Nonlinearity in Structural Dynamics Chapter 08
Nonlinearity in Structural Dynamics Chapter 08
377
378 The Volterra series and higher-order frequency response functions
where
Z +1
y1 (t) = d h1 ( )x(t ) (8.4)
1
Z +1 Z +1
y2 (t) = d1 d2 h2 (1 ; 2 )x(t 1 )x(t 2 ) (8.5)
1 1
Z +1Z +1Z +1
y3 (t) = d1 d2 d3 h2 (1 ; 2 ; 3 )x(t 1 )x(t 2 )x(t 3 ):
1 1 1
(8.6)
The form of the general term is obvious from the previous statements. The
functions h1 ( ); h2 (1 ; 2 ); h3 (1 ; 2 ; 3 ); : : : hn (1 ; : : : ; n ); : : : are generaliza-
tions of the linear impulse response function and are usually referred to as Volterra
kernels. The use of the Volterra series in dynamics stems from the seminal paper
of Barrett [20], in which the series was applied to nonlinear differential equations
for the first time. One can think of the series as a generalization of the Taylor
series from functions to functionals. The expression (8.1) simply represents the
lowest-order truncation which is, of course, exact only for linear systems.
The derivation of the series is beyond the scope of this book, but heuristic
arguments can be found in [261, 25, 221]. Note that these kernels are not
forced to be symmetric in their arguments. In fact, any non-symmetric kernel
can be replaced by a symmetric version with impunity so that h 2 (1 ; 2 ) =
h2 (2 ; 1 ) etc. A formal proof is fairly straightforward; for simplicity, consider
the expression for y 2 (t):
Z +1 Z +1
y2 (t) = d1 d2 h2 (1 ; 2 )2 (1 ; 2 ; t) (8.7)
1 1
with the newly-defined
where
hsym 1
2 (1 ; 2 ) = 2 (h2 (1 ; 2 ) + h2 (2 ; 1 ))
asym
h2 (1 ; 2 ) = 12 (h2 (1 ; 2 ) h2 (2 ; 1 )): (8.10)
hasym asym
2 (v; w)2 (v; w; t) = h2 (w; v)2 (w; v; t) (8.12)
and the overall integral will vanish. This is purely because of the ‘contraction’ or
summation against the symmetric quantity 2 (1 ; 2 ; t). Because hasym2 makes
no contribution to the quantity y 2 (t) it may be disregarded and the kernel h 2 can
be assumed to be symmetric. Essentially, the h 2 picks up all the symmetries of
the quantity 2 . This argument may be generalized to the kernel h n (1 ; : : : ; n ).
In general, for a symmetric kernel, h sym n is obtained by summing all of
the possible permutations of the argument, weighted by an inverse factor which
counts the terms. The following section describes a method of extracting the
kernel transforms directly, which automatically selects the symmetric kernel. This
method will be adopted throughout the remainder of the book. For this reason,
the identifying label ‘sym’ will be omitted on the understanding that all kernels
and kernel transforms are symmetric. For information about other conventions
for kernels, mainly the triangular form, the reader can consult [217].
As previously stated, there exists a dual frequency-domain representation
for nonlinear systems. The higher-order FRFs or Volterra kernel transforms
Hn (!1 ; : : : ; !n ), n = 1; : : : ; 1 are defined as the multi-dimensional Fourier
transforms of the kernels, i.e.
Z +1 Z +1
Hn (!1 ; : : : ; !n ) = ::: d1 : : : dn hn (1 ; : : : ; n )
1 1
e i(Z!1 1 ++!Zn n) (8.13)
1 + 1 + 1
hn (1 ; : : : ; n ) = ::: d!1 : : : d!n Hn (!1 ; : : : ; !n )
(2)n 1 1
e+i(!1 1 ++!n n) : (8.14)
380 The Volterra series and higher-order frequency response functions
where
The first-order FRF of interest is for the process y ! F and this is termed the
dynamic stiffness. A simple calculation yields
F (!)
= H1 (!) = ic1 ! (8.20)
Y (! )
and it follows that the dynamic stiffness varies linearly with ! and the gradient is
the linear damping coefficient. The presence of the imaginary term simply shows
that the displacement and force are in quadrature (90 Æ out of phase).
The first task is to establish H1 (! ). The experimental procedure is a standard
stepped-sine test. The system is excited by a displacement signal, a sinusoid
Y cos(!t) at a given frequency and the amplitude and phase of the force response
F cos(!t ) are recorded. The gain and phase of H 1 (!) are simply F=Y and
as discussed in chapter 1.
In reality it is not quite as simple as this because the damper is nonlinear.
Assuming a polynomial expansion up to third order gives
f (t) = c1 y_ + c2 y_ 2 + c3 y_ 3 : (8.21)
H2 ( ; ) = c2 2 (8.23)
H3 ( ; ; ) = ic3 3 : (8.24)
Figure 8.1. Principal diagonals of the first-, second- and third-order composite HFRFs for
the shock absorber.
Figure 8.2. Nonlinear damping values for the shock absorber estimated from the principal
diagonals.
Figure 8.3. Principal quadrant of the second-order composite HFRF 2 (!1 ; !2 ) for the
shock absorber.
Figure 8.4. Principal quadrant of the third-order composite HFRF 3 (!1 ; !2 ; !3 ) for the
shock absorber.
and
H3 (!1 ; !2; !3 ) = c3 !1 !2 !3 : (8.26)
The second series of tests were with the coil spring in place. These produced
slightly more structured HFRFs due to the internal resonances of the spring.
Using basic elasticity theory, a dynamic stiffness FRF for the spring alone was
estimated and is shown in figure 8.5, the resonances are clear. A monofrequency
test gave the results shown in figure 8.6 for the diagonal composite HFRFs, the
polynomial rise from the damping is combined with the spring resonances. A
bifrequency test yielded the s2 and s3 shown in figures 8.7 and 8.8.
This section has shown how the HFRFs can be estimated using sine-testing
and also how they allow a parametric identification of the damping characteristics
of a shock absorber (although there are easier ways of obtaining estimates of the
c1 , c2 and c3 as discussed in the previous chapter). The figures showing the
386 The Volterra series and higher-order frequency response functions
Figure 8.5. Simulated FRF showing the coil spring’s first four resonant frequencies
calculated from spring theory.
x(t) = ei t : (8.27)
Figure 8.6. Principal diagonal of the first-, second- and third-order composite HFRFs for
the shock absorber and coil spring at an input voltage of 0.5 V.
1 +1
Z
Y (!) = H1 (!)2Æ(! )+ d! H (! ; ! ! 1 )
2 1 1 2 1
2Æ(!1 )2Æ(! !1 )
Z +1 Z +1
1
+ d! d! H (! ; ! ; ! !1 !2 )
(2)2 1 1 1 2 3 1 2
2Æ(!1 )2Æ(!2 )2Æ(! !1 !2 ) +
(8.30)
using the argument-changing property of the Æ -function and carrying out the
388 The Volterra series and higher-order frequency response functions
Figure 8.7. Principal quadrant of the second-order composite HFRF 2 (!1 ; !2 ) for the
shock absorber and coil spring at an input voltage of 0.5 V.
integrals gives
Figure 8.8. Principal quadrant of the third-order composite HFRF 3 (!1 ; !2 ; !3 ) for the
shock absorber and coil spring at an input voltage of 0.5 V.
x(t) = ei 1 t + ei 2 t (8.33)
f2Æ(! !1 1 ) + 2Æ(! !1 2 )g
1 +1
Z Z +1
+ d! d! H (! ; ! ; ! !1 !2 )
(2)2 1 1 1 2 3 1 2
f2Æ(!1 1 )+2Æ(!1 2 )gf2Æ(!2 1 )+2Æ(!2 2 )g
f2Æ(! !1 !2 1 ) + 2Æ(! !1 !2 2 )g + :
(8.35)
On taking the inverse Fourier transform, one obtains the response up to third
order:
The important thing to note here is that the amplitude of the component at
the sum frequency for the excitation, i.e. at 1 + 2 is twice the second-order
FRF H2 ( 1 ; 2 ). In fact, if a general periodic excitation is used, i.e.
x(t) = ei t + + ei n (8.38)
it can be shown that the amplitude of the output component at the frequency
1 + + n is n!Hn ( 1 ; : : : ; n ). This single fact is the basis of the harmonic
probing algorithm. In order to find the second-order FRF of a system for example,
one substitutes the expressions for the input (8.33) and general output (8.37) into
the system equation of motion and extracts the coefficient of e i( 1 + 2 )t ; this
yields an algebraic expression for H 2 .
The procedure is best illustrated by choosing an example. Consider the
continuous-time system
Dy + y + y2 = x(t) (8.39)
Harmonic probing of the Volterra series 391
and
y(t) = y1p (t) = H1 ( )ei t (8.41)
are substituted into the equation (8.39), the result being
and
are used. Note that in passing from the general output (8.37) to the probing
expression (8.46), all second-order terms except that at the sum frequency have
been deleted. This is a very useful simplification and is allowed because no
combination of the missing terms can produce a component at the sum frequency
and therefore they cannot appear in the final expression for H 2 . Substituting
(8.45) and (8.46) into (8.39), and extracting the coefficients of e i( 1 + 2 )t yields
So that
H1 ( 1 )H1 ( 2 )
H2 ( 1 ; 2 ) = = H1 ( 1 )H1 ( 2 )H1 ( 1 + 2 )
i( 1 + 2 ) + 1
1
= (8.48)
(i 1 + 1)(i 2 + 1)(i[ 1 + 2 ] + 1)
on using the previously obtained expression for H 1 .
