3 - Basics of Random Processes and Structural Dynamics
3 - Basics of Random Processes and Structural Dynamics
3
A
If the ensemble consists of N sample functions, out of which there are n realizations
with X(t) < x, then the first-order probability distribution function defined in equation
57
58
Ensemble averaging
X (3)
X (2)
X (1)
Time averaging
t1 t2
Figure 3.1 Sample functions X (K)(t) of a random process.
If n12 is the number of sample functions with X(t1 ) < x1 and X(t2 ) < x2 , for large N
n12
F2 (x1 , t1 ; x2 , t2 ) ≈ .
N
Similarly, the probability distribution function of order n is defined as
Fn (x1 , t1 ; x2 , t2 ; . . . ; xn , tn ) = P X(t1 ) < x1 , X(t2 ) < x2 , . . . , X(tn ) < xn . (3.1.3)
∂ n Fn (x1 , t1 ; x2 , t2 ; . . . ; xn , tn )
pn (x1 , t1 ; x2 , t2 ; . . . ; xn , tn ) = , (3.1.4)
∂x1 ∂x2 · · · ∂xn
i.e.,
pn (x1 , t1 ; x2 , t2 ; . . . ; xn , tn )dx1 dx2 · · · dxn
3.1 random processes 59
= P x1 < X(t1 ) < x1 +dx1 ; x2 < X(t2 ) < x2 +dx2 ; . . . ; xn < X(tn ) < xn +dxn .
for all orders n and any value of τ , the random process is said to be stationary. This
implies that the first-order probability distribution is independent of time, and the
second-order probability distribution depends only on the time difference, i.e.
F1 (x, t) = F1 (x);
(3.1.6)
F2 (x1 , t1 ; x2 , t2 ) = F2 (x1 , x2 ; t2 −t1 ) = F2 (x1 , x2 ; τ ), τ = t2 −t1 .
A random process can be expected to be stationary when the physical factors influenc-
ing it do not change with time. For example, the wind pressure on a building will be
stationary when the wind flow is steady, whereas ground accelerations of an earthquake
are nonstationary random processes.
1 (K)
N
E[ X(t) ] = lim X (t). (3.1.7)
N→∞ N
K=1
Both of these averages can also be defined for any functions of a random process.
The simplest of such functions are the polynomials X K1(t1 )X K2(t2 ) · · · X Kn(tn ), whose
averages are called the moments.
The most important of the moments obtained from the first-order probability dis-
tribution are the mean (or expected value) and the mean-square value defined by
+∞
μX (t) = E[ X(t) ] = x p1 (x, t)dx,
−∞
+∞
(3.1.9)
2
Xrms (t) = E[ X (t) ] =
2 2
x p1 (x, t)dx,
−∞
where Xrms is the root-mean-square value of X(t). The variance of X(t) is defined by
2
σX2 (t) = E X(t) − μ(t) = E[ X 2 (t) ] − μ2X (t), (3.1.10)
which is the expectation of the square of the deviation from the mean. The positive
square root of the variance σX (t) is called the standard deviation. For a random process
with zero mean value, Xrms (t) = σX (t).
For a stationary random process, because p1 (x, t) is independent of t, all of these
averages are also independent of t.
The prefix auto- indicates that the two random variables considered, X(t1 ) and X(t2 ),
belong to the same random process. A related quantity is the covariance function
defined by
KXX (t1 , t2 ) = E[ X(t1 )−μ1 · X(t2 )−μ2 ] = RXX (t1 , t2 ) − μ1 μ2 , (3.1.12)
where μI = E[ X(tI ) ], I = 1, 2.
The counterpart of the autocorrelation function is the cross-correlation function,
defined as
+∞
RXY (t1 , t2 ) = E[ X(t1 )Y(t2 ) ] = x y pXY (x, t1 ; y, t2 )dxdy, (3.1.13)
−∞
where X(t1 ) and Y(t2 ) belong to two different random processes X(t) and Y(t).
Two random processes X(t1 ) and Y(t2 ) are independent if
For a stationary random process, the autocorrelation function depends on the time
difference t2 −t1 only and is usually denoted by RXX (τ ), where τ = t2 −t1 . Without
loss of generality, suppose that X(t) has zero mean value. The autocorrelation function
RXX (τ ) possesses the following properties:
1. RXX (0) = E[ X 2 (t) ] = Xrms
2 (t) = σ 2 , the mean-square value of X(t).
X
2. RXX (τ ) is symmetric about τ = 0, i.e., RXX (τ ) = RXX (−τ ).
3. It can be shown that RXX (0) RXX (τ ). Hence, R XX (0), if it exists, must be zero
and R XX (0), if it exists, is negative.
4. If there is no periodic component, then lim RXX (τ ) = 0.
τ →±∞
5. RXX (τ ) may be expressed in the following form
+∞
F
1
RXX (τ ) = −1
SXX (ω) = SXX (ω) ei ωτ dω, (3.1.14a)
2π −∞
i.e., RXX (τ ) is the inverse Fourier transform of SXX (ω), where SXX (ω) is a real
positive even function of frequency ω. Hence, SXX (ω) is the Fourier transform of
RXX (τ ) given by
+∞
SXX (ω) = F
RXX (τ ) = RXX (τ ) e−i ωτ dτ. (3.1.14b)
−∞
Function SXX (ω) is known as the power spectral density (PSD) function of the
random process X(t). Equations (3.1.14) connect the autocorrelation function and
the power spectral density function of a stationary random process.
The meaning of the term “power spectral density” becomes clear when looking at
the relation
+∞
1 +∞
. E[ X 2 (t) ] = RXX (0) = SXX (ω)dω = SXX ( F )dF, ω = 2πF. (3.1.15)
2π −∞ −∞
RXX (τ ) = A2 cosω0 τ ,
62
RXX(τ) SXX(ω)
0 τ
(a) –ω0 0 ω0 ω
RXX(τ) SXX(ω)
α1
α1 <α2 <α3
α2
α=α1 α3
α2
α3
0 τ (b) 0 ω
RXX(τ) SXX(ω)
A2 δ(τ) A2
0 τ (c) 0 ω
Figure 3.2 Autocorrelation functions and corresponding power spectral density functions.
+∞
SXX (ω) = A2 e−i ωτ cosω0 τ dτ = πA2 δ(ω +ω0 ) + δ(ω −ω0 ) ,
−∞
where δ(·) denotes the Dirac delta function. Thus, all the power is concentrated at
the frequencies ω = ±ω0 as shown in Figure 3.2(a).
2. Exponential autocorrelation function as shown in Figure 3.2(b):
2αA2
RXX (τ ) = A2 e−α|τ | , SXX (ω) = .
α 2 + ω2
3. White noise process:
RXX (τ ) = A2 δ(τ ), SXX (ω) = A2 . (3.1.16)
The power spectral density of a white noise process is constant over all frequencies
as shown in Figure 3.2(c). A white noise process is clearly not physically realizable
because its total power is infinite. However, it is a convenient mathematical ide-
alization for a process whose power spectral density function remains practically
constant over a wide band of frequency.
where Trms is the time interval suitable for calculating the mean-square or root-mean-
square value of X(t), such as the total duration of X(t).
☞ It is appropriate to characterize a stationary process in terms of power, which
is the energy in unit time, and power spectral density function.
☞ It is suitable to describe a transient nonstationary random process in terms of
total energy, Fourier amplitude spectrum, and energy spectral density function.
where, for I = 1, 2,
2 E[ X(t1 )−μ1 X(t2 )−μ2 ]
μI = E[ X(tI ) ], σI2 = E[ X(tI )−μI ], ρ(t1 , t2 ) = ,
σ1 σ2
pn (x1 , t1 ; . . . ; xn , tn ) = , (3.1.23a)
(2π )n/2 1/2
and λIj is the cofactor of the element Ij in . In the matrix form, one has
1
T −1
pn (x, t) = 1/2 exp − 2 (x−m) (x−m) .
