Vous êtes sur la page 1sur 8

Finite Element Model Updating

2.1 Introduction
Precise description of the dynamic behaviour of aerospace, mechanical and civil engineering
structures is very concerned .With the advent of computer technology, pioneering work in the
aerospace industry developed and exploited numerical analysis techniques in the 1950s and
1960s . Due to the fact that experimental analysis is usually expensive and time consuming,
numerical simulation by The Finite Element Method (FEM) (Zienkiewicz 1967) found its
application for the first time in industrial problems and has proven itself as a very flexible
numerical analysis technique to obtain approximate solutions for otherwise intractable
problems. This Computer-based analysis techniques have changed the design and product
development ever since in many other industries. However, the existing differences in
geometry and material between the FE models and real structures may make the simulation
undependable.
To explain the lack of correlation between predictions and observations it is necessary to
consider the likely causes of inaccuracy in numerical models. It should be mentioned that
experimental measurements are not taken without error, but we will return to that issue later.
Here we consider three commonly encountered forms of model error which may give rise to
inaccuracy in the model prediction :
(i)
Model structure errors, which are liable to occur when there is uncertainty
concerning the governing physical equations - such errors might occur typically
in the modelling of neurophysiological process and strongly non-linear behaviour
in certain engineering systems ;
(ii)
Model parameter errors, which would typically include the application of
inappropriate boundary conditions and inaccurate assumptions used in order to
simplify the model;
(iii)
Model order errors, which arise in the discretization of complex systems and can
result in a model of insufficient order- the model order may be considered to be a
part of model structure.
In recent years, a significant amount of work has dealt with evaluating and reducing the
distance between the numerical models and the experimental structures in terms of their
dynamic behaviour. On the other hand, using an updated FE model, damage identification of
structures may be performed.

2.2 Finite Element Model Updating


In contrast to health monitoring or damage detection algorithms, the motivation for finite
element model updating is to improve the accuracy of an initial FE model so that the
predicted dynamic behaviour matches as closely as possible that observed during an
experiment. With increasing reliability and confidence in measurement technology, the need
to improve the numerical model representations initiated the development of model updating
algorithms in the 1970s. Ever since, the interest in systematically adjusting FE models has
produced a wealth of publications on the subject and a good introduction was presented by
Imregun (1992), including a discussion of practical bounds of the algorithms in general
terms. More mathematical and comprehensive surveys, were presented by Natke (1988),
Imregun and Visser (1991), Mottershead and Friswell (1993), Natke et al. (1995) and Friswell
and Mottershead (1995).
Page | 1

2.2.1 Direct FE Model Updating Methods


It is customary to classify model updating techniques into two broad categories. The earliest
generation of algorithms produced the methods often referred to as "direct methods" which
directly solved for updated global system mass and stiffness matrices by contemplating single
matrix equations. The equations defining the differences in spatial parameters between the FE
model and the measured structure usually make use of the measured eigenvectors and
eigenvalues.
Methods using Lagrange multipliers were derived from a strict optimisation point of view
and seek to minimise a predefined objective function along with additional constraints such
as the system symmetry and the orthogonality conditions. Baruch (1978), for instance,
corrects the system stiffness matrix whereas Berman and Nagy (1983) proposed an equation
to identify the differences in the system mass matrix. Both techniques are representative of
this family of methods. The applicability of these methods, however, is limited since the
connectivity pattern is usually destroyed, a limited number of measured eigenvectors is
employed and the measurements are assumed to be complete.
Similarly, matrix mixing methods assume that all modes are measured at all DOFs and use
these to construct the inverses of the global mass and stiffness matrix. Although these data
requirements are difficult to meet in practical applications and other approximations have to
be introduced, Link (1986) and Caesar (1987) used the identified flexibility matrices to
update an initial FE model using the computed and measured eigenvectors.
Other representative examples are the families of error matrix methods. Here the difference
between the initial analytical stiffness matrix and the unknown experimental stiffness matrix
is postulated and the error between both is assumed to be small. After computing pseudo
flexibility matrices using the measured eigenvectors, the error in the global mass and stiffness
matrices can be computed (Sidhu and Ewins 1984).
None of the direct methods, however, gives particularly satisfactory results as they place
almost impossibly high demands on the quantity and quality of experimental data required.

