Nonlinearity in Structural Dynamics Chapter 04
Nonlinearity in Structural Dynamics Chapter 04
Nonlinearity in Structural Dynamics Chapter 04
4.1 Introduction
The Hilbert Transform is a mathematical tool which allows one to investigate
the causality, stability and linearity of passive systems. In this chapter its main
application will be to the detection and identification of nonlinearity. The theory
can be derived by two independent approaches: the first, which is the subject of
this chapter, relies on the decomposition of a function into odd and even parts and
the behaviour of this decomposition under Fourier transformation. The second
method is more revealing but more complicated, relying as it does on complex
analysis; discussion of this is postponed until the next chapter.
The Hilbert transform is an integral transform of the same family as
the Fourier transform, the difference is in the kernel function. The complex
exponential e i!t is replaced by the function 1=i ( !), so if the Hilbert
transform operator is denoted by H, its action on functions 1 is given by2
Z 1
1 G( )
HfG(!)g = G~ (!) = i
PV d
!
(4.1)
1
where P V denotes the Cauchy principal value of the integral, and is needed as
the integrand is singular, i.e. has a pole at ! = . To maintain simplicity of
notation, the P V will be omitted in the following discussions, as it will be clear
from the integrands, which expressions need it. The tilde ~ is used to denote the
transformed function.
1 ()
In this chapter and the following the functions of interest will generally be denoted g t and G !()
to indicate that the objects are not necessarily from linear or nonlinear systems. Where it is important
() () () ( )
to make a distinction h t and H ! will be used for linear systems and t and ! will be used
for nonlinear.
2 This differs from the original transform defined by Hilbert and used by mathematicians, by the
introduction of a prefactor 1 i=i
= . It will become clear later why the additional constant is useful.
127
128 The Hilbert transform—a practical approach
and
godd(t) = g(gjt(jj)t=j)2=;2; tt > 0
<0. (4.6)
That this is only true for causal functions is shown by the simple
counterexample in figure 4.2. It follows immediately from equations (4.5) and
(4.6) that
geven(t) = godd(t) (t) (4.7)
godd(t) = geven(t) (t) (4.8)
where (t) is the signum function, defined by 3
(
1; t>0
(t) = 0; t=0 (4.9)
1; t < 0.
3 The notation sgn(t) is often used.
Basis of the method 129
g(t)
geven(t)
godd(t)
Figure 4.1. Decomposition of a causal function into odd and even parts.
g(t)
1
1 1
3
2
geven(t)
1
godd(t)
2
Equations (4.14) and (4.15) can be brought into the final forms
1
Z 1 Im G( )
Re G(!) = d (4.17)
1 !
1
Z 1 Re G( )
Im G(!) = + d : (4.18)
1 !
It follows from these expressions that the real and imaginary parts of
a function G(! ), the Fourier transform of a causal function g (t), are not
independent. Given one quantity, the other is uniquely specified. (Recall that
these integrals are principal value integrals.) Equations (4.17) and (4.18) can
be combined into a single complex expression by forming G(! ) = Re G(! ) +
i Im G(!), the result is
1 1
Z
G( )
G(!) = d : (4.19)
i 1 !
Now, applying the definition of the Hilbert transform in equation (4.1) yields
G(!) = G~ (!) = HfG(!)g: (4.20)
Basis of the method 131
1 1 Im G( )
Z
Re G(!) = d
1 !
Z 0 Z 1
1 Im G( ) Im G( )
= d + d
1 ! 0 !
Z 1 Z 1
1 Im G( ) Im G( )
= d + d
0 ! 0 !
Z 1 Z 1
1 Im G( ) Im G( )
= d + d
0 ! 0 !
Z 1 Z 1
1 Im G( ) Im G( )
= d + d
0 +! 0 !
Z 1
2 Im G( )
= d 2 !2 (4.23)
0
and similarly
2! 1 Re G( )
Z
Im G(!) = d 2 :
!2
(4.24)
0
These equations are often referred to as the Kramers–Kronig relations [154].
The advantage of these forms over (4.17) and (4.18) is simply that the range of
integration is halved and one of the infinite limits is removed.
132 The Hilbert transform—a practical approach
Unfortunately, log jG(! )j and (! ), as they stand, do not form a Hilbert
transform pair. However, it can be shown that the function (log G(! )
log G(0))=! is invariant under the transform and so the functions (log jG(!)j
log jG(0)j)=! and ((!) (0))=! do form such a pair. If in addition, the
minimum phase condition, (0) = 0, is assumed, the Hilbert transform relations
can be written:
2!2 1
Z
( )
log jG(!)j log jG(0)j = d
( 2 !2 )
(4.29)
0
Z 1
2! log jG(!)j log jG(0)j
(!) = d 2 !2 : (4.30)
0
The effort involved in deriving these equations rigorously is not justified as
they shall play no further part in the development; they are included mainly for
completeness. They are of some interest as they allow the derivation of FRF
phase from FRF modulus information, which is available if one has some means
of obtaining auto-power spectra as
s
jH (!)j = SSyy ((!!)) : (4.31)
xx
4.3 Computation
Before proceeding to applications of the Hilbert transform, some discussion of
how to compute the transform is needed. Analytical methods are not generally
applicable; nonlinear systems will provide the focus of the following discussion
and closed forms for the FRFs of nonlinear systems are not usually available.
Approximate FRFs, e.g. from harmonic balance (see chapter 3), lead to integrals
4 Assuming the principal sheet for the log function.
Computation 133
Im G(ωj)ωj
ω2j ω2i
∆ω
2 ∆ω 2∆ω
ω1 ωi ωn
This, the most direct approach, seeks to estimate the frequency-domain integrals
(4.17) and (4.18). In practice, the Kramers–Kronig relations (4.23) and (4.24)
are used as the range of integration is simplified. Converting these expressions to
discrete sums yields
N
2X Im G(!j )!j
Re G~ (!i ) = !
j=1 !j2 !i2
(4.32)
N
2!i X Re G(!j )
Im G~ (!i ) = !
j=1 !j2 !i2
(4.33)
Figure 4.4. Hilbert transform of a simulated SDOF linear system showing perfect overlay.