1
(3.1.23b)
(2π )
n/2
A Gaussian random process is completely characterized by its mean μ(t) and covari-
ance function K(t1 , t2 ). Many physical processes, resulting from the superimposition
of a large number of random factors, are often modelled as Gaussian. This assump-
tion is a consequence of the central limit theorem, which states that, under very general
conditions, a random variable that occurs as the sum of many smaller independent ran-
dom variables, each of which is almost negligible in itself, is approximately Gaussian,
whatever the distributions of the component variables are.
3.2 properties of random processes 65
t
0
Z(t)=H[ X(t)−u]
1
t
0
Z(t)
t
0
Z(t)H[ X(t)]
t
0
NX+(u,t)
3
2
1
t
0
Figure 3.3 Level crossing and counting process.
Define a process Z(t) = H X(t)−u as shown in Figure 3.3, where H(x) is the
Heaviside step function with H(x) = 1, if x > 0, and H(x) = 0, if x < 0.
Differentiating Z(t) with respect to t yields Ż(t) = δ X(t)−u Ẋ(t), where δ(t) is
the Dirac delta function. Ż(t) is a positive delta function when there is an upcrossing of
level X = u, and a negative delta function when there is a downcrossing of level X = u.
Multiplying Ż(t) by H Ẋ(t) eliminates downcrossing of level X = u. Integrating
Ż(t) H Ẋ(t) results in the counting process NX+ (u, t), which is the total number of
upcrossings in time [0, t]:
t
NX+ (u, t) = Ż(s) H Ẋ(s) ds.
0
Hence, the expected rate of upcrossing level X = u is the expected value of the derivative
of NX+ (u, t), i.e.,
dNX+ (u, t)
[ ] [ ]
νX+ (u, t) =E = E Ż(t) H Ẋ(t) = E δ X(t)−u Ẋ(t) H Ẋ(t)
dt
+∞ +∞
= δ(x−u) ẋ H(ẋ) pXẊ (x, ẋ) dx dẋ
ẋ= − ∞ x= − ∞
+∞ +∞
= ẋ H(ẋ) pXẊ (u, ẋ) dẋ = ẋ pXẊ (u, ẋ) dẋ.
ẋ= − ∞ 0
3.2 properties of random processes 67
Using
pXẊ (x, ẋ) = pX (x) pẊ ẋ X(t) = x
yields the expected rate of upcrossing the level X = u:
+∞ +∞
+
νX (u, t) = ẋ pXẊ (u, ẋ) dẋ = pX (u) ẋ pẊ ẋ X(t) = u dẋ. (3.2.6)
0 0
Between any two upcrossing of the same level u, at least one peak must occur. Hence,
νp νX+ (u, t) for any u and for any process X(t) with continuous time derivatives.
For a narrow-band process, the rate of occurrence of peaks is expected to be only
slightly larger than the rate of upcrossings of the mean μX = E[ X(t) ]. This property is
commonly used to provide a measure of bandwidth through the irregularity factor:
ν+
X (μX , t)
I= , 0 < I 1. (3.2.10)
νp (t)
For a narrow-band process, I tends to 1.
68
X(t)
u
0 t
H[ X(t)]
1
t
0
H[−X(t)]
1
t
0
δ[−X(t)]·[−X(t)]
t
0
t
0
−X(t) · δ[−X(t)]· H[−X(t)]· H[u −X(t)]
t
0
Np(u,t)
3
2
1
t
0
Figure 3.4 Determination of Np (u, t).
[ ]
E − Ẍ(t) · δ − Ẋ(t) · H − Ẍ(t) · H u−X(t)
Fp(t) (u) =
[ ]
E − Ẍ(t) · δ − Ẋ(t) · H − Ẍ(t)
+∞ +∞ +∞
− ẍ(t) δ − ẋ(t) H − ẍ(t) H u−x(t) pXẊẌ (x, ẋ, ẍ)dx dẋ dẍ
= ẍ= − ∞ ẋ= − ∞ x= − ∞
+∞ +∞
− ẍ(t) δ − ẋ(t) H − ẍ(t) pẊẌ (ẋ, ẍ) dẋ dẍ
ẍ= − ∞ ẋ= − ∞
0 u
ẍ pXẊẌ (x, 0, ẍ) dx dẍ
ẍ= − ∞ x= − ∞
= 0 . (3.2.15)
ẍ pẊẌ (0, ẍ) dẍ
ẍ= − ∞
70
Differentiating with respect to u gives the probability density function for the peak
0
ẍ p (u, 0, ẍ) dẍ
XẊẌ
−∞
pp(t) (u) = 0 . (3.2.16)
ẍ p (0, ẍ) dẍ
ẊẌ
ẍ= − ∞
Equations (3.2.15) and (3.2.16) give the probability distribution of peak that occurs
within the vicinity of time t. Other quantities can be easily obtained, such as the mean,
the mean-square value
+∞ +∞
μp = E[ p(t) ] = u pp(t) (u)du, E[ p (t) ] =
2
u2 pp(t) (u)du, (3.2.17)
−∞ −∞
The conditional probability density function of a Gaussian process is also Gaussian, i.e.,
1 (ẍ − μ̂)2
pẌ ẍ X = u = √ exp − ,
2π σ̂ 2 σ̂ 2
where the conditional mean and standard deviation of Ẍ(t) are
σẌ σẊ2
μ̂ = ρXẌ · · u, σ̂ = σẌ 1−ρX2Ẍ , ρXẌ = − = − I. (3.2.18)
σX σX σẌ
Hence,
1 0 (ẍ − μ̂)2
pp (u) = − p (u) ẍ exp − dẍ
σẌ σ̂ X ẍ = − ∞ 2 σ̂ 2
3.2 properties of random processes 71
μ̂2 √
σ̂ μ̂ μ̂
= pX (u) exp − 2 − 2π
−
σẌ 2 σ̂ σẌ σ̂
I 2 u2 √ Iu Iu
= pX (u) 1− I 2 ξ 2 exp − + 2π
√ .
2(1− I )σX
2 2 σX 1− I 2 σX
Substituting the Gaussian form for pX (u) gives the probability density function for the
peaks of a mean zero stationary Gaussian process X(t)
√
1− I 2 u2 Iu u2 Iu
pp. (u) = √ exp − + 2 exp − 2
√ . (3.2.19)
2π σX 2(1− I 2 )σX2 σX 2σX 1− I 2 σX
d d !
Np
pX (x) = FXmax (x) = Fp (x) . (3.2.23)
max dx dx
The expected value of the maximum peak is given by
+∞ +∞
E[ Xmax ] = x · pX (x)dx = x · d Fp (x) Np
max
−∞ −∞
0 +∞ !
Np Np
=− Fp (x) dx + 1 − Fp (x) dx, (3.2.24)
−∞ 0
The peak factor Pf, defined as the ratio of the peak value E[ Xmax ] and root-mean-
square value σX of random process X(t), is given by
!
E[ Xmax ] 1 +∞ x2
Pf = =√ 1 − 1 − I e−θ Np
θ −1/2 dθ , θ= , (3.2.28)
σX 2 0 2σX2
or
!
Pf = [ max ] = 2
E X √ +∞ 2 N x
1 − 1 − I e−z p dz, z= √ . (3.2.29)
σX 0 2 σX
The extreme value distribution is then the distribution of Y(t). Note that even for a
stationary process X(t), Y(t) is generally nonstationary because larger and larger values
of X(t) will generally occur when the period [0, t] is extended.
The cumulative distribution function of Y(t) is
FY(t) (u) = P Y(t) u = P X(s) u for 0 s t . (3.2.31)
Let TX (u) > 0 denote the time at which X(t) has the first upcrossing of the level
u, i.e., X TX (u) = u, Ẋ TX (u) > 0, and there has been no crossing in the interval
0 t < TX (u). For any given u value, TX (u) is a random variable.
