Vous êtes sur la page 1sur 10

A combined genetic and eigensensitivity algorithm for the

location of damage in structures


M.I. Friswell
a,
*, J.E.T. Penny
b
, S.D. Garvey
b
a
Department of Mechanical Engineering, University of Wales Swansea, Swansea SA2 8PP, U.K.
b
Department of Mechanical and Electrical Engineering, Aston University, Birmingham B4 7ET, U.K.
Received 20 January 1997; accepted 28 April 1998
Abstract
Genetic algorithms have been the subject of considerable interest in recent years, since they appear to provide a
robust search procedure for solving dicult problems. Due to the way the genetic algorithm explores the region of
interest it avoids getting stuck at a particular local minimum and locates the global optimum. The genetic algorithm
is slow in execution and is best applied to dicult problems. This paper applies a genetic algorithm to the problem
of damage detection using vibration data. The objective is to identify the position of one or more damage sites in a
structure, and to estimate the extent of the damage at these sites. The genetic algorithm is used to optimize the
discrete damage location variables. For a given damage location site or sites, a standard eigensensitivity method is
used to optimize the damage extent. This two-level approach incorporates the advantages of both the genetic
algorithm and the eigensensitivity methods. The method is demonstrated on a simulated beam example and an
experimental plate example. # 1998 Elsevier Science Ltd. All rights reserved.
Keywords: Damage location; Genetic algorithm; Eigensensitivity
1. Introduction
Genetic algorithms are now frequently applied to
problems of maximising or minimizing a given objec-
tive function, often subject to some constraints.
Genetic algorithms have been applied to a wide range
of optimization problems in engineering which have
this form, such as the design of structures [14], pipe-
line optimization [5], nite element model updating [6
8] and the problem of choosing the number and lo-
cation of actuators for the control of large exible
space structures [9]. This paper approaches the pro-
blem of damage location by formulating an objective
function composed of measured and analytical vi-
bration data and applying a genetic algorithm to opti-
mize it. Other evolutionary strategies [10] and
algorithms such as simulated annealing [7, 11] have
been used for design optimization and model updating.
In recent years, there has been a considerable
demand for more accurate techniques to detect and
locate damage, particularly in large structures.
Damage will cause the stiness distribution in the
structure to change which may be detected by measur-
ing its dynamic response. Doebling et al. [12] gave an
overview of damage detection and location. Three dis-
tinct philosophies have been considered to locate the
damage in a structure using measured vibration data.
The rst group of methods uses techniques of nite el-
ement model updating and error localization to obtain
a corrected set of physical parameters that reproduce
the measured data. Mottershead and Friswell [13] gave
a survey of model updating methods, many of which
can also be used for damage location. The direct
model updating approach, where the whole stiness
matrix is updated, does not require parameterization
of the nite element model [14]. Otherwise, the model
Computers and Structures 69 (1998) 547556
0045-7949/98/$ - see front matter # 1998 Elsevier Science Ltd. All rights reserved.
PII: S0045- 7949( 98) 00125- 4
PERGAMON
* Corresponding author. E-Mail: M.I.Friswell@swansea.
ac.uk.
is parameterized and all the parameters
updated [15, 16]. Generally, the identication will be
underdetermined and some form of regularization
should be employed. Genetic algorithms have also
been used to identify damage in structures by optimiz-
ing the force residual [17] or modal residuals [18].
A second group of methods, called forward
methods, assume a candidate set of possible damage
scenarios, which can include both damage mechanism
and location. The change in dynamic response due to
the damage is predicted, usually in terms of the natural
frequencies. The predicted change in response of the
structure for all the damage scenarios is then compared
to the measured change and the closest damage case is
chosen. Statistical methods are often used to identify
the most likely location of the damage. The dierence
between these methods and those from standard model
updating is that only a subset of parameters are
assumed to be in error. Cawley and Adams [1921]
used this type of method and found that damage in
specimens fabricated from composite materials could
be detected. The method of Cawley and Adams has
the advantage that damage location is identied separ-
ately from damage extent, although this only works
for a single damage site. Friswell et al. [22] proposed a
similar method but used a more rened statistical
analysis. A similar approach is to directly nd the best
subset of parameters that models the damage [23].