2.2.2 Iterative FE Model Updating Methods


The principles of model updating techniques generally described as "iterative methods" are
very different to direct updating formulations. Unlike direct methods, which focus on the
global system matrices, iterative model updating procedures are formulated in line with the
discretised nature of the FE model. FE models are an assembly of individual finite elements
and each finite element is defined by its design parameters, {}, such as its geometry or
material properties.
Iterative methods work together with a parameterised FE model, the "error model", and
introduce changes to a pre-defined number of design parameters on an elemental basis.
The flexibility provided by the parameterisation allows an updating which is physically more
meaningful than that offered by the direct methods.
Typically, the error model is advantageously used to calculate first-order derivatives of a
chosen dynamic property of the system. This linearisation allows the formulations of an often
over-determined linear set of algebraic equations in the form of:

[S] {}={}
where [S] is the sensitivity matrix, {} the changes in updating parameters and {} the
residual, the difference between the measured and predicted dynamic properties. Such a
Page | 2

system of equations is solved for the design parameter changes and the FE model is updated.
The sequence of solving and updating the system has led to the description of these
techniques as "iterative methods".
The system matrices of FE models updated by iterative methods can be uniquely
reconstructed and, unlike direct methods, the connectivity patterns of the modified mass and
stiffness matrices remain intact.

2.3.1 Introduction to Structural Dynamics


The modal properties i.e., natural frequencies and mode shapes were used as a basis for
finite-element-model updating. Thus, these parameters are described in this section. The
modal properties are related to the physical properties of the structure. All elastic structures
may be described in terms of their distributed mass, damping and stiffness matrices in the
time domain through the following expression (Paz and Leigh, 2003):

M X&& C X& K X F
Where, [M] is the mass matrix, [C] is the damping matrix, [K] is the stiffness matrix, {X} is
X&
X&&
the displacement vector,
is the velocity vector,
is the acceleration vector, and {F} is
the applied force vector.
If that Equation is transformed into the modal domain to form an eigenvalue equation for the
i-th mode, then (Ewins, 1995):

M j C K 0
2

1 ; i is the i-th complex eigenvalue; and {0} is the null vector. In this
Equation the real part of {}i corresponds to the normalized mode shape {}i while the
imaginary part of i corresponds to the natural frequency i.
Here, j =

Finite-element-model Updating Using NelderMead Simplex and BFGS Methods ,it may be
deduced that the changes in the mass and stiffness matrices cause changes in the modal
properties of the structure.
Therefore, the modal properties can be identified through the identification of the correct
mass and stiffness matrices. The frequency-response function (FRF) is defined as the ratio of
the Fourier-transformed response to the Fourier-transformed force. The FRF may be written
in terms of the modal properties by using the modal summation equation as follows (Fu and
He, 2001):

2k il i
H kl 2
2
i 1 2 ii j i
N

In Equation Hkl() is an FRF due to excitation at position k and response measurement at


position l, is the frequency point, i is the i-th natural frequency point, N is the number of
modes and i is the damping ratio of mode i.
Page | 3

The excitation and response of the structure and Fourier-transform method (Ewins, 1995) can
be used to calculate the FRF. Through this Equation and modal analysis (Ewins, 1995; Fu and
He, 2001), the natural frequencies and mode shapes can be indirectly calculated from the
FRFs. The modal properties of a dynamic system depend on the mass and stiffness matrices
of the system as indicated by this Equation . Therefore, the measured modal properties can be
reproduced by the model if the correct mass and stiffness matrices are identified.
The finite-element-model updating process is achieved by identifying the correct mass and
stiffness matrices. In the light of the measured data, the correct mass and stiffness matrices
can be obtained by identifying the correct moduli of elasticity of various sections of the
structure under consideration. In this chapter, to identify correctly the moduli of elasticity that
would give the updated finite element model, the following objective function that measures
the distance between measured modal data and finite-element-model calculated modal data,
was minimized (Marwala, 1997):

sn H n 1f xn
Here, N is the number of measured modes; E is the error; and II II is the Euclidean norm.
In this Equation the mass, damping and stiffness matrices are obtained from the finiteelement model, while the natural frequencies and mode shapes are measured. If the natural
frequencies and mode shapes of the system are described by the mass, damping and stiffness
matrices then E is equal to zero.
Therefore, the minimization of E identifies the updated finite-element model. Thus, the
process of finite-element-model updating may be viewed as being an optimization problem.