Note that the extreme value problem X(s) u for 0 s t is equivalent to the
first passage time problem X(0) u, TX (u) t . Taking the probabilities gives
FY(t) (u) = P TX (u) t X(0) u P X(0) u
= P TX (u) t X(0) u FY(0) (u). (3.2.33)
In many problems, the condition in P TX (u) t X(0) u can be neglected:
❧ In some problems, P X(0) u = 1, such as when the system is known to start at
X(0) = 0, and conditioning by a sure event can always be neglected.
❧ In other situations, although the distribution of TX (u) depends on X(0), the effect
of X(0) may be significant only for a small period of time.
Equation (3.2.33) can be written as
FY(t) (u)
P TX (u) t X(0) u = 1 − .
FY(0) (u)
which leads to
1 ∂ 1 FY(t) (u) − FY(t + t) (u)
ηX (u, t) = − FY(t) (u) = lim
FY(t) (u) ∂t t→0 FY(t) (u) t
1 P t TX (u) t+t X(0) u
= lim . (3.2.36)
t→0 t P TX (t) t X(0) u
The event t TX (u) t+t means that the first upcrossing of level u is in the time
interval t, t+t . This event is the intersection of the event that there is no upcrossing
prior to t and the event that there is an upcrossing in the time interval t, t+t , i.e.,
t TX (u) t+t = TX (u) t ∩ Upcrossing in t, t+t .
74
Equation (3.2.36), in which the ratio is the conditional probability, can be written as
P Upcrossing in t, t+t X(0) u ∩ No upcrossing prior to t
ηX (u, t) = lim
t→0 t
E[ Upcrossing in t, t+t X(0) u ∩ No upcrossing prior to t ]
= lim . (3.2.37)
t→0 t
Hence, ηX (u, t) is the conditional rate of upcrossing of level u, given that the initial
condition is below u and that there is no prior upcrossing.
Using the conditional probability density function, ηX (u, t) can be written as
+∞
ηX (u, t) = ẋ pXẊ u, ẋ X(0) u ∩ No upcrossing in t, t+t dẋ.
ẋ=0
However, the conditional probability density function is generally unknown, and some
approximations must be made.
Note that most physical processes have only a finite memory in the sense that X(t)
and X(t−τ ) can generally be considered independent if τ > T for some large T value.
Hence, for t > T,
pXẊ u, ẋ X(0) u ∩ No upcrossing in [0, t] ≈ pXẊ u, ẋ No upcrossing in [t−T, t] .
This limiting behaviour for large t is also applicable if X(t) is a nonstationary process
that has finite memory and that becomes stationary with the passage of time.
Poisson Approximation
The most widely used approximation of the extreme distribution problem is to neglect
the conditioning event X(0) u ∩ No upcrossing prior to t . Hence
If the crossing rate is independent of the past history of the process, the time inter-
vals between upcrossings are independent, which makes the integer-valued counting
process NX+ (u, t) a Poisson process. Hence, the approximate of equations (3.2.39) to
(3.2.41) is commonly called the Poisson approximation of the extreme value or the
first-passage problem.
Substituting equation (3.2.41) into (3.2.34) yields, for stationary process X(t),
1 ∂FY(t) (u) +
pT (t) = − ≈ νX+ (u) e−νX (u) t . (3.2.42)
X FY(0) (u) ∂t
Hence, the Poisson approximation gives an exponential distribution for the first-
passage time of a stationary process X(t). The mean first-passage time can be easily
obtained from the exponential distribution and is given by
1
E[ TX+ (u) ] ≈ , for stationary process X(t). (3.2.43)
νX+ (u)
∂e−ξ dξ dξ
pY(t) (u) = · = e−ξ . (3.2.47)
∂ξ du du
76
4.0
Peak Factor Pf
3.5
3.0
2.5
2.0
ν0 t
1.5
0 100 200 300 400 500 600 700 800 900 1000
Figure 3.5 Peak factor.
When u→ + ∞, ξ →0, and when u→0, ξ →ν0 t. The expected value of Y(t) is
+∞ νt
E[ Y(t) ] = E[ Xmax ] =
0
u pY(t) (u) du = u e−ξ dξ. (3.2.48)
0 0
Pf = [ max ] ≈ 2 lnν0 t + √
E X √ γ
, γ = 0.5772, (3.2.52)
σX 2 lnν0 t
where γ is the Euler number. This result was first obtained in Davenport (1964) and is
plotted in Figure 3.5.
Der Kiureghian (1980) determined an empirical reduced zero-upcrossing rate νe ,
representing an equivalent rate of statistically independent crossings, given by
1.63 q0.45 − 0.38 ν0 , q < 0.69,
νe = (3.2.53)
ν0 , q 0.69,
3.3 single degree-of-freedom system 77
where q is the spectral parameter defined in equation (3.2.5). In the peak factor given
by equation (3.2.52), ν0 is replaced by the reduced rate νe . Equations (3.2.53) and
(3.2.52) with ν0 replaced by νe are applicable for 0.1 q 1 and 5 ν0 t 1, 000,
which are of interest in earthquake engineering. Resulting error in the estimated peak
factor is generally within 3 %.
Double-Barrier Problem
In many engineering applications, it is required to determine large excursions of X(t)
in either the positive or negative direction. For example, for earthquake ground motion
excitation üg(t), the positive or negative sign of üg(t) has no real significance, and it is
important to determine peak ground acceleration üg(t)max .
The event of X(t) remaining between −u and +u is exactly the same as the event
of X(t) remaining below the level u. Following equation (3.2.35), one can write
t
FY(t) (u) = FY(0) (u) exp − η|X| (u, s) ds , Y(t) = max X(s). (3.2.54)
0 0st
The terms double-barrier problem and single-barrier problem are often used to distin-
guish between the upcrossings by X(t) and X(t), respectively.
The Poisson approximation of the symmetric double-barrier problem of equation
(3.2.54) is simply to replace η|X| (u, s) with
+
ν|X| (u, s) = νX+ (u, s) + νX− (−u, s).
+
If the distribution of X(t) and Ẋ(t) is symmetric, this gives ν|X| (u, s) = 2νX+ (u, s).
Let x(t) = u(t)−ug(t) be the relative displacement between the girder and the base. In
terms of the relative displacement x(t), the equation of motion becomes
The equivalent loading on the girder created from ground excitation is −m üg(t), which
is proportional to the mass of the structural system m and the ground acceleration üg(t).
The equation of motion (3.3.1) can be written in the standard form as
1 K c
ẍ(t) + 2ζ0 ω0 ẋ(t) + ω02 x(t) = m F(t), ω02 = m , 2ζ0 ω0 = m , (3.3.2)
where ω0 is the natural circular frequency, ζ0 is the damping ratio, and the forcing
is F(t) = P(t)−m üg(t). In earthquake engineering, it is convenient to use an SDOF
oscillator as illustrated in Figure 3.7 to model an SDOF system under base excitation
ug(t) or üg (t), with P(t) = 0.
xC(t)
a
a e−ζω0 t
ϕ
ωd
t
ae−ζω0 t cos(ωd t − ϕ)
−ae−ζω0 t (Envelope)
−a
Figure 3.8 Response of underdamped free vibration.
t–τ
t t t
τ t τ t –τ
τ
xP
xP(t,τ)
t
τ t
Figure 3.9 Response of underdamped SDOF system due to an impulse.
The characteristic equation is λ2 +2ζ0 ω0 λ+ω02 = 0, giving λ = ω0 −ζ0 ± ζ02 −1 .
Most engineering structures are underdamped with 0 < ζ0 < 1. The roots of the
characteristic equation become
λ = ω0 −ζ0 ± i 1−ζ02 = −ζ0 ω0 ± iωd , ωd = ω0 1−ζ02 ,
where ωd is the damped natural circular frequency.
