You are on page 1of 272

1

Chapter 1 INTRODUCTION

1.1 GENERAL INTRODUCTION
A helicopter structure is normally designed to avoid resonance at the main rotor
rotational frequency. However, very often military helicopters have to be modified
(such as to carry a different weapon system or an additional fuel tank) to fulfill
operational requirements. Any modification to a helicopter structure has the potential of
changing its resonance frequencies and mode shapes. With the potential of changing the
mode shapes, its stress distribution will be changed under continuous vibration
excitation, thus affecting the fatigue life of the helicopter structure. Predicting the
change of natural frequencies and mode shapes is not trivial because of the complexity
of the structure. It is both difficult and time consuming to model the whole structure by
the finite element method (FEM) because of the difficulty in modelling joints, material
properties and complex geometric shapes, etc. and because of the difficulty in validating
the model with experimental measurements. Accurate experimental modal testing of a
complex structure is difficult to achieve as the results are dependent on boundary
conditions and operating conditions. It will, therefore, be advantageous to be able to
predict the new dynamic characteristics of the modified structure based on the
determination of the behaviour of the original structure and a dynamic description of the
modifying structure.

Introduction

2
Methods for predicting the effects of structural modification can be classified into four
categories: (1) Lumped Modification (LM) with no change in the Degrees of Freedom
(DOFs), (2) Lumped Modification with change in DOFs, (3) Distributed Modification
(DM) with no change in the DOFs, and (4) Distributed Modification with change in
DOFs. The problem of LM, related to both the prediction and optimization aspects, has
been solved at the end of 1980s [1-4]. On the contrary, only a few studies have been
devoted to distributed modifications, although most modifications are continuous in
engineering practice, e.g. the attachment of a new missile to a helicopter structure. Even
these limited studies have been mainly directed to distributed modification with no
change in the DOFs [5]. This thesis is focused on predicting the dynamic response due
to distributed structural modification with change in DOFs.
There are two main approaches to the problem of structural modification. The first one
is based on a modal model using modal parameters identified in the original structure.
However the use of this approach is severely limited due to the effect of modal
truncation, and the difficulty in identifying them accurately at high frequencies or in
systems with a high modal density as well as in scaling them [5-7]. Therefore another
approach, based on the direct use of the experimentally determined frequency response
function (FRF), has gained prominence in predicting the effects of structural
modification [3, 5, 8].
1.2 OBJECTIVES
The main purpose of the study presented in this thesis is, therefore, to investigate means
of efficiently assessing the effects of distributed structural modification on the dynamic
properties of a complex structure. The detailed objectives are

Introduction

3
to derive and investigate means of efficiently assessing the effects of
distributed structural modification with additional DOFs on the dynamic
response; and
to derive and investigate means of efficiently assessing the effects of
distributed structural modification with reduced DOFs on the dynamic
response.
1.3 OUTLINE OF THE THESIS
In order to fulfill the above objectives, the thesis is divided into 7 chapters.
In Chapter 2, a literature review on the structural dynamic modification is given.
In Chapter 3, the feasibility of assessing efficiently the effects of structural modification
on the dynamic response is explored. Equations for three types of structural
modifications, with no change in DOFs, with reduced DOFs and with additional DOFs
are derived. Also, based on the quasi-local characteristic of the delta dynamic stiffness
matrix, the method to describe the dynamic effects introduced by the modifying
structure is developed.
In Chapter 4, the structural modification with no change in DOFs on a one-dimensional
structure, namely a simply supported beam, is first studied numerically on a rather
coarse FE model. Because of the unacceptable results from a coarse FE model, a further
study on finer FE model is conducted with success.
The structural modification with additional DOFs on a one-dimensional structure,
namely, a cantilever beam is studied numerically and experimentally with four cases:
Full model, Reduced Model, Simulation for experimental procedure and validation of

Introduction

4
the structural modification method with experimental data. The first two cases, Full
model and Reduced Model are mainly used to demonstrate the effectiveness of the
structural dynamic modification method for predicting the dynamic response of the
modified structure. The third case, Simulation for experiment procedure is to use
simulation to demonstrate the feasibility of the method for the fourth case: validation of
the structural modification method with experimental data.
In chapter 5, structural modification is applied to a two-dimensional structure, namely, a
clamped rectangular plate, in order to illustrate its performance with reduced DOFs.
Four cases are presented in this chapter: three numerical simulations and one
experimental validation. They are the Full model, Reduced Model, Simulation for
experimental procedure and validation of the structural modification method with
experimental data.
In Chapter 6, the structural modification method is applied to a free-free or clamp
supported 3D box frame models which are used to validate the method on 3D structures.
In total, four cases are presented. They are the Full model, Reduced Model, Simulation
for experimental procedure and validation of the structural modification method with
experimental data.
In Chapter 7, the major conclusions are drawn and recommendations for future work are
given.
5
Chapter 2 LITERATURE REVIEW

In order to understand structural dynamic modification method, with and without
changes in degrees of freedom (DOFs), the method, as well as other related
developments, is reviewed in this chapter.
2.1 STRUCTURAL DYNAMIC MODIFICATION
Structural dynamic modification seeks to determine the effects of changes in mass,
stiffness and/or damping on the dynamic response of a mechanical system, based on the
available information for the original system and the modifying part. It falls into two
categories: direct structural modification and inverse structural modification[9]. The
former is treated as a prediction problem used to determine the dynamic response
caused by modifications. The latter is frequently treated as an optimization problem by
which the modifications needed to achieve certain dynamic characteristics, usually in
terms of the desired values for natural frequencies and mode shapes, are determined.
Structural dynamic modification has been studied widely by a number of researchers.
Also, a few surveys have been conducted: on reanalysis techniques in 1976 by Arora
[10], on dynamic reanalysis in 1981 by Wang et al. [11], and on structural modification
in 1986 by Snyder[12]. In 2002, structural modification was reviewed by Avitabile[13].
These surveys, taken together, illustrate the entire development of structural
modification.

Literature Review

6
2.1.1 Direct Problem
As the direct structural modification problem depends on the objectives, it can be
classified into four sub-categories: 1) lumped modification with no change/s in DOFs; 2)
lumped modification with change/s in DOFs; 3) distributed modification with no
change/s in DOFs; and 4) distributed modification with change/s in DOFs. Those with
change/s in DOFs can be further classified into two sub-categories: one with reduced
DOFs and the other with additional DOFs.
2.1.1.1 Lumped modification
The problem of lumped modification, with or without changes in DOFs, as related to
both the prediction and optimization aspects, has been addressed since the 1960s and
remains an active topic today.
In 1968, Weissenburger [14] developed a method for lumped mass and spring
modifications, the local eigenvalue modification procedure, which neglects damping
and uses back-substitution to eliminate additional DOFs. Pomazal [15] and Hallquist
[16] extended this approach by taking damping into consideration and adding DOFs.
In 1973, Hirai et al [17] presented an exact method for calculating the eigenvalues and
eigenvectors of the modified structure using a local modification based on the complete
eigensolution of the original structure but, as this is not available most of the time, he
further developed his method to obtain accurate results from only several eigenvalues
and eigenvectors of the original structure. However, this extended method lacks
applicability when only a few lower eigenvalues and eigenvectors of the original
structure are known, or when higher ordered eigenvalues of the modified structure need
to be determined from a limited number of eigenvalues and eigenvectors of the original
structure.

Literature Review

7
In 1985, Wang et al. [18] experimentally investigated the effects of local modification
on the dynamic characteristics of an existing structure. The frequency response
functions (FRFs) of the modified structure were computed based on the frequency
response data of the original structure and the characteristics of the local modification.
Hence, the dynamic characteristics of the modified structure could be identified using
the calculated transfer functions. This method was applied to a free-free plate with local
mass modification and without additional DOFs.
In 1988, Wallack [19] extended the local modification method, which must be used
repeatedly, i.e., for each local modification, to handle general matrix modification
which can solve multiple modifications simultaneously.
zgven[4] developed a general method, using either theoretically calculated or
experimentally measured FRFs, to analyse a structure subjected to lumped structural
modification in two different cases: with and without additional DOFs. This method has
been successfully applied to several simple structures, e.g., a plate and a propeller by
Tahtali[3].
2.1.1.2 Distributed modification
Structural dynamic modification was a major topic in the world of modal analysis for
several decades until the 1990s, when interest in finite element (FE) model updating
increased and its solution seemed to be able to provide a definitive answer to most of
the issues relevant to modal analysis, including structural modification. In fact, this
opinion is hard to justify given the complex nature of FE model updating and there is,
indeed, a renewed interest in using the structural modification method as a practical tool
for solving vibration and acoustic problems.

Literature Review

8
However, this raises a problem that has neither been solved nor even practically
formulated: the case of distributed structural modification. Few studies have been
devoted to this type of modification, despite most modifications in engineering practice
being continuous, e.g., the attachment of a new missile to a helicopter structure; and
even these rare studies have been directed mainly towards distributed structural
dynamic modification with no changes in DOFs, as in [5, 20].
In 1991, D'Ambrogio [20] investigated the prediction of the FRFs for a system
subjected to structural dynamic modifications. This approach is based on knowledge of
the FRFs of the original structure in combination with a physical description of the
modified structure. It is difficult to couple the theoretical model of the modifying
structure, which includes rotational DOFs, with the FRFs of the original structure when
it is derived from measurements based on translational DOFs only. Therefore, a
condensation procedure is used to eliminate rotational DOFs from the mass and stiffness
matrices of the altered structure. D'Ambrogio was convinced, at that time, that FE
model updating could provide a more definitive answer to the structural modification
problem.
For lumped modifications, the modifying structure can be described independently,
based on the different characteristics of lumped and distributed modifications while, in
cases of distributed modifications, a coupled system must be considered. In 2001,
D'Ambrogio and Sestieri [5] proposed a solution for the problem of computing the
dynamic stiffness matrix
| |
B for distributed structural dynamic modifications by
employing its quasi-local characteristics for no change in DOFs. In his paper, an
appropriate use of a condensation procedure and a careful consideration of the non-local
effects it introduces are addressed.

Literature Review

9
From 2006 to 2009, as a result of this thesis work, Hang et al. [21-24] combined the
theory of lumped structural modification (with additional DOFs) [4] with the method
proposed by DAmbrogio and Sestieri [5] to compute the dynamic stiffness matrix
| |
B of distributed structural dynamic modifications [5]. This combined method is used
to predict the effects of distributed structural modifications with additional DOFs. Also,
Hang et al. [22, 24]developed the distributed structural modification theory to include
distributed structural modifications with reduced DOFs. In order to include the non-zero
effects after the condensation procedure, known as Guyan reduction, the definition of
the interface DOFs is extended from physical to numerical.
2.1.2 Inverse Problem
In the field of structural modification, most of the research to date has been focused on
the inverse structural modification problem because it is widely used in design and
vibration control areas. Since the 1970s, there has been considerable development of
this topic, as summarized in the review by Venkayya [25].
In 1981, Done and Rangacharyulu [26] applied an optimization process using structural
modification to helicopter vibration control. Forced vibration response was used to
identify the most effective parameters for controlling vibration in the crew area. The
application of this method to a simple two-dimensional beam element helicopter
fuselage model demonstrates what can and cannot be done in controlling vibration by
using optimization structural modification.
In 1982, Wang et al. [27] developed numerical methods for the structural modification
of a fuselage structure and for the analysis and design of appendant structures. These
were applied to the problem of alleviation of helicopter vibration while design
techniques were developed to achieve the desired anti-resonance through structural

Literature Review

10
modification. A local modification method was used to analyse both the appendant
structures and the modified system. Since, in this method, the original and additional
systems are treated separately, efficient, repetitive searching can be conducted until the
appendant system produces a meaningful reduction in vibration.
In 1985, Wang et al. [28] introduced an efficient reanalysis procedure for calculating the
eigenvalues and eigenvectors for a modified structure which solves the eigenvalue
problem in a reduced local space instead of in the original global space.
In 1988, Kuang and Chen [29] presented a modified procedure for the optimization of a
linearly damped system in which the eigensolution of the original system is used to
calculate the sensitivities of the eigenvalues and eigenvectors. Numerical and
experimental data are used to illustrate how the method handles real normal and
complex modes in the following two situations: (1) given the locations of the
modifications, it calculates the value of the masses to be added, and (2) given the
number of modifications, it finds the best locations and the values of the added masses.
In 1988, Zhang et al. [30] presented an optimization method for determining both the
critical amount and the best location of mass modification(s) to the structure in order to
achieve certain dynamic requirements. The analysis was made using the perturbation
theory with measured modal data (natural frequency, damping factor and mode shapes).
Concurrently, a best approximation subspace technique was introduced to find the best
combination(s) of mass modification(s) from several possible measurement points.
Sestieri [1] and DAmbrogio [2] developed a method which uses the raw data of
measured FRFs and lumped modifications to achieve the best structural modifications in
order to fulfill the need to vary the structural dynamic characteristics.

Literature Review

11
In 1996, Lisowski et al. [31] performed modal testing of a helicopter prototype under
two types of boundary conditions. Based on the results of these two experiments and
their own engineering experience, they performed a structural modification by
increasing the local mass which not only lowered the value of some natural frequencies
of the helicopter prototype but also changed the considered vibration modes so that the
mast bending became less intensive.
In 1999, Li and He [32] presented a new approach for local structural modification
which determines the mass and stiffness modifications needed to change the dynamic
characteristics of an undamped structural system. This approach formulates a set of
linear equations for the structural modification problem and does not require an
eigenvalue solution. The data input are the FRFs at designated modification points.
In 2000, Park et al. [33, 34] developed a structural dynamic optimization method, based
on measured FRFs, to derive multiple lumped mass and damper and stiffness
modifications for the reallocation of the natural frequencies and mode shapes of an
existing system. Based on FRFs, a substructure coupling concept is used to derive the
system dynamic equations and, finally, a linear algebraic equation for identifying the
necessary structural modifications is obtained.
In 2001, Mottershead et al. [35] presented an inverse structural modification method for
reallocating the natural frequencies and nodes of normal modes of vibration by the
addition of grounded springs and concentrated masses. The method relies entirely on
measured receptances at the coordinates of the nodes and the modifications. There is no
need for measurements at other locations and it does not require an analytical model.
The stiffness and mass parameters are determined by an analysis of the null space of a
matrix containing the measured receptances.

Literature Review

12
In 2004, Kyprianou et al. [36] developed a method for assigning natural frequencies to a
multi-DOF undamped system by adding a mass connected by one or more springs. The
added mass and stiffness are determined using receptances from the original system.
The modifications required to assign a single natural frequency may be obtained by the
non-unique solution of a polynomial equation. If more than one frequency is to be
assigned, then a system of non-linear multivariate polynomial equations must be solved.
Such a modification involves not only an added mass and one or more stiffness terms,
but also an added coordinate. They presented a method which uses Groebner bases [37]
for the solution of the multivariate polynomials and also explains the effect of the
modification on the unassigned natural frequencies and anti-resonances.
In 2005, Kyprianou et al. [9] used a beam connected to the original structure to achieve
the assignment of natural frequencies or anti-resonances. In order for this to be
accomplished, rotational receptances had to be measured, as discussed in the companion
paper [38]. The added beam is cast as an additional forcing term on the original
structure. A system of multivariate polynomials in the parameters of the beams cross-
section is solved to yield the beam parameters that assign the specified vibration
behaviour.
In 2007, Olsson and Lidstrm [39] modified the mass and stiffness matrices to obtain a
desired spectrum and constraints were imposed on the structure. The undamped modal
model of a constrained linear structure is obtained by solving a generalized eigenvalue
problem. The coefficients of the constraint matrix are taken as design variables and a set
of equations defining the inverse structural modification problem is formulated. This
modification problem requires an iterative method for its solution. Each iteration step
involves the calculation of a rectangular Jacobian and the solving of an associated

Literature Review

13
under-determined system of linear equations. The system can be solved using the
MoorePenrose inverse [40] and is demonstrated by numerical examples.
Some other efficient methods have been developed for the inverse problem, e.g., the
convex linearization method (CONLIN) by Fleury [41], the method of moving
asymptotes (MMA) by Svanberg [42] and the sequential quadratic programming
method (SQP) by Schtittkowski [43]. These three approaches depend on convex
approximations. In 1997, Zhang and Fleury [44] made a modification to enhance the
reliability and generality of convex approximations.
The most important part in the optimization procedure is the selection of design
variables and sensitivity analysis will be discussed in the next section.
2.1.2.1 Sensitivity Analysis
Sensitivity analysis of structural data with respect to variations in the values of several
design variables have proved to be a powerful tool leading to a systematic method for
solving structural dynamic optimization problems. The sensitivity data are helpful for
the understanding of the dynamic characteristics of the system. As a result, the
structural refinement with reference to these data becomes much easier.
In 1968, Fox and Kapoor [45] derived the sensitivities of eigenvalues and eigenvectors
with respect to variations in the design parameters of the actual structure indicating that
these derivatives can be used successfully to approximate the results of analysis of new
designs.
In 1988, Sharp and Brooks [46] developed a method to calculate the sensitivities of
FRFs. It estimates the response behaviour of a structure when subjected to
modifications in the frequency domain.

Literature Review

14
In 1989, Chen [47] developed a systematic procedure for performing structural
modifications iteratively in order to achieve specified frequency responses. This goal is
achieved by efficiently computing and reanalysing the sensitivities of frequency
responses in a reduced space.
In 1990, To and Ewins [48] presented a method for calculating a condition number
which would indicate how sensitive the eigenvalues and eigenvectors of a mechanical
structure are in relation to minor modifications. The value of this condition number is
used to determine a limit of applicability for the first-order eigenvalues and eigenvector
sensitivities.
2.1.3 Two Approaches for Structural Modification
Based on data available for the original structure, there are two main approaches for
solving structural modification problems. The first is modal model-based structural
modification which utilizes modal parameters, such as natural frequencies, damping
ratios and mode shapes, as identified in the original structure. The second is the FRF-
based structural modification which utilizes the experimentally measured or numerically
determined FRFs.
2.1.3.1 Modal model-based structural modification
Two approaches are commonly used for modal model-based structural modification: the
eigenvalue modification technique (EMT) and the local eigenvalue modification
procedure (LEMP). However, when the modification involves damping or when the
original modes are complex, a different method, a complex mode solution procedure,
must be applied to ensure accurate results.