Figure 4.4 shows a linear system FRF with the Hilbert transform
superimposed. Almost perfect overlay is obtained. However, there is an important
assumption implicit in this calculation, i.e. that ! 1 = 0 and that !N can be
substituted for the infinite upper limit of the integral with impunity. If the integrals
from 0 to ! 1 or !N to infinity in (4.23) and (4.24) are non-zero, the estimated
Hilbert transform is subject to truncation errors. Figure 4.5 shows the effect of
truncation on the Hilbert transform of a zoomed linear system FRF.
Computation 135
Figure 4.5. Hilbert transform of a simulated SDOF linear system showing truncation
problems.
There are essentially five methods of correcting Hilbert transforms for truncation
errors, they will now be described in order of complexity.
136 The Hilbert transform—a practical approach
N
X i!Ak
HM (!) = 2 2 (4.35)
k=1 !k ! + i2k !k !
where Ak is the complex modal amplitude of the k th mode; ! k is the undamped
natural frequency of the k th mode and k is its viscous damping ratio. By
assuming that the damping is small and that the truncation frequency, ! max , is
much higher than the natural frequency of the highest mode, equation (4.35) can
be reduced to (for ! > ! max )
N
X Ak
HM (!) = i (4.36)
k=1 !
which is an approximation to the ‘out-of-band’ FRF. This term is purely imaginary
and thus provides a correction for the real part of the Hilbert transform via
equation (4.32). No correction term is applied to the imaginary part as the error
is assumed to be small under the specified conditions.
The actual correction is the integral in equation (4.1) over the interval
(!max ; 1). Hence the correction term, denoted C R (!), for the real part of the
Hilbert transform is
2 1
Z N Z 1
Im(G( )) 2X d
C R (!) = d 2 !2 = A 2
k
!2
(4.37)
wmax ! k=1 !max
Computation 137
Im(G(!max )) !2 !4
C R (!) = C R (0) 2 + 4 + (4.40)
!max 2!max
where C R (0) is estimated from
Z
2 wmax Im(G( ))
C R (0) = Re(G(0)) d : (4.41)
0+
Using the same approach, the correction term for the imaginary part, denoted
by C I (! ), can be obtained:
!3 !5
I 2 !
C (!max ) = Re(G(!max )) + 3 + 5 + : (4.42)
!max 3!max 5!max
4.3.2.4 The Simon correction method
This method of correction was proposed by Simon [229]; it allows for truncation
at a low frequency, ! min and a high frequency ! max . It is therefore suitable for
use with zoomed data. This facility makes the method the most versatile so far.
As before, it is based on the behaviour of the linear FRF, say equation (4.35)
for mobility data. Splitting the Hilbert transform over three frequency ranges:
(0; !min), (!min; !max ) and (!max ; 1), the truncation errors on the real part of
the Hilbert transform, B R (! ) at low frequency and the now familiar C R (! ) at
high frequency, can be written as
Z
2 !min Im(G(!))
B R (! ) = d 2 !2 (4.43)
0
138 The Hilbert transform—a practical approach
and
2
Z 1 Im(G(!))
C R (!) = d 2 !2 : (4.44)
!max
If the damping can be assumed to be small, then rewriting equations (4.40)
and (4.44) using the mobility form (4.35) yields
Z
2 !min X N 2 Ak
B R (!) = d 2 2
!k )( 2 !2 )
(4.45)
0 k=1 (
and
2 1 N 2 Ak
Z X
C R (!) = d 2 :
!max k=1 ( 2 !k )( 2 !2 )
(4.46)
ω low ωa ωb ω ri ωc ω d ωhigh
Suppose mode i is the lowest mode in the measured region with resonant
frequency ! ri and therefore has the greatest effect on the low-frequency
truncation error B R (! ), the relevant part of Im H m can be decomposed:
i 1
m (! ) =
X Ak ! Ai !
Im HM 2k !2 )2 + (!i2 !2 )2 (4.52)
k=1 ( !
where the superscript m indicates that this is the mass asymptote of the FRF. In
the lower part of the frequency range !=! k is small and the first term can be
expanded:
" #
i 1 i 1 2
X Ak ! X ! ! !
2 2 2 = A k 1 + + = O (4.53)
k=1 (!k ! ) k=1 !k
!k !k
and neglected, so
m (! ) Ai !
Im HM :
(!i2
!2 )2
(4.54)
and neglected, so
s (!) Aj !
Im HM 2
(!j !2 )2
(4.59)
and aj is estimated by fitting the function (4.59) to the data over the range ! c to
!d (figure 4.6). The high-frequency correction term is obtained by substituting
(4.59) into the Kramers–Kronig relation:
2 1 2 Aj
Z
C R (! ) = d 2
!min ( 2 !j )( 2 !2 )
(4.60)
m (! ) 2Ai i !i !2 ai ! 2
Re HM =
(!i2 !2)2 (!i2 !2 )2
(4.63)
Computation 141
s (!) b1 + b2 !2 + b3 !4
Re HM
(!j2 !2 )2
(4.69)
where
b3 !(2!j2 + !2 ) + b2 ! b1 !j2
j 1 (! ) = ;
2(!j2 !2)
b1 ! b2 + 3b3!j2
j 2 (! ) = ;
4(!j2 !2)
(4.72)
a1! + b2 + b3 !j2
j3 ( ! ) = :
4(!j2 + !2 )
Note that these results only apply to mobility FRFs, substantially different
correction terms are needed for the other FRF forms. However, they are derived
by the same procedure as the one described here.