Neural networks are able to treat damage mechan-
isms implicitly, so that it is not necessary to model the
structure in so much detail [2430]. The method can
easily deal with non-linear damage mechanisms.
Models are still required to provide the training cases
for the networks, and the algorithms should be robust
against systematic errors between the model used for
training and the actual structure. For success, neural
networks require that the essential features in the
damaged structure are represented in the training data.
2. Overview of genetic algorithms
The striking feature of genetic algorithms is that
they are based on ideas from the science of genetics
and the process of natural selection. This cross fertili-
zation from one eld of science to another has led to
stimulating and fruitful applications in many elds and
particularly in computer science.
The genetic algorithm works with an initial popu-
lation which may, for example, correspond to numeri-
cal values of a particular variable. The size of this
population may vary and is generally related to the
problem under consideration. The members of this
population are usually strings of zeros and ones i.e.
binary strings. For example a small initial or rst gen-
eration population may take the form:
1000010
1110000
1010101
1111001
1000001:
In practice, the population may be far larger than this
and the strings longer. The strings themselves may be
the encoded values of a variable or variables that we
are examining. This initial population is generated ran-
domly and we can use the terminology of genetics to
characterize it. Each string in the population corre-
sponds to a chromosome and each binary element of
the string to a gene. A new population must now
develop from this initial population and to do this we
implement the analogue of specic fundamental genetic
processes. These are:
. reproduction based on tness;
. crossover;
. mutation.
A set of chromosomes is selected at the reproduction
stage based on natural selection. Thus, members of the
population are chosen for reproduction on the basis of
their tness dened according to some specied cri-
teria. The ttest are given a greater probability of
reproducing in proportion to the value of their tness.
The actual process of mating is implemented by
using the simple idea of crossover. This means that
two members of the population exchange genes. There
are many ways of implementing this crossover, for
example having a single crossover point or many cross-
over points. These crossover points are selected ran-
domly. A simple crossover is illustrated below for two
chromosomes selected according to tness. Here we
have randomly selected a crossover point after the
fourth digit.
1110 + 000;
1010 + 101:
After crossover this gives
1110 + 101;
1010 + 000:
Applying this procedure to our original population we
produce a new generation.
The nal process is mutation. Here, we randomly
change a particular gene in a particular chromosome.
Thus, 0 may be changed to a 1 or vice versa. The pro-
cess of mutation in a genetic algorithm occurs very
rarely and hence this probability of a change in a
string is kept very low. Mutation ensures genetic diver-
M. Friswell et al. / Computers and Structures 69 (1998) 547556 548
sity is maintained by generating new, random chromo-
somes.
The application of genetic algorithms to optimiz-
ation problems is relatively clear since the test of t-
ness can be based on the function we are trying to
maximize, thus the members of the population which
give the largest values of the objective function are
taken as the ttest. The chromosomes are viewed as a
concatenation of binary strings representing the values
individual variables in the optimization problem.
Continuous variables must be discretized, as described
later. Constraints are usually incorporated into the
objective function using a penalty function which
ensures that during the optimization process the sol-
ution will satisfy the constraints. This penalty must be
carefully chosen to ensure the constraints are satised
and so the correct optimum is reached.
The reason why a genetic algorithm diers from a
simple direct search procedure is that it involves two
special features: crossover and mutation. Thus, starting
from an initial population the algorithm develops new
generations which rapidly explore the region of inter-
est. This is useful for dicult optimization problems
and in particular for those where we wish to nd the
global maximum or minimum of function which has
many local maxima and minima. In this case, standard
optimization methods, such as the conjugate gradient
method, can locate only the local optimum. A genetic
algorithm, however, may locate the global optimum
although this is not guaranteed. This is due to the way
it explores the region of interest avoiding getting stuck
at a particular local minimum. Since we are dealing
with a stochastic process, the solution should be taken
as the best of a number of runs.
Genetic algorithms are a developing area of research
and many amendments could be made to the functions
which we have used to implement a genetic algorithm.
For example, the weighted roulette wheel selection can
be implemented in many dierent ways, crossover can
be changed to multi-point crossover or other alterna-
tives. Elitism, where the best solution is always passed
on to the next generation, is a particular feature for
which good results have been reported. An alternative
coding for the variables of the problem such as Gray
coding have also been used by workers in the eld. It
will often be noticed that a genetic algorithm is slow in
execution, but it is usually applied to problems where
standard optimization techniques perform poorly; for
example those which have multiple optima and where
the global optimum is required. Problems which have
both real valued and integer variables pose particular
problems for standard algorithms. Since standard al-
gorithms often fail in these cases, the extra time taken
by the genetic algorithm is worthwhile. Having
described the basic principles of a genetic algorithm we
will now illustrate how it may be applied by consider-
ing an engineering application.