Literature Review

15
2.1.3.1.1 Eigenvalue Modification Technique (EMT)
This method was developed by Johnson and Synder [49] in 1989. It can be used to
predict modified dynamic characteristics due to structural modifications by
implementing the dynamic characteristics of the original structure.
The advantage of this method is that it reduces the eigensolution from the size of the
physical structure n to the size of the modes m in the modal-space solution; this
reduces the computational power considerably because mis generally significantly
smaller than n . For example, a system under investigation by the finite element (FE)
method, with hundreds of DOFs or n DOFs, is quite common while only the first
several natural frequencies and mode shapes or m modes are of interest to the
investigator. By reducing the hundreds of calculations to several, the computational
power required is reduced dramatically.
The general disadvantage of this model is the over prediction of natural frequencies
compared with the exact values of the modified structure. This is due to a truncation
error; the more severe the truncation error, the higher the natural frequencies it predicts
[50]. The final mode shapes of the modified structure are obtained from linear
combinations of those of the original structure. If a mode, which has a significant
contribution, is not included in the original eigenvector matrix, the predicted modes of
the modified structure could be severely erroneous.
2.1.3.1.2 Local Eigenvalue Modification Procedure (LEMP)
Initially developed by Simpson [51] and Weissenburger [14] in 1968, the LEMP
neglects damping and uses back-substitution to eliminate additional DOFs. This
procedure was further developed by Pomazal [15]and Hallquist [16] who took damping
into consideration and added DOFs. Eormenti and Welaratna [52] utilized this method

Literature Review

16
for a beam modification by grounding the two ends of a free-free beam and Vilmann
and Snyder [53] extended it to include line modification in continuous vibratory systems.
Sydner [12] conducted a survey of this method in 1986.
One advantage of this method is that the problem of solving a 2m order set of
polynomial equations is reduced to msecond-order polynomial equations. m is the
number of modes of interest. Another feature is that the eigenvalues of the modified
system are always bounded by the eigenvalues of the original system. However, this
procedure also has a disadvantage which is discussed by Wang [54]. It only allows one
single modification to be performed each time which means that multiple modifications
need to be made sequentially. This is begun by making the first modification, obtaining
the eigenvalues, updating the system equations and then repeating the procedure until
satisfactory accuracy is achieved.
In 1988, Avitabile [55] compared the results predicted from different modal data bases
by using the LEMP. It was found that the use of "assumed" proportional modal vectors
can introduce errors during structural dynamic modification which are pronounced
when damping modifications are employed. However, errors can also be observed when
mass or stiffness modifications are made using "assumed" proportional modal vectors.
2.1.3.1.3 Complex Mode Solution
The two methods described above were developed based on the assumption of either
proportional modal vectors or no damping. Generally, relatively accurate results can be
obtained for a proportionally damped original system when there are only mass and
stiffness changes involved. However, when the mode shapes of the original system are
complex in nature or the modifications involve damping, only the complex mode

Literature Review

17
solution procedure [56] can be used to obtain the predicted results accurately, so usually,
the state-space solution method is used.
2.1.3.1.4 Perturbation method
The methods mentioned above are very useful for modifying a considerably large
proportion of the original structure. If the modification is quite small compared with the
original structures dimensions, a different method, perturbation, is the most
successfully adopted procedure [57]. According to Ravi et al [58], the perturbation
method is a very useful structural dynamic modification technique for efficient and
reasonably accurate estimations of the dynamic characteristics of a modified structure.
This method expresses the increments in the stiffness and mass matrices of the system
during modification in terms of increments in the eigen-properties.
In 1985, Chen et al. [59] applied the matrix perturbation method to vibration modal
analysis. In this paper, they extended the capability of the matrix perturbation by
introducing a substructuring technique which deals with large complex structures that
have weak internal connections defined by a connection matrix in the physical space.
Each substructure is analysed individually and their eigenvalues and eigenvectors are
used to determine those of the combined structure by using matrix perturbation. The
substructuring technique presented in this paper is more efficient than component modal
synthesis, as evidenced by the example included to illustrate and verify the procedure.
In 1994, Ravi et al. [60] applied eigenvalues reanalysis, based on the multi-step
perturbation method, to a sandwich beam by considering structural modification as a
number of small steps. The limitation of this method is that the errors introduced in
every step accumulate in the final results. In 1995, Ravi et al. [58] further developed the
perturbation method to account for structural modification in one step, called single-step

Literature Review

18
perturbation, by considering the total increments of the stiffness and mass matrices as a
single step.
In 1998, Benfratello and Muscolino [61] developed an unconditionally stable step-by-
step perturbation procedure to evaluate the deterministic responses of linear structures
with dynamic modifications. This technique evaluates the transition matrix, which is the
fundamental operator of the step-by-step solution, through a perturbation approach. It
does not require any computational effort when evaluating the eigen-properties of the
modified structure and its application to a simple case is presented in order to
demonstrate its advantages.
2.1.3.2 Frequency Response Function (FRF)-based structural modification
The use of modal model-based structural modification is severely limited due to the
effects of modal truncation, as shown by Braun and Ram[6] and Elliott and Mitchell[7],
and the difficulty of identifying a modal model in systems with high modal density [62].
In 1985, Deel [63] studied the potential errors of modal models-based structural
modification. He found that the accuracy of the predicted dynamic properties using this
method depends on both the intrinsic limitations of this simplified method, and on the
accuracy and suitability of using the modal model as the database. Some special
considerations which should be taken into account when designing a modal test for the
purpose of generating a structural modification database are discussed. Examples were
presented to show the effects on the accuracy of the structural modification results by
various errors, such as global calibration errors, localized calibration errors, modal
scaling errors and modal truncation errors, etc. in the modal model. Suggestions such as
conducting calibration for the entire measurement system before and after the test,
keeping a consistent unit system and using consistent measurement techniques at all

Literature Review

19
locations, etc. are presented for avoiding these errors and for validating the accuracy of
the data base. Special attention is also given to modal truncation error which results
from a in-complete modal model.
Therefore, another approach, the FRF-based structural modification, has become more
popular for predicting the effects of structural modification [3, 5, 8, 21-24]. A FRF is a
frequency-dependent quantity derived from knowledge of both the magnitude and
phases of a harmonic response and the excitation that causes it [8]. There are a number
of different quantities classed as FRFs, all of which have specific names, as listed in
Table 2.1. Using either measured or calculated FRFs directly in the reanalysis, without
identifying the modal parameters of the original structure, eliminates the modal
truncation problem [4].

Table 2.1 Different types of frequency response functions
Response type Response/Force Force/Response
Displacement x
Receptance
x
f
Dynamic Stiffness
f
x

Velocity
x

Mobility
x
f

Mechanical Impedance
f
x

Acceleration
x

Inertance or Accelerance
x
f

Apparent Mass
f
x


In 1988, Sestieri and DAmbrogio [64] developed a procedure which performs an
optimal structural modification without using modal parameters. An optimal procedure
for structural modification is demonstrated by using the raw experimental data instead
of the modal model derived from the experimentally determined FRF through curve
fitting, in order to avoid the pitfalls mentioned above.

Literature Review

20
As some of the papers dealing with this topic fall into more than one category, those
that have been reviewed before, e.g., Wang et al. [18], Wallack [19], Sestieri[1],
DAmbrogio[2, 5, 20], zgven[4], Park [33, 34] and Hang et al. [21-24] will not be
reviewed again in this section.
2.1.4 Differences between Structural Modification and Substructuring
Substructuring, which is mainly used for assembling components, includes two methods:
the component mode synthesis method and the FRF-based substructure analysis method.
After determination of the modal model and/or the FRF(s) for each component, those of
the assembled structure can be predicted.
Component mode synthesis was presented by Hurthy [65] in 1965, and, since then,
several methods have been developed utilizing this technique. Component mode
synthesis permits dynamic characteristics of small subsystems to be tied together to
form a larger system [12]. In 1975, Hintz [66] and, in 1985, Craig [67] reviewed this
method.
FRF-based substructure analysis can predict the response of complex systems by
combining the FRFs of each substructure as derived from FE analysis or experiments. In
general, the substructure with excitation is separated from the others by rubber bushes to
prevent the transmission of vibration from it to the main structure[68].
Structural modification doesnt require the modal model or FRFs of the modification
part. It only requires the building of FE models for a small part of the original structure
and for the same part of the original structure with its modifying components. This
saves time and effort as the small part of the original structure can be treated as a
simpler structure such as a beam or a plate. However, for substructuring or component

Literature Review

21
mode synthesis, the modal models or FRFs for each individual component need to be
known before the calculations can be performed.
2.1.5 Application of Structural Modification
Structural modification has been widely used in different areas, such as prototype
design validation, vibration control [69], etc.
In 1980, Herbert and Kientzy [70] demonstrated two ways of using structural dynamic
modification: 1) specify a modification and obtain its dynamic properties; or 2) specify
a desired dynamic condition and determine the modifications necessary to achieve it.
In 1987, Eastep et al. [71] presented a method of vibration control of large space
structures by both simultaneous and integrated design and control design. This method
reduces structural responses by using structural modifications of a nominal FE model to
increase the active modal damping factor beyond that of the nominal structure. These
modifications are determined by calculating the sensitivity of the closed-loop
eigenvalues to structural stiffness variations. Improvements in modal damping factors
are specified and a method for the determination of structural modifications is provided.
These modifications, i.e., changes in the cross-sectional areas of the beams, are
achieved by using a non-linear mathematical optimization technique. The application of
the algorithm is illustrated by the structural modification of a nominal ACOSS-FOUR
model with different constraints on the closed-loop eigenvalues.
2.2 DETERMINATION OF DYNAMIC RESPONSES
It is arguable to say that the determination of dynamic responses is no longer a difficult
topic. Before the invention of FE method in 1940s, there were only two methods for this

Literature Review

22
type of work, one is theoretical analysis and the other is modal testing. With the
development of modern computational technologies, the FE method becomes more and
more dominant in the determination of dynamic response while leaving the modal
testing as a source of validation.
2.2.1 Theoretical Analysis
Theoretical analysis of dynamic response determination is only available for some
simple structures such as beams [72, 73] and plates [74]. But, for real-life complex
structures, few applications of theoretical analysis have been found, and is, therefore,
not reviewed here.
2.2.2 Modal Testing
Ewins claimed in his book [62] that:
Modal testing is, in effect, the process of constructing a mathematical model to describe
the vibration properties of a structure based on test data, rather than a conventional
theoretical analysis.
From his statement, it is known that modal testing has two aspects: the test data; and the
construction of a mathematical model to describe the vibration properties of the
structure under test. This has been widely discussed by researchers. Books such as [62,
75-79] can be used as guidelines for conducting modal testing. In addition, numerous
papers such as [80-82] provide discussion points.
2.2.3 Finite Element Analysis
Finite Element Analysis (FEA) has been widely employed since its invention in the
1940s. In its application, the system under investigation is represented by a

Literature Review

23
geometrically similar model consisting of multiple discrete elements which are, linked,
and simplified. Equations of equilibrium, in conjunction with applicable physical
considerations such as compatibility and constitutive relations, are applied to each
element, and a system of simultaneous equations is constructed. The system of
equations is solved for unknown values using the techniques of linear or non-linear
numerical methods. While being an approximate method, the accuracy of the FEA can
be improved by refining the mesh in the model using finer mesh, i.e., more elements
and nodes. Readers who are interested in this topic are referred to conventional books
such as [83-87].
2.3 DEALING WITH ROTATIONAL DEGREES OF FREEDOM
2.3.1 Determination of Rotational Degrees of Freedom
In fact, for a structure, 50% of all the coordinates are rotational DOFs and 75% of all
FRFs involve rotational information [62], as the following equation shows:

| |
/ /
/ /
u F u M
F M
H H
H
H H

(
=
(

(2.1)
where
| |
H = FRF matrix of structure;
u = translation displacement variable;
= rotational displacement variable;
F = force applied on structure; and
M = moment applied on structure.
In practice, a useful modification, such as adding a beam, requires the measurement of
the rotational response as well as the translational response because real modifications

Literature Review

24
transfer both forces and moments [88]. Without experimental rotation information, it is
impossible to conduct such a modification properly. However, it is quite difficult, if not
impossible, to measure the rotational response due mainly to the difficulty of applying a
pure moment of excitation. A number of methods have been tried, including both
experimental and numerical, with limited success as they are still in the developmental
stage.
Throughout the 1980s and 1990s, the need for measurement of rotational responses
intensified for a variety of reasons, including the development of procedures for the
correlation and updating of FE models but, most importantly, for structural modification
and substructure coupling. A typical case of structural modification is when the
dynamic behaviour of a system, modified by an added mass or spring, is determined
from the dynamic properties of the system in its unmodified condition. Such simple
modifications are described entirely by translational coordinates and require the
measurement of only the translational response. If a modification is to include rotational
inertia or rotational stiffness, i.e., by a beam being attached to the original structure, the
measured rotational responses from the original structure will become necessary.
2.3.1.1 Experimental method
For measuring rotational receptance, there are two problems that need to be solved;
firstly, how to measure the rotational response and, secondly, how to generate and
measure the pure moment excitation. The first is less difficult and there are many papers
on rotational measurements as well as a number of techniques being invented, e.g.,
using a T-block [38], a laser vibrometer [89-92], and PZT sensors [93]. Also, attempts
have been made to apply pure moment excitation onto the structure but with limited
success.

Literature Review

25
In 1969, Smith [94] used a configuration of two electromagnetic shakers, capable of
applying either a pair of equal forces or a pair of opposite forces, and applied it to a
specially designed fixture. The requirement to control the amplification of random
experimental errors was discussed.
In 1972, Ewins and Sainsbury [95] and, in 1975, Ewins and Gleeson [96] investigated
the measurement of rotational response. Using a rigid attachment, such as the T-block,
they showed that the full matrix of receptances could be expressed in terms of the
measured translational response, a coordinate transformation matrix and the mass
matrix of the attachment. In 1997, Cheng and Qu [97] and Qu et al. [98] used a similar
approach, but based it on the use of a rigid L-shaped attachment.
In 1980, Ewins [99] showed that all the elements of the full N x N mobility matrix can
be derived from the measurement of the elements in just one row or one column of that
matrix which seems to suggest that the most difficult rotational responses are
recoverable from other measurements. However, Ewins pointed out the impracticability
of his statement because, in practice, the number of measured natural frequencies is
much less than the number of natural frequencies of the system and, therefore, the
residuals due to the out-of-range (lower or higher) modes cannot be counted.
In 1984, Yasuda et al. [100] and, in 1986, Kanda et al. [101] also used a rigid
attachment but their studies concentrated on the measurement of rotational
displacements and did not extend to the determination of rotational receptances.
In 1987, Petersson [102] conducted some experimental work on moment mobility
measurement by using two giant magnetostrictive alloy rods, excited out of phase with
each other, to produce a couple on a small T-block, assumed to be a pure moment,
which is a little encouraging. In 1993, Gibbs and Petersson [103] adapted the principle

Literature Review

26
of magnetostrictive rods to produce a smaller moment actuator which was used to
demonstrate the importance of rotational vibrations to structure-born sound. This
methodology, together with the equations required to estimate the moment mobility
using the magnetostrictive actuator, is detailed in Ref. [104]. Su and Gibbs [105], who
used the two-shaker configuration and the magnetostrictive actuator, concluded that the
main drawback is from the mismatching of forces from the two shakers.
In 1994, Bokelberg et al. [91, 92] described the operational principle of the
measurement of rotational responses by means of a multi-dimensional laser vibrometer.
This measurement system can determine three translational and three rotational
displacements of a vibrating structure by sensing the positions of three light beams
reflected by a tetrahedral target. A prototype system is presented, the practical
implementation of this measurement technique is discussed and the dynamic
measurement capabilities are evaluated via a structural impact test of a cantilever beam.
In 1995, Sanderson and Fred [106] and Sanderson [107] investigated the bias error in a
moment mobility measurement when using a simple straightforward two-channel
technique. It is composed of two parts: error in the rotation caused by unwanted
excitations; and error in the measurement of the moment. The former can be further
separated into two parts: the sensitivity of the structure to false excitation; and the
quality of the moment excitation. The error in unwanted excitation of the rotation can be
minimized in two steps: by reducing the total mass attached to the structure; and by
matching the forces from two separated exciters. The error in the measurement of the
moment depends not only on the design of the moment-excitation fixture, but also on
the structure itself, and can be minimized by applying a correction for the rotational
inertia of the moment exciters. Sanderson [107] employed moment exciter

Literature Review

27
configurations, in the form of a T-block and an I-block, and compared them in order to
determine their effectiveness in performing moment mobility measurements. He stated
that the greatest obstacle to overcome in the measurement of moment mobility is the
size of the moment exciter's mass.
In 1999, Stanbridge and Ewins [108] discussed the measurement of vibration responses
using a laser Doppler vibrometer (LDV) in which both the angular and translational
vibrations can be obtained by demodulating the LDV vibration output.
In 2000, Duarte and Ewins [109] considered the experimental derivation of the
rotational DOF parameters from the transformation matrices for closely spaced
accelerometers by using the finite-difference formulation. The residual compensation in
the rotational DOFs derivation was considered carefully, together with the best order of
the approximations used.
In 2000, Yoshimura and Hosoya [110] determined the rotational receptances by solving
an over-determined least-square problem using measured time-domain data from a rigid
T-block. This method suffers from a poorly conditioned set-up due to: (1) rotational
motion being produced by the small differences among translations from the
accelerometers on the rigid fixture; and (2) the excitation points on the rigid fixture
being close in comparison to the wavelengths of the modes excited in the structure and,
therefore, the receptances excited on the attachment from points in close proximity
being very similar.
Due to the difficulty of measuring rotational DOFs and the impossibility of generating
and measuring pure moment excitation, either condensation or expansion is required
before coupling the structures together. In 2000, DAmbrogio [93] analysed and
compared a few possible options for providing an efficient formulation in the case of

Literature Review

28
distributed modifications and also suggested the use of PZT sensors for measuring the
rotational DOFs.
In 2002, Trethewey and Sommer [111] presented a different way of generating a pure
moment by utilizing a set of geared eccentric masses driven by a DC motor. In order to
cancel an undesirable moment in an orthogonal direction, they added a second identical
counter-rotating system of masses on top of the previous one. This prototype pure
moment exciter performed as predicted by the underlying theory and the experimental
moment amplitudes were within 5% of the theoretical values across the evaluated
frequency range.
In 2002, Dong and McConnel [112] relaxed the requirement for a rigid fixture. They
determined the full inertance matrix by using a FE inertance matrix of the T-block
together with inertances measured by an instrument cluster (including rotational
accelerometers) on the T-block itself.
In 2005, Mottershead [38] investigated the ill-conditioning problem of the widely used
T-block technology in which the T-block is attached to the structure at the
modification point. A force applied to the T-block generates a moment, together with a
force, at the connection point between the T-block and the structure. Forces and linear
displacements measured on the T-block, together with a mass and stiffness model of the
T-block itself, allow the problem to be considered as a special case of excitation by
multiple inputs. The generally ill-conditioned resulting equations are regularized by
using a small number of independent measurements. The methodology and signal
processing techniques required to estimate the rotational receptances are described in
the paper and an experimental example is adopted to demonstrate the practical
application of the method.