Although the Ahmed correction procedure is rather more complex than the
others, it produces excellent results. Figure 4.7 shows the Hilbert transform
in figure 4.5 recomputed using the Ahmed correction terms; an almost perfect
overlay is obtained.
4.3.2.6 Summary
None of the correction methods can claim to be faultless; truncation near to a
resonance will always give poor results. Considerable care is needed to obtain
satisfactory results. The conversion to receptance, Fei and Haoui techniques are
only suitable for use with baseband data and the Simon and Ahmed corrections
require a priori curve-fitting. The next sections and the next chapter outline
approaches to the Hilbert transform which do not require correction terms and
in some cases overcome the problems.
Note also that the accelerance FRF tends to a constant non-zero value as
! ! 1. As a consequence the Hilbert transform will always suffer from
truncation problems, no matter how high ! max is taken. The discussion of this
problem requires complex analysis and is postponed until the next chapter.
1 1
Z
G( )
HfG(!)g = G~ (!) = i 1
d
!
: (4.73)
Figure 4.7. Hilbert transform with Ahmed’s correction of zoomed linear data.
it is clear that
i
G~ (!) = G(!) : (4.75)
!
Now, a basic theorem of Fourier transforms states that
where (t) is the signum function defined in (4.9). (Ff(t)g = i=(!) is proved
in appendix D.)
It immediately follows from (4.77) that
The disadvantages of the method arise from the corrections needed. Both
result from the use of the FFT, an operation based on a finite data set.
The first problem arises because the FFT forces periodicity onto the
data outside the measured range, so the function (t) which should look like
figure 4.8(a), is represented by the square-wave function sq(t) of figure 4.8(b).
This means that the function G(! ) is effectively convolved with the function
i cot(!) = Ffsq(t)g instead of the desired i=(!). (See [260] for the
appropriate theory.) The effective convolving functions is shown in figure 4.9(b).
Computation 145
(a)
ε(t)
+1
-1
(b)
sq(t) +1
-1
Figure 4.8. Effect of the discrete Fourier transform on the signum function.
(a)
F [ ε(t) ] = -i
πω
(b)
F [ sq(t) ] = -i cot πω
Figure 4.9. Desired Hilbert window and periodic form from the discrete FFT.
!=2 and !=2 has been discarded. This can be alleviated by transferring
energy to the neighbouring lines and adopting the definition
8
> 0; k=1
>
>
> 3i
>
> ; k=2
>
> 2
>
> i N
< ; 3k
Uk = (k 1) 2 (4.81)
>
> i N
>
>
> ; +1k N 1
>
>
> (N + 1 k ) 2
: 3i ;
>
>
k = N.
2
Computation 147
1.5
1.2
1.0
Windoe Magnitude
0.8
0.5
0.2
0.0
0 64 128 192 256 320 384 448 512
Window Index
1.2
1.0
Windoe Magnitude
0.8
0.5
0.2
0.0
0 64 128 192 256 320 384 448 512
Window Index
inefficient if the zoom range is small or at high frequency and will clearly lead to
errors if low-frequency modes have been discarded.
Of the correction methods described in section 4.4.2, the only one applicable
is conversion to receptance and this should be stressed. This is only effective for
correcting the high-frequency error. However, as previously discussed, the data
should always be baseband in any case.
In summary then, the modified Fourier method 1 proceeds as follows.
(1) Convert the measured 12 N -point positive-frequency FRF G(! ) to an N -point
positive-frequency FRF by translation, reflection and padding.
(2) Complete the FRF by generating the negative-frequency component. The
real part is reflected about ! = 0, the imaginary part is reflected with a sign
inversion. The result is a 2N -point function.
(3) Take the inverse Fourier transform of the discretized i=(! ) on 2N points.
This yields the Hilbert window hi(t).
(4) Take the inverse Fourier transform of the 2N -point FRF. This yields the
impulse response g (t).
(5) Form the product g (t) hi(t).
(6) Take the Fourier transform of the product. This yields the desired Hilbert
transform G ~ (!).
Computation 149
G (ω)
1
πω
Fourier method 1 was discussed as it was the first Hilbert transform method to
exploit Fourier transformation. However, it is rather complicated to implement
and the method discussed in this section is to be preferred in practice.
The implementation of this method is very similar to Fourier method 1;
however, the theoretical basis is rather different. This method is based on the
properties of analytic 6 signals and is attributed to Bendat [24]. Given a time
6 This terminology is a little unfortunate, as the word analytic will have two different meanings in
this book. The first meaning is given by equation (4.82). The second meaning relates to the pole-zero
structure of complex functions—a function is analytic in a given region of the complex plane if it has
no poles in that region. (Alternatively, the function has a convergent Taylor series.) The appropriate
meaning will always be clear from the context.
150 The Hilbert transform—a practical approach
G(ω)
1
πω
B
Circular convolution
component
Range of convolution
Now, recall that the Hilbert transform factors into Fourier operations. The
decomposition depends on whether the operator acts on time- or frequency-
domain functions. The appropriate factorization in the frequency domain is given
by (4.79). Essentially the same derivation applies in the time domain and the
result is
H = F 1 Æ 2 ÆF : (4.84)
7 This definition differs from convention
G(ω)
0 1 N N 3N 2N
Ω πω 2 2
-N N
Figure 4.14. Solution to the circular convolution problem using translation and
zero-padding.
so 8
< 2G(! ); !>0
A(!) = G(!); !=0 (4.86)
:
0; !<0
thus, the spectrum of an analytic signal depends only on the spectrum of the real
part. This fact is the basis of the method.
Any function of frequency has a trivial decomposition
zoom range
the result is
8
< 2GR( ); >0
G ( ) = : GR( ); = 0 (4.90)
0; <0
where
GR( ) = F fRe G(!)g (4.91)
transform of the real part8 . This fact provides a means of computing the FRF
imaginary part from the real part. In principle, three steps are required:
(1) Take the Fourier transform F of the FRF real part Re G(! ), i.e. GR( ).