3. Penalty functions for damage location
The location of damage in a structure using vi-
bration data is very dicult. We will make the
assumption that damage causes a change in the sti-
ness of the structure. In many cases this is a gross
assumption and the distinction between strength and
stiness should be made. For example, in a concrete
highway bridge the concrete is held in compression by
steel tendons and the stiness of the bridge comes pre-
dominantly from the concrete. The strength of the
bridge comes from the steel tendons, since concrete has
little strength or stiness in tension. Thus, as a bridge
is damaged its dynamic response will change very little
until just before failure.
The simplied problem to be addressed by this
paper is to identify the location and extent of a limited
number of stiness reductions that best reproduce the
measured dynamics. The stiness reductions will be
modelled by a reduction in the Young's modulus for a
given element, and its location is specied by the el-
ement number. Localization of the damage is depen-
dent on a suciently rened mesh. A very ne mesh
could be used with a tness penalty included to ensure
that adjacent elements are chosen. The examples that
follow only include beam and plate examples, although
any structure with its associated nite element model
may be used. A maximum of two damage sites will be
allowed, although the extension to more sites is
straightforward. The measured data will be the natural
frequencies and mode shapes of the structure.
We next consider the objective function which is
based on the measured data. This will consist of three
terms: a term relating to the error in natural frequen-
cies, J
o
, a term relating to the error in mode shapes,
J
f
, and a term to weight against two damage sites.
These terms will be weighted to give a total error of
J = W
o
J
o
W
f
J
f
W
ns
d
ns
(1)
where W
o
, W
f
, W
ns
are weighting factors and
d
ns
=
0 if one site is damaged,
1 if more than one site is damaged.

The term added to weight against too many damage


sites is very important. Suppose damage only occurs at
one site. Due to measurement noise and model inade-
quacies, the modal model will not be reproduced
exactly, even if the damage is correctly identied.
Introducing a second damage site may make a minor
improvement in the correlation between test and analy-
sis, although in reality one site is damaged. The ten-
M. Friswell et al. / Computers and Structures 69 (1998) 547556 549
dency will always be to nd damage at every site.
Thus, a penalty is introduced to weight against an
increased number of damage sites.
3.1. The frequency error
The obvious objective function for frequency errors
is just a weighted sum of squares of the dierences in
the natural frequencies, provided the modes are paired
correctly. Any natural frequencies corresponding to
analytical and experimental modes that do not pair
with sucient condence are simply omitted from the
objective function. Thus, one possibility for J
o
is,
J
o
=
X
r
j=1
W
oj
o
mj
o
aj
o
mj
!
2
(2)
where o
mj
and o
aj
are the jth measured and analytical
natural frequency of the damaged structure, and r is
the number of measured modes. If required, the indi-
vidual frequencies may be given dierent weights,
denoted by W
oj
, although often these weights are
taken as unity. The analytical natural frequencies, o
aj
,
could be used in place of the measured frequencies in
the denominator in Eq. (2), although the dierences in
these frequencies should be small and using the ana-
lytical frequencies makes the sensitivity calculations
much more dicult.
The problem with Eq. (2) is that the model for the
undamaged structure should correlate well with the
measured dynamics if the structure was undamaged. If
periodic measurements of a structure are undertaken,
so that measurements of the undamaged structure
exist, then the model of the undamaged structure may
be updated to correspond closely to the undamaged
structure. If this updating is not performed, then the
algorithm to identify the location and extent of the
damage will try to compensate for both the original
modelling error and the change due to the damage.
Obviously, the identication of damage in this case is
very dicult.
A dierent method to reduce the eect of modelling
errors is to consider changes in frequency from the
undamaged to the damaged structure. Of course, this
assumes that the natural frequencies of the undamaged
structure were measured. In this case, J
o
is given by
J
o
=
X
r
j=1
W
oj
do
mj
do
aj
o
mj
!