Literature Review

29
2.3.1.2 Numerical method
OCallahan et al. [113, 114]and Avitabile et al. [115] used FE modal data as a mapping
transformation matrix to expand the set of measured translational DOFs, thereby
producing an estimate of the unmeasured rotations as well as of some unmeasured
translations.
In 1988, Mitchell-Dignan and Pardoen [116] investigated the estimation of rotational
DOFs by using FE shape functions or interpolation polynomials in conjunction with the
modal parameters of the unmodified structure. One- and two-dimensional shape
functions are used to estimate the translational and rotational DOFs at any given
location on the structure. A limited development of structural modification and shape
function theory is presented. Although the proposed method, using shape functions for
estimating rotational DOFs, is by no means an exact technique, its simplicity and
versatility make it a viable approach for use with experimental data.
In 1997, in a series of papers by Silva, Maia and Ribeiro [117-119], a new technique for
modal analysis, the Mass Uncoupling Method (MUM), was proposed; it allows for the
cancellation of added transducers masses and enables the estimation of the complete
FRF inertance matrix from either one row or one column of the measurements in the
matrix. The MUM technique works well for the estimation of unmeasured rotational
FRFs with exact theoretical data, but fails when applied directly to real cases or even to
theoretical data with limited noise pollution. In 1999, Silva et al. [120] presented a
refined version of the MUM technique to improve its performance when used with
noise-polluted data. Subsequently, in 2000, Silva et al. [121] applied it to a real case, i.e.,
an experiment, and stated that there are still problems to be overcome and it needs
improvement before being applied for practical purposes.

Literature Review

30
In 2003, Avitabile [122] compared several techniques for expanding experimentally
derived mode shapes and residual modes to include the necessary rotational DOFs. The
FRFs involving translational to rotational DOFs are developed as well as those for
rotational to rotational DOFs.
2.3.2 Reduction of Degrees of Freedom
Model reduction is a commonly used algorithm to effectively estimate some low
eigenvalues and corresponding eigenvectors of structures by reducing the order of the
original structural model to a smaller one. This strategy removes some DOFs (slave
DOFs) of the original FE model and retains a much smaller set of DOFs (master DOFs)
and, subsequently, solves the eigen-function of the reduced model and approximates the
eigen-solutions of the original model [123]. In accordance with its original definition,
the model reduction technique is used to remove the rotational DOFs in research
discussed in this thesis.
2.3.2.1 Static condensation
The most famous and widely used technique for model reduction is Guyan reduction
which was developed by both Guyan [124] and Irons [125] independently.
Consider an N DOFs mechanical system subjected to free harmonic motion:

( )
2
0 K M U = (2.2)
where K is the symmetric stiffness matrix with size N N ; and M is the symmetric
mass matrix with size N N . The eigenvalue problem (2.2) has a set of N eigenvalues,
r
and associated eigenvectors,
r
U . Generally, only the first ( ) m m N << eigen-
solutions will be of interest. According to the chosen master and slave DOFs, Eq. (2.2)
can be re-written as:

Literature Review

31

2
0
0
mm ms mm ms m
sm ss sm ss s
K K M M U
K K M M U

| |
( ( ( (
=
|
( ( ( (

\ .
(2.3)
The subscripts m and s represent the number of master and slave DOFs respectively
and the total number of DOFs is m s N + = . From the second set of the above equation,
the relationship between
m
U and
s
U is provided by:

( ) ( ) | |
2 2
0
sm sm m ss ss s
K M U K M U + = (2.4)
or

( ) ( )
1
2 2
s ss ss sm sm m
U K M K M U

= (2.5)
Then, the vector U can be expressed as:

m
m sm m
s sm
U I
U U T U
U t
( (
= = =
( (

(2.6)
where
( ) ( )
1
2 2
sm ss ss sm sm
t K M K M

= (2.7)
If the inertia effect is ignored, Eq. (2.7) becomes:

1
sm ss sm
t K K

=
(2.8)
Substituting Eq. (2.8) into Eq. (2.6) leads to a reduced eigen-problem of order m, in the
form of:

( )
2
0
c c m
K M U =
(2.9)
where
T
c sm sm
K T KT = ; and
T
c sm sm
M T MT = .
Eq. (2.9) yields Guyan reduction which is a static condensation method and is only
accurate at zero frequency because it takes the transformation matrix in the form of Eq.

Literature Review

32
(2.8). It is observed that the frequencies obtained by this method are normally
satisfactory in the domain of
| |
0, 0.3
s
f , where
s
f is the smallest positive eigenvalue
of the structure with all the master DOFs grounded, as shown in Bouhaddis two papers
[126, 127]. This reduction process is not exact and will produce frequencies higher than
those of the full model. As the mode number increases, so does the discrepancy which is
mainly due to the fact that significant inertia terms of the slave DOFs, critical to the
accuracy of higher modes of the system, are discarded. In the higher frequency domain,
errors in the results are large and sometimes unacceptable. To extend the frequency
domain and, therefore, the validity of this technique, the master DOFs must be selected
very carefully in order to increase the value of
s
f , as discovered by some researchers,
e.g., Penny [128].
2.3.2.2 Dynamic condensation
Guyan reduction, although accurate for the reduction of static problems, introduces
large errors when applied to the reduction of dynamic problems. Dynamic condensation
considers the effects of the inertia terms of the slave DOFs. Because these terms are
associated with the inverse of the dynamic stiffness matrix, they cannot be obtained
directly so several methods have been developed to estimate their effects.
In 1984, Paz [129] presented a method based on dynamic condensation which requires
only elementary operations for use in solving a linear system of equations for the
reduction of dynamic systems. Its application to structural problems produces nearly
exact eigenvalues and eigenvectors for all the modes considered in the reduced eigen-
problem. In 1989, Paz [130] presented a modification of the dynamic condensation
method for the solution of the eigen-problem in structural dynamics; it substantially

Literature Review

33
reduces the number of numerical operations required for its application but still provides
accurate solutions for higher order modes of the system.
In 1992, Suarez and Singh [131] presented an iterative approach for eigen-problems by
taking advantage of the orthogonality conditions of the eigenvectors to upgrade the
condensation matrix. Gaussian elimination or matrix inversion is avoided. However,
like other condensation methods, a significant drawback is that special attention should
be paid to the selection of the master DOFs; in other words, the success of this method
depends on the successful selection of the master DOFs. The guidelines for the selection
of master DOFs are summarized below,
The total number of master DOFs should be at least twice big as the number
of modes of interest;
Master DOFs should be in the direction in which the structure is expected to
vibrate;
Master DOFs should be at locations having relatively large mass or rotary
inertia and relatively low stiffness;
If the primary interest is in bending modes, the rotational DOF can be
neglected; and
For a forced analysis, master DOFs should be at locations where forces are
to be applied.
In 1995, Zhang [132] presented a procedure for determining the condensed mass and
stiffness matrices of a structure in which the dynamic contributions from all the original
DOFs of the structure are considered. The reduced model is constructed so that: 1) its
natural frequencies are equal to those of the original structure; and 2) its mode shapes

Literature Review

34
are equal to those of the master DOFs in the corresponding mode shapes of the original
structure. The natural frequencies, the corresponding modes and the dynamic response
used for the condensation are obtained from FEA of the original structure. This model
can be used for additional purposes, such as structural dynamic reanalysis. Though the
proposed method can retain the original dynamic characteristics accurately, its
limitation is that it needs the FEA of the original structure first which is not practical for
complex structures.
In 2000, Qu and Fu [133] presented a new iterative scheme for dynamic condensation.
Comparing it with other iterative methods developed to date, this iterative algorithm has
the advantage that convergence is faster, especially when the eigensolution of the
reduced model is close to the actual solution, because it does not require the stiffness
matrix, the mass matrix and the eigensolution to be calculated at every iterative step. In
this paper, numerical examples are provided to demonstrate the efficiency of the
proposed scheme. Qu et al. [134-136] further developed this iterative algorithm for
application to the model reduction of viscously damped and non-classically damped
vibration systems. The dynamic condensation matrix, which represents the displacement
relationship between the master and slave DOFs, is defined in state-space. A governing
equation for the dynamic condensation matrix is derived from the dynamic equations of
the full model in state-space. The reduced model obtained from the proposed method
can represent the full model when used for different purposes, such as vibration control
and test-analysis model correlation.
In 2003, Lin and Xia [123] proposed a new and effective eigensolution technique, via
iterated dynamic condensation. It retains all the inertia terms associated with the
removed DOFs in an iterated form which generates the reduced mass matrix with a

Literature Review

35
frequency-dependent perturbed term. The corresponding eigenvalues and eigenvectors
of interest are obtained by this iterative procedure combined with an eigen-sensitivity-
based method. The proposed technique is successfully applied to two numerical cases.
Kim [137-139] presented an iterative method for the efficient calculation of eigen-
solutions in dynamic problems. It does not have the drawbacks of other dynamic
condensation methods in which the eigen-pairs are obtained one by one and require
repeated matrix decompositions in the original space. Based on the dynamic
condensation, this method is improved through the use of a modified subspace iteration
and dynamic reduction. The approximate eigenvectors obtained in the dynamic
condensation are used as a starting point for the next step of simplified inverse iteration
and can be improved through orthogonalization using the eigensolution in a subspace.
For better accuracy, an expanded subspace can be formed in terms of master DOFs and
generalized coordinates. Actually, the proposed method is a combination of dynamic
condensation, subspace iteration and modal reduction. Comparisons made with other
iteration methods show that, even when the eigenvalues are closely distributed, good
results can be obtained.
Bouhaddi [127] presented a simplified dynamic condensation method based on the Ritz
approach which considerably improves Guyan reduction method in terms of direct
dynamic analysis or substructuring. This method takes the dynamic effects of the slave
DOFs into consideration and includes a change of variables. The proposed simplified
variant leads to improved precision in the computed eigen-solutions in a well-defined
frequency domain. This method is also easier to apply numerically than Guyan
reduction technique and can be easily adapted to dynamic substructuring. On the other
hand, it also has its own limitations: firstly, like other condensation methods, its success

Literature Review

36
depends on the selection of the master DOFs; and secondly, it becomes inefficient when
the structure is complex.
In 1999, Takewaki [140] developed a new redesign technique based on domain
decomposition and dynamic condensation in order to enhance computational efficiency
in the redesign of large damped structural systems. A structure is treated as an
assemblage of substructures in which the nodes in each substructure are classified as
boundary and interior. The design variables are determined so that the undamped
fundamental natural frequency of the total system and the deformation component ratios
in the resonant damped steady-state vibration will attain the target values. It is shown
that new design sensitivity expressions of the lowest eigenvalue can be derived in terms
of a condensed coordinate system including only boundary nodes. The validity and
order of approximation of the proposed method are demonstrated by examples using a
six-degree-of-freedom FE rod model and a 100-degree-of-freedom FE rod model.
Bouhaddi [141, 142] stated that one of the problems of substructuring methods using
either dynamic condensation or the technique of constrained interfaces is that, to
achieve the assembly, all the junction DOFs among the substructures are conserved.
This can lead to a condensed problem of high dimension when the structure is complex.
The method presented in these two papers is an improvement in substructuring using
linearized dynamic condensation. It consists of reducing the final condensed problem by
a modal projection technique followed by a linearization using the frequency bands of
each substructure. An example is presented of a structure where the initial FE model
(FEM) leads to 4000 DOFs; the condensed problem is reduced to 40 DOFs and the first
30 eigen-modes of the complete structure are calculated with sufficient precision. This
example shows the practical value of the proposed method.

Literature Review

37
In 1997, Qiu et al. [143] presented a new modal synthesis method for finding natural
frequencies and modes of structures. This method expresses the exact residual constraint
modes in two parts, based on an incomplete set of normal modes of substructures with
fixed interfaces: the static constraint modes; and a residual term representing the effect
of normal modes not being retained. This method, which uses a Ritz procedure and an
iterative technique, is presented for solving the non-linear equations of modal synthesis.
Its advantage is that the results produced by its first iteration are identical to the final
results obtained using the Craig-Bampton-Hurty modal synthesis method[144] which
leads to the rapid convergence of subsequent iterations. Numerical examples show that
using the proposed method yields superior accuracy when only a few lower modes are
used.
In a series of papers [145-150], Flippen developed a new condensation approach called
dynamic condensation model reduction (CMR) theory. It utilizes projection operators to
reduce DOFs but still retains the integrity of the response prediction for semi-discrete
models. This method is a general-purpose approach which allows dynamic DOF
reduction in a straightforward fashion without the introduction of additional
approximations. Guyan reduction can be expressed as a special example of this method.
The CMR algorithm avoids the restrictions of both the periodic media and the global
boundary layer. Dyka et al. [151] implemented this theory in a linear FE environment
with transient two-dimensional elasticity problems. The algorithm is permissible but can
only be applied to second-order linear differential equations in the time domain and its
application to deterministic, heterogeneous materials, such as composites, is limited.
Though most of the dynamic condensation techniques can produce accurate results
efficiently in terms of natural frequencies and mode shapes from the reduced model,

Literature Review

38
after condensation, the magnitude of the elements in the reduced stiffness and mass
matrix is not preserved whereas it is in the full stiffness and mass matrix. For example,
from our experience with calculations by using the methods in [129] and [133],
although the order of magnitude of the largest element in the two full stiffness matrices,
a x 10
8
and b x 10
8
, is comparable between the original and modified structures, after
dynamic condensation, in the reduced stiffness matrices, c x 10
18
and d x 10
14
, it is not.
Due to this limitation, it cannot be used in structural dynamic modification for
computation of the delta dynamic stiffness matrix. However, the magnitude of the
elements in the reduced model using Guyan reduction is still preserved.
2.3.2.3 System Equivalent Reduction Expansion Process (SEREP)
In 1989, OCallahan et al. [152] presented a technique called the System Equivalent
Reduction Expansion Process (SEREP) which is formulated as a global mapping
technique to estimate rotational DOFs for experimental modal data. SEREP has several
advantages, the greatest being that the natural frequencies and mode shapes of the
reduced system are exactly the same as those (for the selected modes) of the full system.
However, the drawback of this technique, which limits its usage, is that it requires
conducting an eigensolution of the full system first.
In 2003, Sastry et al. [153] presented an iterative SEREP for the extraction of high
frequency responses from a reduced model under frequency band-limited excitation. To
alleviate the drawback of the original SEREP an iterative method, based on the Sturm
sequence check, is proposed. This method uses eigenvalue separation properties on the
excitation frequency band to identify the optimal number of eigen-pairs required to
capture accurate responses.

Literature Review

39
2.3.2.4 Improved Reduced System (IRS)
In 1987, OCallahan [154] proposed the improved reduced system (IRS) by adding an
extra term to the transformation matrix of Guyan reduction.
The generalized eigen-problem of a system with N DOFs is described, according to the
selected master DOFs and slave DOFs, as:

| |
mm ms mm mm ms mm
mm
sm ss sm sm ss sm
K K M M
K K M M
( ( ( (
=
( ( ( (


(2.10)
where
| |
K and
| |
M are the N N symmetric stiffness and mass matrices respectively;
| |
mm
is the first m mass-normalized eigenvector matrix; and
| |
is the m m
diagonal matrix containing the corresponding eigenvalues, ( 1, 2,..., )
i
i m = . Only the
first m modes are retained in the above equation. The subscripts m and s represent the
number of master and slave DOFs respectively and the total number of DOFs is
m s N + = . Without loss of generality, the eigenvalues are arranged in ascending order,
1 2
...
m
. From the second set of Equation (2.10):

sm mm ss sm sm mm mm ss sm mm
K K M M + = +
(2.11)
Hence,
sm
can be expressed as:

1 1
( )
sm ss sm mm ss sm mm mm ss sm mm
K K K M M

= + +
(2.12)
Let

sm mm
t =
(2.13)
where
| |
t is the transformation matrix between
| |
mm
and
| |
sm
and it can be
expressed as:

Literature Review

40

1 1 1
( )
ss sm ss sm ss mm mm mm
t K K K M M t

= + +
(2.14)
If the second term in Eq. (2.14) is ignored, the transformation matrix
| |
t becomes:

1
ss sm
t K K

=
(2.15)
which is exactly the same as Guyan reduction transformation matrix. Then, the mass-
normalized eigenvector matrix
| |
Nm
becomes:

I
mm mm
Nm mm mm
sm
T
t
( (
= = =
( (


(2.16)
where
| |
I
mm
is the m m unit matrix.
Substituting Eq. (2.16) into Eq. (2.10), a reduced eigen-problem of order m is obtained:

R mm R mm mm
K M =
(2.17)
where
T
R
K T KT = and
T
R
M T MT = are the reduced stiffness and mass matrices
respectively. The eigenvalues of the reduced system are the approximation of
the lowest meigenvalues of the full system so Eq. (2.17) can be re-written as:

1 1
mm mm mm R R
M K

=
(2.18)
Substituting Eq. (2.18) into Eq. (2.14), the transformation matrix
| |
t can be obtained by:

1 1 1
( )
ss sm ss sm ss R R
t K K K M M t M K

= + +
(2.19)
A direct solution for the transformation matrix
| |
t from Eq. (2.19) is impossible. In
1995, Friswell et al. [155] proposed an iterated IRS (IIRS) technique and the
convergence was proved later in Ref. [156] in which the iterative scheme is:

1
( ) 1 1 ( 1) ( 1) ( 1)
( )
k k k k
ss sm ss sm ss R R
t K K K M M t M K


( = + +

(2.20)

Literature Review

41

( )
( )
I
mm k
k
T
t
(
=
(

(2.21)

( ) ( ) ( )
T
k k k
R
K T KT ( =

(2.22)

( ) ( ) ( )
T
k k k
R
M T MT ( =

(2.23)
where the superscript k denotes the k th ( 2 k ) iteration. When 1 k = ,
(1) 1
ss sm
t K K

= ,
which is Guyan reduction, as in Eq. (2.15), and when 2 k = , it is the standard IRS
method proposed by OCallahan [154]. In 1996, Friswell et al. [157] applied this IIRS
method successfully to systems with a local non-linearity.
In 2001, Kim and Kang [158] presented an accelerated method for the IIRS. The
reduction procedures are supplemented with second-order approximations in the series
expansion of the system transformation matrix. The reduced equation of an equivalent
system and the transformation matrix are then updated in an iterative manner. Series
expansion can be considered as repeated updates of the transformation matrix through
inverse iteration. The limitation of this method is that the accuracy of the solution is
sensitive to the selection of the master DOFs; poor selection causes failure of the update
method. However, when this happens, hybrid dynamic condensation can be used.
Unfortunately, the convergence speed of Friswells IIRS method cannot be compared
with that of the subspace iteration method, though some improvements have been made
to increase its convergence speed [158]. Also dependence on the selection of the master
DOFs prevents this method from becoming a more popular eigen-solver in engineering.
In 2004, Xia and Lin [159] presented an improvement on Friswells IIRS method by
modifying the iterative formula of the transformation matrix thereby achieving faster
convergence. Concurrently, the connection between the present iterative scheme and the

Literature Review

42
SIM, and proof of the convergence property are demonstrated. Two numerical examples
demonstrate that the proposed method can obtain the lowest eigen-solutions of
structures more accurately and efficiently than Friswells IIRS method.
2.4 CONCLUSION
In this chapter, structural modification, including both direct and inverse problems, was
reviewed. Also discussed were: for the direct problem, four sub-categories of structural
modification, including their status and the methods used to solve the problem; and, for
the associated topic, the inverse problem, a sensitivity analysis with its status and
methods involved.
From the above literature review, it can be seen lumped structural modification has been
researched extensively, while studies on distributed parameter structural modification
have been sparse and largely incomplete. Thus there is a gap in knowledge on
methodologies for predicting the effects of structural modification in distributed
parameter systems, especially with additional degrees of freedom. This thesis will focus
on addressing this issue. Methods for solving this problem will also be presented. To
determine the FRFs of the original system, the following three methods will be used:
1) For a simple system, such as spring-mass systems, theoretical analysis ;
2) For a complex system, where a theoretical model is not practical, FEA; and
3) As validation of the structural modification method proposed in this thesis,
experimental testing on both the original and modified structures.