The response of free vibration is
xC (t) = e − ζ0 ω0 t (A cosωd t + B sinωd t), (3.3.4)
where constants A and B are determined from the initial conditions x(0) = x0 and
ẋ(0) = v0 , resulting in
v +ζ ω x
xC (t) = e − ζ0 ω0 t x0 cosωd t + 0 0 0 0 sinωd t , 0 ζ0 < 1, (3.3.5)
ωd
= ae − ζ0 ω0 t cos(ωd t − ϕ), (3.3.6)
where v +ζ ω x 2 v +ζ ω x
0 0 0 0 −1 0 0 0 0
a= x20 + , ϕ = tan .
ωd ωd x0
The response of free vibration of an underdamped system with 0 < ζ0 < 1 is shown
in Figure 3.8, which decays exponentially and approaches zero as t→∞ and is called
transient response. Because its value becomes negligible after some time, its effect is
small and is not important in practice.
80
where H(t) is the unit impulse response function of the SDOF system given by
sinωd t
H(t) = e−ζ0 ω0 t , ωd = ω0 1−ζ02 . (3.3.10)
m ωd
Integral of the form (3.3.9) is called a convolution integral or Duhamel integral. For
lightly damped system, ζ0
1, ωd = ω0 1−ζ02 ≈ ω0 .
where h(t) is referred to as the impulsive response function with respect to base excitation
in this book and is given by
sinωd t
h(t) = e − ζ0 ω0 t . (3.3.13)
ωd
☞ The difference between functions H(t) and h(t) is the mass term m. h(t) is used
in structures under base excitation (earthquake) because the mass term m in
H(t) is cancelled with the mass in the equivalent earthquake load −m üg (t).
Equation (3.3.12) can be rewritten as
t
1
x(t) = e−ζ0 ω0 (t−τ ) sinωd (t−τ ) üg(τ ) dτ. (3.3.14)
ωd 0
Substituting equations (3.3.14) and (3.3.16) into (3.3.2) yields the absolute acceleration
where
t
I s
(t) = e−ζ0 ω0 (t−τ ) sinωd (t−τ ) üg(τ ) dτ ,
0
(3.3.18)
t
I c
(t) = e −ζ0 ω0 (t−τ )
cosωd (t−τ ) üg(τ ) dτ.
0
t
ẋ(t) ≈ e−ζ0 ω0 (t−τ ) cosω0 (t−τ ) üg(τ ) dτ ≈ ω0 h ∗ üg , (3.3.16 )
0
t
ü(t) ≈ −ω0 e−ζ0 ω0 (t−τ ) sinω0 (t−τ ) üg(τ ) dτ = −ω02 h ∗ üg . (3.3.17 )
0
82
If dynamic effect is not considered, i.e., if only static terms are considered in equation
(3.3.2), one obtains xstatic = P0 /K, which is the static displacement of the structure
under static force P0 .
The dynamic magnification factor (DMF) is defined by
x (t) 1 ω
DMF =
- P max
= , r= ω , (3.3.21)
xstatic (1−r 2 )2 +(2ζ0 r)2 0
where r is the frequency ratio. D-MF is plotted in Figure 3.10 for various values of the
damping ratio ζ0 ; it is one of the most important quantities describing the dynamic
behavior of an underdamped SDOF system under harmonic excitation.
❧ When r→0 (ω
ω0 ), D-MF →1. The dynamic excitation is effectively a static force
and the amplitude of dynamic response approaches the static displacement.
❧ When r→∞ (ω ω0 ), DMF →0 or the dynamic response approaches zero.
-
1 1
DMFmax
-
≈ DMF r=1 =
-
= . (3.3.22)
(1−r ) +(2ζ0 r) r=1
2 2 2 2ζ 0
Hence, the smaller the damping ratio, the larger the amplitude of dynamic response.
❧ When ζ0 = 0 and ω = ω0 , the system is in resonance, and the amplitude of the
response grows linearly with time.
5
ζ=0.1
3
ζ=0.2
1
r ≈1 – ζ2 ζ=0.3 r = ωω
0
5
ζ=0.1
3
ζ=0.2
2
ζ =0.3
1 r ≈1 + ζ2
r = ωω
0
0 0.5 1 1.5 2 2.5
Figure 3.11 DMF of SDOF system under ground excitation.
where H(ω) is the complex frequency response function with respect to base excitation
∞ 1
H(ω) = h(t) e−i ωt dt = . (3.3.24)
−∞ (ω02 −ω2 ) + i2ζ0 ω0 ω
The D-MF, characterizing the magnification of the dynamic displacement response am-
plitude x(t)max of the SDOF oscillator in terms of the ground displacement amplitude
u0 , is defined as
x(t) r2 ω
DMF =
- max
= , r= ω . (3.3.25)
u0 (1−r 2 )2 +(2ζ0 r)2 0
84
The D-MF is plotted in Figure 3.11 for various values of the damping ratio ζ0 .
❧ When r→0 (ω
ω0 ) or when the SDOF oscillator is very stiff, D-MF →0. The
SDOF oscillator moves with the ground as a rigid body, and the relative displacement
between the mass and the ground approaches 0.
❧ When r→∞ (ω ω0 ) or when the SDOF oscillator is very flexible, D-MF →1. The
mass m does not move, and the relative displacement between the mass and the
ground approaches the ground displacement.
❧ When r ≈ 1 (ω ≈ ω0 ),DMF tends to large values for small damping.
-
r2 1
DMFmax
-
≈ DMF r=1 =
-
= . (3.3.26)
(1−r ) +(2ζ0 r) r=1
2 2 2 2ζ 0
Seeking a solution of the form x(t) = ϕ e i ωt and substituting into equation (3.4.2) yield
an eigenvalue problem
(K−ω2 M)ϕ = 0. (3.4.3)
To have nonzero solutions for ϕ, the determinant of the coefficient matrix must be zero
det(K−ω2 M) = 0. (3.4.4)
The Ith root (eigenvalue) ωI (ω1 < ω2 < · · · < ωN ) is the natural frequency of the Ith
mode of the system or the Ith modal frequency.
3.4 multiple degrees-of-freedom systems 85
is the Ith eigenvector or the Ith mode shape. Construct the modal matrix as
⎡ ⎤
ϕ11 ϕ12 · · · ϕ1N
⎢ϕ ϕ22 · · · ϕ2N ⎥
= ϕ1 ϕ2 · · · ϕN = ⎢ ⎣ .
21
. .. .. .. ⎥
⎦, (3.4.6)
. . . .
ϕN1 ϕN2 · · · ϕNN
where the first subscript I of element ϕI j refers to the node number and the second
subscript j corresponds to the mode number. has the following orthogonal relations
T M = diag m̄1 , m̄2 , . . . , m̄N = m̄, (3.4.7)
T K = m̄2 = diag m̄1 ω12 , m̄2 ω22 , . . . , m̄N ωN , 2
(3.4.8)
where = diag ω1 , ω2 , . . . , ωN , and m̄1 , m̄2 , . . . , m̄N are the modal masses.
It is usually assumed that
T C = diag c̄1 , c̄2 , . . . , c̄N , c̄n = 2ζn ωn m̄n , (3.4.9)
i.e., the structure has classical damping and the modal matrix can diagonalize the
damping matrix.
or
Ln
q̈n + 2ζn ωn q̇n + ωn2 qn = −n üg , n = , (3.4.10)
m̄n
where Ln , called the earthquake excitation factors, are given by
Ln = ϕ Tn M I.
n
t
qn (t) = − ω Vn (t), Vn (t) = e−ζn ωn (t−τ ) sin ωn (t−τ ) üg(τ )dτ. (3.4.11)
n 0
86
The elastic forces associated with the relative displacements are given by
where Fe,n (t) is the elastic force at the nth floor given by
Having obtained the elastic forces Fe,n (t) at any time t during the earthquake, any
desired force results, such as base shear and overturning moment, can be determined.