The next example is a little more interesting. Consider the asymmetric
Duffing equation
Note that the constant k 2 multiplies the whole expression for H 2 , so that if
the square-law term is absent from the equation of motion, H 2 vanishes. This
reflects a quite general property of the Volterra series; if all nonlinear terms in
the equation of motion for a system are odd powers of x or y , then the associated
Volterra series has no even-order kernels. As a consequence it will possess no
even-order kernel transforms.
In order to obtain H 3 , the required probing expressions are
and
4y + y + y2 = x(t): (8.55)
interval for a discrete-time system is scaled to unity. This yields a unit sampling
frequency and Nyquist frequency of 0.5.) In the usual notation for difference
equations, (8.55) becomes
yi 1 + yi + yi2 = xi : (8.56)
However, the form containing 4 allows the most direct comparison with
the continuous-time case. It is clear from the previous argument that the only
differences for harmonic probing of discrete-time systems will be generated by
the fact that the operator 4 has a different action on functions e i!t to the operator
D. This action is very simple to compute, as shown in chapter 1 3,
4ei!t = ei!(t 1) = e i! ei!t : (8.57)
It is now clear that one can carry out the harmonic probing algorithm for
(8.55) exactly as for the continuous-time (8.39); the only difference will be that
the 4 operator will generate a multiplier e i! wherever D generated a factor i! .
As a consequence H 1 and H2 for (8.55) are easily computed.
1
H1 (!) =
e i! + 1
(8.58)
H1 (!1 )H1 (!2 )
H2 (!1 ; !2) = = H1 (!1 )H1 (!1 )H1 (!1 + !2 ):
e i(! +! ) + 1
(8.59)
1 2
y y
y_i i i 1 (8.61)
t
yi+1 2yi + yi 1
yi
t2
(8.62)
2m + ct + kt2
m
yi = yi 1 y
m + ct m + ct i 2
k2 t2 k3 t2 t2
y 2 3
y + x :
m + ct i 1 m + ct i 1 m + ct i 1
(8.63)
Figure 8.9. Comparison between simulated Duffing oscillator data and the prediction by a
NARX model.
(a)
Gain (dB)
Frequency (Hz)
Phase (degrees)
(b)
Normalised Frequency
(c)
Gain (dB)
Normalised Frequency
(d)
Phase (degrees)
Figure 8.10. H1 (f ) for the Duffing oscillator system: (a) exact magnitude; (b) exact
phase; (c) NARX model magnitude; (d) NARX model phase.
Validation and interpretation of the higher-order FRFs 397
(a) (b)
(d)
(c)
Figure 8.11. H2 (f1 ; f2 ) surface for the Duffing oscillator system: (a) exact magnitude;
(b) exact phase; (c) NARX model magnitude; (d) NARX model phase.
The comparison between the exact H 2 and that from the NARMAX model
is given in figure 8.11. The same comparison using the contour maps for the
functions is shown in figure 8.12; again the agreement is very good. Note that
because H2 contains factors H1 (2f1 ) and H2 (2f2 ) it would be meaningless
to plot it outside the ranges corresponding to f 1 100; f2 100. Further,
H2 also contains a factor H1 (2(f1 + f2 )) so that the plots should not extend
past the area specified by f 1 + f2 100. Rather than plot irregularly shaped
regions, the H2 figures presented in this book include information beyond this last
bound, which is indicated by the full line in the model contour maps in figure 8.12;
information presented outside this region on any H 2 plot should not be regarded
as meaningful.
The comparison between the exact H 3 and model H 3 is given in figure 8.13,
and in contour map form in figure 8.14. Unfortunately, the whole H 3 surface
cannot be plotted as it exists as a three-dimensional manifold embedded in a
four-dimensional space over the (! 1 ; !2 ; !3 )-‘plane’. However, one can plot
two-dimensional submanifolds of H 3 , and this is the approach which is usually
adopted. Figures 8.13 and 8.14 show H 3 (!1 ; !2 ; !1 ) plotted over the (! 1 ; !2 )-
plane. The region of validity of the H 3 surface is a little more complicated in
398 The Volterra series and higher-order frequency response functions
Gain (dB)
(a)
f1
f2
Phase
(b)
f1
f2
Gain (dB)
(c)
f1
f2
Phase
(d)
f1
f2
Figure 8.12. H2 (f1 ; f2 ) contours for the Duffing oscillator system: (a) exact magnitude;
(b) exact phase; (c) NARX model magnitude; (d) NARX model phase.
Validation and interpretation of the higher-order FRFs 399
(a) (b)
(d)
(c)
Figure 8.13. H3 (f1 ; f2 ; f1 ) surface for the Duffing oscillator system: (a) exact magnitude;
(b) exact phase; (c) NARX model magnitude; (d) NARX model phase.
this situation. In all cases, agreement between the exact H n and those obtained
from the NARMAX model appears impressive. For a less passive comparison,
figure 8.15 shows the gain and phase of the output components y 1 , y2 and y3
obtained from the systems defined by the exact and model FRFs when excited by
a unit sinusoid at various frequencies. Again, agreement looks excellent. Note
that the plot for second harmonic in figure 8.15 contains a peak at f r =2. This
is due to the fact that the diagonal HFRF contains a factor H 1 (2! ) as shown by
equation (8.51).
Having established that a NARX model can yield good representations of
the FRFs from a continuous system, the next question which must be addressed
concerns the correspondence between frequency-domain representations of
different yet exactly equivalent NARX models. (Non-uniqueness is actually a
problem with most methods of modelling, it is not specific to NARX). Suppose
one has obtained as an accurate discretization of a continuous system, the ARX
model,
yi = a1 yi 1 + a2 yi 2 + b1 xi 1 : (8.65)
As this expression holds for all values of i (away from the initial points), it
400 The Volterra series and higher-order frequency response functions
Gain (dB)
(a)
f1
Phase
f2
(b)
f1
f2
Gain (dB)
(c)
f1
f2
Phase
(d)
f1
f2
Figure 8.14. H3 (f1 ; f2 ; f1 ) contours for the Duffing oscillator system: (a) exact
magnitude; (b) exact phase; (c) NARX model magnitude; (d) NARX model phase.
Validation and interpretation of the higher-order FRFs 401
Frequency (Hz)
(a)
Amplitude (dB)
Phase (deg)
(b)
Frequency (Hz)
Normalised Frequency
(c)
Amplitude (dB)
Phase (deg)
(d)
Normalised Frequency
Figure 8.15. H1 , H2 and H3 components for the Duffing oscillator response excited by
a unit sinusoid: (a) exact magnitude; (b) exact phase; (c) NARX model magnitude; (d)
NARX model phase.
402 The Volterra series and higher-order frequency response functions
yi 1 = a1 yi 2 + a2 yi 3 + b1 xi 2 (8.66)
which is exactly equivalent to (8.65) yet contains different terms. This type of
ambiguity will occur for any system which regresses the present output onto
past values of output. It is a reflection of a type of ambiguity for continuous-
time systems; one can always differentiate the equation of motion to obtain a
completely equivalent system. The only thing which changes is the set of objects
for which initial conditions are required. Harmonic probing of (8.65) yields (in
symbolic notation where 4 = e i! )
b1 4
H1(8:65) =
1 a1 4 a2 42
(8.68)
b1 4 +a1 b1 42
H1(8:67) = :
1 (a21 + a2 ) 42 a1 a2 43
(8.69)
b1 4 (a1 4 +1) b1 4
H1(8:67) = = = H1(8:65):
(a1 4 +1)(1 a1 4 a24 ) 1 a1 4 a2 42
2
(8.70)
The final type of non-uniqueness is generated by the fact that NARMAX
models can be approximately equivalent. As an illustration consider the simple
system
yi = yi 1 + xi 1 : (8.71)
If is small, a simple application of the binomial theorem gives
(1 4)yi = xi 1 =) yi = (1 4) 1xi 1
=) yi = (1 + 4)xi 1 + O( 2 ) (8.72)
So the system
yi = xi 1 + xi 2 (8.73)
is equivalent to the system in (8.71) up to O( 2 ). Now, harmonic probing of
system (8.71) yields the FRF
1
H1(8:71)(!) =
e i!
(8.74)
1
Validation and interpretation of the higher-order FRFs 403
yi = xi 1 + xi 2 + + n 1 xi n (8.76)
The power of the NARX and higher-order FRF approaches can be demonstrated
by the following example used in chapter 6 where force and velocity data were
obtained from a circular cylinder placed in a planar oscillating fluid flow in a large
U-tube [199]. The standard means of predicting forces on cylinders used by the
offshore industry is to use Morison’s equation (6.121) which expresses the force
as a simple nonlinear function of the instantaneous flow velocity and acceleration.
For one particular frequency of flow oscillation, Morison’s equation gave the force
prediction shown in figure 8.16(a) compared with the measured force. Morison’s
equation is inadequate at representing the higher-frequency components of the
FRFs and Hilbert transforms: sine excitation 405
(a)
(b)
Figure 8.16. Morison equation fit to experimental U-tube data: (a) model-predicted
output; (b) correlation tests.
(a)
(b)
Figure 8.17. NARX model fit to experimental U-tube data: (a) model-predicted output;
(b) correlation tests.