Literature Review

43
Guyan reduction will be used to eliminate the rotational DOFs in the FE model when
coupling the experimentally determined FRFs of the original structure with the
numerically determined delta dynamic stiffness matrix.
The method used for determining the delta dynamic stiffness matrix for the distributed
modifications will be discussed in Chapter 3.
44
Chapter 3 METHODOLOGY

In this chapter, methods used to predict the effects of structural dynamic modifications
on dynamic responses are reviewed and developed. Also numerically reviewed and
developed is the description of a modifying structure. The theory of structural
modification is validated on a spring-mass system with lumped modification.
3.1 STRUCTURAL MODIFICATION
In this section, three types of structural dynamic modifications are presented: structural
modification with no change in DOFs; structural modification with reduced DOFs; and
structural modification with additional DOFs.
3.1.1 Structural Modification with No Change in DOFs
The dynamic behaviour of a structure can be expressed through its FRFs, without using
modal parameters such as natural frequencies and mode shapes [160]. Consider a
system represented by the receptance FRF matrix,
| |
0
H , for a harmonic excitation,
{ } ( ) f , where the response, { }
0
( ) u , at frequency,

, can be written as [4]

{ } | |{ }
0 0
( ) ( ) u H f =
(3.1)
or in terms of the dynamic stiffness matrix, which is the inverse of the receptance
matrix,
| | | |
1
0 0
B H

=

Methodology

45

{ } | |{ }
0 0
( ) ( ) f B u =
(3.2)
Now, let the system be subjected to modifications but without changing the DOFs.
Then, the response, { }
1
( ) u , of the modified structure is described by

{ } | |{ }
0 1
( ) ( ) ( ) f B B u = +
(3.3)
where
| |
B is the delta dynamic stiffness matrix (difference between original and
modified dynamic stiffness matrix) introduced by the modifying structure. It is obvious
that

| | | | | | | |
1
1
1 1 0
H B B B

( = = +

(3.4)
or

| | | | | |
1 1
1 0
H H B

= +
(3.5)
By pre-multiplying with
| |
0
H and post-multiplying with
| |
1
H , then Equation (3.5)
becomes

| | | | | | | | | |
0 0 1
H I H B H ( = +

(3.6)
By matrix manipulation,
| |
1
H can be obtained as

| | | | | | | | | |
1
1 0 0
H I H B H

( = +

(3.7)
Equation (3.7) gives the receptance FRF matrix of the modified structure as a function
of the original receptance FRF matrix and the delta dynamic stiffness matrix
| |
B introduced by the modifying part. Also, the order of the matrix inversion is equal
to the number of DOFs of either the original or the modified structure, as it is assumed
that the modification does not change the number of DOFs. If the DOFs of the original

Methodology

46
structure can be divided into two parts: those belonging only to the original structure
(indicated by subscript a); and the interface DOFs through which the modifying part
is connected to the original structure (indicated by subscript b), then

| |
0 0
0
0 0
aa ab
ba bb
H H
H
H H
(
=
(

(3.8)

| |
1 1
1
1 1
aa ab
ba bb
H H
H
H H
(
=
(

(3.9)

| | | | | |
1 0
0 0
0
bb
B B B
B
(
= =
(


(3.10)
By substituting Equations (3.8) to (3.10) into Equation (3.7) and matrix manipulation,
Equation (3.7) is expanded into the following four equations:

| | | | | | | | | |
| | | |
1
1 0 0
1
0
bb bb bb bb
bb
H I H B H
H

( = +

=

(3.11)

| | | | | | | || |
| | | || |
1 0 0 1
0 1
ab ab ab bb bb
ab bb
H H H B H
H H
=
=

(3.12)

| | | |
1 1
T
ba ab
H H =
(3.13)

| | | | | | | || |
| | | || |
1 0 0 1
0 1
aa aa ab bb ba
aa ba
H H H B H
H H
=
=

(3.14)
where
| | | | | |
0ab bb
H B =
(3.15)

| | | | | | | |
0bb bb
I H B = +
(3.16)
Though more equations have to be solved to obtain the FRFs of the modified structure,
the order of matrix inversion is reduced from the DOFs of the original structure to the
interface DOFs which are much fewer.

Methodology

47
3.1.2 Structural Modification with Reduced DOFs
When structural modification reduces the number of DOFs of the original system, the
receptance FRF matrix of the original structure,
| |
0
( ) H , can be partitioned into three
parts[24]:
Remained DOFs means the DOFs which are remained on the original
structure after modification but exclusive of the Interface DOFs (indicated
by subscript a);
Interface DOFs means the DOFs from which the modifying structure is cut
off from the original structure (indicated by subscript b); and
Eliminated DOFs means the DOFs which are belonging to the modifying
structure only but exclusive of the Interface DOFs (indicated by subscript
c).
For structural modification with reduced DOFs, the following equations can be written
for both the original and modified structures:

| | | |
1
0 0 0
1
0 0 0 0 0
0 0 0
aa ab ac
ba bb bc
ca cb cc
H H H
B H H H H
H H H

(
(
= =
(
(

(3.17)

| | | |
1
1
0 0 0
1 1
1 0 0 0
1 1
0 0 0
0
0
0 0 0
aa ab ac
aa ab
ba bb bc
ba bb
ca cb cc
H H H
H H
B H H H B
H H
H H H

(
(
(
(
(
(
= = +
(
(
(
(


(3.18)
where
| | | | | |
1 0
0 0 0
0
0
bb bc
cb cc
B B B B B
B B
(
(
= = (
(
(
(



(3.19)

Methodology

48
By substituting the Equations (3.17) to (3.19) into Equation (3.7) and matrix
manipulation, the following four equations can be obtained to yield the FRFs of the
modified structure:

| | | | | | | | | | | | | |
| | | |
1
1 0 0 0
1
0
bb bb bb bc cb bb
bb
H I H B H B H
H

( = + +

=

(3.20)

| | | | | | | | | | | | | |
| | | || |
1 0 0 0 1
0 1
ab ab ab bb ac cb bb
ab bb
H H H B H B H
H H
( = +

=

(3.21)

| | | |
T
1 1 ab ab
H H =
(3.22)

| | | | | | | | | | | | | |
| | | || |
1 0 0 0 1
0 1
aa aa ab bb ac cb ba
aa ba
H H H B H B H
H H
( = +

=

(3.23)
where
| | | | | | | | | |
0 0 ab bb ac cb
H B H B = +
(3.24)

| | | | | | | | | | | |
0 0 bb bb bc cb
I H B H B = + +
(3.25)
Only one matrix inversion is needed for the computation of the complete receptance
FRF matrix in the above equations. The order of matrix inversion is equal to the number
of interface DOFs (b), which is usually much fewer than the total number of DOFs of
the modified system.
3.1.3 Structural Modification with Additional DOFs
When structural modifications introduce additional DOFs into the system, the
receptance FRF matrix of the modified structure,
| |
1
( ) H , can be partitioned into three
parts [4, 22, 23]:
Original DOFs means the DOFs which are belonging to the original
structure only but exclusive of the Interface DOFs (indicated by subscript
a);

Methodology

49
Interface DOFs means the DOFs through which the modifying structure is
connected to the original structure (indicated by subscript b); and
Passenger DOFs means the DOFs which are belonging to the modifying
structure only but exclusive of the Interface DOFs (indicated by subscript
c).
Then, the following equations can be written for both the original and modified
structures:

| | | |
1
1
0 0
0 0
0 0
aa ab
ba bb
H H
B H
H H

(
= =
(

(3.26)

| | | |
| |
| |
1
1
1 1 1
1
0
1 1 1 1 1
1 1 1
0
0
0 0 0
aa ab ac
ba bb bc
ca cb cc
H H H
H
B H H H H B
H H H

( (
( (
= = = +
( (
( (

(3.27)
where
| | | | | |
1 0
0 0 0
0
0
bb bc
cb cc
B B B B B
B B
(
(
= = (
(
(
(



(3.28)
By substituting the Equations (3.26) to (3.27) into Equation (3.7) and matrix
manipulation, the receptance FRF matrix of the modified structure can be obtained as a
function of the original receptance FRF matrix and the delta dynamic stiffness matrix of
the modifying part, as follows [4]:

| |
1
1 0
1
0
ba ba
ca
H H
H

( (
=
( (

(3.29)

| |
1
1 1 0
1 1
0
0
bb bc bb
cb cc
H H H
H H I

( (
=
( (

(3.30)

| | | | | || |
1
1 0 0
1
0
ba
aa aa ab
ca
H
H H H B
H
(
=
(

(3.31)

Methodology

50

| | | | | | | |
1 1
1 1 0
1 1
0
bb bc
ab ac ab
cb cc
H H
H H H I B
H H
(
(
=
(
(


(3.32)
where
| | | |
0
0 0
0 0 0
bb
I H
B
I

(
( (
= +
(
( (


(3.33)
Only one matrix inversion is needed for the computation of the complete receptance
FRF matrix in the above equations. The order of the matrix inversion is equal to the
number of DOFs of the modifying structure (b+c) which is usually much fewer than the
total number of DOFs of the modified system (a+b+c). Following the same method as
in section 3.1.1, after further matrix manipulation, the receptance FRF matrix of the
modified structure is given as follows:

| | | | | | | | | | | || | | | | |
| | | |
1
1
1 0 0 0
1
0
bb bb bb bb bc cc cb bb
bb
H I H B H B B B H
H

(
= +

=

(3.34)

| | | | | || |
| || |
1
1 1
1
cb cc cb bb
bb
H B B H
H

=
=
(3.35)

| | | | | | | | | | | || | | | | |
| | | |
1
1
1 0 0 0
1
0
ba bb bb bb bc cc cb ba
ba
H I H B H B B B H
H

(
= +

=

(3.36)

| | | | | || |
| || |
1
1 1
1
ca cc cb ba
bb
H B B H
H

=
=
(3.37)

| | | |
T
1 1 bc cb
H H =
(3.38)

| | | |
T
1 1 ab ba
H H =
(3.39)

| | | |
T
1 1 ac ca
H H =
(3.40)

| | | | | | | || |
| | | || |
1 1
1 1
1
cc cc cc cb bc
bc
H B B B H
H

=
=
(3.41)

Methodology

51

| | | | | || || | | || || |
1 0 0 1 0 1 aa aa ab bb ba ab bc ca
H H H B H H B H =
(3.42)
where
| | | |
1
cc
B

=
(3.43)

| | | | | |
| || |
1
cc cb
cb
B B
B

=
=
(3.44)

| | | | | || | | || || | | |
| | | || | | || || |
1
0 0
0 0
bb bb bb bc cc cb
bb bb bb bc
I H B H B B B
I H B H B

= +
= +
(3.45)
Two matrix inversions are involved in the computation of the complete receptance FRF
matrix in the above equations. The orders of the two matrix inversions, (

B
cc
| |
1
and

| |
1
), is equal to the numbers of interface (b) and passenger DOFs (c), respectively,
which are usually much fewer than the total number of DOFs (a+b+c) of the modified
system. Although two matrix inversions are required (Equations (3.34) and (3.43)), the
calculation speed will be much faster compared with Equation (3.29) where the order of
the matrix to be inverted is equal to the number of DOFs (b+c) of the modifying
structure. Also, our experience with calculations using equations from (3.29) to (3.33)
for the cases presented in Chapter 4 indicates that they are more likely to be ill-
conditioned compared with Equations (3.34) and (3.43).
3.2 DELTA DYNAMIC STIFFNESS MATRIX
The delta dynamic stiffness matrix,
| |
B , introduced by the modifying structure,
represents the differences between the dynamic stiffness matrices of the modified and
original structures, i.e., [5]

| | | | | |
1 0
B B B =
(3.46)

Methodology

52
If FE models for both the original and modified structures are available, the difference
between them will generate a matrix whose non-zero elements do not extend beyond the
DOFs of the modifying structure (including the interface and passenger DOFs). It might
be argued that, if such FE models are available, there is no need to use the structural
modification method to predict the FRFs of the modified structure. However, it would
be computationally faster to predict the FRFs of the modified structure using the
structural modification method than by analysing the modified structure from scratch
(i.e., solving a new eigenvalue problem and calculating the FRFs). The saving in
computational time can be quite significant, especially when the system is large and the
modification local. For example, for the modified structure presented in Chapter 5,
computation of the whole FRF matrix within frequency range 0 to 800Hz with a 0.25Hz
frequency resolution in ANSYS would take a HP workstation with a Core 2 Duo E6300
processor about 36 hours to finish while by using structural modification, it only takes
the same computer less than 3 hours to finish.
Normally, in practice, a good FE model is not available for a complex structure so it is
more realistic to assume that only the experimentally determined FRFs are known. On
the contrary, a FE model of the modifying structure can be easily obtained because the
modifying structure is simpler to model. However, these two different models
(experimentally determined FRFs of the original structure and a FE model of the
modifying structure) should be coupled to predict the FRFs of the modified structure.
Usually, while both the translational and rotational DOFs are contained in the FE model,
only the former will be obtained from the experimental FRFs of the original structure
due to the difficulty of measuring the rotational responses subjected to torque excitation,
although there have been some studies reported in this area [38].

Methodology

53
A condensation procedure, such as Guyan reduction [124], the dynamic condensation
[129, 133, 144] or the SEREP[152], can be used to overcome this obstacle by reducing
the number of DOFs in the FE model to that of the experimental measurements.
According to the original definition of the matrix [ ] B , the condensation procedure
should be performed on the FE models of both the original and modified structures to
describe the effects of modification. Thus, the matrix [ ] B , which is reduced to the
experimental DOFs, is given by

| | | | | |
1 0 c c c
B B B =
(3.47)
where
1
[ ]
c
B and
0
[ ]
c
B represent the condensed dynamic stiffness matrices of the
modified and original structures, respectively. After the condensation procedure,
1
[ ]
c
B
and
0
[ ]
c
B , as well as [ ]
c
B , become full matrices. For accurate computation, the FE
models of both the whole original and the modified structures are needed but this is not
practical, as mentioned above. However, the quasi-local character of the matrix [ ] B
can be applied to provide an acceptable procedure for approximately calculating the
matrix [ ]
c
B . For a detailed discussion on the quasi-local character of the matrix [ ]
c
B ,
see [5].
3.3 VALIDATION OF STRUCTURAL MODIFICATION ON SPRING-
MASS SYSTEM
In this section, validation of structural modification on a spring-mass system without
damping will be conducted for three cases: with no change in DOFs; with additional

Methodology

54
DOFs; and with reduced DOFs. Without loss of generalisation, all the springs in the
spring-mass system have the same stiffness and all the masses have the same mass.
3.3.1 Structural Modification with No Change in DOFs
3.3.1.1 Original FRFs of a two-DOF spring-mass system
A two-DOF spring-mass system is illustrated in Figure 3.1 with its FRF matrix
expressed as

11 12
0
21 22
( )
H H
H
H H

(
=
(

(3.48)

Figure 3.1 Two-DOF spring-mass system

If the first DOF is excited by a harmonic force,
1 1
sin f F t = , the differential equation
of motion for the spring-mass system can be obtained as

1 1 1
2 2
0 2
0 0
x x m k k f
x x m k k
( (
+ =
` ` `
( (

) ) )

(3.49)
Assuming the solution to be

1 1
2 2
sin
x X
t
x X


=
` `
) )

and substituting this solution into Equation (3.49), we obtain
k

m

x
2
k

m

x
1

Methodology

55

2
1 1
2
2
2
0
X F k m k
X k k m

(
=
` ` (

) )
(3.50)
By matrix manipulation, the solution can be obtained as

2
1
1 2 2 2
( )
(2 )( )
F k m
X
k m k m k

=



1
2 2 2 2
(2 )( )
Fk
X
k m k m k
=


or

2
1 1
11 2 2 2
1 1
( )
(2 )( )
x X k m
H
f F k m k m k

= = =



2 2
21 2 2 2
1 1
(2 )( )
x X k
H
f F k m k m k
= = =


Similarly, we obtain

12 2 2 2
(2 )( )
k
H
k m k m k
=



2
22 2 2 2
(2 )
(2 )( )
k m
H
k m k m k

=


Therefore, the FRF matrix of this spring-mass system is

0 2
( )
1
( )
(2 ) (2 )( )
k k
H
k k k k k


(
=
(


(3.51)
where
2
m = .
3.3.1.2 Adding a mass, m, to the first mass of the two-DOF spring-mass system
Now, let the spring-mass system be subjected to modification by adding another mass,
m, on the first DOF, as illustrated in Figure 3.2.