Taking the expectation of both sides of equation (3.5.2), the mean response is
+∞
E[ X(t) ] = H(τ ) E[ F(t−τ ) ] dτ.
−∞
If F(t) has zero mean, i.e., mF = 0, then the mean response is zero, i.e., mX = 0.
The auto-correlation function of X(t) is
+∞ +∞
RXX (τ ) = E[ X(t) X(t+τ ) ] = E H(τ1 )F(t−τ1 )dτ1 H(τ2 )F(t+τ −τ2 )dτ2
−∞ −∞
+∞ +∞
= H(τ1 ) H(τ2 ) RFF (τ +τ1 −τ2 ) dτ1 dτ2 , (3.5.4)
−∞ −∞
+∞
SXX (ω) = RXX (τ ) e−i ωτ dτ
−∞
+∞ +∞ +∞
= H(τ1 ) H(τ2 ) RFF (τ +τ1 −τ2 ) e−i ωτ dτ dτ1 dτ2 .
−∞ −∞ −∞
where H(ω) is the frequency response function given by equation (3.3.20), H∗ (ω)
is the complex conjugate of H(ω), and SFF (ω) is the power spectral density (PSD)
function of the excitation. Thus,
SXX (ω) = H(ω)2 SFF (ω). (3.5.5)
88
If the excitation F(t) is a stationary Gaussian process, it can be shown that the response
X(t) is also Gaussian; the mean and auto-correlation function are then sufficient to
describe the response random process completely.
For the SDOF system described by equation (3.5.1), for which H(ω) is given by
(3.3.20), the response power spectral density is, by (3.5.5),
1
SXX (ω) = SFF (ω). (3.5.8)
m2 (ω02 −ω2 )2 + (2ζ ω0 ω)2
The auto-correlation function, given by equation (3.5.6), is
+∞
1 1
S (ω) e
i ωτ
RXX (τ ) = dω, (3.5.9)
2π −∞ m (ω0 −ω ) + (2ζ ω0 ω)2 FF
2 2 2 2
Suppose that F(t) is a white noise process with SFF (ω) = S0 and zero mean. Then
the mean response given by equation (3.5.3) is mX = 0. The auto-correlation function
and the mean-square response given by equations (3.5.9) and (3.5.10) are
S0 +∞ e i ωτ
RXX (τ ) = dω, (3.5.11)
2πm2 −∞ (ω02 −ω2 )2 + (2ζ ω0 ω)2
S0 +∞ 1
σX = E[ X (t) ] =
2 2
dω. (3.5.12)
2πm −∞ (ω0 −ω ) + (2ζ ω0 ω)2
2 2 2 2
The integrals in equations (3.5.11) and (3.5.12) may be evaluated by the method of
residues (see Section 3.8, in particular equation (3.8.11)) to give
S0 π S0 S0
σX2 = · =⇒ σX2 = = . (3.5.13)
2πm2 2ζ ω03 4 m2 ζ ω03 2cK
and
ζ ω0
RXX (τ ) = σX2 e−ζ ω0 τ cosωd τ + sinωd τ . (3.5.14)
ωd
3.5 stationary response to random excitation 89
For a general (coloured noise) case, the integration in (3.5.9) may have to be per-
2
formed numerically. However, if the damping is light with ζ
1, the function H(ω)
is sharply peaked near the frequency ω = ω0 (see Figure 3.12). If further SFF (ω) does
not vary too rapidly in the neighbourhood of ω = ω0 as in Figure 3.13, the excitation
may be approximated by a white noise with spectral density equal to SFF (ω0 ). One may
then write
S (ω ) +∞ 2 S (ω ) S (ω )
σX2 ≈ FF 0 H(ω) dω = FF2 0 3 = FF 0 . (3.5.15)
2π −∞ 4 m ζ ω0 2cK
H(ω) 2
1
2ζω0
(2kζ)2
1 1
2(2kζ)2 k2
−ω0 ω0 ω
H(ω) 2
White noise
approximation PSD of excitation
SFF(ω)
H(ω) 2
White noise
approximation
SFF(ω)
H(ω) 2
u(t)
t (s)
Stationary
where
sinωn, d t
hn (t) = e − ζn ωn t , ωn, d = ωn 1−ζn2 , (3.5.17)
ωn, d
is the impulse response function with respect to base excitation of the nth mode, and
its Fourier transform is the complex frequency response function with respect to base
excitation given by
∞
1
Hn (ω) = hn (τ ) e−i ωτ dτ = . (3.5.18)
−∞ (ωn −ω ) + i2ζn ωn ω
2 2
Mean Response
Taking the expectation of both sides of equation (3.5.16), the mean response is
∞
E[ qn (t) ] = −n hn (τ ) E[ üg (t−τ ) ] dτ = 0, (3.5.19)
−∞
because üg (t) is stationary with mean zero, i.e., E[ üg (t) ] = 0.
Covariance of Response
The covariance of responses produced by modes m and n is given by
E[ qm (t)qn (t+τ ) ]
∞ ∞
=E −m üg (t−τ1 ) hm (τ1 )dτ1 −n üg (t+τ −τ2 ) hn (τ2 )dτ2
−∞ −∞
∞ ∞
= m n hm (τ1 ) hn (τ2 ) E[ üg (t−τ1 ) üg (t+τ −τ2 ) ] dτ1 dτ2
−∞ −∞
∞ ∞
= m n hm (τ1 ) hn (τ2 )Rüg üg (τ +τ1 −τ2 ) dτ1 dτ2 , (3.5.20)
−∞ −∞
where Rü (τ ) = E[ üg (t) üg (t+τ ) ] is the auto-correlation function of üg (t). Taking
g üg
Fourier transform of both sides yields
∞
Sqm qn (ω) = E[ qm (t)qn (t+τ ) ] e−i ωτ dτ
−∞
∞ ∞ ∞
= m n hm (τ1 ) hn (τ2 ) Rüg üg (τ +τ1 −τ2 ) e−i ωτ dτ1 dτ2 dτ.
−∞ −∞ −∞
92
where Süg üg(ω) is the power spectral density of earthquake excitation üg (t), which is
the Fourier transform of Rü (τ ). Hence, taking the inverse Fourier transform gives
g üg
∞
1
E[ qm (t)qn (t+τ ) ] = Sqm qn (ω) e i ωτ dω
2π −∞
∞
= m n Hm∗ (ω) Hn (ω) Süg üg(ω) e i ωτ dω, (3.5.22)
2π −∞
and, by setting τ = 0,
∞
m n
E[ qm (t)qn (t) ] = Hm∗ (ω) Hn (ω) Süg üg(ω) dω. (3.5.23)
2π −∞
Snn n2
E[ q2n (t) ] = , (3.5.29)
4 ζn ωn3
= αmn ρmn σqm σqn , αmn = m n = sgn(m n ). (3.5.31)
m n
where coefficients An are known for the structural system under consideration. Squar-
ing both sides of equation (3.5.32) yields
N
N
z2 (t) = Am An qm (t) qn (t).
m=1 n=1
94
N
N
E[ z2 (t) ] = σz2 = Am An E[ qm (t) qn (t) ].
m=1 n=1
The maximum value of the response z(t) is, according to Section 3.2,
z(t)
max
= Pfz · σz , Pfz = Peak factor. (3.5.34)
Similarly,
q (t)
n max
= Pfqn · σqn , Pfqn = Peak factor, n = 1, 2, . . . , N. (3.5.35)
❧ CQC Method. The combination method of (3.5.37) for evaluating maximum total
response from the individual maxima of modal responses is known as the complete
quadratic combination (CQC) method. When the major contributing modes have
frequencies close together, the corresponding cross terms in equation (3.5.37) can be
very significant and should be retained.