2
(3)
where do
mj
and do
aj
are the changes, due to damage,
in the jth measured and analytical natural frequencies.
Eq. (3) gives the penalty function that will be used in
subsequent sections.
One possible disadvantage with Eqs. (2) and (3) is
that the range of the penalty function is not known.
Using the penalty function
J
o
=
X
r
j=1
W
oj
o
2
mj
o
2
aj
o
2
mj
o
2
aj
!
2
(4)
in place of Eq. (2) overcomes this problem since the in-
dividual natural frequency error terms are between 0
and 1. This may not be a great advantage, since
Eqs. (2) and (4) are approximately the same when the
dierences between the measured and analytic frequen-
cies are small. Calculating the sensitivities for Eq. (4) is
also dicult. The problem of the relative weighting
between the frequencies and mode shapes will be
addressed later. Expressions similar to Eq. (4), but
using frequency changes, may also be dened, and
instead of squaring the terms the absolute value may
be taken.
3.2. The mode shape error
The error in the mode shapes may be incorporated
into a penalty function as
J
f
=
X
r
j=1
W
fj
(f
mj
f
aj
)
T
(f
mj
f
aj
) (5)
where f
mj
and f
aj
are the jth measured and analytical
mode shapes. Only those degrees of freedom that are
measured are picked out of the analytical mode shapes.
To apply Eq. (5) correctly, both sets of mode shapes
should be normalized consistently, and typically mass
normalization is used. Alternatively, the modes may be
normalized using the modal scale factor (MSF) [31].
Using the MSF also makes sure that no modes are
1808 out of phase. Of course the modes must also be
real if the nite element model does not include damp-
ing. The terms in Eq. (5) can also be weighted using
the analytical mass or stiness matrix, reduced to the
measured degrees of freedom if necessary.
An alternative method that does not require the
mode shape scaling, is to use the modal assurance cri-
terion (MAC) [31]. Thus, the new penalty function is
J
f
=
X
r
j=1
W
fj
[1 MAC(f
mj
; f
aj
)] (6)
where MAC(f
mj
, f
aj
) is the MAC value for the jth
paired mode. This penalty function has the advantage
that each term is between 0 and 1, since the MAC
values are between 0 and 1. It is dicult to use penalty
functions based on the MAC with eigensensitivity
methods, since the derivative of the penalty function is
dicult and costly to obtain. The [1 MAC(f
mj
, f
aj
)]
terms may also be square rooted.
M. Friswell et al. / Computers and Structures 69 (1998) 547556 550
3.3. The relative weighting of frequencies and mode
shapes
How much weight should be given to the errors in
the natural frequencies, compared to the errors in the
mode shapes? Although the mode shapes can provide
valuable information, one has to be careful. The mode
shapes are measured with less accuracy than the natu-
ral frequencies, and generally the mode shapes are less
sensitive to damage in the structure. As an example,
consider a uniform cantilever beam with 15 elements.
Let the stiness of the fourth element from the
clamped end be reduced to simulate damage. Figs. 1
and 2 give the change in the natural frequency and
mode shape (through the MAC value) for the rst ve
modes as the element stiness is reduced. Obviously,
the stiness reduction has a relatively large eect on
the natural frequencies, as compared with the MAC
values. Because of the insensitivity of the modes to the
damage, only the natural frequency penalty function
will be used in the following examples.
4. The application of the genetic algorithm
Let us rst consider the parameters to be optimized
by the genetic algorithm. One advantage of the genetic
algorithm is the ability to optimize a combination of
discrete and continuous variables. In this application
we have both discrete variables, the location of the
damage given by the element number, and continuous
variables, the extent of the damage given as a percen-
tage reduction in stiness.
These continuous variables must be expressed in
terms of a xed number of binary digits. Thus, in rea-
lity, genetic algorithms optimize discrete variables, and
continuous variables are approximated to the required
accuracy. For example, the damage extent variables
are in the range 0100% and if 8 bits are allocated to
represent these variables, the maximum resolution is
100/102310.1%. The location variables, or element
numbers, are already discrete variables. Zero will be
included in this variable set to denote no damage.
Thus, if there are 15 elements, or possible damage lo-
cations, in the nite element model then the location
variable will range from 0 to 15, requiring a 4 bit
binary number. If the model contains less elements,
then some numbers would have to be assigned a zero
tness. For example, if there were only 12 elements,
then the objective function would be zero for the vari-
able values 13, 14 and 15.