+ AB 2 H3 ( a ; b ; b)ei( a +2 b )t + B 3 H3 ( b ; b ; b )e3i b t +
(8.79)
408 The Volterra series and higher-order frequency response functions
(a) (b)
Gain (dB)
(c)
f1
f2
Phase
(d)
f1
f2
Figure 8.18. H3 (f1 ; f2 ; f1 ) from NARX fit to U-tube data: (a) magnitude; (b) phase; (c)
magnitude contours; (d) phase contours.
to third order.
Now, for the response to a cosinusoid
X X X2
y(t) = H1 ( )ei t + H1 ( )e i t + H2 ( ; )e2i t
2 2 4
X2 X2 X3
+ H2 ( ; ) + H2 ( ; )e 2i t + H3 ( ; ; )e3i t
2 4 8
3X 3 3 X 3
+ H ( ; ; )ei t + H ( ; ; )e i t
8 3 8 3
X3
+ H3 ( ; ; )e 3i t + : (8.81)
8
Making use of the reflection properties H 1 ( ) = H ( ) etc, and applying
de Moivre’s theorem in the form
yields
which shows again that the response contains all odd and even harmonics. The
component of the response at the forcing frequency is
3X 2
s ( ) = H1 ( ) + H ( ; ; ) +
4 3
(8.85)
or
3X 2 5X 4
s ( ) = H1 ( ) + H( ; ; )+ H( ; ; ; ; )+
4 3 8 5
(8.86)
to the next highest order. Again, it is useful to take the Duffing oscillator (8.49)
as an example. Equation (8.54) with k 2 = 0, gives
H3 (!1 ; !2 ; !3) = k3 H1 (!1 )H1 (!2 )H1 (!3 )H1 (!1 + !2 + !3 ) (8.87)
410 The Volterra series and higher-order frequency response functions
1
H5 (!1 ; !2 ; !3 ; !4 ; !5) = H1 (!1 + !2 + !3 + !4 + !5 )
5!
( 3!3!k3 fH3(!1 ; !2 ; !3)H1 (!4 )H1 (!5 )
+ H3 (!1 ; !2 ; !4)H1 (!3 )H1 (!5 )
+ H3 (!1 ; !3 ; !4)H1 (!2 )H1 (!5 ) + H3 (!2 ; !3 ; !4 )H1 (!1 )H1 (!5 )
+ H3 (!1 ; !2 ; !5)H1 (!3 )H1 (!4 ) + H3 (!1 ; !3 ; !5 )H1 (!2 )H1 (!4 )
+ H3 (!2 ; !3 ; !5)H1 (!1 )H1 (!4 ) + H3 (!1 ; !4 ; !5 )H1 (!2 )H1 (!3 )
+ H3 (!2 ; !4 ; !5)H1 (!1 )H1 (!3 ) + H3 (!3 ; !4 ; !5 )H1 (!1 )H1 (!2 )g)
(8.89)
and thence
H5 (!; !; !; !; !) = 103 k32 (3H1 (!)4 H1 (!)3 + 6H1 (!)5 H1(!)2
+ H1 (!)4 H1 (!)2 H1 (3!)): (8.90)
Substituting (8.90) and (8.88) into (8.86) gives the FRF up to O(X 4 )
3X 2 3X 4 2
s ( ) = H1 ( ) k3 H1 (!)3 H1 (!) + k (3H (!)4 H1 (!)3
4 16 3 1
+ 6H1 (!)5 H1 (!)2 + H1 (!)4 H1 (!)2 H1 (3!)) + O(X 6 ): (8.91)
(Amongst other places, this equation has been discussed in [236], where it
was used to draw some conclusions regarding the amplitude dependence of the
stepped-sine composite FRF.)
In order to illustrate these expressions, the system
was chosen. Figure 8.19 shows the FRF magnitude plots obtained from (8.91)
for X = 0:01 (near linear), X = 0:5 and X = 0:75. At the higher amplitudes,
the expected FRF distortion is obtained, namely the resonant frequency shifts
up and the magnitude at resonance falls. Figure 8.20 shows the corresponding
Nyquist plots. (Note the unequal scales in the Real and Imaginary axes; the plots
are effectively circular.) Figure 8.21 shows the O(X 4 ) FRF compared with the
‘exact’ result from numerical simulation. There is a small degree of error near
resonance which is the result of premature truncation of the Volterra series.
FRFs and Hilbert transforms: sine excitation 411
0.00050
0.00040
FRF magnitude
0.00030
0.00020
0.00010
0.00000
70.0 80.0 90.0 100.0 110.0 120.0 130.0
Frequency (rad/s)
Figure 8.19. Distortion in the magnitude plot of s1 (! ) computed from the Volterra series
for different levels of excitation.
and the distortion suffered in passing from the FRF to the Hilbert transform is
given by the simple relation
-0.00020
FRF Imaginary Part
-0.00040
-0.00060
-0.00040 -0.00020 0.00000 0.00020 0.00040
FRF Real Part
Figure 8.20. Distortion in the Nyquist plot of s1 (! ) computed from the Volterra series
for different levels of excitation. (Note that the Real and Imaginary axes do not have equal
scales.)
This section presents a technique which allows the Hilbert transform
distortion to be derived term by term from a Volterra series expansion of the
system FRF, the expansion parameter being X , the magnitude of the applied
sinusoidal excitation. It is illustrated on the Duffing oscillator (8.49), and the
basic form of FRF used is the O(X 4 ) approximation given in (8.91). If the FRF
is known, the Hilbert transform follows from the distortion (8.95). In order to
obtain the distortion, the pole–zero form of the FRF is needed.
0.00050
0.00040
FRF Magnitude
0.00030
0.00000
70.0 80.0 90.0 100.0 110.0 120.0 130.0
Frequency (rad/s)
Figure 8.21. Comparison between FRFs s1 (! ) computed from the Volterra series and
from numerical simulation.
and this is the required ‘pole–zero’ expansion. Note that p 1 and p2 are both in
the upper half-plane so H 1 (! ) = H1+ (! ) and the Hilbert transform is therefore
the identity on H 1 (! ) as required. However, the expression for s (! ) in (8.91)
contains terms of the form H 1 (! ) with poles p1 and p2 ; these are in the lower
half-plane and are the cause of the Hilbert transform distortion for s (! ). In
pole–zero form (8.91) becomes
1
s (! ) =
m(! p1 )(!
p2 )
3X 2 1
k3 4
4 m (! p1 ) (! p2 )3 (! p1 )(! p2 )
3
3 2 4 3
k3 X
16 m (! p1 ) (! p2 )4 (! p1 )3 (! p2 )3
7 4
6
+ 7
m (! p1 )5 (! p2 )5 (! p1 )2 (! p2 )2
1
+ 7
m (! p1 )4 (! p2 )4 (! p1 )2 (! p2 )2 (3! p1 )(3! p2 )
(8.98)
A1 A2 A3 A4
+ 2 + 3 +
(! p1 ) (! p1 ) (! p1 ) (! p1 )
B1 B2 B3 B4
+ + + +
(! p2 ) (! p2 )2 (! p2 )3 (! p2 )
(8.99)
where
1
A4 =
( p1 + p1 ) ( p1 + p2 )3 ( p1 + p2 )
3 (8.100)
1
A3 =
( p1 + p1 )( p1 + p2 )3 (p1 p2 )
(8.101)
and finally
N1
A1 = (8.103)
D1
with
3X 2k3
s (!) =
2m4(p1 p1 )3 (p2 p1 )3 (p1 p2 )(! p1 )
+ (p1 ! p2 ):
(8.106)
FRFs and Hilbert transforms: sine excitation 415
-0.00020
-0.00040
-0.00060
-0.00040 -0.00020 0.00000 0.00020 0.00040
FRF Real Part
Figure 8.22. Comparison between the numerical estimate of the Hilbert transform and the
O(X 2 ) Volterra series estimate for the Duffing oscillator under sine excitation. (Note that
the Real and Imaginary axes do not have equal scales.)
0.00000
FRF Imaginary Part
-0.00020
-0.00040
-0.00060
-0.00040 -0.00020 0.00000 0.00020 0.00040
FRF Real Part
Figure 8.23. Comparison between the numerical estimate of the Hilbert transform and the
O(X 4 ) Volterra series estimate for the Duffing oscillator under sine excitation. (Note that
the Real and Imaginary axes do not have equal scales.)
in many respects, there are important differences. This section is this book’s
only real foray into the realm of random vibration. If the reader would like to
study the subject in more depth, [198] is an excellent example of an introductory
textbook. A considerably more advanced treatment can be found in [52], which
treats nonlinear random vibration amongst other topics.
There have been a number of related calculations over the years. The
simplest method of approximating an FRF for a nonlinear system is based on
equivalent linearization [54]. This approach estimates the parameters of the linear
system which is closest (in a statistical sense) to the original nonlinear system.