2.4 Methods for Comparing Data


The most vital characteristic of modal testing is a comparison between the computed dynamic
properties and those actually observed in practice (Ewins, 1995 Marwala, 1997; Ewins,
2001). This procedure is frequently called validating a theoretical model and it involves a
number of stages. The first stage is to compare the specific dynamic properties, as measured
against the predicted ones. The second stage is to measure the degree of the discrepancies or
similarities between the two sets of data. The third stage is to bring the theoretical model
closer to the measured data. When this is accomplished, the theoretical model is said to have
been updated. In this section we closely study the computational methods used in the first,
second and third stages.
In most situations, much endeavor goes into deriving the theory-based model and the
experimentally derived model. For this reason, it is prudent to compare on as many different
levels as possible. The dynamic model of a structure may be classified into spatial, modal,
and response models (Ewins, 2001).
It is, at this time, appropriate to revisit this classification and attempt to compare the
experimental and the theoretical model at each of these classifications. Consequently, a
comparison of response properties as well as modal properties will be made. Comparisons
between spatial properties are complex and, for that reason, will not be reflected on. In using
any medium of comparison, the model must be developed comprehensively from the original
form.

Page | 4

2.4.1 Direct Comparison


A) Comparisons of Natural Frequencies
The simplest method of comparison between the experimental and theoretical model is by
comparing the natural frequencies (Ewins, 2001). This may be achieved by tabulating the
experimental and theoretical natural frequencies. The most convenient way of comparison is
to plot the graph of the experimental natural frequencies against the analytical ones for all
available modes. If the gradient of the best straight line passing through the points is close to
zero, then the correlation between the experimental and computed model is good. If the
points lie spread out extensively about a straight line, then there is a severe failure of the
model in representing the theoretical models capacity to estimate the measured natural
frequencies and, therefore, the theoretical model should be re-evaluated. If the positions
diverge to some extent from the straight line, in a systematic fashion, this implies that there is
a particular characteristic responsible for the deviation.
B) Comparisons of Mode Shapes
The mode shapes can also be compared by plotting the analytical modes against the
experimental ones. For a simple structure with well-separated modes, this technique of
comparison may be used with ease. Nevertheless, for a complicated structure with modes that
are close to one another, this technique frequently becomes tricky to employ. Therefore, it is
appropriate to make comparisons of mode shapes at the same time as those of the natural
frequencies.
In the situation where we have more data to handle for each mode, the comparison may be
conducted by plotting the deformed shape for each model, experimental and theoretical, and
overlaying one plot on the other. The disadvantage of this technique is that, even though the
difference is evident, the plots are not easy to understand and they are usually confusing
because there is so much information. A suitable method of comparison, which is along the
lines of the natural-frequency plot, is to plot each element in the mode shape vector,
experimental and theoretical, on an x-y plot (Ewins, 2001).
The individual points on this graph relate to modal coordinates, and it is expected that they
should lie close to a straight line. If the mode shape vectors are mass-normalized, this straight
line should have a slope of 1. If the points lie close to a straight line with a slope that is not 1,
then one of the mode shapes is not mass normalized or there is scaling error in the data. If the
points are widely spread out about the line, then there is inaccuracy in one set or the other set.
If the dispersion is too great, then it may be that the eigenvectors that are being compared do
not relate to the same mode.
The slope of the best straight line is called the modal scale factor (MSF) and is defined as
(Ewins, 2001):


MSF a , m a T m *
a a
T

Here, a is the analytical modes; m is the measured modes; and * is the complex conjugate.
Page | 5

The MSF parameter gives no indication of the quality of the measured points with respect to
the straight-line fit.