❧ SRSS Method. If the frequencies of the contributing modes are well separated, the
cross terms in equation (3.5.37) are negligible, i.e., ρmn
1, m = n. In this case,
equation (3.5.37) reduces to
N 2
z(t) = A2n qn (t)max . (3.5.38)
max
n=1
It is very important to note that using the CQC or the SRSS combination
method must always be the last step in evaluating the maximum value of any
response quantity. In other words, one cannot use the maximum value of one
response quantity, obtained using the CQC and the SRSS method, to evaluate
the maximum value of another response quantity.
where ⎧ ⎫
⎧ ⎫ x ⎪ ⎧ I⎫ ⎧ ⎫
⎪ x ⎪ ⎪
⎪ δ ⎪
⎨ x1 ⎪
⎪ ⎬ ⎪ n,1 ⎪
⎨ ⎪ ⎪1⎪
⎨1I ⎪
⎪ ⎪
⎨δI1 ⎪
⎪
xn,2 ⎬ I
⎬
I I2
⎬
x = .2 , xn = . , I = . ,
. 1 = . ,
.⎪
(3.6.2)
⎪ . ⎪ ⎪ . ⎪ ⎪ .⎪ ⎪
⎩ . ⎪
⎪ ⎭ ⎪
⎩x . ⎪
⎪ ⎪
⎭ ⎩ I⎪
⎪ ⎭ ⎩.⎪
⎪ ⎭
xN 1 δI6
n,6
M, C, and K are, respectively, the mass, damping, and stiffness matrices of dimension
6N×6N, xn is the relative displacement vector of node n, I I is the influence vector of
seismic excitation in direction I, and δIj denotes the Kronecker delta function, i.e.,
δIj = 0, if I = j, and δIj = 1, if I = j.
Let x = x1 + x2 + x3 , where x I is the relative displacement vector due to earthquake
excitation ugI (t) in direction I. Hence, xn,I j = un,I j − ugI δIj , where xn,I j and un,I j
96
are, respectively, the relative and absolute displacements of node n in direction j due to
earthquake excitation in direction I. Because the system is linear, from equation (3.6.1),
x I is governed by
where m̄1 , m̄2 , . . . , m̄6N are the modal masses. Assume that the structure has classical
damping so that the modal matrix can also diagonalize the damping matrix
T C = diag c̄1 , c̄2 , . . . , c̄6N T , c̄K = m̄K · 2ζK ωK . (3.6.6)
Forced Vibration
Apply the transformation
N
6
6N
xn,I j = ϕn, j; 6(ν−1)+δ ν,I δ qν,I δ = ϕn, j; K KI qKI , (3.6.9)
ν=1 δ=1 K=1
I
Ln, T I
j = ϕ 6(n−1)+j M I , or LKI = ϕ TK M I I , (3.6.10)
I
I
Ln, j LKI ϕ TK M I I
n, j = , or KI = = . (3.6.11)
m̄ 6(n−1)+j m̄K ϕ TK Mϕ K
For ease of presentation, the two-subscript-notation n, j (node, direction) and the
one-subscript-notation K = 6(n−1)+j are used interchangeably; the former is advan-
tageous in describing the meaning of the quantity in terms of node and direction, and
I
the latter gives the position of the quantity in the corresponding vector. Ln, j is the
earthquake excitation factor, quantifying the contribution of earthquake excitation in
the Ith direction to the modal response qn,I . I is the modal participation factors; if I
j K K
is small, then the contribution of mode ϕ K to the structural response due to excitation
in the Ith direction is small.
Substituting equation (3.6.7) into (3.6.3) and multiplying T from the left yield
(T M) Q̈I (t) + (T C) Q̇I (t) + (T K)QI (t) = −T M I I ügI (t).
m̄K · KI q̈KI + m̄K 2ζK ωK · KI q̇KI + m̄K ωK2 · KI qKI = − LKI ügI (t),
or
q̈KI (t) + 2ζK ωK q̇KI (t) + ωK2 qKI (t) = − ügI (t), K = 1, 2, . . . , 6N, I = 1, 2, 3. (3.6.12)
The seismic response history analysis (SRHA) procedure is concerned with the
calculation of structural response as a function of time when the system is subjected to
a set of tridirectional ground acceleration ügI (t) (I = 1, 2, 3). For illustration purpose,
the nodal inertia forces of the 6N-DOF system are computed using SRHA based on the
modal superposition method.
From equations (3.6.7) and (3.6.14), the acceleration vector of modal masses (DOF)
of the system relative to the ground is given by
2N
ẍ I (t) = ϕ K KI q̈KI (t), I = 1, 2, 3, (3.6.15)
K=1
where q̈KI (t) is the acceleration of the Kth mode of the system relative to the ground.
The vector of the earthquake-induced ground motion in the Ith direction can be
written as
2N
2N
ügI (t) = I I ügI (t) = ϕ K KI ügI (t), II = ϕ K KI . (3.6.16)
K=1 K=1
The nodal inertia forces of the 6N-DOF system subjected to ground acceleration in the
Ith direction is given by
F I (t) = −M ügI (t) + ẍ I (t) , (3.6.17)
where ügI (t)+ ẍ I (t) is the vector of the nodal absolute accelerations of the system
subjected to ground acceleration in the Ith direction. Substituting equations (3.6.15)
and (3.6.16) into equation (3.6.17) gives
2N
FI (t) = −M ϕ K KI ügI (t)+ q̈KI (t) , (3.6.18)
K=1
which is a function of time. The vector of time-histories of the nodal inertial forces
of the system due to the tridirectional ground acceleration can be obtained by alge-
braic summation of the response time-histories at each time step due to the ground
3
acceleration in individual direction, i.e., F(t) = F I (t).
I=1
3
M ẍ(t) + C ẋ(t) + Kx(t) = p(t), p(t) = −M I I ügI (t) (3.6.19)
I=1
with equivalent earthquake loading p(t) and initial conditions x = x(0), ẋ = ẋ(0) at
t = 0. The solution will provide the displacement vector x(t) as a function of time.
By direct time integration, the equation of motion (3.6.19) is solved using numerical
integration schemes. For linear response analysis of multiple DOF systems, central
difference method and Newmark’s method are two popular direct time integration
methods, which are detailed in Chopra (2012). For nonlinear systems, the solution
algorithm involves an iteration process to converge at each time step. Direct time
integration is usually done using a commercial finite element analysis package, such as
STARDYNE and ANsys.
where
t
S I
V (ζK , ωK ) = V I (t)
K max
= e−ζK ωK (t−τ )
sin ωK (t−τ ) üg (τ )dτ
I
(3.6.21)
0 max
S VI (ζK , ωK )
SAI (ζK , ωK ) = ωK , SDI (ζK , ωK ) = ωK S VI (ζK , ωK ) (3.6.22)
In undamped free vibration, the elastic forces can be expressed in terms of the equiva-
lent inertial forces
K x I = M ẍ I =⇒ K = M2 . (3.6.24)
Hence,
f sI (t) = KQ I (t) = M2 Q I (t). (3.6.25)
In general, for a response quantity z I (t) due to seismic excitation in direction I
6N
z I (t) = AIK qKI (t), (3.6.26)
K=1
I I
αK K = KI KI = sgn(KI KI ), (3.6.28)
K K
8 ζK ζK (ζK +rζK )r3/2 ω
ρK K = , r = ωK , (3.6.29)
(1−r 2 )2 + 4 ζK ζK r (1+r 2 ) + 4 (ζK2 +ζK2 )r 2 K
or using SRSS as
I
6N 2
z (t) = (AIK )2 qKI (t)max . (3.6.30)
max
K=1
The directional maximum absolute value z I (t)max , I = 1, 2, 3, are combined using
SRSS to obtain z(t)max as
2 2
z(t) = z1 (t)2 + z 2 (t)max + z 3 (t)max . (3.6.31)
max max
Example
Consider a frame ABC with a rigid right-angle at B and clamped to ground at support
A as shown in Figure 3.16(a). It supports a lumped mass m at end A with three DOF
u1 , u2 , and u3 . The weight of the frame is negligible. The flexural rigidity of both
members AB (in both directions 1 and 2) and BC (in both directions 2 and 3) is EI.