The FRF of the linearized system is computed. In [75], statistical linearization
was combined with perturbation analysis [68], in order to calculate the spectral
response of a Duffing oscillator to white noise excitation. (This is equivalent
to the FRF calculation up to a multiplicative constant.) It was shown that the
FRF exhibits a secondary peak at three times the natural frequency, a result
which is unavailable from statistical linearization alone. An approach based on
perturbation theory alone is described in [147] and the calculation is carried to
first order in the perturbation parameter. A number of studies of spectra have
appeared based on the use of the Fokker–Planck–Kolmogorov equation (FPK)
[55, 15, 137, 138, 284]. The latter two references actually examine the Duffing
oscillator system which is studied in the current work. Good representations of
the spectra were obtained; however, to the order of approximation pursued, the
approach was unable to explain the presence of the secondary peak described
418 The Volterra series and higher-order frequency response functions
From (8.124) the expected value of the product of inputs, i.e. the fourth-order
moment of the input, reduces to the following product of second-order moments,
Using this equation and taking advantage of the symmetry of the Volterra
kernels leads to
Z +1 Z +1 Z +1
y3 x ( ) = 3 d1 d2 d3 h3 (1 ; 2 ; 3 )
1 1 1
E [x(t 1 )x(t 2 )]E [x(t 3 )x(t )]
Z +1 Z +1 Z +1
=3 d1 d2 d3 h3 (1 ; 2 ; 3 )
1 1 1
xx (2 1 )xx ( 3 ): (8.127)
3Sxx(!) +1
Z
Sy3 x (!) = d!1 H3 (!1 ; !1; !)Sxx(!1 ): (8.128)
2 1
This result is already available in the literature [25]. Its presence here is
justified by the fact that the derivation of the general term is a simple modification.
FRFs and Hilbert transforms: random excitation 421
(2n)!Sxx(!) +1
Z Z +1
Sy2n 1 x (!) = : : : d!1 : : : d!n 1
n! 2n (2)n 1 1 1
H2n 1 (!1 ; !1 ; : : : ; !n 1 ; !n 1; !)
Sxx(!1 ) : : : Sxx (!n 1 ): (8.129)
Now, given that the input autospectrum is constant over all frequencies for
a Gaussian white noise input (i.e. S xx (! ) = P ), the composite FRF for random
excitation follows. Substituting (8.129) into (8.116) gives
=1
nX
(2n)!P n 1 +1
Z Z +1
r (!) = n n 1 1 : : : 1 d!1 : : : d!n 1
n=1 n! 2 (2 )
H2n 1 (!1 ; !1; : : : ; !n 1 ; !n 1; !): (8.130)
This equation will be used to analyse the effect of a Gaussian white noise input
on the SDOF Duffing oscillator system.
The first term of this equation needs no further work but the others require
expressions for the HFRF terms as functions of the H 1 s and k3 . The results for
H3 and H5 are given in (8.87) and (8.89) respectively, the specific forms needed
for (8.131) are
+ H1 ( !1 + !2 + !) + H1 ( !1 !2 + !)g
3k2
= 3 H1 (!)2 jH1 (!1 )j2 jH1 (!2 )j2 f2H1 (!) + H1 (!1 ) + H1 ( !1)
10
+ H1 (!2 ) + H1 ( !2 ) + H1 (!1 + !2 + !) + H1 (!1 !2 + !)
+ H1 ( !1 + !2 + !) + H1 ( !1 !2 + !)g (8.133)
S 3 x (!)
So only one integral needs to be evaluated for Syxx (!) compared to nine for
Sy5 x (!)
Sxx (!) .
S 3 x (!)
Substituting (8.132) into the Syxx (!) term of (8.131) gives
3P k3H1 (!)2 +1
Z
Sy3 x (!)
= d!1 jH1 (!1 )j2 : (8.134)
Sxx (!) 2 1
This integral may be found in standard tables of integrals used for the
calculation of mean-square response, e.g. [198]. However, the analysis is
instructive and it will allow the definition of notation for the integrals which
follow.
Consider the common expression for the linear FRF (8.97). In terms of this,
2
the integral in (8.134), jH 1 (!1 )j may be written
1
jH1 (!1 )j2 = m2 (! p1 )(!1 p2 )(!1 p1 )(!1 p2 )
(8.135)
1
where p1 (respectively p2 ) is the complex conjugate of p 1 (respectively p2 ).
The integral is straightforwardly evaluated using the calculus of residues 5 . The
approach is well known and details can be found in numerous textbooks, e.g. [6].
The appropriate contour is given in figure 8.24. (In the calculation, the radius of
the semicircle is allowed to go to infinity.) The result of the calculation is the
standard
Z +1
d! jH (!)j2 = = :
2m2 !n (!d2 + 2 !n2 ) ck1
(8.136)
1
Substituting this expression into (8.134) gives
3 i ζ ωn
2 i ζ ωn
p p1
2
i ζ ωn α1
α2
-3 ω d -2 ω d - ωd ωd 2 ωd 3 ωd
- i ζ ωn
p* p1*
2
-2 i ζ ω n
-3 i ζ ω n
Figure 8.24. The pole structure of jH1 (!1 )j2 with the integration contour shown closed in
the upper-half of the !1 -plane.
S 5 x (!)
Substituting (8.133) into the Syxx (!) term of (8.131) gives the integrals
Z +1 Z +1
Sy5 x(!) 9P 2 k32 H1 (!)2
= 2 H1 (! ) d!1 d!2 jH1 (!1 )j2 jH1 (!2 )j2
Sxx(!) 8 2 1 1
Z +1 Z +1
+ d!1 d!2 H1 (!1 )jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ d!1 d!2 H1 ( !1 )jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ d!1 d!2 H1 (!2 )jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ d!1 d!2 H1 ( !2 )jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ d!1 d!2 H1 (!1 + !2 + !)jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ d!1 d!2 H1 (!1 !2 + !)jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ d!1 d!2 H1 ( !1 + !2 + !)jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
+ 2
d!1 d!2 H1 ( !1 !2 + !)jH1 (!1 )j jH1 (!2 )j : 2
1 1
(8.138)
424 The Volterra series and higher-order frequency response functions
3 i ζ ωn
2 i ζ ωn
p p1
2
i ζ ωn Double pole
-3 ω d -2 ω d - ωd ωd 2 ωd 3 ωd
α2 α1
- i ζ ωn
p* p1*
2
-2 i ζ ω n
-3 i ζ ω n
Figure 8.25. The pole structure of H1 (!1 ) H1 (!1 ) 2 with the integration contour shown
j j
closed in the lower-half of the !1 -plane.
S 3 x (!)
and this has a similar pole structure to the Syxx (!) integrand, except the poles at
p1 and p2 are now double poles (figure 8.25). The contour should be closed in the
lower half of the ! 1 -plane as there are only simple poles there. The result is
Z +1
Re d!1 H1 (!1 )jH1 (!1 )j2 = 2 =
3 2 2 2 2ck12
(8.145)
1 4m !n (!d + !n )
and this combined with (8.136) yields
Z +1 Z +1 2
Re d!1 d!2 H1 (!1 )jH1 (!1 )j2 jH1 (!2 )j2 = :
2c2 k13
(8.146)
1 1
S 5 x (!)
The third and final integral of the Syxx (!) expression is slightly more
complicated as the integrand does not factorize.
Z +1 Z +1
I (!) = d!1 d!2 H1 (!1 + !2 + !)jH1 (!1 )j2 jH1 (!2 )j2
1 1
Z +1 Z +1
= d!2 jH1 (!2 )j 2 d!1 H1 (!1 + !2 + !)jH1 (!1 )j2 :
1 1
(8.147)
The second integral must be solved first. In terms of poles,
H1 (!1 + !2 + !)jH1 (!1 )j2
1
= 3
m (!1 p1 )(!1 p2 )(!1 p1 )(!1 p2 )(!1 q1 )(!1 q2 )
(8.148)
426 The Volterra series and higher-order frequency response functions
3 i ζ ωn
2 i ζ ωn
q2 p2 q1 p1
i ζ ωn
ω + ω2 ω + ω2
-3 ω d -2 ω d - ωd ωd 2 ωd 3 ωd
α2 α1
- i ζ ωn
p* p1*
2
-2 i ζ ω n
-3 i ζ ω n
Figure 8.26. The pole structure of H1 (!1 + !2 + !3 ) H1 (!1 ) 2 with the integration
j j
contour shown closed in the lower-half of the !1 -plane.
where p1 , p2 , p1 and p2 are the same as before and q1 and q2 are the poles of
H1 (!1 + !2 + !)
q1 = p1 ! !2 = (!d ! !2 ) + i!n
q2 = p2 ! !2 = ( !d ! !2 ) + i!n : (8.149)
For simplicity, the contour is again closed on the lower half of the ! 1 -plane
(figure 8.26). The result is
Z +1
d!1 H1 (!1 + !2 + !)jH1 (!1 )j2
1
= [ (! + !2 4i!n )][2m3 !n (!d2 + 2 !n2 )(! + !2 2i!n )
(! + !2 + 2!d 2i!n )(! + !2 2!d 2i!n )] 1
= [ (! + !2 4i!n )][mck1 (! + !2 2i!n )
(! + !2 + 2!d 2i!n )(! + !2 2!d 2i!n )] 1 : (8.150)
The expression above is then substituted into (8.147) and the integral over
!2 evaluated. The integrand is this expression multiplied by jH 1 (!2 )j2 . In terms
of poles it is
(! + !2 4i!n )
3
m ck1 (!2 p1 )(!2 p2 )(!2 p1 )(!2 p2 )(!2 q1 )(!2 q2 )(!2 r)
(8.151)
where p1 , p2 , p1 and p2 are as before and
q1 = ! + 2!d + 2i!n ; q2 = ! 2!d + 2i!n
FRFs and Hilbert transforms: random excitation 427
3 i ζ ωn
q2 q1
r 2 i ζ ωn
ω ω p1 ω
p2
i ζ ωn
-3 ω d -2 ω d - ωd ωd 2 ωd 3 ωd
α2 α1
- i ζ ωn
p* p1*
2
-2 i ζ ω n
-3 i ζ ω n
Figure 8.27. The pole structure of equation (8.151) with the integration contour shown
closed in the lower-half of the !1 -plane.
r = ! + 2i!n : (8.152)
The contour is again closed on the lower half of the ! 2 -plane (figure 8.27).