The objective is to minimize the function given in
Eq. (1), whereas genetic algorithms try to nd the t-
test members of the population, or maximize an objec-
tive function. This is easily rectied by subtracting J,
dened in Eq. (1) from a suitable number. Negative t-
ness values will be taken to be zero. The choice of this
number is critical as it gives the range of tness values
for the algorithm, which is used to generate the prob-
ability that a particular member of a generation will be
selected for the next generation. In general, the most
satisfactory results are obtained when the population
has a good spread of tness values, from zero to the
ttest. The problems that occur are easily shown by an
example. Suppose the objective function we wish to
minimize takes values between 0 and 1. If we subtract
these values from 100 we have tness values between
99 and 100. The tness values are approximately
equal, and all members of a generation will have ap-
proximately equal probability of survival to the next
generation. This is not desirable, as we want only the
ttest to survive. If, instead, we subtract the value of
the objective function from 1 we obtain tness values
between 0 and 1, and a more satisfactory performance
of the algorithm. The direct application of genetic al-
gorithms to damage detection was discussed by the
authors in an earlier paper [32].
Fig. 1. The eect of damage on the rst ve natural frequen-
cies of the cantilever beam (mode 1 w; 2 r; 3 q; 4 X; 5 +).
Fig. 2. The eect of damage on the MAC for the rst ve
modes of the cantilever beam (mode 1 w; 2 r; 3 q; 4 X; 5
+).
M. Friswell et al. / Computers and Structures 69 (1998) 547556 551
5. Combining eigensensitivity and genetic algorithms
Genetic algorithms may be used to directly minimize
the penalty function given in Eq. (1), by forcing the
continuous variables to take on a nite number of dis-
crete values. Genetic algorithms are best used to opti-
mize dicult problems. In damage location and extent
estimation most of the diculty arises from trying to
locate the damage. Given one or two damage lo-
cations, then the optimum values for the extent of the
damage are easily calculated from sensitivity type
methods, based on a Taylor series expansion of the
natural frequencies and mode shapes in terms of the
unknown extent variables (see for example, Ref. [33]).
In this way the length of the binary string representing
each member of the generation is considerably
reduced. In practice the penalty functions given by
Eq. (3) for frequency and Eq. (5) for mode shapes, are
most appropriate for the eigensensitivity method.
6. A simulated cantilever beam example
The location of damage using the combined eigen-
sensitivity and genetic algorithm method was demon-
strated using a simulated steel, cantilever beam of
length 1 m and cross section 2550 mm. Bending in
the plane of the thinner dimension only, was con-
sidered, that is in the most exible of the two bending
planes. The changes in the rst 5 natural frequencies
were used to locate the damage, and thus the penalty
function is given by Eq. (3).
Four cases were tested: case 1 demonstrated the per-
formance of the method to locate a single damage site;
in case 2 the damage was to an element near the free
end of the cantilever and hence dicult to locate; in
case 3 the damage was located at two sites; case 4
demonstrated the performance of the method to sys-
tems with systematic errors, namely the addition of
extra mass into the ``measured'' system.
In all cases the population had 10 members, 60% of
which were mated at every generation. Each binary bit
had a 0.5% chance of mutation at every generation. In
all cases, the nite element model used to locate the
damage contained 15 elements, or 30 degrees of free-
dom. These elements were numbered from the xed
end. Only one run of the genetic algorithm was used
for each optimization. Mating was accomplished using
a single point crossover, with a randomly generated
crossover point. The variables used for the genetic al-
gorithm optimization were the location of two damage
sites. For a given set of damage sites the extent of the
damage was optimized using the eigensensitivity
approach. The genetic algorithm has been modied so
that the most t member of a population survives into
the next generation, without mating or mutation. This
is called elitism and ensures that the ``best'' solution is
maintained from generation to generation. The average
tness is also shown in the plots, although care must
be exercised in interpreting these plots. The optimum
was always selected as the best member of a gener-
ation, and the convergence of the average tness
should not be taken as any sort of convergence to the
optimum. The increasing average tness values do indi-
cate a reduction in the genetic diversity of the popu-
lation.
