

    to a Gaussian distribution for large values of T.

    In the special case of a stationary Gaussian process crossing

    its own mean value, as in zero-crossings by a mean-zeroprocess, the general integral expressions become much simpler

    and some of them can be evaluated in closed form. In particular,

    the dual crossing rate can be obtained in closed form, and

    this allows investigation of such quantities as the variance of

    N and the probability of two crossings in a very small time

    interval. Limited numerical results have been presented for this

    situation to illustrate the general relationships developed in this

    work.

    References

    [1] Rice SO. Mathematical analysis of random noise. In: Wax N, editor.

    Selected papers on noise and stochastic processes. New York: Dover; 1954.

    Reprinted from: Bell System Technical Journal 19441945: 2324.

    [2] Cramer H, Leadbetter MR. The moments of the number of crossings of a

    level by a stationary normal process. The Annals of Mathematical Statistics

    1965;30(6):165663.

    [3] Geman D. On the variance of the number of zeros of a stationary Gaussian

    process. The Annals of Mathematical Statistics 1972;43(3):97782.

    [4] Cuzick J. Conditions for finite moments of the number of zero crossings

    for Gaussian processes. The Annals of Probability 1975;3(5):84958.

    [5] Cuzick J. A central limit theorem for the number of zeros of a stationary

    Gaussian process. The Annals of Probability 1976;4(4):54756.

    [6] Cramer H, Leadbetter MR. Stationary and related stochastic processes:

    Sample function properties and their applications. Mineola (NY): Dover;

    2004. Reprinted from New York: Wiley; 1967.

    [7] Leadbetter MR, Lindgren G, Rootzen H. Extremes and related properties

    of random sequences and processes. New York: Springer-Verlag; 1983.

    [8] Grigoriu M. Applied non-Gaussian processes. Englewood Cliffs (NJ):

    Prentice Hall; 1995.

    [9] Lin YK. Probabilistic theory of structural dynamics. Melbourne (FL):

    Krieger Publishing; 1976. Reprinted from: New York: McGraw-Hill; 1967.