Finally
Substituting (8.153), (8.146) and (8.142) into (8.141) gives the overall expression
S 5 x (!)
for Syxx (!) :
Sy5 x (!) 9P 2k32 H1 (!)3 9P 2 k32 H1 (!)2 9P 2 k32 H1 (!)2
= + + I (! )
4c2 k12 4c2 k13 22
(8.154)
Sxx (!)
for the classical Duffing oscillator. This equation shows that the first two terms
do not affect the position of the poles of the linear system. The first term does,
however, introduce triple poles at the position of the linear system poles. The term
of greatest interest, though, is the final one which has introduced four new poles
at
The pole structure to this order is shown in figure 8.28. The poles at 3! d +3i!n
explain the secondary observed peak at three times the resonant frequency in the
output spectra of nonlinear oscillators [285].
428 The Volterra series and higher-order frequency response functions
3 i ζ ωn
2 i ζ ωn
i ζ ωn
Figure 8.28. The pole structure of the first three terms of r (! ) for the classical Duffing
oscillator.
0.0004
FRF Magnitude
0.0003
0.0002
0.0001
0.0000
50 60 70 80 90 100 110 120 130 140 150
Circular Frequency (rad/s)
Figure 8.29. Composite FRF r (! ) to order O(P 2 ) for the classical Duffing oscillator.
Combining (8.154) and (8.137) into (8.131) yields an expression for the
composite FRF r (! ) up to O(P 2 ). The magnitude of the composite FRF is
plotted in figure 8.29 for values of P equal to 0 (linear system), 0.01 and 0.02.
The Duffing oscillator parameters are m = 1, c = 20, k 1 = 104 and k3 = 5 109.
FRFs and Hilbert transforms: random excitation 429
The evaluation of these integrals was carried out as before. However, the
results will not be included here. The important point is whether the calculation
introduces new poles. By extrapolating from the poles in the first three terms of
r (!), a fourth term might be expected to introduce poles at
!d + 5i!n ; !d + 5i!n ; 3!d + 5i!n ;
3!d + 5i!n ; 5!d + 5i!n ; 5!d + 5i!n: (8.157)
However, after evaluating the integrals in (8.156), again by contour integration,
it was found that no new poles arose. Instead, three of the integrals resulted
in simple poles at the locations given in equation (8.155), whilst another three
integrals resulted in double poles at these locations.
Due to the rapidly increasing level of difficulty associated with the addition
S 9 x (!)
of further terms to r (! ) it was not possible to completely examine the Syxx (!)
term. It was possible, however, to consider one integral which would be included
S 9 x (!)
in the overall expression. The expression for Syxx (!) is given by
Sy9 x (!) 945P 4 +1 +1 +1 +1
Z Z Z Z
= d! d! d! d!
Sxx(!) (2)4 1 1 1 1 1 2 3 4
H9 (!1 ; !1; !2 ; !2; !3 ; !3; !4 ; !4 ; !): (8.158)
Validity of the Volterra series 431
This integral was evaluated as before and was found to have triple poles at
the locations given in equation (8.155), simple poles at the locations given in
equation (8.157) and also at the following locations
Although it is possible that these contributions cancel when combined with other
S 9 x (!)
integrals from Syxx (!) , it can be conjectured that including all further terms would
result in FRF poles for this system being witnessed at all locations a! d + bi!n
where a b are both odd integers. Note that this implies the existence of an
infinite sequence of FRF peaks, each one associated with an odd multiple of the
natural frequency.
The restriction of a and b to the odd integers might be expected from the fact
that only odd-order HFRFs exist for systems which only contain odd polynomial
nonlinearities. In that case the introduction of an even nonlinearity results in both
odd- and even-order HFRFs. This is discussed in appendix K where a Duffing
oscillator with additional quadratic spring stiffness is considered. Appendix K
also shows an extension of the analysis to a simple MDOF system.
An interesting feature of this analysis is that the multiplicity of the poles
increases with order P . The implication is that the poles will become isolated
essential singularities in the limit. This implies rather interesting behaviour in the
!-plane near the poles as Picard’s theorem [6] asserts that the FRF will take all
complex values in any neighbourhood of the poles regardless of the size of the
neighbourhood. This behaviour need not, of course, be visible from the real line.
For the random case,Pthey established that if x(t) is stationary with bounded
1
statistical moments then i=1 yi converges in the mean if
n
X
lim ak < 1
n!1
(8.165)
k=1
and
lim a = 0:
k!1 k
(8.166)
oscillator using the criteria developed by Barrett [21] 6 . The first step is to convert
the equation of motion (8.49) with k 2 = 0, to the normalized form
0
y0 + 2 y_ 0 + y0 + y 3 = x0 (t0 ) (8.167)
and has the usual definition. Once in this coordinate system, convergence of the
Volterra series is ensured as long as [21]
yb
mb
fd ks
yt
mt
kt
y0
where mt , kt and yt are the mass, stiffness and displacement of the tyre—the
unsprung mass. m b and yb are the mass and displacement of the body or the
sprung mass (this is usually taken as one-quarter of the total car body mass).
fd is the characteristic of the nonlinear damper and k s is the (linear) stiffness
Harmonic probing for a MDOF system 435
of the suspension. A cubic characteristic will be assumed for the damper, i.e.
fd (z ) = c1 z + c2 z 2 + c3 z 3 . y0 is the displacement at the road surface and acts
here as the excitation.
Each of the processes y 0 ! yt and y0 ! yb has its own Volterra series
and its own HFRFs and it will be shown later that the responses will depend on
the kernel transforms from both series. The notation is defined by
1 1
Z
Yt (!) = H1t (!)Y0 (!) + d! H t (! ; ! !1 )Y0 (!1 )Y0 (! !1 )
(2) 1 1 2 1
Z 1Z 1
1
+ d! d! H t (! ; ! ; ! !1 !2 )
(2)2 1 1 1 2 3 1 2
Y0 (!1 )Y0 (!2 )Y0 (!Z !1 !2 ) + (8.175)
1 1
Yb (!) = H1b(!)Y0 (!) + d! H b (! ; ! !1 )Y0 (!1 )Y0 (! !1 )
(2) 1 1 2 1
Z 1Z 1
1
+ d! d! H b (! ; ! ; ! !1 !2 )
(2)2 1 1 1 2 3 1 2
Y0 (!1 )Y0 (!2 )Y0 (! !1 !2 ) + : (8.176)
y0p = ei t (8.177)
ytp = H1t ( )ei t + (8.178)
and
ybp = H1b( )ei t + (8.179)
are used. These expressions are substituted into the equations of motion (8.174),
and the coefficients of e i t are extracted as before. The resulting equations are, in
matrix form
mt 2 + ic1 + kt + ks i c1 ks H1t ( ) = kt :
ic1 ks mb 2 + ic1 + ks H1b( ) 0
(8.180)
This 2 2 system has a straightforward solution:
kt ( mb!2 + ic1 ! + ks )
H1t (!) =
( mt !2 + ic1 ! + kt + ks )( mb !2 + ic1 ! + ks ) (ic1 ! + ks )2
(8.181)
kt (ic1 ! + ks )
H1b (!) = :
( mt ! + ic1 ! + kt + ks )( mb !2 + ic1 ! + ks ) (ic1 ! + ks )2
2
(8.182)
436 The Volterra series and higher-order frequency response functions
y0p = ei 1 t + ei 2 t (8.184)
ytp = H1t ( 1 )ei 1 t + H1t ( 2 )ei 2 t + 2H2t( 1 ; 2 )ei( 1 + 2 )t + (8.185)
and
These expressions are substituted into the equations of motion (8.174) and
the coefficients of e i( 1 + 2 )t are extracted. The resulting equations are, in matrix
form
mt( 1 + 2 )2 + ic1 ( 1 + 2 ) + kt + ks ic1( 1 + 2 ) ks
ic1 ( 1 + 2 ) ks mb( 1 + 2 )2 + ic1 ( 1 + 2 ) + ks
t
HH2b(( 11;; 22))
1
= 11 c2 1 2 [H1t ( 1 ) H1b( 1 )][H1t ( 2 ) H1b( 2 )] (8.187)
so
H2t (!1 ; !2 ) = (! + ! )c ! ! [H t (! ) H b (! )][H t (! ) H b (! )]:
H1b (!1 ; !2 ) 1 2 2 1 2 1 1 1 1 1 2 1 2
(8.188)
The calculation of the third-order kernel transforms proceeds as before,
except a three-tone probing expression is used:
y0p = ei 1 t + ei 2 t + ei 3 t : (8.189)
yb
mb
fd
ks
yt
mt
kt
y0
P
where C denotes a sum over cyclic permutations of ! 1 , !2 and !3 . Also
3
Y
G(!1 ; !2; !3 ) = ic3 !n[H1t (!n ) H1b(!n )]: (8.192)
n=1
Note that to obtain Y t (! ) (respectively Yb (! )) requires a specification of H 3t
(respectively H3b ) and H1t , H1b , H2t and H2b .