(2) Form the transform G ( ) using (4.89).
(3) Take the inverse Fourier transform F 1 of G ( ). The result is G(! ), i.e. the
desired Hilbert transform, Re~G(! ), has been obtained as the imaginary part.
The structure used to obtain the experimental data was a composite 23 1 scale
aircraft wing used for wind tunnel tests. The wing was secured at its root to a
rigid support, effectively producing a cantilever boundary condition. Excitation
of the wing was via an electrodynamic exciter attached to the wing via a push
rod (stinger) and a force transducer. The excitation was a band-limited random
signal in the range 0–512 Hz. The response of the wing was measured using
lightweight accelerometers. (Note that random excitation is not optimal for
nonlinear structures—this will be discussed later. This study is intended to show
how the Hilbert transform is computed, and one can only validate the method on
a linear structure.)
Figure 4.16 shows the accelerance FRF measured by the experiment. At least
seven modes are visible. For information, the resonance at 76 Hz was identified
as first wing bending, that at 215 Hz was identified as first wing torsion.
8 Z Z 1
Note that
1
Re G(!) = dt cos(!t)g(t) = dt e i!tgeven (t)
1 1
so,
Z
1 Z 1
GR( ) = d! e i ! dt e i!tgeven (t)
1
Z 1
1
Z 1 Z 1
= dt geven (t) d! e i!( +t) = 2 dt geven (t)Æ( + t)
1 1 1
= 2geven ( ):
So GR is essentially the even component of the original time signal. This fact does not help with
the development of the algorithm. However, it does show that the terminology ‘pseudo spectrum’ for
GR, which is sometimes used, is probably inappropriate.
154 The Hilbert transform—a practical approach
Figure 4.16. A typical experimental cross-accelerance FRF measured from a scaled wing
model.
The first step in the procedure is to correct for truncation; the FRF is
converted to receptance by dividing by ! 2 (avoiding the division at ! = 0).
The result is shown in figure 4.17. To further reduce truncation errors, the FRF
was extended to 2N points by padding with zeroes (figure 4.18).
The next stage was the completion of the FRF, i.e. the conversion to a
double-sided form. The negative frequency parts were obtained by assuming even
symmetry for the real part and odd symmetry for the imaginary part. The double-
sided signals are given in figure 4.19.
The function GR( ) was formed by Fourier transforming the real part
(figure 4.20(a)). This was converted to G ( ) by zeroing the negative- component
and doubling the positive- part. The = 0 line was left untouched. Taking the
inverse Fourier transform then gave Re] G as the imaginary part.
The function GI ( ) was formed by Fourier transforming the imaginary part
of the FRF (figure 4.20(b)). This was also converted to the full G ( ) as before.
Taking the inverse FFT gave Im ] G as the real part.
Both the real and imaginary parts of the Hilbert transform have now been
obtained. The next stage was simply to convert back to the accelerance form.
In order to evaluate the results, the Hilbert transform is shown overlaid on
the original FRF in figure 4.21, the two curves should match. Both the Bode
magnitude and Nyquist plots are given. The somewhat poor quality of the Nyquist
Computation 155
Figure 4.17. Receptance FRF converted from accelerance FRF in figure 4.16.
Figure 4.19. (a) Double-sided (even function) real part of the FRF of figure 4.18. (b)
Double-sided (odd function) imaginary part of the FRF of figure 4.18.
Figure 4.20. (a) Pseudo-spectrum from the Fourier transform of the curve in figure 4.19(a).
(b) Pseudo-spectrum from the Fourier transform of the curve in figure 4.19(b).
Figure 4.21. Overlay of the experimental (——) and Hilbert transformed (– – –) data in
(a) Bode plot, (b) Nyquist plot.
In the Nyquist plot, the characteristic circle is rotated clockwise and elongated
into a more elliptical form. Figure 4.24 shows the FRF and transform in a more
extreme case where the FRF actually shows a jump bifurcation. The rotation and
elongation of the Nyquist plot are much more pronounced.
The characteristic distortions for a number of common nonlinearities are
summarized next (in all cases the FRFs are obtained using sine excitation).
Figure 4.23. Hilbert transform of a hardening cubic spring FRF at a low sine excitation
level.
Figure 4.24. Hilbert transform of a hardening cubic spring FRF at a high sine excitation
level.
The FRF and Hilbert transform are given in figure 4.27. In the Bode plot
164 The Hilbert transform—a practical approach
Figure 4.25. Hilbert transform of a softening cubic spring FRF at a high sine excitation
level.
the peak of the Hilbert transform curve stays at the same frequency as in the
FRF, but decreases in magnitude. In the Nyquist plot, the characteristic circle is
compressed into an ellipse along the imaginary axis.
Note that in the case of Coulomb friction, the nonlinearity is only visible if
the level of excitation is low. Figure 4.28 shows the FRF and transform at a high
level of excitation where the system is essentially linear.
Choice of excitation 165
Figure 4.27. Hilbert transform of a Coulomb friction system FRF at a low sine excitation
level.
to the most non-causal time functions. Recalling the discussion of chapter 2, the
most distorted FRFs are obtained from stepped-sine excitation and, in fact, it will
be proved later that such FRFs for nonlinear systems will generically show Hilbert
transform distortions. (The proof requires the use of the Volterra series and is
therefore postponed until chapter 8 where the appropriate theory is introduced.)
Choice of excitation 167
Figure 4.28. Hilbert transform of a Coulomb friction system FRF at a high sine excitation
level.
The non-causal power ratio (NPR) is then defined as the ratio of non-causal
power Pn to the total system power P as encoded in the FRF
R0
Pn 1 dt jgn(t)j2
NPR =
P
= R1
1 dt jg(t)j
2: (4.101)
Indicator functions 169
. .
k3 y3 k3 > 0 c2 y y
.
k3 y3 k3 < 0 Fc sgn( y)
Figure 4.29. Non-causal power ratio plots for various SDOF nonlinear systems.