2.4.2 The Modal Assurance Criterion (MAC)


The modal assurance criterion (MAC) compares the measured and the computed mode
shapes. Stetson (2008) calculated the MAC from electronic holography data, while Allemang
(2003) reviewed the use of the modal assurance criterion over a period of 20 years. Yuan et
al. (2009a) used the MAC optimally place a sensor on a cable-stayed bridge, whereas
Caponero et al. (2002) used an interferometer and the MAC for identification of component
modes. Brechlin et al. (1998) introduced the scaled modal assurance criterion to analyze a
system with rotational degrees of freedom, whereas Lars (1998) used the modal assurance
criteria to analyze two orthogonal modal vectors. Finally, Desforges et al. (1996) used the
MAC for tracking modes during flutter testing while Heylen and Janter (1989) applied the
modal assurance criterion for dynamic model updating. The MAC can be mathematically
summarized by the following equation (Allemang and Brown, 1982):

cr dr*
MACcdr
T
T
cr cr * dr dr*
2

Here, MAC is modal assurance criterion; c is for reference; d is the degrees-of freedom; r is
the mode; T is the transpose; * is the complex conjugate; and {} is a vector.
The MAC is a measure of the least-squares deviation of the points from a straight-line
correlation. A value close to 1 suggests that the two mode shapes are well correlated, while a
value close to 0 indicates that the mode shapes are not correlated.

2.4.3 The Coordinate Modal Assurance Criterion (COMAC)


The COMAC technique is based on the same principle as the MAC and is, in essence, a
measure of the correlation between the measured and the computed mode shapes for a given
common coordinate. Meo and Zumpano (2008) used the COMAC for damage estimation on
plate structures, while Zhao and DeWolf (2007) applied the COMAC for damage detection in
a cracked I-shaped steel beam. The COMAC for coordinate j is given by (Lieven and Ewins,
1988):

COMAC j

r 1

r 1

j
ar jmr *

ar

r 1

mr *

Unlike the MAC, the COMAC does not have any difficulty in comparing modes that are
close in frequency or that are measured at insufficient transducer locations. L is the total
number of well-correlated modes as indicated by the MAC. A value close to 1 suggests a
good correlation. If the mode shape vectors are used then the COMAC becomes a vector. In
Page | 6

this chapter, the MAC and direct comparison of mode natural frequencies are used to evaluate
the effectiveness of the finite element- model updating.

2.5 Optimization Methods


2.5.1 Quasi-Newton BroydenFletcherGoldfarbShanno (BFGS) Algorithm
The quasi-Newton optimization method is a successful, robust and quadratically convergent
optimization method that uses a gradient. The technique is a derivation of the Newton
Raphson method, with the difference being that the inverse of the second derivative is
updated through using a one-dimensional or multi-dimensional Hessian estimation method.
Ghosal and Chaki (2009) used the quasi-Newton method for estimating and optimizing the
depth of penetration in CO2 Laser-MIG welding. The NewtonRaphson algorithm is
mathematically represented as follows (Ransome, 2006):

Fig: An easy reflection of a two-dimensional Simplex (Ransome, 2006)

f& xn
xn1 xn
&
f& x
n

Here,
&
f& xn

f xn

is the objective function; and

f& xn

is the Jacobian (first derivative), and

is the Hessian (second-derivative).

For a single-dimensional function,

&
f& xn

can be updated by using the following equation:

s
&
f& xn 1 n

yn

Where
Page | 7

sn f& xn 1 f& xn

qn f& xn 1 f& xn
The most popular multi-dimensional technique for estimating the Hessian is the Broyden
FletcherGoldfarbShanno (BFGS) technique (Broyden, 1970; Fletcher, 1970; Goldfarb,
1970; Shanno, 1970). Sun et al. (2009) applied the BFGS technique for optimizing a
machining allowance. Tan et al. (2009) applied the BFGS method for Stokes flows with fixed
or moving interfaces and rigid boundaries, while Du et al. (2009) applied the BFGS method
for optimizing the distribution of fibrous insulation. Further applications of the BFGS method
include studies in aerodynamics (Papadimitriou and Giannakoglou, 2009), in solving
equations (Yuan et al., 2009b) and in non-convex problems (Xiao et al., 2009). The BFGS
estimation of the inverse Hessian Hn+1 is given by:

qn qnT H n T snT sn H n
H n 1 H n T
qn sn
snT H n sn
From an initially estimated 0 x and Hessian matrix, H0, repeat the following steps
(Nocedal and Wright, 2006):

1.Obtain a step

sn by solving

sn H n 1f xn

and do a line search to discover an optimal

step size n in the direction obtained in the initial step, and then update xn = xn + n sn;
2. qn = f(xn+1) f(xn) ;
3. Now we have to calculate the

qn qnT H nT snT sn H n
H n 1 H n T
qn sn
snT H n sn

Page | 8

Vous aimerez peut-être aussi