This theory is easily adapted to the case of the ‘sky-hook’ suspension model
(figure 8.31) [128]. The equations of motion are
mt yt + (kt + ks )yt ks yb = kt y0
mb yb + fd (y_b) + ks (yb yt ) = 0: (8.193)
The H1 functions for this system are given by
kt ( mb!2 + ic1 ! + ks )
H1t (!) =
( mt !2 + kt + ks )( mb!2 + ic1 ! + ks ) ks2
(8.194)
kt + ks
H1b(!) = :
( mt !2 + kt + ks )( mb !2 + ic1 ! + ks ) ks2
(8.195)
where
mt !2 + kt + ks ks
1 0
(!) = ks mb!2 + ic1 ! + ks 1 : (8.197)
438 The Volterra series and higher-order frequency response functions
The quadratic stiffness coefficient k 2 enters as a linear multiplier for the product
of H1 s. Most importantly, H 1 is now known as the linear parameters have been
identified at the first stage. This useful property also holds for the H 3 , which has
the form
or
H3 (!1 ; !2; !3 ) = k2 F1 [H1 ; H2 ] + k3 F2 [H1 ] (8.205)
N
X
H2rs (!1 ; !2 ) = (!1 !2 c2mm k2mm )H1sm (!1 + !2 )H1rm (!1 )H1rm(!2 )
m=1
NX1 X
N
(!1 !2c2mn k2mn )[H1sm (!1 + !2 ) H1sn (!1 + !2 )]
m=1n=m+1
[H1rm(!1 ) H1rn (!1 )][H1rm (!2 ) H1rn(!2 )] (8.206)
where c2mn is the quadratic velocity coefficient for the connection between mass
m and mass n (c2mm is the connection to ground). k 2mn is the corresponding
stiffness coefficient. H2rs etc. refer to the FRFs between DOF r and DOF s.
Note that despite the considerable increase in complexity, the expressions are still
linear in the parameters of interest. H 3 is more complex still:
N
2X
H3rs (!1 ; !2 ; !3 ) = H sm (! + !2 + !3 )
3 m=1 1 1
[[!1 (!2 + !3 )c2mm k2mm ]H1rm (!1 )H2rm (!2 ; !3 )
+ [!2 (!3 + !1 )c2mm k2mm ]H1rm (!2 )H2rm(!3 ; !1)
+ [!3 (!1 + !2 )c2mm k2mm ]H1rm (!3 )H2rm(!1 ; !2)]
2 NX1 X N
[H sm (! + !2 + !3 ) H1sn (!1 + !2 + !3 )]
3 m=1n=m+1 1 1
[[!1 (!2 + !3 )c2mn k2mn ][H1rm (!1 ) H1rn(!1 )]
[H2rm (!2 ; !3 ) H2rn(!2 ; !3 )][!2 (!3 + !1 )c2mn k2mn ]
[H1rm (!2 ) H1rn(!2 )][H2rm(!3 ; !1 ) H2rn (!3 ; !1 )]
[!3 (!1 + !2 )c2mn k2mn ][H1rm (!3 ) H1rn(!3 )]
[H2rm (!1 ; !2 ) H2rn(!1 ; !2 )]]
N
X
+ (i!1 !2 !3 c3mm k3mm )
m=1
H sm (!
1 1 + !2 + !3 )H1rm(!1 )H1rm (!2 )H1rm (!3 )
440 The Volterra series and higher-order frequency response functions
NX1 X
N
(i!1 !2 !3 c3mn k3mn )
m=1n=m+1
[H1sm (!1 + !2 + !3 ) H1sn (!1 + !2 + !3 )]
[H rm (! ) H rn(! )][H rm(! ) H rn (! )][H rm (!
1 1 1 1 1 2 1 2 1 3 ) H1rn (!3 )]
(8.207)
and again, despite the complexity of the expression, the coefficients c 2mn , k2mn ,
c3mn and k3mn enter the equation in a linear manner.
In principle then, the availability of measurements of the HFRFs up to order
n would allow the identification of the parameters in the differential equations of
motion up to the nth-power terms. The problem is to obtain the HFRFs accurately
and without bias. Gifford used a random excitation test which is described in the
next section along with an illustrative example.
Recall, that (8.208) assumes that the input x(t) is white Gaussian. This
is also an assumption used in the derivation of (8.209). The most important
7 The arguments of this section do not directly follow [110]. The analysis there makes extensive
use of the Wiener series, which is essentially an orthogonalized version of the Volterra series. The
important facts which are needed here are simply that the correlations used approach the HFRFs in
the limit of vanishing P and these can be demonstrated without recourse to the Wiener series.
Higher-order modal analysis: hypercurve fitting 441
Charge
Amplifier
Accelerometer
m1 m2 m3
Nonlinear
Circuit
Power
Amplifier
Excitation Damper
Sy0 xx (!1 ; !2 )
= H2 (!1 ; !2 ) + O(P ) (8.211)
Sxx (!1 )Sxx (!2 )
so as P ! 0, the correlation function tends towards H 2 .
The theory described here was validated by experiment in [110]. The
structure used for the experiment was a forerunner of the beam used in
section 7.5.3 and is shown in figure 8.32. The main difference between this and
the rig used to validate the restoring force surface approach is that Gifford’s beam
was fixed–fixed and the nonlinear damping force was applied to the third lumped
mass instead of the second. Figure 8.33 shows a first-order FRF for the system
(the feedback circuit which provides the nonlinear damping force is switched off),
the fourth mode is sufficiently far removed from the first three to make this a
credible 3DOF system as the excitation force is band-limited appropriately.
Figure 8.34 shows the form of the feedback circuit used to provide the cubic
velocity force on mass m 3 . After collecting input v 1 , and output voltage v 2 , data
from the nonlinear circuit alone, a polynomial curve-fit established the circuit
characteristics
v2 = 1:34v1 + 1:25v12 + 0:713v13: (8.212)
The overall gain of the feedback loop was obtained by measuring the FRF
between input v 1 and output v 4 when the circuit was in linear mode; this is
shown in figure 8.35. This FRF is very flat at all the frequencies of interest.
Using the gain from the plot and all the appropriate calibration factors from
the instrumentation, the linear force–velocity characteristics of the nonlinear
feedback loop were obtained as
Ff = 120y_ 3 (8.213)
442 The Volterra series and higher-order frequency response functions
Figure 8.33. First-order frequency response function for experimental nonlinear beam
structure under low-level random excitation.
When the experiment was carried out, the level of excitation was set low
enough for distortion effects on the H 1 and H2 to be minimal. Figure 8.36 shows
a driving-point FRF for the system with and without the nonlinear part of the
circuit switched in; the effect of the nonlinearity is apparently invisible.
In order to estimate the H 2 functions, the autocorrelation function
Accelerometer
& v1 = 31.6 y
Charge Amplifier
v1
Nonlinear
Circuit
v2
Power
Amplifier
v3
Shaker
F
Force Link
& v4 = .316 F
Charge Amplifier
v4
Figure 8.34. Block diagram of all the elements that make up the linear damping system
with all their calibration factors in the configuration in which the results were measured.
form), together with the results of a standard modal analysis global curve-fitter;
the results are very close indeed, even by linear system standards.
In order to carry out the LS step which fits the model (8.206), it is necessary
to select the data from the (! 1 ; !2 )-plane. By analogy with common H 1 curve-
fitters, the data are chosen from around the peaks in order to maximize the signal-
to-noise ratio. Figure 8.39 shows how the data is selected to fit H 2 .
Gifford fitted three models to the H 2 measurements: one assuming a
quadratic stiffness characteristic, one assuming quadratic damping and one
assuming both. Figure 8.40 shows the H 2 reconstruction from the two ‘pure’
444 The Volterra series and higher-order frequency response functions
models compared to the measured data. It is clear that the damping model (on the
right) is the appropriate one. The coefficients obtained from a local curve-fit to
the second FRF H212 were as given in table 8.2.
By far the dominant contribution comes from the c 233 term as expected. The
estimated value compares well with the calibration value of 3540 ( 15.8% error).
When a global LS fit was carried out using data from all three H 2 curves at once,
the coefficient estimate was refined to 3807, giving a percentage error of 10.3%.
The results of this exercise validate Gifford’s method and offer a potentially
powerful means of using the Volterra series and HFRFs for parametric
identification.
Figure 8.36. First-order FRF from point 1 of the nonlinear rig in both its linear and
nonlinear modes.
Figure 8.37. Contour plots of the moduli of the three second-order HFRFs of the nonlinear
rig operating in its nonlinear mode.
The equivalent forms of (8.206) and (8.207) for a general N DOF system
form the basis of Storer’s approach to identification discussed in detail in
[237, 238, 253] 9 .
Storer’s experimental procedures were based on sinusoidal testing and he
made essentially the same assumptions as Gifford regarding the need for low
excitation levels. In the case of a SDOF system, equation (8.86) confirms that
Figure 8.38. The three first-order FRFs of the system formed from time-domain
correlation measurements (dotted).