6.1. Case 1
The simulated data was obtained from a nite el-
ement model with 15 elements. The model used to
locate the damage was the same as that used to simu-
late the measurements. Damage was simulated as a
30% reduction in the stiness of element 4. Table 1
shows the damaged and undamaged natural frequen-
cies. The weights given to the terms in the objective
function were
W
oj
= 100; W
ns
= 0:25
and the objective function, Eq. (1), was subtracted
from 1, giving tness values between 0 and 1.
Remember, negative values of tness are taken as zero.
Fig. 3 shows the tness of the most t member of
the population, and also the average tness of the
population. The damage was correctly located at el-
ement 4 at the sixth generation. The average tness
increased with the generation number.
Table 1
The damaged and undamaged natural frequencies for cases 1 and 4
Natural frequencies (Hz)
Without added mass (case 1) With added mass (case 4)
Mode number Undamaged Damaged Undamaged Damaged
1 21.0 20.4 20.5 20.0
2 131.3 131.2 131.3 131.2
3 367.7 362.2 365.5 360.1
4 720.7 705.7 709.2 694.5
5 1191.7 1181.7 1176.0 1165.5
M. Friswell et al. / Computers and Structures 69 (1998) 547556 552
6.2. Case 2
Damage was simulated by a 30% reduction in sti-
ness at element 14. The cantilever beam model has
only 15 elements, and the damage was very close to
the free end of the cantilever beam. Changes in sti-
ness in this region have very little inuence on the
lower modes of the beam. Fig. 4 shows the results in
this case. The genetic algorithm correctly located the
damage, although the high average tness even at the
rst generation shows that many estimated damage lo-
cations tted the measured data reasonably well.
6.3. Case 3
Damage was simulated at two locations, by 30%
stiness reductions at elements 4 and 12, and the
results are shown in Fig. 5. The algorithm has less suc-
cess here but still located the damage at elements 5
and 12, which was very close to the actual damage lo-
cations. Note that the maximum tness expected in
this case was 0.75, since damage was located at 2 sites.
6.4. Case 4
This case demonstrated the eect of a systematic
error in the ``measurements''. Substantial discretization
errors in the nite element model are often found in
practice, although in our example these were relatively
small because the original nite element model was
suciently rened. The systematic error introduced
into the 15 element cantilever beam model was a mass
of 0.2 kg added at node 12. Table 1 shows the error in
the rst 5 natural frequencies of the damaged and
undamaged model, and shows the level of systematic
error introduced.
The evolution of the tness values is shown in Fig. 6.
The damage was correctly located in element 4, despite
the systematic error in the ``measured'' data. The per-
formance in many respects was better than that in case
1, without the systematic error. This is probably
because only a single run of the algorithm was per-
formed.
Fig. 3. Fitness values for the simulated beam problem, case
1(most t member ; average tness - - - - -).
Fig. 4. Fitness values for the simulated beam problem, case
2(most t member ; average tness - - - - -).
Fig. 5. Fitness values for the simulated beam problem, case
3(most t member ; average tness - - - - -).
Fig. 6. Fitness values for the simulated beam problem, case
4(most t member ; average tness - - - - -).
M. Friswell et al. / Computers and Structures 69 (1998) 547556 553
7. An experimental example
The second example consists of a 3 mm steel plate
which was clamped along one edge. The exposed plate
measured 305 mm 357 mm and was clamped along
the shorter side. Fig. 7 shows the nite element mesh
used to model the undamaged plate. This mesh
resulted in a 48 degree of freedom model and produced
the estimated natural frequencies given in Table 2.
Table 2 also gives the natural frequencies of the plate
measured experimentally using impact excitation. The
12th mode was not measured because the acceler-
ometer was placed at a node of this mode. The analyti-
cal and experiment modes were easily paired in this
case using the displayed mode shapes. Clearly, the
nite element mesh was very coarse and there were
considerable modelling and discretization errors pre-
sent. Any damage location method should be robust to
such errors, and, to increase the robustness, the dier-
ence in natural frequencies between the undamaged
and damaged plate was used to locate the damage.
The plate was damaged by 4 saw cuts each of length
40 mm in the shape of a cross centred on element 4 of
the 48 degree of freedom nite element model, as
shown in Fig. 7.