Both members AB and BC are rigid in the axial direction and in torsion.
The frame is subject to tridirectional ground excitations üg1 (t), üg2 (t), and üg3 (t). The
ground response spectra (GRS) follow USNRC R.G. 1.60 (USNRC, 2014), with GRS in
both horizontal directions being the same and anchored at PGA = 0.3g and vertical
GRS anchored at PGA = 0.2g. The modal damping is 5 % for all modes .
3.6 seismic response analysis 101
ug2(t) M
ug3(t) 2
L,EI
A
ug1(t) δB = ML θB = ML
2EI EI
Figure 3.16 A frame under tridirectional seismic excitations.
2
(1) f11 f11 f11 L3 θ= L
2EI
3EI θ lL2
−f31 F=1
F=1 l 2EI
f11
L,EI
F=1 F=1
(2) L3 F=1 l, EI 3
f22 3EI l
3EI
L,EI
l,EI l3
A EI
Use the parameters L = 2L, L = 4 m, m = 5 kg, and EI = 106 N · m2 . Using the re-
sponse spectrum method, determine the maximum absolute values of the relative
displacements x1 = u1 −u1g , x2 = u2 −ug2 , and x3 = u3 −ug3 and the maximum over-
turning moments M1 and M2 in directions 1 and 2.
Because there is only one lumped mass at end A, the mass matrix is
M = diag m, m, m .
It is easy to determine the flexibility matrix using the definition of flexibility: F I j is the
displacement along DOF I due to a unit force applied along DOF j, while forces along
102
all other DOF being zero. Some useful results for bending of cantilever are shown in
Figure 3.16(b). Referring to Figure 3.17, apply a unit force F = 1 along each of DOF in
turn, the elements of flexibility matrix F are
L3 LL2
F11 = , F21 = 0, F31 = − ,
3EI 3EI
L3 +L 3
F12 = 0, F22 = , F32 = 0,
3EI
LL2 L2L L3
F13 = − , F23 = 0, F33 = + .
3EI EI 3EI
Using L = 2L, the stiffness matrix is given by
⎡ ⎤
21 9
0
EI ⎢ ⎢ 20 10 ⎥
1 ⎥
K = F−1 = ⎢0 0 ⎥.
L3 ⎣ 3 ⎦
9 6
0
10 5
Due to the tridirectional ground excitations, the inertial forces applied on the mass m
are FI = −m(ẍI + ügI ), I = 1, 2, 3. Using the flexibility matrix, the relative displacements
xI = uI −ugI , I = 1, 2, 3, due to the inertial forces are
xI = FI1 −m(ẍ1 + üg1 ) + FI2 −m(ẍ2 + üg2 ) + FI3 −m(ẍ3 + üg3 ) ,
The critical points of USNRC R.G. 1.60 GRS are listed in Table 3.1; each spectrum
is obtained by connecting the critical points linearly in the log-log scale. Using linear
interpolation, the spectral values at the modal frequencies can be determined, with
ζ1 = ζ2 = ζ3 = 5 %:
1 0.8957 0.8723
0.5862 0.6761
0.5729
Vertical 0.2
0.1
I I
x = Q I2 max , x = Q I 2 + Q I 2 ,
2 max 3 max 1 max 3 max
which give
1 2 3
x x x
1 max = 0.006872 m, 1 max = 0, 1 max = 0.006329 m,
1 2 3
x x x
2 max = 0, 2 max = 0.008206 m, 2 max = 0,
1 2 3
x x x
3 max = 0.004143 m, 3 max = 0, 3 max = 0.003817 m.
3.7 nonlinear systems 105
E[X1 X2 X3 ] = 0,
In general, for m = 2, 3, . . . ,
E[X1 X2 · · · X2m−1 ] = 0, E[X1 X2 · · · X2m ] = E[XI Xj ] E[XK XL ], (3.7.5)
where the summation involves (2m)!/(2m m!) terms and is to be taken over all
different ways by which 2m elements can be grouped into m distinct pairs.
2. If X(t) is a stationary process, then
RX X (τ ) = E[ X(t) X(t+τ ) ],
If the error term is zero, the response of the linear system (3.7.2) is given by (3.5.10)
+∞
1 SFF (ω)
E[ X (t) ] =
2
dω. (3.7.4)
2π −∞ (ω0 −ω2 )2 + (βω)2
2
The parameters β and ω02 are chosen in such a way that some average function of the
error E is minimized.
Let E1 (X) = α X− g(X), E2 (Ẋ) = β Ẋ− F(Ẋ), α = ω02 . The usual choice is to mini-
mize the mean-square errors
i.e.,
Equations (3.7.4) and (3.7.7) are solved to determine the mean-square response E[ X 2 ],
and parameters β and ω02 = α.
As an example, consider the following SDOF system with nonlinear damping
3
Ẍ + β (Ẋ + ε Ẋ ) + ω02 X = F(t), (3.7.8)
and, from equation (3.5.13), the mean-square response of system (3.7.2) is given by
+∞
H(ω)2 S dω = S0 .
1
E[ X ] = σX = RXX (0) =
2 2
0 (3.7.11)
2π −∞ 2βω02
ω02 2 − 1 βτ
RXX (τ ) = RX Ẋ (τ ) = − σ e 2 sinωD τ , (3.7.12)
ωD X
RXX (0) = −RẊ Ẋ (0) = −ω02 σX2 . (3.7.13)
Minimizing the mean-square error E[ E22 (X) ], one obtains, from equation (3.7.7),
S0
E[ Ẋ F(Ẋ) ] = β E[ Ẋ ]=
2
. (3.7.16)
2
Because
!
E[ Ẋ F(Ẋ) ] = E[ Ẋ · β (Ẋ + ε Ẋ ) ] = β E[ Ẋ ] + ε · E[ Ẋ4 ]
3 2
108
!
= β E[ Ẋ ] + ε · 3 E[ Ẋ 2 ] E[ Ẋ 2 ]
2
, (3.7.17)
in which equation (3.7.5) is used. Substituting equation (3.7.14) into (3.7.17) gives
S0
β ω02 σX2 + ε · 3 · ω02 σX2 · ω02 σX2 = , (3.7.18)
2
which yields the mean-square response
S0 1 S0 S0 3 S0
σX2 = ≈ (1−ε · 3ω0
2 2
σX ) = 1−ε . (3.7.19)
2βω02 1+ε · 3ω02 σX2 2βω02 2βω02 2β
☞ Although this result has been obtained for a Gaussian white noise excitation, it
can be expected to yield a reasonable approximate solution when the spectral
density of the excitation is slowly varying in the neighbourhood of ω02 and when
the damping in the system is light.