Methodology

56


Figure 3.2 Modified two-DOF spring-mass system with mass, m, added on the first
DOF

3.3.1.2.1 FRFs of modified structure by structural modification method
Due to the added mass, m
,
on the first DOF, the delta dynamic stiffness matrix can be
obtained from Equation (3.46) as

| | | |
1 0
2 2
2 2
2
2 2 2
0
0 0
0
0 0
B B B
k m k k m k
k k m k k m
m

=
( (
=
( (


(
=
(

(
=
(

(3.52)
By using Equation (3.7), the FRF matrix of the modified system can be obtained as

1
1 0 0
1
0 2
2 2
2 2
2 2
2 2
[ ]
1 0 0
1
0 1 2 0 0 (2 )( )
1
4 2( )
1 0
4 2( )
2
(2 )( ) (2 )( )
1
2( ) 2( )
H I H B H
k k
H
k k k k k
k k
k k
k k k
k k
k k k k k k
k k
k k k k


= +
( ( (
= +
`
( ( (

)
=
+
(
(
(
+
(
(


(


(
=
(



k

m

x
2
k

m

x
1
m


Methodology

57
or

1 2 2
1
2( ) 2( )
k k
H
k k k k


(
=
(


(3.53)

3.3.1.2.2 FRFs of modified structure by normal method
If we treat the modified system as a whole system then the FRF matrix for the whole
system can be obtained from the following two equations:

)
`

=
)
`


+
)
`

0
2
0
0 2
1
2
1
2
1
f
x
x
k k
k k
x
x
m
m




)
`

=
)
`


+
)
`

2 2
1
2
1
0 2
0
0 2
f x
x
k k
k k
x
x
m
m



Similarly, by matrix manipulation, the FRF matrix of the modified system can be
obtained as

2 2
1
2( ) 2( )
m
k k
H
k k k k


(
=
(


(3.54)
Equation (3.54) is exactly the same as Equation (3.53) and similar results can be
obtained for an added mass on the second DOF. Therefore, structural modification has
been validated on a two-DOF spring-mass system for the case of adding mass on the
DOFs with no change in the number of DOFs.
3.3.1.3 Adding a spring, k , on the second mass of the two-DOF spring-mass
system
Now, let the spring-mass system be subjected to modification by adding another spring,
k , on the second DOF, as illustrated in Figure 3.3.

Methodology

58


Figure 3.3 Modified two-DOF spring-mass system with added spring, k, on second
DOF

3.3.1.3.1 FRFs of modified structure by structural modification method
Due to the added spring, k , on the second DOF, the delta dynamic stiffness matrix can
be obtained as

(

=
k
B
0
0 0
(3.55)
By using Equation (3.7), the FRF matrix of the modified system is obtained by
k

m

x
2
k

m

x
1
k


Methodology

59

1
1 0 0
1
0 2
2
1
2
2
2 2
2
2 2
2 2
[ ]
1 0 0 0
1
0 1 2 0 (2 )( )
1
(2 )( )
1
(2 )( )
2
3 4 ( )
0
(2 )( )
1
3 4 ( )
3 4 ( )
(2
H I H B H
k k
H
k k k k k k
k k k
k
k k k k k
k k
k k
k k k
k k
k k
k

= +

( ( (
= +
`
( ( (


)
=

(
(
(
(
(
(
+

(


=
+
+

2
2 2
2 2
)( ) (2 )( )
2
0 1
2
1
2 (2 )
k
k k
k k k k k
k k
k k
k k k k


(

(
(

(
(


(

(
=
(



or

1 2 2
2
1
( )
2 (2 )
k k
H
k k k k


(
=
(


(3.56)
3.3.1.3.2 FRFs of modified structure by normal method
If we treat the modified system as a whole system, the FRF matrix can be obtained
through the following two equations:

)
`

=
)
`


+
)
`

0 2
2
0
0
1
2
1
2
1
f
x
x
k k
k k
x
x
m
m




)
`

=
)
`


+
)
`

2 2
1
2
1
0
2
2
0
0
f x
x
k k
k k
x
x
m
m



Similarly, by matrix manipulation, the FRF matrix of the modified system is obtained as

2 2
2
1
2 (2 )
m
k k
H
k k k k


(
=
(


(3.57)

Methodology

60
Equation (3.57) is exactly the same as Equation (3.56). Similar results can be obtained
by adding a spring on the first DOF. Therefore, structural modification has been
validated on a two-DOF spring-mass system in the case of a spring being added on the
DOFs with no change in the number of DOFs.
3.3.1.4 Adding a mass, m, and a spring, k , on the second mass of the two-DOF
spring-mass system
Let the spring-mass system be subjected to modification by adding another mass, m,
and a spring, k , on the second DOF, as illustrated in Figure 3.4.

Figure 3.4 Modified two-DOF spring-mass system with added mass, m, and spring,
k, on second DOF

3.3.1.4.1 FRFs of modified structure by structural modification method
Due to the added mass, m, and the spring, k , on the second DOF, the delta dynamic
stiffness matrix can be obtained as

22
B k =

0 0
0
B
k
(
=
(


(3.58)
By using Equation (3.7), the FRF matrix of the modified system can be obtained as
k

m

x
2
k

m

x
1
k
m


Methodology

61

1
1 0 0
1
0 2
2
1
2
2
2
[ ]
1 0 0 0
1
0 1 2 0 (2 )( )
1
(2 )( )
( )
1
(2 )( )
2 (2 )( ) (2 )( )
0
(2 )( )
1
2( )(2
H I H B H
k k
H
k k k k k k
k k k
k k
k k k k k
k k k k k k k
k k k
k k

= +
( ( (
= +
`
( ( (

)
=

(
(
(
(
(
( +

(


=

2
2 2 2
2
2 2
2
)
2 2 3 ( )
( )
(2 )( ) (2 )( )
2
0 1
2( )
1
2 2( )(2 )
k
k k k
k k
k k
k k k k k k
k k
k k
k k k k k


( ( +


( (
(
(


(

(
=
(



or

1 2
2( )
1
( )
2 2( )(2 )
k k
H
k k k k k


(
=
(


(3.59)
3.3.1.4.2 FRFs of modified structure by normal method
If we treat the modified system as a whole system, the FRF matrix can be obtained
through the following two equations:

1 1 1
2 2
0 2
0 2 2 0
x x m k k f
x x m k k
( (
+ =
` ` `
( (

) ) )



1 1
2 2 2
0 0 2
0 2 2
x x m k k
x x f m k k
( (
+ =
` ` `
( (

) ) )




Methodology

62
Similarly, by matrix manipulation, the FRF matrix of the modified system is obtained as

2
2( )
1
( )
2 2( )(2 )
m
k k
H
k k k k k


(
=
(


(3.60)
Equation (3.60) is exactly the same as Equation (3.59). Similar results can be obtained
for adding a mass and a spring on the first DOF. Thus, structural modification has been
validated on a two-DOF spring-mass system in the case of both a mass and a spring
being added on the DOFs with no change in the number of DOFs.
3.3.2 Structural Modification with Additional DOFs
3.3.2.1 FRFs of modified structure by structural modification method
The two-DOF spring-mass system in the above section is subjected to modification with
an additional DOF, as illustrated in Figure 3.5.


Figure 3.5 Two-DOF spring-mass system after structural modification with an
additional DOF

The FRF matrix of the original two-DOF spring-mass system from the above section is
repeated:

| |
0 2
1
2 ( )(2 )
k k
H
k k k k k


(
=
(



Due to the added mass, m, and spring, k , on the second DOF, the delta dynamic
stiffness matrix can be obtained as
k

m

x
2
k

m

x
1
k

m

x
3
original system
modifying part

Methodology

63

| | | | | |
1 0
2 0 2 0
2 0
0 0 0 0
0 0 0
0
0
B B B
k k k k
k k k k k
k k
k k
k k

=
( (
( (
=
( (
( (

(
(
=
(
(

(3.61)
By using Equations (3.29) to(3.32), the FRF matrix of the modified system is obtained
as

1
0 0
1
2 2
2
3 2 2
0 0
0 0 0 0
2
0 0
( )(2 ) ( )(2 )
0 0
0 0
( )(2 )
6 5
ba bb bc bb ba
ca cb cc
H B B I H H
H B B I
k k
I k k
k k k k k k
k k
I
k k k
k k k

( ( ( ( ( (
= +
(
( ( ( ( (



( ( (
( (
(
( (
= +
( (
(
( (


(
( (



=
+
2
2
3 2 2
2
2 3 2 2 3
(2 )
( )(2 )
( )(2 )
(2 )
0
( )(2 )
( )
1
6 5
k k
k
k
k k k
k k k
k k
k k k
k k
k k k k

(
(

(
(

(
(
(
(


(
=
(
+

(3.62)

Methodology

64

1
0 0
2
3 2 2 3
2
2
2 2
2
0 0 0
0 0 0 0
( )(2 )
6 5
(2 )
2
0 ( )(2 )
( )(2 )
(2 )
0 1
( )(2 )
bb bc bb bc bb bb
cb cc cb cc
H H B B I H H
H H B B I I
k k k
k k k
k k
k
k
k k k
k k k
k k
k
k k k

( ( ( ( ( (
= +
( ( ( ( ( (



=
+
(


(

(

(
(

2 2 3 2 2 3
( )(2 ) (2 )
1
(2 ) (2 ) 6 5
k k k k
k k k k k k k


(
(
(
(

(
=
(
+

(3.63)


| | | | | |
0 0
2
2 2 3 2 2 3
2
3 2 2 3
0
( )(2 )
( )
1
0
( )(2 ) 6 5
1
( )(2 )
6 5
bb bc ba
aa aa ab
cb cc ca
B B H
H H H
B B H
k
k k k
k k k k
k
k k k k k k k k k
k k k
k k k




( (
=
( (


(
=
(


( ( (
( ( (
+

( =

+
(3.64)

Methodology

65

| | | | | |
0
2
2 2 3 2 2 3
2
0
0
( )(2 )
1 0 ( )(2 ) (2 )
1
0 1 (2 ) (2 ) 6 5
0
( )(2 )
1 0
0 1
bb bc bb bc
ab ac ab
cb cc cb cc
B B H H
H H H I
B B H H
k
k k k
k k k k k k
k k k k k k k k k
k
k k k




( ( (
=
( ( (


(
=
(


( ( ( (

( ( ( (
+

(
=
(

3 2 2
3 2 2 3 3 2 2 3
2
3 2 2 3
( )(2 ) 3
6 5 6 5
0 1
1
( )
6 5
k k k k k
k k k k k k
k k k
k k k


( ( +
(
( (

+ +
(
( (

( (

( =

+
(3.65)
or

| |
1 3 2 2 3
2 2
2 2 2
1
6 5
( )(2 ) ( )
( ) ( )(2 ) (2 )
(2 ) (2 )
H
k k k
k k k k k k
k k k k k k
k k k k k




=
+
(
(

(
(

(3.66)
3.3.2.2 FRFs of modified structure by normal method
If we treat the modified system as a whole system, the matrix format equation of a
three-DOF spring-mass system can be expressed as

(
(
(




3
2
1
3
2
1
2
2
2
0
2
0 2
F
F
F
X
X
X
m k k
k m k k
k m k

(3.67)
Let
2
m = , then

(
(
(

k k
k k k
k k
B
0
2
0 2
] [ (3.68)

Methodology

66
The inverse of matrix | | B is the receptance FRF matrix of the three-DOF spring-mass
system. So, the matrix inverse is

| |
3 2 2 3
2 2
2 2 2
1
6 5
( )(2 ) ( )
( ) ( )(2 ) (2 )
(2 ) (2 )
m
H
k k k
k k k k k k
k k k k k k
k k k k k




=
+
(
(

(
(

(3.69)
This is exactly the same as the FRF matrix obtained by the structural modification
method in Equation (3.66). Therefore, structural modification has been validated on a
two-DOF spring-mass system in the case of additional DOFs.
3.3.3 Structural Modification with Reduced DOFs
Treating the modified system with additional DOF in section 3.3.2 as the original
system and the original system with the two DOFs in section 3.3.2 as the modified
system would make the validation of structural modification with reduced DOFs as a
simple task. The FRF matrix of the original system, a three-DOF spring-mass system, is
repeated.

| |
0 3 2 2 3
2 2
2 2 2
1
6 5
( )(2 ) ( )
( ) ( )(2 ) (2 )
(2 ) (2 )
H
k k k
k k k k k k
k k k k k k
k k k k k




=
+
(
(

(
(

(3.70)
and the delta dynamic stiffness matrix is

Methodology

67

| | | | | |
1 0
2 0 2 0
0 2
0 0 0 0
0 0 0
0
0
B B B
k k k k
k k k k k
k k
k k
k k

=
( (
( (
=
( (
( (

(
(
=
(
( +

(3.71)
By using Equations (3.20) to (3.25), the FRF matrix of the modified system is obtained
as

| | | | | | | | | |
| || | | |
0 0
2
3 2 2 3 3 2 2 3
2
3 2 2 3
1 1
( )
6 5 6 5
1
6 5
ab bb ac cb
H B H B
k k k k k
k k k k k k
k
k k k


= +
( = +

+ +
( =

+



| | | | | | | | | | | |
| | | || |
| || |
0 0
3 2 2 3
3 2 2 3
3 2 2 3
3 2 2 3
1
1 ( )(2 )
6 5
1
(2 )
6 5
1
4 4
6 5
bb bb bc cb
I H B H B
k k k
k k k
k k k
k k k
k k k
k k k




= + +
= +
+
+
+
( = +

+



| | | | | |
| |
1
1 0
3 2 2 3
2
1
( )(2 )
4 4
1
(2 )
( )(2 )
bb bb
H H
k k
k k k
k
k k k

=
=
+
=



| | | | | || |
| |
1 0 1
3 2 2 3
2
3 2 2 3 2
2
1
( )
6 5
1 1
(2 )
6 5 ( )(2 )
1
( )(2 )
ab ab bb
H H H
k k
k k k
k k
k k k k k k
k
k k k





=
=
+
(

+
=



Methodology

68

| | | |
T
1 1
2
1
( )(2 )
ba ab
H H
k
k k k
=
=




| | | | | || |
1 0 1
2
3 2 2 3
2
3 2 2 3 2
2
1
( )(2 )
6 5
1 1
6 5 ( )(2 )
1
( )
( )(2 )
aa aa ba
H H H
k k k
k k k
k k
k k k k k k
k
k k k


=
( =

+
(

+
=


or

| |
1 1
1 2
1 1
1
2 ( )(2 )
aa ab
ba bb
H H k k
H
H H k k k k k


( (
= =
( (


(3.72)
Equation (3.72) is exactly the same as the FRF matrix for the original structure with the
two DOFs in Equation (3.51) in section 3.3.1.1. Thus, structural modification has been
validated on a three-DOF spring-mass system in the case of reduced DOFs.
3.4 CONCLUSION
The methodology used for structural modification has been derived for three different
cases; with no change in DOFs; with reduced DOFs; and with additional DOFs. It has
also been validated on a spring-mass system with lumped modification.

69
Chapter 4 STRUCTURAL MODIFICATION ON
1D STRUCTURE (BEAM)

In this chapter, structural modification is applied to a beam as a one-dimensional (1D)
structure. Two beam models are considered: one is a simply-supported beam and the
other is a cantilever beam. The former is used to illustrate structural modification with
no change in DOFs, and the latter structural modification with additional DOFs.
4.1 SIMPLY-SUPPORTED BEAM
In this section, numerical examples of a simply-supported beam subjected to structural
modification, which are used mainly to investigate the effects of the number of elements
on the structural modification with no change in DOFs, are presented. The original
beam is a beam with uniform cross-sectional area and the modified beam consists of
two identical beams attached to the top and bottom side of the original beam. Two FE
models with different numbers of elements for the original beam are adopted and are
referred to as the coarse and fine models.
4.1.1 Coarse FE Model

The first example is a simply-supported beam subjected to structural modification with
no change in DOFs and without condensation. The geometry and material properties of

Structural Modification on 1D Structure (Beam)

70
the original and modifying beams are the same as those in [5] and are listed in Table 4.1.
The modified length is 50% of the length of the original beam. Figure 4.1 and Figure
4.2 show the coarse FE models of the original and modified beams respectively. A trial
run of structural modification is conducted.

Table 4.1 Geometry and material parameters of the original simply-supported and
modifying beams
Original beam Modifying beams
Length 1000mm 500mm
Width 7mm 7mm
Thickness 6mm 3mm on each side
Youngs modulus 210 GPa
Poissons Ratio 0.3
Density 7850.0 kg/m
3



Figure 4.1 Coarse FE model of the original beam

Figure 4.2 Coarse FE model of the modified beam


Structural Modification on 1D Structure (Beam)

71
In these two FE models, the one meter long beam is divided into 16 2-D beam elements
with 3 DOFs per node. The node sequence is 1 to 17 from the left end to the right end.
By using the full harmonic analysis in ANSYS V10, the FRF matrix of the original
beam is obtained within the frequency range from 0.125Hz to 800Hz with a frequency
resolution 0.125Hz. An example of the calculated FRFs is shown in Figure 4.3. The
term, 4FY4UY, refers to the FRFs obtained by measuring the displacement response at
node 4 in the Y (Vertical) direction when that node is subjected to a harmonic force
excitation in the same direction. The FRF is plotted in terms of Receptance against
frequency, ie, the measured or calculated FRF is the displacement over the force [8] and
will be used in other figures, also.

Figure 4.3 Calculated FRFs of the original beam at 4FY4UY



Structural Modification on 1D Structure (Beam)

72
As the delta dynamic stiffness matrix, [ ] B , is frequency-dependent, it must be
calculated at each frequency point by

| | | | | |
2
B K M = (4.1)
where is the circular frequency. In Equation(4.1), [ ] K and [ ] M are respectively
the difference in the stiffness and the mass matrices between the original and the
modified beams. Initially, the value of [ ] K and [ ] M are obtained by comparing the
whole FE models of the original beam and the modified beam. It will be shown
subsequently that it is not necessary to model the beams completely and that [ ] K and
[ ] M (and hence [ ] B ) can be determined by comparing partial models of both beams.
The predicted FRFs of the modified beam from the complete FE models of both beams
are compared with the results obtained from full harmonic analysis of the modified
beam for two different locations, node 4 and node 11 respectively, as well as the FRFs
of the original beam, in Figure 4.4 and Figure 4.5. These two locations are selected due
to their reasonable responses in the frequency range interested.