448 The Volterra series and higher-order frequency response functions
Frequency f2
f1 = fr
000
111
111
000
11
00
000
11100
11
000
111
00
11
11
00 000
111
0
1
000
111
1
0 11
00
000
11100
11
000
11111
00
00
11 00
11
f2 = fr
00
11
00
110
1 0
1
000
111
0
1
00
1100
110
1
0
1
0
1
00
11 0
1
00
11
000
111
11
00
11
00
00
11
11
00 11
00
00
11
000
111 00
11
000
11100
11
000
111
Frequency f1
00
11
000
111 00
11
000
111
00
11
000
11100
11
000
111
00
11
11
00 11
00
011
11
11
0011
0000
11
000
111
0111
000
000
111
00
11 00
11
00
0
1
000
111000
111
00
11
000
111
000
111
f2 = -f
11
00 10111
000
000
111
r
000
1111111
0000
00
11
000
111 11
00
000
111
000
111
000
111
000
111000
111
000
111
000
111
000
111
000
11100
11
000
111
11
00
000
111
000
111
00
11
11
00
f1
+
f2
=
fr
Figure 8.39. Curve-fitting to 2 (!1 ; !2 ) takes place in the black area where the bands
containing the poles overlap.
Y (2 ) = 12 X 2 H2 ( ; ) + 12 X 4 H4 ( ; ; ; )
+ 15 6
32 X H6 ( ; ; ; ; ; ) + (8.220)
Y (3 ) = 14 X 3 H3 ( ; ; ) + 165 X 5 H5 ( ; ; ; ; ) + : (8.221)
Figure 8.40. Comparison of contour plots of the modulus of the second-order HFRF for
point 1 and both stiffness (left) and damping (right) curve-fits.
1 X12
0 1 4 15 X16 1
0
Y1 (2 ) 1 2 2 X1 32 0 1
B Y2 (2 ) C B 1 X2 1 4 15 X26 C H2 ( ; )
B .. C=B
B
2 2 2 X2 32 C C@ H4 ( ; ; ; ) A
@ A @ .. .. .. A
. . . . H6 ( ; ; ; ; ; )
YN (2 ) 1 XN2 1 XN4 15 XN6
2 2 32
(8.223)
450 The Volterra series and higher-order frequency response functions
0.0005
Magnitude (m/N)
0.0004
0.0003
0.0002
0.0001
0.0000
0.0
Phase (deg)
-50.0
-100.0
-150.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.41. Comparison of the theoretical (dotted) against calculated (full) H1 (! ) from
the Duffing oscillator system sinusoidally excited with amplitudes of 0.1, 0.15 and 0.2 N.
0
Y1 (3 ) 1 1 X13 5 X15 1
0
4 16
B Y2 (3 ) C B 1 X3 5 5 C
B 4 2 16 X2 C H3 ( ; ; )
H5 ( ; ; ; ; ) :
B .. C=B C (8.224)
@ A @ .. .. A
. . .
YN (3 ) 1 X3 5 X5
4 N 16 N
These systems of equations can be solved by standard LS methods.
The illustration for this method will be taken from [56]. A Duffing oscillator
(8.49) was chosen with coefficients m = 1, c = 20, k = 10 4 , k2 = 0
and k3 = 5 109 . The response to a stepped-sine excitation was computed
using fourth-order Runge–Kutta. The magnitudes and phases of the harmonic
components were extracted using an FFT after the transients had died down
at each frequency. Data for equations (8.222) to (8.224) were assembled for
up to H4 at amplitudes 0.1, 0.15 and 0.2 N and the first three diagonal FRFs
were calculated. The results are shown in figures 8.41–8.43 compared with the
analytical results; the agreement is impressive.
0.00008
Magnitude (m/N^2)
0.00006
0.00004
0.00002
0.00000
200.0
Phase (deg)
100.0
0.0
-100.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.42. Comparison of the theoretical (dotted) against calculated (full) principal
diagonal of H2 (!1 ; !2 ) from the Duffing oscillator system sinusoidally excited with
amplitudes of 0.1, 0.15 and 0.2 N.
Figure 8.43. Comparison of the theoretical (dotted) against calculated (full) principal
diagonal of H2 (!1 ; !2 ; !3 ) from the Duffing oscillator system sinusoidally excited with
amplitudes of 0.1, 0.15 and 0.2 N.
delay neural network (TDNN) are rather simple functions of the network weights.
The work is discussed here together with the extension to NARX networks and
452 The Volterra series and higher-order frequency response functions
u01
xi
w1
xi-1
yi
wn k
xi-N
uNn k
where hn is the usual nth-order Volterra kernel and the i are the sampling
intervals which can all be taken equal if desired. The requirement of causality
allows the lower index to be replaced by zero and the effect of damping is to
impose a finite memory T on the system, so the discrete series becomes
T
X
y(t) = h0 + h1 (1 )x(t 1 )
1 =0
T X
X T
+ 2 h2 (1 ; 2 )x(t 1 )x(t 2 ) +
1 =02 =0
T
X T
X
+ ::: n hn (1 ; : : : ; n )x(t 1 ) : : : x(t n ) + :
1 =0 n =0
(8.227)
The first step of the Wray–Green method is to expand the activation function
of the neural network—the hyperbolic tangent as a Taylor series, then
1(
X 1)n+1 Bn (24n 22n ) 2n 1
tanh(z ) = z (8.228)
n=1 (2n)!
where the Bn are the Bernoulli numbers defined by
2(2n)! X1 1
Bn = :
(2)2n h=1 h2n
(8.229)
The first three non-zero coefficients for this series are 1 = 1, 3 = 1=3
and 5 = 2=15. The function is odd so all the even coefficients vanish. In
practice, in order to deal with the bias b j for each of the hidden nodes, Wray
and Green expand the activation function not around zero, but around b j of each
454 The Volterra series and higher-order frequency response functions
and equating the coefficients of (8.234) and (8.226) yields a series of expressions
for the Volterra kernels
Xnh
h0 = wj a0j (8.235)
j =1
Xnh
h1 (m) = 1 wj a1j umj (8.236)
j =1
Xnh
h2 (m; l) = 2 wj a2j umj ulj (8.237)
j =1
Higher-order FRFs from neural network models 455
0.00010
0.00005
Amplitude
0.00000
-0.00005
-0.00010
0.00 0.10 0.20 0.30 0.40 0.50
Time (s)
Figure 8.45. First-order time-domain Volterra kernel from the Duffing oscillator obtained
from a neural network model by the Wray–Green method.
and, in general,
nh
X
hn (m1 : : : mn ) = n wj anj um1 j um2 j : : : umnj : (8.238)
j =1
In order to illustrate this approach, time data for an asymmetric Duffing
oscillator (8.49) were simulated with the usual coefficients m = 1, c = 20,
k = 104, k2 = 107 and k3 = 5 109. A TDNN with 20 hidden units was
fitted using an MLP package and the resulting predictions of the first- and second-
order impulse responses are shown in figures 8.45 and 8.46. Wray and Green
consider that this method produces considerably better kernel estimates than the
Toepliz matrix inversion method which they cite as the previous best method of
kernel estimation [150] and their results are confirmed by an independent study
by Marmarelis and Zhao [172].
Figure 8.46. Second-order time-domain Volterra kernel from the Duffing oscillator
obtained from a neural network model by the Wray–Green method.
all network inputs are normalized to lie in the interval [ 1; 1] and the output is
normalized onto the interval [ 0:8; 0:8]. Equation (8.239) actually holds between
the normalized quantities. The transformation back to the physical quantities,
once the HFRFs have been calculated, is derived a little later.)
The NARX models considered here arise from models of dynamical systems
and a further simplification can be made. It is assumed that the effects of all
the bias terms cancel overall, as the systems being modelled will not contain
constant terms in their equations of motion. In dynamical systems this can always
be accomplished with an appropriate choice of origin for y (i.e. the equilibrium
position of the motion) if the excitation x is also adjusted to remove its d.c. term.
Because of the presence of the lagged outputs in (8.239), the correspondence
between the network and Volterra series does not hold as it does for the TDNN,
and an alternative approach is needed. It will be shown here that the Volterra
kernels are no longer accessible; however, the HFRFs are.
The method of deriving the HFRFs is to apply the standard method of
harmonic probing. Because of the complex structure of the model however, the
algebra is not trivial. In order to identify H 1 (! ), for example, the system is probed
with a single harmonic
xpi = ei t (8.240)
and the expected response is as usual
where
ny X
X ny
Aj = vjk vjl H1 ( 1 )H1 ( 2 )(e i 1 kÆt e i 2 lÆt + e i 2 kÆt e i 1 lÆt )
k=1 l=1
(8.249)
x 1 nX
nX x 1
Bj = ujk ujl (e i 1 kÆt e i 2 lÆt + e i 2 kÆt e i 1 lÆt ) (8.250)
k=0 l=0
x 1
ny nX
X
Cj = 2 vjk ujl (H1 ( 1 )e i 1 kÆt e i 2 lÆt + H1 ( 2 )e i 2 kÆt e i 1 lÆt )
k=1 l=0
(8.251)
and
nh
X ny
X
D=1 wj tanh(1) (bj ) vjk e i( 1 + 2 )kÆt :
(8.252)
j =1 k=1
Derivation of H 3 is considerably more lengthy and requires probing with
three harmonics. The expression can be found in [57].
zi = kxi ci k: (8.254)
This is then passed through a basis function which decays rapidly with its
argument, i.e. it is significantly non-zero only for inputs close to c i . The overall
Higher-order FRFs from neural network models 459
output of the RBF network is therefore the summed response from several locally-
tuned units. It is this ability to cover selectively connected regions of the input
space which makes the RBF so effective for pattern recognition and classification
problems. The RBF structure also allows an effective means of implementing the
NARX model for control and identification [61, 62].