P(x)
where the sum is taken over all poles of that lie in the upper half-plane
Q(x)
U = z : Im (z) > 0 .
where
∞
R (ωm 2 −ω2 )(ω2 −ω2 ) + 4ζ ζ ω ω ω2
n m n m n
Imn = dω, (3.8.3)
−∞
2 −ω2 )2 + (2ζ ω ω)2 (ω2 −ω2 )2 + (2ζ ω ω)2
(ωm m m n n n
∞
I 2ω ζm ωm (ωn2 −ω2 ) − ζn ωn (ωm
2 −ω2 )
Imn = dω. (3.8.4)
−∞
2 −ω2 )2 + (2ζ ω ω)2 (ω2 −ω2 )2 + (2ζ ω ω)2
(ωm m m n n n
Because
2
F(ω) = (ωm −ω2 )2 + (2ζm ωm ω)2 = ω4 − 2ωm
2
(1−2ζm2 )ω2 + ωm
4
,
the roots of F(ω) = 0, solved by treating the equation as a quadratic equation in ω2, are
2 2ωm 2 (1−2ζ 2 )± 4ω4 (1−2ζ 2 )2 − 4 · 1 · ω4
m m m m 2
ω = = ωm 1−2ζm2 ± i 2ζm 1−ζm2
2
2
= ωm ( cosθ ± i sinθ), cosθ = 1−2ζm2 , sinθ = 2ζm 1−ζm2 .
θ 1+ cosθ 1 + (1−2ζm2 )
cos 2 = = = 1−ζm2 ,
2 2 2
θ 1− cosθ 1 − (1−2ζm2 )
sin 2 = = = ζm2 .
2 2 2
Hence, for both ImnR and I I , there are four of total eight poles that lie in the upper
mn
half-plane given by
ω = ωm ± 1−ζm2 + iζm , ωn ± 1−ζn2 + iζn . (3.8.5)
Using equation (3.8.1) and a symbolic computation software package, such as Maple,
R and I I can be easily evaluated to yield
the integrals Imn mn
I
Imn = 0, (3.8.6)
R 4π(ωm ζm + ωn ζn )
Imn =
(ωm −2ωm ωn +ωn ) + 4ωm ωn ζm ζn (ωm
4 2 2 4 2 +ω2 ) + 4ω2 ω2 (ζ 2 +ζ 2 )
n m n m n
110
1 ωn
ζm +
4π · ζ
ωm3 ωm n
=
ω2 ω4 ω ω2 ω2
1−2 2n + 4n + 4 n ζm ζn 1+ 2n + 4 2n (ζm2 +ζn2 )
ωm ωm ωm ωm ωm
4π ζm + rζn ωn
= · , r= . (3.8.7)
ωm (1−r ) + 4ζm ζn r(1+r 2 ) + 4(ζm2 +ζn2 )r 2
3 2 2 ωm
where
8 ζm ζn (ζm + rζn )r3/2 ωn
ρmn = , r= . (3.8.9)
(1−r 2 )2 + 4ζm ζn r(1+r 2 ) + 4(ζm2 +ζn2 )r 2 ωm
Maple Program
> restart:
I and I R , expressed in
Q is the denominator of the integrands of integrals Imn mn
terms of poles.
> Q:=(w-omega[m1])*(w-omega[m2])*(w-omega[m3])*(w-omega[m4])
*(w-omega[n1])*(w-omega[n2])*(w-omega[n3])*(w-omega[n4]);
Q := (w−ωm1 )(w−ωm2 )(w−ωm3 )(w−ωm4 )(w−ωn1 )(w−ωn2 )(w−ωn3 )(w−ωn4 )
I of the integrand of integral I I .
IP is the numerator Pmn mn
3.8 appendix − method of residue 111
> IP:=2*w*(zeta[m]*omega[m]*(omega[n]^2-w^2)-zeta[n]*omega[n]
*(omega[n]^2-w^2));
IP := 2w(ζm ωm (ωn2 −w2 ) − ζn ωn (ωn2 −w2 ))
> FI:=IP/Q; I .
FI is the integrand of integral Imn
2w(ζm ωm (ωn2 −w2 ) − ζn ωn (ωn2 −w2 ))
FI :=
(w−ωm1 )(w−ωm2 )(w−ωm3 )(w−ωm4 )(w−ωn1 )(w−ωn2 )(w−ωn3 )(w−ωn4 )
R of the integrand of integral I R .
RP is the numerator Pmn mn
> RP:=(omega[m]^2-w^2)*(omega[n]^2-w^2) +4*zeta[m]*zeta[n]*omega[m]
*omega[n]*w^2;
RP := (ωm
2 −w2 )(ω2 −w2 ) + 4ζ ζ ω ω w2
n m n m n
> FR:=RP/Q; R .
FR is the integrand of integral Imn
(ωm
2 −w2 )(ω2 −w2 ) + 4ζ ζ ω ω w2
n m n m n
FR :=
(w−ωm1 )(w−ωm2 )(w−ωm3 )(w−ωm4 )(w−ωn1 )(w−ωn2 )(w−ωn3 )(w−ωn4 )
I and I R .
Q is the denominator of the integrands of both integrals Imn mn
> Q:=((omega[m]^2-w^2)^2 +(2*zeta[m]*omega[m]*w)^2)*((omega[n]^2-w^2)^2
+(2*zeta[n]*omega[n]*w)^2);
Q := ((ωm
2 −w2 )2 + (2ζ ω w)2 )((ω2 −w2 )2 + (2ζ ω w)2 )
m m n n n
I , IR .
omega[m1],..., omega[n4] are eight poles of integrands of integrals Imn mn
> omega[m1]:=omega[m]*(sqrt(1-zeta[m]^2)+I*zeta[m]);
ωm1 := ωm ( 1−ζm2 + iζm );
> omega[m2]:=omega[m]*(sqrt(1-zeta[m]^2)-I*zeta[m]);
ωm2 := ωm ( 1−ζm2 − iζm );
> omega[m3]:=omega[m]*(-sqrt(1-zeta[m]^2)+I*zeta[m]);
ωm3 := ωm (− 1−ζm2 + iζm );
> omega[m4]:=omega[m]*(-sqrt(1-zeta[m]^2)-I*zeta[m]);
ωm4 := ωm (− 1−ζm2 − iζm );
> omega[n1]:=omega[n]*(sqrt(1-zeta[n]^2)+I*zeta[n]);
ωn1 := ωn ( 1−ζn2 + iζn );
> omega[n2]:=omega[n]*(sqrt(1-zeta[n]^2)-I*zeta[n]);
ωn2 := ωn ( 1−ζn2 − iζn );
> omega[n3]:=omega[n]*(-sqrt(1-zeta[n]^2)+I*zeta[n]);
ωn3 := ωn (− 1−ζn2 + iζn );
112
> omega[n4]:=omega[n]*(-sqrt(1-zeta[n]^2)-I*zeta[n]);
ωn4 := ωn (− 1−ζn2 − iζn );
> RIm1:=residue(FI,w=omega[m1]): Evaluate the residue of FI at ωm1 .
> RIm3:=residue(FI,w=omega[m3]): Evaluate the residue of FI at ωm3 .
> RIn1:=residue(FI,w=omega[n1]): Evaluate the residue of FI at ωn1 .
> RIn3:=residue(FI,w=omega[n3]): Evaluate the residue of FI at ωn3 .
I .
Use Theorem A.1 to determine the integral Imn
> INTI:=simplify(2*Pi*I*(RIm1+RIm3+RIn1+RIn3));
INTI :=0
> RRm1:=residue(FR,w=omega[m1]): Evaluate the residue of FR at ωm1 .
> RRm3:=residue(FR,w=omega[m3]): Evaluate the residue of FI at ωm3 .
> RRn1:=residue(FR,w=omega[n1]): Evaluate the residue of FI at ωn1 .
> RRn3:=residue(FR,w=omega[n3]): Evaluate the residue of FI at ωn3 .
R .
Use Theorem A.1 to determine the integral Imn
> INTR:=simplify(2*Pi*I*(RRm1+RRm3+RRn1+RRn3)):
❧ ❧
In this chapter, fundamentals of random processes and structural dynamics that are
the theoretical foundations for a number of topics in this book were presented.
❧ A random process is described by its various probability distribution functions. If
the probability distribution functions are invariant under a change of time origin,
the random process is stationary. A random process can be practically modelled as
stationary when the physical factors influencing it do not change with time.
3.8 appendix − method of residue 113