Structural Modification on 1D Structure (Beam)

73

Figure 4.4 Comparison of FRFs of simply-supported beams at 4FY4UY

Figure 4.5 Comparison of FRFs of simply-supported beams at 11FY11UY


Structural Modification on 1D Structure (Beam)

74
As can be seen in Figure 4.4 and Figure 4.5, while the agreement between predicted
results from structural modification and the FRFs obtained from the full harmonic
analysis is reasonable, it is not as good as expected from the results of Chapter 3. One
possible reason could be that the FE models of the original and modified beams are too
coarse to be used for obtaining their modal properties accurately. Thus modal analysis
for both the original and modified beams with finer element sizes and hence a greater
number of elements are carried out. The computed natural frequencies for FE models
with 16 elements, 40 elements, 100 elements and 200 elements for both the original and
modified beams are compared in Table 4.2 and Table 4.3 respectively. In both tables,
the error for other FE models is calculated using the FE model with 200 elements as the
benchmark reference.

Table 4.2 Comparison of natural frequencies from different element sizes of FE
model for original beam
Modes
FE model
with 200
elements
FE model with 100
elements
FE model with 40
elements
FE model with 16
elements
Natural
frequencies
Natural
frequencies
Error
(%)
Natural
frequencies
Error
(%)
Natural
frequencies
Error
(%)
1 14.072 14.072 0 14.072 0 14.072 0
2 56.285 56.285 0 56.285 0 56.286 0.002
3 126.63 126.63 0 126.63 0 126.64 0.008
4 225.10 225.10 0 225.10 0 225.16 0.027
5 351.67 351.67 0 351.68 0.003 351.89 0.063
6 506.32 506.32 0 506.34 0.004 506.97 0.128
7 689.03 689.03 0 689.07 0.006 690.65 0.235
8 899.76 899.76 0 899.85 0.010 903.31 0.395



Structural Modification on 1D Structure (Beam)

75
Table 4.3 Comparison of natural frequencies from different element sizes of FE
model for modified beam
Modes
FE model
with 200
elements
FE model with 100
elements
FE model with 40
elements
FE model with 16
elements
Natural
frequencies
Natural
frequencies
Error
(%)
Natural
frequencies
Erro
r (%)
Natural
frequencies
Erro
r (%)
1 19.903 19.903 0 19.903 0 19.903 0
2 63.537 63.537 0 63.537 0 63.537 0
3 179.04 179.04 0 179.04 0 179.06 0.011
4 330.6 330.6 0 330.61 0.003 330.73 0.039
5 461.35 461.35 0 461.36 0.002 461.69 0.074
6 675.31 675.31 0 675.34 0.004 676.32 0.150
7 988.57 988.58 0.001 988.66 0.009 991.85 0.332
8 1201.4 1201.4 0 1201.5 0.008 1201.8 0.033

From Table 4.2 and Table 4.3, a conclusion can be drawn that the FE models with 16
elements for both beams are too coarse to be used for obtaining the natural frequencies.
As matrix manipulation, including matrix inversion, is involved in structural
modification, small errors in the elements of the matrices may accumulate and magnify,
leading to inaccurate results from structural modification. The error in the computed
natural frequencies of the FE model with 40 elements is less than 1% for the first 8
frequencies compared with those of the FE model with 200 elements. This model is,
therefore, considered to be adequate for investigation of the structural modification
method, without having to incur excessive computing resources (speed, memory and
storage).

Structural Modification on 1D Structure (Beam)

76
4.1.2 Fine FE Model
4.1.2.1 Full model simply-supported beam without condensation
The geometry and material properties of the original and modifying beams are the same
as the one in the coarse model (Table 4.1). Figure 4.6 and Figure 4.7 show the FE
models of the original and modified beams respectively.

Figure 4.6 FE model of original beam


Figure 4.7 FE model of modified beam

In these two FE models, the 1m-long beams are divided into 40 2-D beam elements with
3 DOFs per node. The node sequences for both the original and modified beams are 1 to
41 from the left end to the right end. By using the full harmonic analysis in ANSYS
V10, the FRF matrix of the original beam is obtained within the frequency range of
0.125Hz to 800Hz with a frequency resolution 0.125Hz. An example of the calculated
FRFs is shown in Figure 4.8.

Structural Modification on 1D Structure (Beam)

77

Figure 4.8 Calculated FRFs of original beam at 10FY10UY

Comparing the FE models of the original and modified beams gives the delta stiffness
matrix, [ ] K , and the delta mass matrix, [ ] M . The delta dynamic stiffness matrix,
[ ] B , is calculated at each frequency point by Equation(4.1).
The predicted FRFs of the modified beam using the complete FE models of both beams
to determine [ ] K and [ ] M are compared with the results obtained from the full
harmonic analysis of the modified beam, as well as the FRFs of the original beam, and
are presented in Figure 4.9 and Figure 4.10 for excitation and response at node 10 and
15 respectively.

Structural Modification on 1D Structure (Beam)

78

Figure 4.9 Comparison of FRFs of simply-supported beams at 10FY10UY

Figure 4.10 Comparison of FRFs of simply-supported beams at 15FY15UY

Structural Modification on 1D Structure (Beam)

79
As can be seen from Figure 4.9 and Figure 4.10, the results predicted by the structural
modification method are in very good agreement with the results computed from the FE
model of the modified beam. Thus it appears that a mesh size of 40 elements is
sufficiently fine to provide accurate results for structural modification and will be
employed in future analyses. Until now, the delta stiffness and delta mass matrices are
calculated by comparing the complete FE model of the modified beam to the complete
FE model of the original beam. If both structures need to be modelled in full in FE, then
the procedure for structural modification is hardly required to determine the FRFs of the
modified structure. They can be obtained from FE model directly. Hence, taking
advantage of the quasi-local characteristic of the delta dynamic stiffness matrix [ ] B ,
the feasibility of using partial models of the original and the modified structure to obtain
[ ] K and [ ] M matrices is investigated below Three partial FE models are used; their
lengths are 0.7m, 0.8m and 0.9m, corresponding to 70%, 80% and 90% of the original
beam, named as partial FE model 1, 2 and 3 respectively. By partially modelling both
beams with the same length, which covers the modifying area, and comparing the two
FE models, the matrices, [ ] K and [ ] M , are obtained. The matrices, [ ] B , are
obtained from Equation (4.1) at each frequency point for the different FE models.
Figure 4.11 and Figure 4.12 show the comparison of the predicted FRFs from different
FE models with different size with the calculated FRFs from the full harmonic analysis
of the modified beam, at nodes 10 and 15 respectively.

Structural Modification on 1D Structure (Beam)

80

Figure 4.11 Comparison of predicted FRFs from different-sized FE models at
10FY10UY

Figure 4.12 Comparison of predicted FRFs from different-sized FE models at
15FY15UY

Structural Modification on 1D Structure (Beam)

81
It can be seen from Figure 4.11 and Figure 4.12 that identical FRF curves can be
obtained from the above computation for the modified beam whether the matrix, [ ] B ,
is obtained from either the whole or the partial FE models.
A good indication of the agreement between the FRFs predicted by the structural
modification method and by direct Finite Element Analysis (FEA) can be examined by
calculating the Frequency Response Assurance Criterion (FRAC) values for all nodes.
The FRAC values are calculated using the following Equation [161]:

{ } { }
{ } { }
( )
{ } { }
( )
2
'
' '
( ) * ( )
( ) * ( ) ( ) * ( )
cal ij prd ij
ij
cal ij cal ij prd ij prd ij
f f
FRAC
f f f f


=

(4.2)
where ( )
cal ij
f is the FRF from ANSYS harmonic analysis (full or reduced), which is a
1 n vector;
( )
prd ij
f is the FRF from structural modification prediction, which is a 1 n
vector;
is the circular frequency;
* stands for matrix multiplication;
n is the frequency points included in the FRF
Figure 4.13, to Figure 4.16 show the distributions of the FRAC values among the FRFs
predicted from the whole FE model and the partial FE models 1 (0.7m), 2 (0.8m) and 3
(0.9m), and the FRFs obtained from the full harmonic analysis of the modified beam,
with only the transverse direction being taken into consideration, i.e., only the FRFs
iFYjUY (i = 2,3, , 40, j = 2, 3, , 40). The higher the FRAC value, the better the
correlation between the predicted FRF and the calculated FRF. A FRAC value of 1
indicates perfect correlation while a value of 0 indicates no correlation.

Structural Modification on 1D Structure (Beam)

82

Figure 4.13 FRAC values between predicted FRFs from whole FE model and
calculated FRFs from full harmonic analysis of the modified beam

Figure 4.14 FRAC values between predicted FRFs from partial FE model 1 (0.7m)
and calculated FRFs from full harmonic analysis of the modified beam
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

83

Figure 4.15 FRAC values between predicted FRFs from partial FE model 2 (0.8m)
and calculated FRFs from full harmonic analysis of the modified beam

Figure 4.16 FRAC values between predicted FRFs from partial FE model 3 (0.9m)
and calculated FRFs from full harmonic analysis of the modified beam
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

84
Table 4.4 shows the maximum, minimum and average FRAC values of the predicted
FRFs from the different-sized FE models compared with the calculated FRFs from the
full harmonic analysis of the modified beam. It can be seen that the agreement is
excellent, with an average FRAC value of over 0.995 in all cases. The fact that the
maximum, minimum and average FRAC values are exactly the same also indicates that,
in this full model case, the whole and partial FE models generate the same delta
dynamic stiffness matrices. As stated in Chapter 3, If FE models for both the original
and modified structures are available, the difference between them will generate a delta
dynamic stiffness matrix whose non-zero elements do not extend beyond the DOFs of
the modifying structure (including the interface and passenger DOFs). In other words,
the effects introduced by the modifying structure on the delta dynamic stiffness matrix
are self contained by the elements in the matrix corresponding to the modifying
structure itself. In this case, no matter how big the partial model is, as long as it covers
the modifying area, it will generate a delta dynamic stiffness matrix exactly the same as
the whole model. Although not presented here, this has been verified for partial models
of lengths down to 0.55 m.

Table 4.4 Maximum, minimum and average FRAC values from different-sized FE
models compared with FRFs from full harmonic analysis
FE models Maximum Minimum Average
Whole 1 0.966 0.9959
Partial FE model 3(0.9m) 1 0.966 0.9959
Partial FE model 2(0.8m) 1 0.966 0.9959
Partial FE model1 (0.7m) 1 0.966 0.9959

It must be stated that no condensation procedure has been applied in this example as it is
used mainly to validate the strategy for predicting the effects of structural modification.

Structural Modification on 1D Structure (Beam)

85
4.1.2.2 Reduced model simply-supported beam with condensation
This example is of the simply-supported beam subjected to structural modification with
no change in DOFs but with condensation. The geometry and material properties of the
original and modifying beams used in the model are the same as the ones used in the
previous section, and the same FE models with 40 elements along the length of the
beams are adopted.
By using the reduced harmonic analysis in ANSYS V10, the reduced FRF matrix of the
original beam is obtained containing information only in the transverse direction, Y,
within the frequency range of 0.25Hz to 800Hz and a frequency resolution of 0.25Hz.
However, as the FE models of the original and modified beams contain information in
both the translational and rotational directions, Guyan reduction procedure [124] is
employed to reduce the FE models to only considering the transverse direction, Y, for
both the original beam matrices,
| |
0
K and
| |
0
M and the modified beam matrices,
| |
1
K
and
| |
1
M . After Guyan reduction,
| |
0
K ,
| |
0
M ,
| |
1
K and
| |
1
M become
| |
0c
K ,
| |
0c
M ,
| |
1c
K and
| |
1c
M . Comparing the reduced stiffness matrix and the mass matrix gives the
matrices,
| |
c
K and
| |
c
M . The matrix,
| |
c
B is obtained by

| | | | | |
2
c c c
B K M = (4.3)
where is the circular frequency.
One whole, and three partial, FE models (partial FE models 1 (0.7m), 2 (0.8m) and 3
(0.9m)) are employed to obtain the matrix
| |
c
B .
It must be stated that before Guyan reduction, the non-zero elements of the matrix
| |
B
do not extend beyond the DOFs of the modifying part. But after Guyan reduction, the
non-zero elements extend to the whole matrix of the partially modelled modified beam;

Structural Modification on 1D Structure (Beam)

86
that is, not only are the DOFs at the physical interface nodes (the attachment points)
non-zero, so are the DOFs belonging to the partial FE models. Therefore, all the DOFs
included in this matrix which belong to the original structure should also be treated as
interface DOFs. It should be noted that the definition of interface DOFs has changed
from physical to numerical. The results of the original beam and the modified beam
using full and reduced harmonic methods at 10FY10UY are compared with those for
the modified beam using the structural modification method for the whole beam with
Guyan reduction in Figure 4.17. It can be seen that the results using the structural
modification with Guyan reduction for the modified beam are virtually identical to those
obtained by the reduced harmonic analysis of the modified beam.

Figure 4.17 Comparison of FRFs of simply-supported beam at 10FY10UY

Structural Modification on 1D Structure (Beam)

87

Figure 4.18 Comparison of predicted FRFs from different-sized FE models at
10FY10UY by coupling reduced harmonic analysis results and Guyan-reduced
delta stiffness matrix

Figure 4.18 shows that by coupling the reduced FRF matrix of the original beam and the
four matrices,
| |
c
B , the predicted FRFs of the modified beam using the structural
modification method with different-sized FE models, yield virtually identical results
with the calculated FRFs from the reduced harmonic analysis of the modified beam.
Figure 4.19, to Figure 4.22 show the distribution of the FRAC values among the FRFs
predicted from coupling the reduced harmonic results of the original beam and the
Guyan reduced delta dynamic stiffness matrices from different-sized FE models, and the
FRFs calculated by the reduced harmonic analysis of the modified beam, with a 2Hz
frequency band. Table 4.5 shows the maximum, minimum and average FRAC values,
with a 2Hz frequency band.

Structural Modification on 1D Structure (Beam)

88

Figure 4.19 FRAC values between predicted FRFs from whole FE model and
calculated FRFs from reduced harmonic analysis of the modified beam

Figure 4.20 FRAC values between predicted FRFs from partial FE model 1 (0.7m)
and calculated FRFs from reduced harmonic analysis of the modified beam
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

89

Figure 4.21 FRAC values between predicted FRFs from partial FE model 2 (0.8m)
and calculated FRFs from reduced harmonic analysis of the modified beam

Figure 4.22 FRAC values between predicted FRFs from partial FE model 3 (0.9m)
and calculated FRFs from reduced harmonic analysis of the modified beam

Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

90
Table 4.5 Maximum, minimum and average FRAC values from different-sized FE
models compared with FRFs from reduced harmonic analysis, with 2Hz frequency
band
FE models Maximum Minimum Average
Whole 0.9885 0.9549 0.9724
Partial FE model 3 (0.9m) 0.9900 0.9667 0.9786
Partial FE model 2 (0.8m) 0.8284 0.7017 0.7513
Partial FE model 1 (0.7m) 0.5644 0.2510 0.4838

It can be seen that although the average FRAC value of the predicted FRFs from partial
FE model 1 (0.7m) is only 0.48, the agreement is generally good, with an average
FRAC value of over 0.75 for the other 3 models. An additional numerical process was
carried out to determine what could be the main reason for the poor predicted results
from partial FE model 1: the area covered by it or the number of elements in it.
4.1.2.3 Additional numerical evaluation of partial FE model 1
In this example, the area covered by partial FE model 1 has been re-meshed as shown in
Figure 4.23 and Figure 4.24. The crossed-out areas in the figures are the modifying
areas.



Figure 4.23 Comparison of FE models of original beam in partial FE model 1

(a) FE model of the original beam before re-meshing in partial FE model 1
(b) FE model of the original beam after re-meshing in partial FE model 1

Structural Modification on 1D Structure (Beam)

91



Figure 4.24 Comparison of FE models of the modified beam in partial FE model 1

The re-meshed partial FE model 1 covers the same area as the partial FE model 1 in the
previous section, but with double the number of elements. The FRFs of the original
beam are re-calculated using the reduced harmonic analysis in ANSYS for the re-
meshed original beam. Structural modification is carried out by coupling the re-
calculated FRF matrix and the Guyan reduced delta dynamic stiffness matrix,
| |
c
B ,
from the re-meshed partial FE model 1. Figure 4.25 and Figure 4.26 show the predicted
FRFs compared with the calculated FRFs for the modified beam.
(a) FE model of the modified beam before re-meshing in partial FE model 1
(b) FE model of the modified beam after re-meshing in partial FE model 1

Structural Modification on 1D Structure (Beam)

92

Figure 4.25 Comparison of predicted FRFs from re-meshed partial FE model 1
with calculated FRFs for the modified beam at 13FY13UY

Figure 4.26 Comparison of predicted FRFs from re-meshed partial FE model 1
with calculated FRFs for modified beam at 23FY23UY

Structural Modification on 1D Structure (Beam)

93
As shown in Figure 4.25 and Figure 4.26, good agreement has been obtained between
the predicted FRFs from the re-meshed partial model 1 and the calculated FRFs for the
modified beam. Figure 4.27 shows the distribution of the FRAC values between the
predicted FRFs from the re-meshed partial model 1 and the calculated FRFs for the
modified beam, with a 2Hz frequency band. Table 4.6 shows the maximum, minimum
and average FRAC values.

Table 4.6 Maximum, minimum and average FRAC values for re-meshed partial
FE model 1 compared with calculated FRFs from reduced harmonic analysis, with
2Hz frequency band
FE models Maximum Minimum Average
Partial FE model 1 (0.7m)
before re-meshing
0.5644 0.2510 0.4838
Partial FE model 1 (0.7m) after
re-meshing
0.9877 0.9839 0.9861


Figure 4.27 Distribution of FRAC values of predicted FRFs from re-meshed
partial FE model 1 compared with calculated FRFs for modified beam
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

94
Comparing Figure 4.27 with Figure 4.20, it can be seen that the agreement between the
predicted FRFs and the calculated FRFs has improved dramatically from that in Section
4.1.2.2 by the re-meshing of partial FE model. From this additional numerical process,
the main reason for the poor predicted results from partial FE model 1 in Section 4.1.2.2
is determined by the number of elements. By double the number of elements in this
partial FE model, the agreement between the predicted FRFs and the calculated FRFs
has improved dramatically.
4.2 CANTILEVER BEAM
In this section, four examples are presented: a cantilever beam subjected to structural
modification with additional DOFs and without condensation (the full model); the same
cantilever beam subjected to structural modification with additional DOFs and with
condensation (the reduced model); a simulation for the experimental procedure; and
validation of the structural modification method with experimental data.
4.2.1 Full Model -- Cantilever Beam without Condensation
In this example, a cantilever beam subjected to structural modification is analysed with
additional DOFs and without condensation. The geometry and material properties of the
original and modifying beams used in the model, as listed in Table 4.7, are those of the
actual aluminum alloy beams employed in the experiments described in section 4.2.4.
Figure 4.28 and Figure 4.29 show the FE models of the original and modified beams
respectively. The Youngs modulus and the density of the beam were determined by FE
model updating using FEMTools [162, 163]. As it can be seen from Figure 4.29, the

Structural Modification on 1D Structure (Beam)

95
modified beam consists of the original beam to which the modifying beam is attached,
thus introducing additional DOFs to the original beam in Figure 4.28.