R0 2
NPR =
Pn
= 1 R1 1 dt jgn(t)j :
2 1 d! jG(!)j2
(4.102)
P
4.6.2 Corehence
This measure of nonlinearity, based on the Hilbert transform, was introduced in
[213] as an adjunct to the coherence function described in chapter 2. The basis of
the theory is the operator of linearity P , defined by 9
(!)2 =
jE fG~ (!)G(!) gj2 :
E fjG~ (!)j2 gE fjG(!)j2 g
(4.104)
(!)2 =
jG~ (!)G(!) j2 :
jG~ (!)j2 jG(!)j2
() ~( )
However, the expectation operators are implied; if the G ! and G ! are themselves expectations,
expression (4.104) collapses to unity. There, is therefore, an implicit assumption that the form of
excitation must be random as it is in the case of the coherence. Now, it is stated above that the Hilbert
transform of an FRF obtained from random excitation does not show distortions. This does not affect
f ~ g 6= f ~ g f g
the utility of the corehence as that statement only applies to the expectation of the FRF, i.e. the FRF
after averaging. Because E GG E G E G , the corehence departs from unity for nonlinear
systems.
Indicator functions 171
in 1
so Z
dn x n
n X (! ) = i M (n)
= d! ! (4.107)
dtn t=0 2 1 2
where M (n) the nth moment integral of X (! ) or the nth spectral mo-
R 1 denotes
ment— 1 d! ! n X (! ). Now it follows from the Taylor’s series
1 1 dn x 1 (it)n
n= 1
X X
x(t) = n t M (n) (4.108)
n=1 n! dt t=0 2 n=1 n!
that the function x(t) is specified completely by the set of spectral moments. As a
result, X (! ) is also specified by this set of numbers. The moments offer a means
of characterizing the shape of the FRF or the corresponding Hilbert transform
in terms of a small set of parameters. Consider the analogy with statistical
theory: there, the mean and standard deviation (first- and second-order moments)
of a probability distribution establish the gross features of the curve. The third-
and fourth-order moments describe more subtle features—the skewness and the
‘peakiness’ (kurtosis). The latter features are considered to be measures of the
distortion from the ideal Gaussian form. The zeroth moment is also informative;
this is the energy or area under the curve.
Assuming that the moments are estimated for a single resonance between
!min and !max , the spectral moments of an FRF G(!) are
Z !max
(
MG = n ) d! !nG(!): (4.109)
!min
Note that they are complex, and in general depend on the limits; for consistency,
the half-power points are usually taken. The moments are approximated in
practice by
!X
MG(n)
max
( n ) MG(~n) MG(n)
HTD = 100
MG(n)
(4.111)
and these are simply the percentage differences between the Hilbert transform
moments and the original FRF moments.
In practice, only the lowest-order moments have been investigated; in the
terminology of [145], they are
real energy ratio (RER) = Re HTD (0)
imaginary energy ratio (IER) = Im HTD (0)
real frequency ratio (RFR) = Re HTD (1) :
172 The Hilbert transform—a practical approach
Figure 4.30. The variation in Hilbert transform describers (HTDs) for various SDOF
nonlinear systems.
N~ NG
imaginary amplitude ratio (IAR) = Im 100 G
NG
where Z !max
NG = d! G(!)2 (4.112)
!min
(which is essentially the centroid of the FRF about the ! -axis).
Measurement of apparent damping 173
Figure 4.30 shows the plots of the HTD statistics as a function of applied
force for several common nonlinearities. The parameters appear to separate
stiffness and damping nonlinearities very effectively. Stiffness nonlinearity is
identified from the changes in the RFR and IAR, while damping nonlinearity
is indicated by changes in the energy statistics without change in the other
describers. Note that the describers tend to zero at low forcing for the polynomial
nonlinearities as expected; G ~ ! G in this region. For the discontinuous
nonlinearities, clearance and friction, the describers tend to zero at high forcing
as the behaviour near the discontinuities becomes less significant. The describers
therefore indicate the level of forcing at which the FRF of the underlying linear
system can be extracted.
and
Hfe t cos(t)g = ie t sin(t) (4.114)
1 1
ah (t) = h(t) h~ (t) = e !n t sin(!dt) i e !nt cos(!d t)
m!d m!d
i ( !n +i!d )t
= e : (4.116)
m!d
174 The Hilbert transform—a practical approach
and this provides a new time-domain algorithm for estimating the damping of a
system, given the linear system FRF H (! ):
(1) Take the inverse Fourier transform of H (! ) to get the impulse response h(t).
(2) Take the Hilbert transform of h(t) and form the analytic impulse response
ah (t) as in (4.116).
(3) Plot the log magnitude of a h (t) against time; the gradient (extracted by a
linear regression) is = p! n . p
(4) If !d is measured, !n = 2 !n2 + !d2 = 2 + !d2 and = =!n .
at low and high levels of excitation. Figure 4.31 shows the corresponding log
envelopes. Extremely clear results are obtained in both cases. In contrast, the
corresponding FRFs with curve-fits are shown in figure 4.32. The high excitation
FRF is significantly noisier.
11 Note that using the conventional definition of analytic signal and Hilbert transform given in
footnote 4.6, equation (4.116) is modified to
Figure 4.31. Impulse response and envelope function for a nonlinear system under random
excitation: (a) low level; (b) high level.
characteristics of SDOF systems. There are essentially two approaches, one based
on free vibration FREEVIB and one on forced vibration FORCEVIB. They will
be discussed separately. Note that Feldman uses the traditional definition of the
analytic signal and time-domain Hilbert transform throughout his analysis.