For the calculation given here, a Gaussian basis function is assumed as this
is by far the most commonly used to date. Also, following Poggio and Girosi
[206], the network is modified by the inclusion of direct linear connections from
the input layer to the output. The resulting NARX model is summarized by
nh y n x 1
nX
X 1 X
yi = s + wj exp (y vjk )2 + (xi m ujm )2
j =1 2j2 k=1 i k m=0
n
X y x 1
nX
+ aj yi j + bj xi j (8.255)
j =1 j =0
| {z }
from linear connections
where the quantities v jk and ujm are the hidden node centres and the i is the
standard deviation or radius of the Gaussian at hidden node i. The first part of
this expression is the standard RBF network.
As with the MLP network the appearance of constant terms in the exponent
will lead to difficulties when this is expanded as a Taylor series. A trivial
rearrangement yields the more useful form
nh n
y
X 1 X
yi = s + wj j exp (y2 2vjk yi k )
j =1 2j2 k=1 i k
x 1
nX
+ (x2i m 2ujm xi m )
m=0
ny
X x 1
nX
+ aj yi j + bj xi j (8.256)
j =1 j =0
where
n
y nx 1
1 X 2 + X u2
= exp v :
j
2j2 k=1 jk m=0 jm
(8.257)
Now, expanding the exponential and retaining only the linear terms leads to
the required expression for obtaining H 1 :
nh ny x 1
nX
X wj j X
yi = vjk yi k + ujm xi m : (8.258)
j =1 j k=1 m=0
460 The Volterra series and higher-order frequency response functions
where
ny nh n y
X X 1 X
D=1 aj e i( 1 + 2 )jÆt j wj e i( 1 + 2 )kÆt :
j2 k=1
(8.261)
j =1 j =1
The neural network program used here, MLP, calculates the scaled x values
as
x xmin 1
xs = 1:6
xmax xmin 2
1:6x 0:8(xmin + xmax )
= (8.263)
xmax xmin xmax xmin
which gives
1:6
= (8.264)
xmax xmin
and similarly
1:6
a= : (8.265)
ymax ymin
The dimensions of the nth kernel transform follow from
sF [y(t)jn ] sF [ay + b]
Hn ( ) =
F [x(t)j ]n = F [ x + ]n (8.266)
saF [y(t)]
= n
F [x(t)]n (8.267)
a
Hns ( 1 ; : : : ; n ) = n Hn ( 1 ; : : : ; n ): (8.268)
(y y )
Hn ( 1 ; : : : ; n ) = 1:6n 1 max min Hns ( 1 ; : : : ; n ): (8.269)
(xmax xmin)n
The network is therefore scale specific and may not perform well with data
originating from very different excitation levels. The solution to this problem is
for the training set to contain excitation levels spanning the range of interest.
462 The Volterra series and higher-order frequency response functions
0.0005
Magnitude (m/N)
0.0004
0.0003
0.0002
0.0001
0.0000
0.0
Phase (deg)
-50.0
-100.0
-150.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.47. Exact (dashed) and estimated (full) H1 from the Duffing oscillator, estimate
from probing of 10:2:1 NARX MLP network.
Figure 8.48. Comparison between the measured data and the NARX MLP network
estimate for figure 8.47.
in figure 8.49 along its leading diagonal ( 1 = 2 ). This was calculated from a
10:4:1 network trained to an MPO error of 0.27. The corresponding H 1 ( 1 ) from
the same network, shown in figure 8.50, shows a little discrepancy from theory.
This is unfortunate from a system identification point of view as one would like a
correct representation of the system at all orders of nonlinearity. That this happens
appears to be due to the fact that the network can reproduce the signals with some
confusion between the different ordered components y n . This is discussed a little
later.
The RBF networks were trained and tested using the force and displacement data.
A spread of results was observed, the best H 1 ( ) was given by a 6:2:1 network,
trained to a 0.48 MPO error as shown in figure 8.51. A 4:4:1 network produced the
best H2 ( 1 ; 2 ) after training to a 0.72 MPO error; this is shown in figure 8.52.
464 The Volterra series and higher-order frequency response functions
0.00010
0.00008
Magnitude (m/N^2)
0.00006
0.00004
0.00002
0.00000
200.0
Phase (deg)
100.0
0.0
-100.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.49. Exact (dashed) and estimated (full) principal diagonal for H2 from the
Duffing oscillator; an estimate from probing of a 10:4:1 NARX MLP network.
0.0005
Magnitude (m/N)
0.0004
0.0003
0.0002
0.0001
0.0000
0.0
Phase (deg)
-50.0
-100.0
-150.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.50. Exact (dashed) and estimated (full) H1 from the Duffing oscillator; the
estimate from probing of a 10:4:1 NARX MLP network.
8.11.5.3 Discussion
0.0005
Magnitude (m/N)
0.0004
0.0003
0.0002
0.0001
0.0000
0.0
Phase (deg)
-100.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.51. Exact (dashed) and estimated (full) H1 from the Duffing oscillator; the
estimate from probing of a 6:2:1 NARX RBF network.
0.00010
0.00008
Magnitude (m/N^2)
0.00006
0.00004
0.00002
0.00000
200.0
Phase (deg)
100.0
0.0
-100.0
-200.0
0.0 100.0
Frequency (rad/s)
Figure 8.52. Exact (dashed) and estimated (full) principal diagonal for H2 from the
Duffing oscillator; the estimate from probing of a 4:4:1 NARX RBF network.
the error on the training data continues to decrease but the error over a testing
set begins to increase. Neural network users have long known this and always
use a training and testing set for validation, and often use a training, testing and
validation set. This may be necessary in this approach to system identification.
Often the MPO error is taken as conclusive.
A principled approach to model testing may be able to inform the modeller
that the model is indeed over-parametrized. The problem remaining is to remove
the over-parametrization. It may be possible to accomplish this in two ways.
First, there is a possibility of regularization [40], which seeks to ensure that the
network weights do not evolve in an unduly correlated manner. The ideal situation
would be if an analogue of the orthogonal LS estimator discussed in appendix E
could be derived for neural networks. This appears to be a difficult problem and
there is no current solution. The second possibility is pruning, where the network
nodes and weights are tested for their significance and removed if they are found
to contribute nothing to the model. This reduces the numbers of weights to be
estimated for a given network structure.
Z +1 Z +1
( j )
y1 (t) = (j : a) ( a)
d h1 ( )x (t ) + d h(1j:b) ( )x(b) (t ) (8.270)
1 1
The multi-input Volterra series 467
and
Z +1 Z +1
( j )
y2 (t) = d1 d2 h(2j:aa) (1 ; 2 )x(a) (t 1 )x(a) (t 2 )
1 1
Z +1 Z +1
+ d1 d2 fh(2j:ab) (1 ; 2 ) + h(2j:ba) (2 ; 1 )g
1 1
x(a) (t 1 )x(b) (t 2 )
Z +1 Z +1
+ d1 d2 h(2j:bb) (1 ; 2 )x(b) (t 1 )x(b) (t 2 ):
1 1
(8.271)
m1 y(1) + c1 y_ (1) + k11 y(1) + k12 y(2) + k0 y(1)2 + k00 y(1)3 = x(1) (t) (8.272)
m2 y(2) + c2 y_ (2) + k21 y(1) + k22 y(2) = x(2) (t): (8.273)
x(1)
p (t) = 0; x(2) p (t) = ei!t
yp(1) (t) = H1(1:2) (!)ei!t ; yp(2) (t) = H1(2:2) (!)ei!t (8.275)
(a)
Frequency (rad/s)
(b)
Frequency (rad/s)
Figure 8.53. Comparison between HFRFs from a continuous-time system (full) and a
(1:1) (1:2)
discrete-time model (dashed): (a) H1 (! ); (b) H1 (! ).
(a)
(b)
(c)
(1:11)
Figure 8.54. H2 magnitude surface: (a) from the continuous-time equation of motion;
(b) from a NARX model; (c) difference surface.
(a)
(b)
(a)
(b)
(c)
(1:12)
Figure 8.56. H2 magnitude surface: (a) from the continuous-time equation of motion;
(b) from a NARX model; (c) difference surface.
The multi-input Volterra series 473
(a)
(b)
order Runge–Kutta scheme. The two excitations x (1) (t) and x(2) (t) were two
independent Gaussian noise sequences with rms 5.0, band-limited on the interval
(0; 200) Hz using a Butterworth filter. The time step was 0.001 corresponding
to a sampling frequency of 1000 Hz. Using the resulting discrete force and
displacement data, a NARX model was fitted. The model structure was easily
guessed as
and the parameters were fitted using a linear LS algorithm. The actual values of
the i coefficients will not be given here as they have little meaning. The model
was tested by stepping it forward in time using the same excitations x (1) and x(2)
and comparing the results with those from the original system. Figure 8.58 shows
the comparisons for each DOF; it is impossible to distinguish the two curves (the
NARX prediction is the broken line, the original data are the full line).
The HFRF extraction algorithm is almost identical to that for the continuous-
time system, the same probing expressions and coefficient extraction procedures
apply; the only difference is that where the derivative operator produces a
prefactor i! when it meets the harmonic e i!t , the lag operator extracts a
prefactor ei!t , where t is the sampling interval.
The H1 matrix for the discrete-time model is found to be
The H1 s are shown in figure 8.53 as the broken lines, in comparison with
the H1 s from the original continuous-time model. As would be expected from the
accuracy of the NARX models, the two sets of H 1 s agree almost perfectly.
The multi-input Volterra series 475
(a)
(b)
Figure 8.58. Comparison between the measured data and the NARX model-predicted
output.
476 The Volterra series and higher-order frequency response functions
where