Figure 4.28 FE model of original cantilever beam



Figure 4.29 FE model of modified cantilever beam

Table 4.7 Geometry and material parameters of original cantilever and modifying
cantilever beams
Original beam Modifying beam
Length 825mm
*
500mm
Width 50.275mm 50.275mm
Thickness 5.921mm 5.921mm
Youngs modulus 64.811GPa
Poissons Ratio 0.33
Density 2622.7 kg /m
3

* Full length of beam 1100 mm; clamped length 275mm

In the FE model, the 825mm-long beam is divided into 33 3-D beam elements with 6
DOFs per node. The node sequence is 1 to 34 from the clamped end to the free end with
nodes 15 and 33 corresponding to the attachment points of the modifying beam which
Fixed end
Fixed end

Structural Modification on 1D Structure (Beam)

96
are located at 350mm and 800mm respectively from the clamped end. The 500 mm-long
modifying beam corresponding to 60.6% of the length of the original beam is divided
into 20 3-D beam elements with 6 DOFs per node and node sequence from 35 to 55.
The modifying beam is attached to the original beam through short beam elements with
full metal-to-metal contact. By using the full harmonic analysis in ANSYS V10, the
FRF matrix of the original beam is obtained within the frequency range of 0.25Hz to
800Hz and a frequency resolution of 0.25Hz. Figure 4.30 shows the calculated
receptance FRF of the original beam at 9FZ9UZ. Comparing the two FE models for the
original and modified beams gives the matrices [ ] K and [ ] M . The matrix, [ ] B , is
calculated at each frequency point by Equation (4.1).

Figure 4.30 Calculated FRFs of original beam at 9FZ9UZ from full harmonic
analysis


Structural Modification on 1D Structure (Beam)

97
The FRFs of the modified beam predicted by the structural modification method are
compared with the results obtained from the full harmonic analysis of the modified
beam in Figure 4.31 and Figure 4.32 for 9FZ9UZ and 9FZ34UZ respectively.

Figure 4.31 Comparison of FRFs of cantilever beams at 9FZ9UZ


Structural Modification on 1D Structure (Beam)

98

Figure 4.32 Comparison of FRFs of cantilever beams at 9FZ34UZ

As can be seen in Figure 4.31 and Figure 4.32, the predicted results from the whole FE
model using the structural modification method are in good agreement with the results
computed from the FE model of the modified beam. Now, two partial FE models of the
original and modified beams are used to obtain the [ ] K and [ ] M because of the
quasi-local characteristic of the matrix [ ] B ; their lengths are 0.6m and 0.7m
corresponding to 72.7% and 84.8% of the original beam respectively. By partially
modelling both the original and modified beams with the same length, which covers the
modifying area, and comparing these two FE models, the matrices [ ] K and [ ] M are
obtained. The matrices [ ] B are obtained from Equation (4.1). Figure 4.33 and Figure
4.34 show the comparison of the predicted FRFs from different-sized FE models with
the calculated FRFs for the modified beam at 9FZ9UZ and 9FZ34UZ respectively.

Structural Modification on 1D Structure (Beam)

99

Figure 4.33 Comparison of predicted FRFs from different-sized FE models at
9FZ9UZ for modified beam

Figure 4.34 Comparison of predicted FRFs from different-sized FE models at
9FZ34UZ for modified beam

Structural Modification on 1D Structure (Beam)

100
It can be seen in Figure 4.33 and Figure 4.34 that identical FRF curves can be obtained
from the above computation for the modified beam whether the matrix, [ ] B is
obtained from the whole or the partial FE models. Figure 4.35 shows the distribution of
the FRAC values of the predicted receptance FRFs from whole FE model compared
with the calculated receptance FRFs for the modified beam, taking into consideration
only the transverse direction, i.e., the FRF iFZjUZ(i = 2,3, , 55, j = 2, 3, , 55).
Because the magnitude of the first mode is dominant in the whole frequency range and
the difference between the predicted FRFs and the calculated FRFs is large on the
magnitude plot at this mode (Figure 4.33 and Figure 4.34), the FRAC value plot is not
as good as expected with maximum value at 0.8636, minimum value at 0.0046 and
average value at 0.5245. If the first mode is excluded in the FRAC value calculation, the
maximum value, minimum value and average value are increased to 0.9992, 0.8871 and
0.9824 respectively. As expected, the FRAC value distribution plots from Figure 4.36 to
Figure 4.38 with the first mode excluded show excellent agreement. Since normally the
inertance FRFs are measured in practice, the FRAC value is re-calculated for the whole
FE model, partial FE model 1 and 2 by comparing the inertance FRFs. Figure 4.39,
Figure 4.40 and Figure 4.41 show the distribution of the FRAC values calculated from
inertance FRFs.
Table 4.8 shows the maximum, minimum and average FRAC values by comparing
inertance FRFs from different-sized FE models with calculated FRFs for modified beam.
It can be seen that the agreement is much improved with an average FRAC value of
about 0.8 for all cases because the discrepancies at the first mode are less dominant in
the inertance than in the receptance.

Structural Modification on 1D Structure (Beam)

101

Figure 4.35 Distribution of FRAC values of predicted receptance FRFs from whole
FE model compared with calculated FRFs for the modified beam

Figure 4.36 Distribution of FRAC values of predicted receptance FRFs from whole
FE model compared with calculated FRFs for the modified beam with the first
mode excluded
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

102

Figure 4.37 Distribution of FRAC values of predicted receptance FRFs from
partial FE model 1 compared with calculated FRFs for the modified beam with the
first mode excluded

Figure 4.38 Distribution of FRAC values of predicted receptance FRFs from
partial FE model 2 compared with calculated FRFs for the modified beam with the
first mode excluded
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

103

Figure 4.39 Distribution of FRAC values of predicted inertance FRFs from whole
FE model compared with calculated FRFs for the modified beam

Figure 4.40 Distribution of FRAC values of predicted inertance FRFs from partial
FE model 1 compared with calculated FRFs for the modified beam
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

104

Figure 4.41 Distribution of FRAC values of predicted inertance FRFs from partial
FE model 2 compared with calculated FRFs for the modified beam

Table 4.8 Maximum, minimum and average FRAC values by comparing inertance
FRFs from different-sized FE models with calculated FRFs for modified beam
FE models Maximum Minimum Average
Whole FE model 0.9862 0.2138 0.7985
Partial FE model 1 (0.7m) 0.9862 0.2138 0.7985
Partial FE model 2 (0.6m) 0.9862 0.2138 0.7985

In the example in this section, no condensation procedure was applied as it is used
mainly to validate the strategy for predicting the effects of distributed structural
modification with additional DOFs. The exact same FRAC values from the whole and
partial FE models confirm the finding in Section 4.1.2.1.
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

105
4.2.2 Reduced Model -- Cantilever Beams with Condensation
In this section, the same cantilever beam is subjected to structural modification with
additional DOFs and with condensation. The geometry and material properties of the
original and modifying beams used in the models are the same as the ones used in the
previous numerical example (Table 4.7) while the same FE models of the original and
modified beams are adapted.
By using the reduced harmonic analysis in ANSYS V10, the reduced FRF matrix of the
original beam, containing only the DOFs in the transverse direction, Z, within the
frequency range of 0.25Hz to 800Hz and a frequency resolution of 0.25Hz, is obtained.
An example of the calculated FRFs is shown in Figure 4.42.

Figure 4.42 Calculated FRFs of original beam at 9Z9Z from ANSYS harmonic
analysis


Structural Modification on 1D Structure (Beam)

106
As the reduced FRF matrix contains information only in the transverse direction, Z, and
the FE models of the original and modified beams contain information in both the
translational and rotational directions, Guyan reduction procedure is employed to reduce
the FE models to the transverse direction, Z, for both the original beam matrices
| |
0
K
and
| |
0
M and the modified beam matrices
| |
1
K and
| |
1
M . After Guyan reduction,
| |
0
K ,
| |
0
M ,
| |
1
K and
| |
1
M become
| |
0c
K ,
| |
0c
M ,
| |
1c
K and
| |
1c
M . Comparing the
reduced stiffness matrix with the mass matrix gives the matrices
| |
c
K and
| |
c
M . The
matrix
| |
c
B is obtained by Equation (4.3).
One whole, and two partial, FE models are employed to obtain the matrix
| |
c
B . The
previous definition of the interface DOFs changing from physical to numerical still
holds; that is, before Guyan reduction, the non-zero elements of the matrix
| |
B do not
extend beyond the DOFs of the modifying structure but, after applying the reduction
procedure, the non-zero elements extend to the whole matrix of the partially modelled
modified beam. In other words, not only are the DOFs at the physical interface nodes
(the attachment points) non-zero so are the DOFs belonging to all the nodes of the
original structure used in the partial FE models. Therefore, all the DOFs included in this
matrix which belong to the original structure should also be treated as interface DOFs.
As shown in Figure 4.43 and Figure 4.44 for 9FZ9UZ and 34FZ34UZ respectively, the
agreement between the FRFs of the modified beam calculated using the full harmonic
analysis and the reduced harmonic analysis is very good up to 600 Hz beyond which
there are significant discrepancies. These discrepancies are due to Guyan reduction. On
the other hand, the FRFs of the whole FE model for the modified beam calculated by
the structural modification method with Guyan reduction are almost indistinguishable

Structural Modification on 1D Structure (Beam)

107
from those obtained by the reduced harmonic analysis of the modified beam for the full
frequency range of interest up to 800 Hz.

Figure 4.43 Comparison of FRFs of cantilever beams at 9FZ9UZ

Structural Modification on 1D Structure (Beam)

108

Figure 4.44 Comparison of FRFs of cantilever beams at 34FZ34FZ

By coupling the reduced FRF matrix of the original beam and the three matrices,
| |
c
B ,
the FRFs for the modified beam using the structural modification method with Guyan
reduction are predicted for the whole FE model and two partial models (0.6 m and 0.7).
As shown in Figure 4.45 and Figure 4.46, these FRFs are virtually identical to those
calculated by the reduced harmonic analysis of the modified beam. The fact that the
predicted FRFs using the structural modification method and the calculated FRFs from
the reduced harmonic analysis do not agree with the FRFs calculated by full harmonic
analysis at the higher order modes beyond 600 Hz is due to Guyan reduction procedure.
Guyan reduction procedure is never exact and will always produce frequencies higher
than those of the full model for natural frequencies greater than 0.3f
s
, where f
s
stands for
the first flexible mode when all the master DOFs are constrained. As the mode number
increases, the discrepancy increases. This is mainly due to the fact that significant

Structural Modification on 1D Structure (Beam)

109
inertia terms of the eliminated DOFs, which are critical to the accuracy of the higher
modes of the system, are discarded [126, 128].

Figure 4.45 Comparison of predicted FRFs from different-sized FE models at
9FZ9FZ by coupling reduced harmonic analysis results and Guyan reduced delta
dynamic stiffness matrices

Structural Modification on 1D Structure (Beam)

110

Figure 4.46 Comparison of predicted FRFs from different-sized FE models at
34FZ34FZ by coupling reduced harmonic analysis results and Guyan reduced
delta dynamic stiffness matrices

Figure 4.47, Figure 4.48 and Figure 4.49 show the distribution of the FRAC values
of the predicted FRFs from different-sized FE models compared with the
calculated FRFs of the modified beam from the reduced harmonic analysis with
the original frequency resolution of the FRFs which is 0.25 Hz. A further check
by comparing the calculated FRFs from reduced harmonic analysis to the
calculated FRFs from full harmonic analysis for the modified beam reveals the
minimum FRAC value is 0.8176, maximum FRAC value is 1 and average FRAC
value is 0.9993. As in Section 4.2.1, the magnitude of the first mode is dominant
in the whole frequency range and the difference between the predicted FRFs
and the calculated FRFs is large on the magnitude plot at this mode (Figure 4.45

Structural Modification on 1D Structure (Beam)

111
and Figure 4.46), the FRAC value plot is not as good as expected with maximum
value at 0.9437, minimum value at 0.0321 and average value at 0.7077 for the
whole FE model. If the first mode is excluded in the FRAC value calculation, the
maximum value, minimum value and average value are increased to 0.9902,
0.5041 and 0.8921 respectively for whole FE model. As expected, the FRAC
value distribution plots in Figure 4.50 and Figure 4.52 with the first mode
excluded show excellent agreement for whole FE model and partial FE model 2
respectively, while Figure 4.51 shows no correlation between the predicted FRFs
and the calculated FRFs from partial FE model 1 at all. The FRAC value is re-
calculated for the whole FE model, partial FE model 1 and 2 by comparing the
inertance FRFs. Figure 4.53, Figure 4.54 and Figure 4.55 show the distribution of
the FRAC values calculated from inertance FRFs. Table 4.9 shows the maximum,
minimum and average FRAC values with the original frequency resolution of
the FRFs which is 0.25 Hz. It can be seen that the agreement is good with an
average FRAC value of over 0.65 in all cases from the receptance FRFs FRAC
value calculation.

Structural Modification on 1D Structure (Beam)

112

Figure 4.47 Distribution of FRAC values of predicted FRFs from whole FE model
compared with calculated FRFs of the modified beam from reduced harmonic
analysis

Figure 4.48 Distribution of FRAC values of predicted FRFs from partial FE model
1 (0.6 m) compared with calculated FRFs of the modified beam from reduced
harmonic analysis
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

113

Figure 4.49 Distribution of FRAC values of predicted FRFs from partial FE model
2 (0.7 m) compared with calculated FRFs of the modified beam from reduced
harmonic analysis

Figure 4.50 Distribution of FRAC values of predicted receptance FRFs from whole
FE model compared with calculated FRFs of the modified beam from reduced
harmonic analysis with the first mode excluded
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

114

Figure 4.51 Distribution of FRAC values of predicted receptance FRFs from
partial FE model 1 compared with calculated FRFs of the modified beam from
reduced harmonic analysis with the first mode excluded

Figure 4.52 Distribution of FRAC values of predicted receptance FRFs from
partial FE model 2 compared with calculated FRFs of the modified beam from
reduced harmonic analysis with the first mode excluded
Node number
N
o
d
e

n
u
m
b
e
r

N
o
d
e

n
u
m
b
e
r

Node number

Structural Modification on 1D Structure (Beam)

115


Figure 4.53 Distribution of FRAC values of predicted inertance FRFs from whole
FE model compared with calculated FRFs for the modified beam

Figure 4.54 Distribution of FRAC values of predicted inertance FRFs from partial
FE model 1 compared with calculated FRFs for the modified beam
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

116

Figure 4.55 Distribution of FRAC values of predicted inertance FRFs from partial
FE model 2 compared with calculated FRFs for the modified beam

Table 4.9 Maximum, minimum and average FRAC values from different-sized FE
models compared with calculated receptance FRFs of the modified beam from
reduced harmonic analysis
FE models Maximum Minimum Average
Whole FE model
Receptance 0.9437 0.0321 0.7077
Inertance 0.9825 0.1182 0.6851
Partial FE model 2 (0.7m)
Receptance 0.9512 0.0284 0.7107
Inertance 0.9827 0.1226 0.7008
Partial FE model 1 (0.6m)
Receptance 0.9836 0.0004 0.6628
Inertance 0.9256 0.76826e-5 0.2771

4.2.3 Simulation for Experimental Procedure
In Sections 4.2.1 and 4.2.2, the whole FRF matrix was calculated from the FE model of
the original beam. In practice, usually the elements of only one row or one column of
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

117
the whole FRF matrix array are measured in modal testing. The measured FRFs are then
imported into modal identification software (here, STAR Modal version 5 [164, 165]
was used) to identify the modal parameters, such as the natural frequencies, damping
ratios and mode shapes. These modal parameters can then be used to create the whole
FRF matrix by using a FRF synthesis procedure.
In order to minimize the residual effects of the unknown higher order modes, a
frequency range from 0 Hz to 4800 Hz with a resolution of 0.5Hz was used. One
column of the FRFs, due to the force excitation at node 9 in the Z direction as calculated
from the FE model, was imported into the software STAR Modal version 5. Table 4.10
shows the identified natural frequencies of the original beam from STAR Modal version
5 compared with the calculated natural frequencies from ANSYS. Totally there are 16
modes within frequency range from 0 Hz to 4800 Hz. Table 4.11 shows the first 10
identified mode shapes of the original beam from STAR Modal version 5. Mode shapes
for modes 11 to 16 are not shown because the number of measurement points in the
beam is not sufficient to display the true mode shapes for these higher order modes. As
shown in Table 4.10, the differences between the natural frequencies for the first 16
modes calculated from ANSYS and those identified in STAR Modal are insignificant.