178 The Hilbert transform—a practical approach
Figure 4.34. Data from the nonlinear beam in non-impacting condition: (a) measured
FRF; (b) calculated impulse response; (c) calculated envelope.
Identification of nonlinear systems 179
Figure 4.35. Data from the nonlinear beam impacting condition: (a) measured FRF;
(b) calculated impulse response; (c) calculated envelope.
180 The Hilbert transform—a practical approach
Figure 4.36. Results of the estimated natural frequency and apparent damping ratio for the
impacting cantilever: (a) linear regime; (b) nonlinear regime.
4.8.1 FREEVIB
Consider a SDOF nonlinear system under free vibration:
and
p
A(t) = y(t)2 y~(t)2 (4.124)
y~(t)
(t) = tan 1 : (4.125)
iy(t)
So both envelope and phase are available as functions of time if y (t) is known
and y~(t) can be computed. The derivatives can also be computed, either directly
or using the relations
" #
y(t)y_ (t) y~(t)y~_ (t) Y_ (t)
A_ (t) = p 2 2 = A(t) Re (4.126)
y(t) y~(t) Y (t)
" #
i(y ( t ) _ (t) y_ (t)~y (t))
y
~ _ (t)
Y
!(t) = _ (t) = = Im
y(t)2 y~(t)2
(4.127)
Y (t)
where ! (t) is the instantaneous frequency, again a real signal. The last two
equations can be used to generate the first two derivatives of the analytic signal
" #
A_ (t)
Y_ (t) = Y (t) + i!(t) (4.128)
A(t)
" #
A(t) 2 A_ (t)!(t)
Y (t) = Y (t) !(t) + 2i + i!_ (t) : (4.129)
A(t) A(t)
Now, consider the equation of motion (4.120), with h(y_ (t)) = h(t) and
!02 (y(t)) = !02 (t) considered purely as functions of time (there is a slight abuse
of notation here). Because the functions h and ! 02 will generally be low-order
polynomials of the envelope A, they will have a lowpass characteristic. If the
resonant frequency of the system is high, y (t) will, roughly speaking, have a
highpass characteristic. This means that h and y can be considered as non-
overlapping signals (see appendix C) as can ! 02 and y . If the Hilbert transform
182 The Hilbert transform—a practical approach
is taken of (4.120), it will pass through the functions h and ! 02 . Further, the
transform commutes with differentiation (appendix C again), so
Adding (4.120) and (4.130) yields a differential equation for the analytic
signal Y , i.e.
Y + h(t)Y_ + !02 (t)Y = 0 (4.131)
or, the quasi-linear form,
A_ !_
h(t) = 2 (4.134)
A !
A A_
!02 (t) = !2 h (4.135)
A A
or
A A_ 2 A_ !_
!02 (t) = !2 +2 2 + (4.136)
A A A!
and these are the basic equations of the theory.
On to practical matters. Suppose the free vibration is induced by an impulse,
the subsequent response of the system will take the form of a decay. y (t) can
be measured and y~ can then be computed 12. This means that A(t) and ! (t) are
available by using (4.124) and (4.125) and numerically differentiating (t).
Now, consider how the damping function is obtained. h(t) is known from
(4.134). As A(t) is monotonically decreasing (energy is being dissipated), the
inverse function t(A) is single-valued and can be obtained from the graph of
A(t) against time (figure 4.37). The value of h(A) is simply the value of h(t)
at t(A) (figure 4.38). Similarly, the stiffness function is obtained via the sequence
A ! t(A) ! !02 (t(A)) = !02 (A). The inverse of the latter mapping A(!)
is sometimes referred to as the backbone curve of the system. (For fairly simple
systems like the Duffing oscillator, the backbone curves can be calculated [41].)
12 As in the frequency-domain case, there are a number of methods of computing y~, the decomposition
H = F 1 Æ 2 ÆF provides one.
Identification of nonlinear systems 183
Envelope
t (A) t
Figure 4.37. Envelope used in Feldman’s method.
h ( t (A)) = h(A)
A
t (A) t
Figure 4.38. Damping curve for Feldman’s method.
Once h(A) and !02 (A) are known, the damper and spring characteristics
fd (A) and fs (A) can be obtained trivially
fd (A) = !(A)Ah(A) (4.137)
fs (A) = A!02 (A): (4.138)
Note that as there are no assumptions on the forms of f d and fs , the method
is truly non-parametric. However, once the graphs A ! f d etc have been
obtained, linear least-squares methods (as described in chapter 6) suffice to
estimate parameters.
The method can be readily illustrated using data from numerical
simulation13 . The first system is a Duffing oscillator with equation of motion
a
0.4
0.2
y(t), A(t)
−0.2
−0.4
0 0.2 0.4 0.6 0.8 1 1.2
b
30
25
f(t), Hz
20
15
10
0 0.2 0.4 0.6 0.8 1 1.2
Time, s
c d
0.45 0.45
0.4 0.4
0.35 0.35
0.3 0.3
0.25 0.25
A
0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
10 15 20 25 3 4 5 6 7
Frequency, Hz Damping coef., 1/s
Figure 4.39. Identification of cubic stiffness system: (a) impulse response; (b)
envelope; (c) backbone curve; (d) damping curve; (e) stiffness characteristic; (f ) damping
characteristic.
Identification of nonlinear systems 185
4
x 10 e f
1 800
0.8
600
0.6
400
0.4
200
0.2
Damping force
Spring force
0 0
−0.2
−200
−0.4
−400
−0.6
−600
−0.8
−1 −800
−0.5 0 0.5 −100 −50 0 50 100
Displacement Velocity
and initial condition y_ (0) = 200. Figure 4.39(a) shows the decaying displacement
and the envelope computed via equation (4.124). Figure 4.39(b) shows the
corresponding instantaneous frequency obtained from (4.127). The backbone and
damping curve are given in figures 4.39(c) and (d) respectively. As expected
for a stiffening system, the natural frequency increases with the amplitude of
excitation. Apart from a high-frequency modulation, the damping curve shows
constant behaviour. Using equations (4.138) and (4.139), the stiffness and
damping curves can be obtained and these are shown in figures 4.39(e) and (f ).