The population had 10 members, 60% of which
were mated at every generation. Each binary bit had a
0.5% chance of mutation at every generation. The
weights given to the terms in the objective function
were
W
oj
= 1000; W
ns
= 0:25
and the objective function, Eq. (1), was subtracted
from 1, giving tness values between 0 and 1. Negative
values of tness were taken as zero. A maximum of 2
damage locations were assumed, and each location
variable was 4 bits long; zero was taken as no damage,
integers 112 as damage in the corresponding element
and integers 13, 14 and 15 were given a zero tness
value. Fig. 8 shows the tness of the best member of
Fig. 7. The experimental cantilever plate example, showing
the nite elements and the position of the saw cut.
Table 2
Analytical and experimental natural frequencies and frequency changes due to damage for the experimental cantilever plate
example
Mode number
Analytical undamaged natural
frequency (Hz) (48 degrees of
freedom model)
Experimental undamaged
natural frequency (Hz)
Experimental natural frequency
change due to damage (Hz)
1 19.1 20.0 0.4
2 55.7 56.7 1.0
3 119.7 124.6 3.8
4 192.8 198.1 4.7
5 214.6 212.6 3.0
6 341.6 353.9 10.9
7 367.6 380.8 3.0
8 405.6 427.5 12.0
9 544.7 530.1 9.0
10 602.3 639.9 17.5
11 663.6 690.7 24.5
12 665.8 Not measured Not measured
13 705.7 774.5 25.2
Fig. 8. Fitness values for the experimental cantilever plate
example(most t member ; average tness - - - - -).
M. Friswell et al. / Computers and Structures 69 (1998) 547556 554
the population in each generation. The correct location
of the damage (element 4) was found in the 8th gener-
ation, showing the approach is robust to modelling
and measurement errors.
8. Conclusions
A combined genetic algorithm and eigensensitivity
method has been used to identify the location and
magnitude of damage from measured vibration data.
Essentially, the genetic algorithm is used to identify
the damage located and the eigensensitivity is used to
identify the damage extent. Damage at one and two
sites have been successfully located in the simulated
example of a cantilever beam. Damage was also suc-
cessfully location in an experimental cantilever plate.
The algorithm is robust to systematic errors in the
measured data, demonstrated in simulation by the ad-
dition of a discrete mass and experimentally by using a
crude model for the plate. One feature of the numeri-
cal example given is that genetic diversity was reduced
as the generation number increased. Alternatively, this
could be viewed as the convergence to a uniform ``op-
timum'' population, although it would not be guaran-
teed that this population would be the global
optimum. This may be a feature of the small number
of genes associated with each member of the popu-
lation.
Acknowledgements
Dr Friswell gratefully acknowledges the support of
the Engineering and Physical Sciences Research
Council through the award of an advanced fellowship.
References
[1] Rajeev S, Krishnamoorthy CS. Discrete optimization on
structures using genetic algorithms. Journal of Structural
Engineering 1992;118(5):123349.
[2] Jenkins WM. Plane frame optimum design environment
based on genetic algorithm. ASCE Journal of Structural
Engineering 1992;118(11):310312.
[3] Jenkins WM. Towards structural optimization via the
genetic algorithm. Computers and Structures
1991;40(5):13217.
[4] Keane AJ. Passive vibration control via unusual geome-
tries: the application of genetic algorithm optimization to
structural design. Journal of Sound and Vibration
1995;185(3):44153.
[5] Goldberg DE, Kuo CH. Genetic algorithms in pipeline
optimization. Journal of Computing in Civil Engineering
1987;1(2):12841.
[6] Larson CB, Zimmerman DC. Structural model rene-
ment using a genetic algorithm approach. 11th
International Modal Analysis Conference, Kissimmee,
Florida, 1095101, 1993.
[7] Levin RI, Lieven NAJ. Dynamic nite element model
updating using simulated annealing and genetic algor-
ithms. Mechanical Systems and Signal Processing
1998;12(1):91120.
[8] Dunn SA. Modied genetic algorithm for the identi-
cation of aircraft structures. Journal of Aircraft
1997;34(2):2513.
[9] Zimmerman DC. A Darwinian approach to the actuator
number and placement problem with non-negligible
actuator mass. Mechanical Systems and Signal
Processing 1993;7(4):36374.
[10] Cai JB, Thierauf G. Evolution strategies for solving dis-
crete optimization problems. Advances in Engineering
Software 1996;25(23):17783.