The second example shows the utility of the method for non-parametric
system identification. The system has a stiffness deadband, the equation of motion
is
y + 5y_ + fs (y) = 0 (4.140)
where 8
< 104 (y 0:1); y > 0:1
fs (y ) = 0;
: 4
jyj < 0:1 (4.141)
10 (y + 0:1); y < 0:1
and the motion began with y_ (0) = 200 once more. The sequence of figures
4.40(a)–(f ) show the results of the analysis. The backbone curve (figure 4.40(c))
shows the expected result that the natural frequency is only sensitive to the
nonlinearity for low levels of excitation. The stiffness curve (figure 4.40(e))
shows the size of the deadband quite clearly. (This is useful information, if
the clearance is specified, the parameter estimation problem becomes linear and
186 The Hilbert transform—a practical approach
0.5
y(t), A(t)
−0.5
20
f(t), Hz
15
10
5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Time, s
c d
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
A
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
Figure 4.40. Identification of backlash system: (a) impulse response; (b) envelope; (c)
backbone curve; (d) damping curve; (e) stiffness characteristic, (f ) damping characteristic.
Identification of nonlinear systems 187
e f
8000 600
6000
400
4000
200
2000
Damping force
Spring force
0 0
−2000
−200
−4000
−400
−6000
−8000 −600
−1 −0.5 0 0.5 1 −100 −50 0 50 100
Displacement Velocity
simple methods suffice to estimate the stiffness function.) Note that because the
initial displacement did not decay away completely, there are gaps in the stiffness
and damping functions at low amplitude.
The final example shows a damping nonlinearity. The system has equation
of motion
y + 300 sgn(y_ ) + 104y = 0 (4.142)
so Coulomb friction is present. The decay began with the same initial conditions
as before and the resulting anlysis is shown in figures 4.41(a)–(f ). Note the
characteristic linear decay envelope for this type of nonlinear system as shown
in figure 4.41(a). In this case, the backbone (figure 4.41(c)) shows no variation of
natural frequency with amplitude as expected. The coefficient of friction can be
read directly from the damping function (figure 4.41(f )).
Further examples of nonlinear systems can be found in [93, 95]. A practical
application to a nonlinear ocean mooring system is discussed in [120].
All of these examples have viscous damping models. It is a simple matter to
modify the theory for structural (hysteretic) damping, the equation of motion for
the analytic signal becomes
i
Y + !02(A) 1 + Æ(A) Y = 0 (4.143)
where Æ (A) is the loss factor or logarithmic decrement. The basic equations are
A
!02 (t) = !2 (4.144)
A
188 The Hilbert transform—a practical approach
1
y(t), A(t)
−1
−2
20
f(t), Hz
15
10
2.5 2.5
2 2
A
1.5 1.5
1 1
0.5 0.5
10 15 20 2 4 6
Frequency, Hz Damping coef., 1/s
Figure 4.41. Identification of Coulomb friction system: (a) impulse response; (b)
envelope; (c) backbone curve; (d) damping curve; (e) stiffness characteristic; (f ) damping
characteristic.
Identification of nonlinear systems 189
4
x 10 e f
3
600
2 400
1 200
Damping force
Spring force
0 0
−1 −200
−2 −400
−600
−3
−4 −2 0 2 4 −200 −100 0 100 200
Displacement Velocity
and
_
2A! !_
Æ(t) = :
A!02 !02
(4.145)
The method described here is only truly suitable for monocomponent signals,
i.e. those with a single dominant frequency. The extension to two-component
signals is discussed in [96].
4.8.2 FORCEVIB
The analysis for the forced vibration case is very similar to FREEVIB; the
presence of the excitation complicates matters very little. Under all the same
assumptions as before, the quasi-linear equation of motion for the analytic signal
can be obtained:
X
Y + h(A)Y_ + !02 (A)Y = : (4.146)
m
Carrying out the same procedures as before which lead to equations (4.134) and
(4.135) yields
(t) A_ !_
h(t) = 2 (4.147)
!m A !
and
(t) (t)A_ A A_ 2 A_ !_
!02 (t) = !2 + +2 2 + (4.148)
m A!m A A A!
190 The Hilbert transform—a practical approach
where (t) and (t) are, respectively, the real and imaginary parts of the
input/output ratio X=Y , i.e.
[C ] = [A][][A]T (4.151)
where [] is diagonal. (Singular value decomposition can be used for this step
[209].) The transformation to principal components is then
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
combinations of the original, yet still encode all the dynamics. This was the idea
of Leuridan.
In terms of sampled data, there would be N samples of fy (t)g taken at
regular intervals t. These will be denoted fy (t i )g; i = 1; : : : ; N . The signals
observed from structures are usually zero-mean, so the covariance matrix for the
system is
X N
[] = fy(ti )gfy(ti )gT : (4.153)
i=1
It is not particularly illuminating to look at the principal time signals.
Visualization is much simpler in the frequency domain. The passage from time
to frequency is accomplished using the multi-dimensional version of Parseval’s
Principal component analysis (PCA) 193
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
Z 1
[] = dt fy(t)gfy(t)gT: (4.154)
1
Z 1
1 1
Z Z 1
1
[] = dt i
d! e fY (!1 )g
! 1t
d! e i !2t
fY (!2 )g T
1 2 1 1 2 1 2
(4.155)
194 The Hilbert transform—a practical approach
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
where the reality of the time signals has been used. Rearranging yields
1 1 1
Z Z Z 1
T 1 i( 1 !2 )t
[] = d! d! fY (!1 )gfY (!2 )g dte ! :
2 1 1 1 2 2 1
(4.156)
Now, using the integral representation of the Æ -function from appendix D, one
finds
1 1 1
Z Z
[] = d! d! fY (!1 )gfY (!2 )gT Æ(!1 !2 )
2 1 1 1 2
(4.157)
Principal component analysis (PCA) 195
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
Figure 4.46. Corrected principal FRF P H1 for symmetric 2DOF linear system.