[11] Topping BHV, Khan AI, Leite JPD. Topological design
of truss structures using simulated annealing. Structural
Engineering Review 1996;8(23):30114.
[12] Doebling SW, Farrar CR, Prime MB, Shevitz DW.
Damage identication and health monitoring of struc-
tural and mechanical systems from changes in their vi-
bration characteristics: a literature review. Los Alamos
National Laboratory report LA-13070-MS. Electronic
version available on WWW: http://lib-www.lanl.gov/la-
pubs/00285981.pdf, 1996.
[13] Mottershead JE, Friswell MI. Model updating in struc-
tural dynamics: a survey. Journal of Sound and
Vibration 1993;167(2):34775.
[14] Zimmerman DC, Kaouk M. Structural damage detection
using a minimum rank update theory. ASME Journal of
Vibration and Acoustics 1994;116:22231.
[15] Hemez FM, Farhat C. Structural damage detection via a
nite element model updating methodology. Modal
Analysis: The International Journal of Analytical and
Experimental Modal Analysis 1995;10(3):15266.
[16] Topole KG, Stubbs N. Nondestructive damage evalu-
ation in complex structures from a minimum of modal
parameters. Modal Analysis: The International Journal
of Analytical and Experimental Modal Analysis
1995;10(2):95103.
[17] Mares C, Surace C. An application of genetic algorithms
to identify damage in elastic structures. Journal of Sound
and Vibration 1996;195(2):195215.
[18] Ruotolo R, Surace C. Damage assessment of multiple
cracked beams: numerical results and experimental vali-
dation. Journal of Sound and Vibration 1997;206(4):567
88.
[19] Cawley P, Adams RD, Pye CJ, Stone BJ. A vibration
technique for non-destructively assessing the integrity of
structures. Journal of Mechanical Engineering Science
1978;20(2):93100.
[20] Cawley P, Adams RD. The location of defects in struc-
tures from measurements of natural frequencies. Journal
of Strain Analysis 1979;14(2):4957.
[21] Cawley P, Adams RD. A vibration technique for non-
destructive testing of bre composite structures. Journal
of Composite Materials 1979;13:16174.
M. Friswell et al. / Computers and Structures 69 (1998) 547556 555
[22] Friswell MI, Penny JET, Wilson DAL. Using vibration
data and statistical measures to locate damage in struc-
tures. Modal Analysis: The International Journal of
Analytical and Experimental Modal Analysis
1994;9(4):23954.
[23] Friswell MI, Penny JET, Garvey SD. Parameter subset
selection in damage location. Inverse Problems in
Engineering, 1997; 5(3):189215.
[24] Wu X, Ghaboussi J, Garrett JH. Use of neural networks
in detection of structural damage. Computers and
Structures 1992;42:64959.
[25] Tsou P, Shen MHH. Structural damage detection and
identication using neural networks. AIAA Journal
1994;32(1):17683.
[26] Elkordy MF, Chang KC, Lee GC. Application of neural
networks in vibrational signature analysis. ASCE Journal
of Engineering Mechanics 1994;120(2):25065.
[27] Szewczyk ZP, Hajela P. Damage detection in structures
based on feature-sensitive neural networks. ASCE
Journal of Computing in Civil Engineering
1994;8(2):16378.
[28] Worden K. Structural fault detection using a novelty
measure. Journal of Sound and Vibration
1997;201(1):85101.
[29] Elkordy MF, Chang KC, Lee GC. A structural damage
neural network monitoring system. Microcomputers in
Civil Engineering 1994;9:8396.
[30] Kirkegaard P, Rytter A. Use of neural networks for
damage assessment in a steel mast. 12th International
Modal Analysis Conference, Honolulu, Hawaii,
1994:112834.
[31] Allemang RJ, Brown DL. A correlation coecient for
modal vector analysis. 1st International Modal Analysis
Conference, Orlando, Florida, 1982:1106.
[32] Friswell MI, Penny JET, Lindeld G. The location of
damage from vibration data using genetic algorithms.
13th International Modal Analysis Conference,
Nashville, Tennessee, USA, 1995:16406.
[33] Friswell MI, Mottershead JE. Finite element model
updating in structural dynamics. Dordrecht: Kluwer,
1995.
M. Friswell et al. / Computers and Structures 69 (1998) 547556 556

Vous aimerez peut-être aussi