1 1
Z
[] = d! fY (!1 )gfY (!1 )gT
2 1 1
(4.158)
and the transformation which decorrelates the time signals also decorrelates the
spectra. (In (4.158) the overline refers to the complex conjugate and not the mean.
In order to avoid confusion with complex quantities, the mean will be expressed
in the rest of this section using the expectation operator, i.e. x = E [x].)
Now suppose the system is excited at a single point with a white excitation so
196 The Hilbert transform—a practical approach
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
Figure 4.47. Corrected principal FRF P H2 for symmetric 2DOF linear system.
0.0002
Receptance
0.0001
0.0000
0.0 10.0 20.0 30.0 40.0 50.0 60.0
Frequency (Hz)
0.0003
X = 1.0
X = 5.0
X = 10.0
0.0002
Receptance
0.0001
0.0000
0.0 10.0 20.0 30.0 40.0 50.0 60.0
Frequency (Hz)
Figure 4.49. FRF 1 for symmetric 2DOF nonlinear system at low medium and high
excitation.
X2
N=
[] = fH (!i )gfH (!i )gT (4.162)
i=1
198 The Hilbert transform—a practical approach
0.0003
X = 1.0
X = 5.0
X = 10.0
0.0002
Receptance
0.0001
0.0000
0.0 10.0 20.0 30.0 40.0 50.0 60.0
Frequency (Hz)
Figure 4.50. FRF 2 for symmetric 2DOF nonlinear system at low medium and high
excitation.
0.0004
X = 1.0
0.0003 X = 5.0
X = 10.0
Receptance
0.0002
0.0001
0.0000
0.0 10.0 20.0 30.0 40.0 50.0 60.0
Frequency (Hz)
Figure 4.51. Principal FRF P 1 for symmetric 2DOF nonlinear system at low medium
and high excitation.
the results are as shown in figures 4.44 and 4.45. The decomposition appears to
have almost produced a transformation to modal coordinates, both FRFs are only
mildly distorted versions of SDOF FRFs. In fact in this case, the distortions are
simple to explain.
The previous argument showed that the principal component transformation
for time data also decorrelated the FRF vector. However, this proof used integrals
Principal component analysis (PCA) 199
0.0003
X = 1.0
X = 5.0
X = 10.0
0.0002
Receptance
0.0001
0.0000
0.0 10.0 20.0 30.0 40.0 50.0 60.0
Frequency (Hz)
Figure 4.52. Principal FRF P 2 for symmetric 2DOF nonlinear system at low medium
and high excitation.
with infinite ranges. In practice, the covariance matrices are computed using
finite summations. In the time-domain case, this presents no serious problems
in applying (4.153) as long as the records are long enough that the means of the
signals approximate to zero. However, in the frequency domain, the FRFs are
not zero-mean due to the finite frequency range. This means that the covariance
matrix in (4.162) is inappropriate to decorrelate the FRF vector. The remedy is
simply to return to equation (4.150) and use the covariance matrix
X2
N=
[] = (fH (!i )g E [fH (!i )g])(fH (!i )g E [fH (!i )g])T : (4.163)
i=1
Using this prescription gives the principal FRFs shown in figures 4.46 and
4.47. This time the principal component transformation has produced modal
FRFs. Unfortunately, this situation is not generic. It is the result here of
considering a system with a high degree of symmetry; also the mass matrix is
unity and this appears to be critical. Figure 4.48 shows the principal FRFs for
a system identical to (4.160) and (4.161) except that the two equations have
different mass values—the decoupling property has been lost even though the
modal transformation can still achieve this. However, throughout the development
of the PCA method it was hoped that the principal FRFs would generally exhibit
some simplification.
In terms of nonlinear systems, the aim of PCA (or as it is sometimes called—
the Karhunen–Loeve expansion [257]) is to hopefully localize the nonlinearity in
a subset of the responses. By way of illustration consider the system in (4.160)
and (4.161) supplemented by a cubic stiffness nonlinearity connecting the two
200 The Hilbert transform—a practical approach
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
Figure 4.53. Principal FRF P 1 for symmetric 2DOF nonlinear system with Hilbert
transform.
masses
The FRFs for the system at a number of different levels of excitation are
given in figures 4.49 and 4.50. The distortion is only shown on the second mode
as this is the only nonlinear mode (as discussed in section 3.1). When the principal
FRFs are computed (figures 4.51 and 4.52), only the second principal FRF shows
the distortion characteristic of nonlinearity. Again one should not overemphasize
these results due to the high symmetry of the system.
Principal component analysis (PCA) 201
Magnitude
Frequency (Hz)
Phase (rad)
Frequency (Hz)
Figure 4.54. Principal FRF P 2 for symmetric 2DOF nonlinear system with Hilbert
transform.
The reason for the presence of this section in this chapter is that any test
for nonlinearity can be applied to the principal FRFs including of course the
Hilbert transform. This has been studied in the past by Ahmed [7] amongst
others. Figures 4.53 and 4.54 show the result of applying the Hilbert transform
to the principal FRFs for the system discussed earlier. As one might expect, the
nonlinearity is only flagged for the second mode.
With that brief return to the Hilbert transform the chapter is concluded.
The Hilbert transform has been seen to be a robust and sensitive indicator of
nonlinearity. It is a little surprising that it has not yet been adopted by suppliers
of commercial FRF analysers. The next chapter continues the Hilbert transform
theme by considering an approach to the analysis which uses complex function
theory.