Table 4.10 Identified natural frequencies of original beam from STAR Modal
version 5 compared with calculated natural frequencies from ANSYS
Modes
Calculated from ANSYS Identified from STAR Modal
Natural frequencies (Hz) Natural frequencies (Hz) Error (%)
1 6.9858 6.9820 -0.05483
2 43.777 43.777 -0.00032
3 122.57 122.565 -0.00413
4 240.15 240.146 -0.00158
5 396.91 396.91 0.00051

Structural Modification on 1D Structure (Beam)

118
6 592.80 592.80 0.00012
7 827.78 827.78 -0.00020
8 1101.8 1101.81 0.00091
9 1414.9 1414.88 -0.00141
10 1767.0 1766.94 -0.00340
11 2158.1 2158.04 -0.00278
12 2588.2 2588.20 0
13 3057.5 3057.50 0
14 3565.9 3564.50 -0.03926
15 4113.6 4114.20 0.01459
16 4700.8 4702.03 0.02617

Table 4.11 First 10 identified mode shapes of the original beam from STAR Modal
version 5
Modes Mode shapes Modes Mode shapes
1

2

3

4

5

6


Structural Modification on 1D Structure (Beam)

119
7

8

9

10


After identification of the modal parameters, the whole FRF matrix was synthesized
within a frequency range of 0 Hz to 800 Hz and a resolution of 0.25Hz by using
equation [164]

*
*
1
( )
2 ( ) 2 ( )
N
k ij k ij
ij
k
k k
r r
H s
j s p j s p
=
(
= +
(

(

(4.4)
where
k k k
p j = + , the pole location in the s-place for the kth mode;

k ij
r is the element, ij , in the complex residue matrix of the kth mode;

k
is the damping coefficient of the kth mode, in this case, the damping value is
from the measured case and proportional damping is assumed;
k
is the damped natural frequency of the kth mode;
s is the Laplace variable;
* denotes the complex conjugate; and
N is the number of modes included in the synthesize procedure.
Figure 4.56 and Figure 4.57 show the synthesized FRFs compared with the calculated
FRFs of the original beam at different locations (9FZ9UZ and 9FZ34UZ). Note, most

Structural Modification on 1D Structure (Beam)

120
comparisons are done for nodes 9 and 34 because node 9 is the driving point for the
experiment described in section 4.2.4 and node 34 is the free end of the beam. From
Figure 4.56 and Figure 4.57, it can be seen that the FRFs obtained from FRF synthesis
are in great agreement with the calculated FRFs from ANSYS reduced harmonic
analysis. Only one apparent discrepancy can be observed in the phase plot in Figure
4.57 around 330Hz. The apparent discrepancy is due to the fact that is the same
phase angle, and where jumps occur is somewhat arbitrary and due to bit errors. A
FRAC value plot is shown in Figure 4.58. Since only the FRFs subjected to the
excitation at node 9 are available from calculation, the FRAC value plot is a bar plot
with the response nodes along the x axis.

Figure 4.56 Comparison of synthesized FRFs with calculated FRFs of the original
beam from reduced harmonic analysis at 9FZ9UZ

Structural Modification on 1D Structure (Beam)

121

Figure 4.57 Comparison of synthesized FRFs with calculated FRFs of the original
beam from reduced harmonic analysis at 9FZ34UZ

Figure 4.58 Distribution of FRAC values of synthesized FRFs compared with
calculated FRFs of the original beam from reduced harmonic analysis

Structural Modification on 1D Structure (Beam)

122
From Figure 4.58, it can be seen that the synthesized FRFs are in good agreement with
the calculated FRFs with most of the FRAC values greater than 0.8 except one value at
9FZ31UZ which is about 0.75.
By using the same procedure as in section 4.2.2, the FRFs of the modified structure
were predicted using the synthesized FRFs of the original structure. Figure 4.59 and
Figure 4.60 show the FRFs predicted by the structural modification method with FE
models of different sizes compared with the calculated FRFs from the reduced harmonic
analysis at different locations (9FZ9UZ and 9FZ34UZ).

Figure 4.59 Comparison of FRFs of the modified beam at 9FZ9FZ

Structural Modification on 1D Structure (Beam)

123

Figure 4.60 Comparison of FRFs of the modified beam at 9FZ34FZ

From Figure 4.59 and Figure 4.60, it can be seen that the predicted results from
structural modification, by using the synthesized results and the Guyan reduced delta
dynamic stiffness matrices from different-sized FE models, are in reasonable agreement
with those calculated by the reduced harmonic analysis of the modified beam in terms
of frequencies and amplitudes. There are, however, some apparent discrepancies in the
phase results due to the same reason as the fact that is the same phase angle, and
where jumps occur is somewhat arbitrary and due to bit errors. Figure 4.61, Figure 4.62
and Figure 4.63 show the distribution of the FRAC values between the FRFs predicted
from different-sized FE models using synthesized FRFs of the original beam and the
FRFs calculated by reduced harmonic analysis of the modified beam, with a 4Hz
frequency bandwidth by only taking the magnitude of the FRFs into consideration.
Table 4.12 shows the maximum, minimum and average FRAC values, with a 4Hz

Structural Modification on 1D Structure (Beam)

124
frequency bandwidth by only taking the magnitude of the FRFs into consideration. It
can be seen that the agreement is good with an average FRAC value of over 0.81 in all
cases.

Figure 4.61 FRAC values between predicted FRFs from whole FE model and
calculated FRFs from reduced harmonic analysis of the modified beam by only
taking the magnitude of the FRFs into consideration
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

125

Figure 4.62 FRAC values between predicted FRFs from partial FE model 1 (0.6m)
and calculated FRFs from reduced harmonic analysis of the modified beam by
only taking the magnitude of the FRFs into consideration

Figure 4.63 FRAC values between predicted FRFs from partial FE model 2 (0.7m)
and calculated FRFs from reduced harmonic analysis of the modified beam by
only taking the magnitude of the FRFs into consideration
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on 1D Structure (Beam)

126
Table 4.12 Maximum, minimum and average FRAC values from different-sized
FE models compared with FRFs from reduced harmonic analysis by only taking
the magnitude of the FRFs into consideration, with 4 Hz frequency bandwidth
FE models Maximum Minimum Average
Whole FE model 0. 9844 0. 5889 0. 8164
Partial FE model 1 (0.6m) 0. 9913 0. 6621 0. 8617
Partial FE model 2 (0.7m) 0. 9843 0. 5885 0. 8168

4.2.4 Structural Modification with Experimental Data
In this section, structural modification with experimental data is conducted on a
cantilever aluminum beam subjected to modification by attaching a short aluminum
beam, with the same material parameters as of the original beam, to the original beam.
4.2.4.1 Modal testing
Experimental modal testing was conducted on both the original and modified beams
which were clamped to a very heavy frame with two G clamps. This set-up was used to
simulate cantilever conditions for vibration in the transverse direction. The dimensions
of the original beam were 110050.2755.921mm with 45 measurement points equally
spaced along the length of the beam while the area from points 34 to 45 was clamped to
the frame, as shown in Figure 4.64 and Figure 4.65. In order to be consistent with the
simulation in section 4.2.3, the node sequence was re-arranged from 1 at the clamped
end to 34 at the free end. Therefore, the valid length of the beam was only 825mm with
33 measurement points.
Shaker excitation, using a B&K 4810 electromagnetic shaker, was employed at node 9
in the transverse direction, Z, and the response measurements in the same direction were
made using B&K type 4374 accelerometers mounted on the beam using beeswax. The
input was burst random noise generated by the HP3566/67 FFT analyzer. In order to

Structural Modification on 1D Structure (Beam)

127
minimize the residual effects of the unknown higher order modes, the modal testing was
conducted four separate times with: a frequency range of 0 Hz to 800 Hz and a
resolution of 0.25 Hz; a frequency range of 800 Hz to 1600 Hz and a resolution of 0.25
Hz; a frequency range of 1600 Hz to 3200 Hz and a resolution of 0.5 Hz; and a
frequency range of 3200 Hz to 4800 Hz and a resolution of 0.5 Hz. Force measurements
were obtained from a B&K type 8001 impedance head connected between the beam and
through a 40mm-long stinger to the shaker. The shaker was supported by a retort stand
and the impedance head was fixed to the beam using a threaded stud, as shown in
Figure 4.64 and Figure 4.65. The shadow area in Figure 4.64 stands for the clamped
area by the G-clamp in Figure 4.65.



Figure 4.64 Schematic of experimental set-up for original beam


Structural Modification on 1D Structure (Beam)

128

Figure 4.65 Photograph of modal testing of original beam

For burst random noise excitation, up to 50 averages were taken at each measurement
point. No window was used because there was no leakage problem due to the natural
characteristic of burst random noise excitation [166]. Several trial measurements were
taken to set the noise amplitude, which was determined by maximizing the coherence
between the force and the response signals. Figure 4.66 and Figure 4.67 show the
comparison between measured FRFs and calculated FRFs at both the driving point and
the free end of the original beam.

Structural Modification on 1D Structure (Beam)

129

Figure 4.66 Comparison between measured FRF and calculated FRF of original
beam at driving point (9FZ9UZ) from shaker tests

Figure 4.67 Measured FRFs of original beam at free end (9FZ34UZ) from shaker
tests

Structural Modification on 1D Structure (Beam)

130
From Figure 4.66 and Figure 4.67, we can see that on the magnitude plot, for the first 4
modes or within 300 Hz, the calculated FRFs matched the measured FRFs very well.
There was a slight difference for the 5
th
mode and a significant difference for the 6
th

mode. On the other hand, for the phase plot, the agreement between the measured FRFs
and the calculated FRFs is very poor. The measured natural frequencies of the original
beam from the modal test with shaker excitation did not agree well with the natural
frequencies from the FE analysis, presumably because the measured natural frequencies
were affected by the neglect the spring lines of out-of-band higher modes and a small
mass loading effect of the force transducer between the shaker and the beam. So, a
modal test using an impact hammer, B&K type 8202 with B&K type 8200 force
transducer integrated, was used to measure the natural frequencies of the original beam.
Figure 4.68 and Figure 4.69 show comparison between measured FRFs from the impact
hammer test and calculated FRFs at different locations of the original beam. Because of
the tapped hole at node 9 to mount the impedance head in the shaker test, it is hard to
mount the accelerometer close to node 9 in the impact hammer test. So the
accelerometer was mounted at different locations. From Figure 4.68 and Figure 4.69, we
can see that on the magnitude plot, for the 6 modes within frequency range of interest,
the calculated FRFs matched the measured FRFs very well. However, for the phase plot,
the agreement between the measured FRFs and the calculated FRFs is also very poor
like the shaker test. Table 4.13 shows the comparison of the natural frequencies between
the modal tests using an impact hammer and those using a shaker, with a FE model
analysis within the frequency range of 0Hz to 800Hz. It can be seen from Table 4.13
that the modal testing using impact hammer results in frequencies that are in better
agreement (less than 1%) with the FE values.

Structural Modification on 1D Structure (Beam)

131

Figure 4.68 Comparison between measured FRFs from the impact hammer test
and calculated FRFs of original beam at 17FZ6UZ

Figure 4.69 Comparison between measured FRFs from the impact hammer test
and calculated FRFs of original beam at 26FZ6UZ

Structural Modification on 1D Structure (Beam)

132
Table 4.13 Comparison of natural frequencies of original beam obtained by
different methods (0Hz to 800Hz)
Mode
Modal test
(impact hammer)
Modal test (shaker) FEM
Frequency (Hz)
Frequency
(Hz)
Error (%)
Frequency
(Hz)
Error (%)
1 7.03 7.00 -0.339 6.99 -0.569
2 43.73 44.76 2.363 43.78 0.114
3 122.62 121.95 -0.544 122.57 -0.041
4 240.12 238.45 -0.694 240.15 0.012
5 396.36 392.93 -0.867 396.91 0.139
6 593.70 574.26 -3.274 592.80 -0.152

Since the entire FRF matrix,
0
[ ] H is needed for structural modification, it is necessary
to perform 3333 = 1089 FRF measurements. Actually,
0
[ ] H is symmetric due to
reciprocity and, therefore, only its lower (or upper) triangular part needs to be
experimentally determined; this leads to (1+33) 33/2 = 561 FRF measurements - still a
considerable number. However, knowing that all the dynamic properties, such as the
natural frequencies and mode shapes, are contained in each row or each column of the
FRF matrix,
0
[ ] H , only the elements of one row or one column were measured here.
The remaining terms in the matrix were obtained through a FRF synthesis by combining
the natural frequencies and damping ratios from modal testing using an impact hammer
with the mode shapes from modal testing using a shaker. The mode shapes determined
by impact hammer test were not used because of two reasons: 1) several points in the
mode shapes determined by impact hammer were not quite right; and 2) the main
purpose of impact hammer test was to determine the natural frequencies; with the
difficulty to mount the accelerometer at node 9, the measurement at node 9 was omitted
in impact hammer test. Besides reducing the number of measurements, this procedure

Structural Modification on 1D Structure (Beam)

133
has the advantage of smoothing the data. However, it has the drawback of neglecting the
residual effects of the higher order modes in the synthesized FRFs.
4.2.4.2 Structural modification
The synthesized matrix,
0
[ ] H , contains only the transverse direction data but, the matrix
[ ] B from the FE model contains data in all six DOFs. The Guyan reduced matrices,
[ ] B , in the previous numerical example in section 4.2.3 are used here to predict the
FRFs of the modified structure. They are compared with the FRFs experimentally
measured, at two different locations on the modified structure, the driving point (node 9)
and the free end (node 34), respectively, in Figure 4.70 and Figure 4.71. It should be
noted that the modified beam is also excited at node 9 in the Z direction.

Figure 4.70 Comparison of FRFs of modified cantilever beam (measurements at
driving point, node 9)

Structural Modification on 1D Structure (Beam)

134

Figure 4.71 Comparison of FRFs of modified cantilever beam (measurements at
free end, node 34)

It can be seen from these figures that the predicted FRFs are in reasonably good
agreement with the measured FRFs of the modified structure in terms of frequencies
and amplitudes for the first 5 modes up to about 300 Hz. There are some significant
discrepancies in phase because the phase discrepancies between the measured FRFs and
the calculated FRFs are inherited and magnified by the matrix manipulation involved in
the FRF synthesis and structural modification. The discrepancies between the
predictions and measurements increase in the higher order modes, for example, the 6th,
7th and 8th modes do not agree very well, mostly due to the neglect of the spring lines
of out-of-band higher modes and the effects of the joints. The mass loading effects of
the shaker and the force transducer may also have contributed, to a lesser degree, to the
observed discrepancies.

Structural Modification on 1D Structure (Beam)

135
4.3 CONCLUSION
The effects of distributed structural modification on the dynamic response of a beam as
a 1-D structure with two different boundary conditions, namely simply-supported and
cantilever, with no change in DOFs and with additional DOFs have been determined
efficiently and successfully. Also, the structural dynamic modification method with no
change of DOFs and with additional DOFs for coupling the experimentally measured
FRFs and the numerical models has been validated. For example, for the modified
structure presented in Section 4.2.1, computation of the whole FRF matrix within the
frequency range 0 to 800Hz with a 0.25Hz frequency resolution in ANSYS would take
a HP workstation with a Core 2 Duo E6300 processor about 12 hours to finish while by
using structural modification, it only takes the same computer less than 2 hours to finish.
This is because in the full harmonic analysis in ANSYS, the matrix inversion is
involved whose order is (a+b+c), while in structural modification the order of matrix
inversion is only b and c, where a, b and c holds the same meaning as in Section 3.1.3.
For both cases, structural modification with no change in DOFs and with additional
DOFs, by using the numerically determined FRFs of the original structure, from full
harmonic analysis and matrices, [ ] B , of different-sized FE models without reduction,
virtually identical FRF curves for the modified structure can be obtained.
In the reduced model of structural modification with no change in DOFs, good
agreement of the FRF curves for the modified structure can be obtained from whole FE
model, partial FE model 3 (corresponding to 90% of the original beam) and partial FE
model 2 (corresponding to 80% of the original beam). The predicted results from partial
FE model 1 (corresponding to 70% of the original beam) did not agree with the
calculated FRFs well. A further study by doubling the elements in this partial model

Structural Modification on 1D Structure (Beam)

136
reveals that the main reason for the poor prediction is due to the number of elements in
the partial model. In the reduced model of structural modification with additional DOFs,
good agreement of the FRF curves for the modified structure can be obtained
irrespective of whether the delta dynamic stiffness, [ ] B , is obtained from a whole FE
or a partial FE model.
Good agreement of the magnitude of the FRF curves for the modified structure is
obtained from numerical simulation for the experimental procedure, regardless of
whether the delta dynamic stiffness matrix, [ ] B , is obtained from a whole or a partial
FE model. The phase of the FRF curves from structural modification does not match the
calculated FRFs for the modified beam because the phase discrepancies in the
synthesized FRFs are inherited and magnified by the matrix manipulation involved in
the structural modification.
By using the experimentally determined FRFs of the original structure and the Guyan
reduced matrices, [ ] B , from different-sized FE models, good agreement is obtained
between the predicted and measured FRFs of the modified structure at frequencies of up
to 300Hz. The agreement is not ideal at higher frequencies for three reasons: the
discrepancy between the FE models of the beams and the physical beams such as
welding joints between the small beam and the modifying beam not being modelled in
the FE simulation; Guyan reduction and the mass loading effects introduced by the
shaker test.
The definition of the interface DOFs has been extended from physical to numerical to
include the non-zero effects after Guyan reduction procedure.

137
Chapter 5 STRUCTURAL MODIFICATION ON
A 2D STRUCTURE (PLATE)

In this chapter, the structural modification method developed in Chapter 3 is applied to
a clamped rectangular plate as an example of a 2-D structure, in order to illustrate its
performance with reduced DOFs. Four cases are presented in this chapter: three
numerical simulations and one experimental validation.
5.1 NUMERICAL SIMULATIONS
5.1.1 Full Model Clamped Rectangular Plate without Condensation

This first numerical example of structural modification with reduced DOFs is an
aluminum plate clamped along four sides with a rectangular cut-out in the center
(Figure 5.1 and Figure 5.2). The dimensions of the original and modified plates are
listed in Table 5.1. The material properties are: Youngs modulus=70 GPa; Poissons
ratio=0.33; and density=2800 kg/m
3
.

Structural Modification on a 2D Structure (Plate)

138

Figure 5.1 FE model of clamped rectangular plate without cut-out


Figure 5.2 FE model of clamped plate with rectangular cut-out

Structural Modification on a 2D Structure (Plate)

139
Table 5.1 Dimensions of original and modified plates

In each of these two FE models, the plate is divided into 20 shell elements per side with
6 DOFs per node. For the original plate, the node sequence starts from 1 at the upper
right corner and, in a clockwise spiral manner, ends at 441 in the center. For the
modified plate, the node numbering follows the same sequence as the original plate but
is from 1 to 432. The cut-out area is from nodes 417 to 441 with nodes 417 to 432 as the
physical interface nodes. The clamped boundary condition is set on nodes 1 to 80. By
using the full harmonic analysis in ANSYS V10, the FRF matrix of the original plate is
obtained within the frequency range of 0.25 Hz to 800 Hz and a 0.25 Hz frequency
resolution. Figure 5.3 and Figure 5.4 show the calculated FRFs of the original plate at
different locations (271FZ271UZ and 401FZ271UZ). These two locations were selected
because they have reasonable responses for all the modes within the frequency range of
interest.
Original plate Modified plate
Length 490mm 490mm
width 410mm 410mm
Thickness 4mm
4mm(with a 98mmX82mm window cut-out in
the center, ie, 4% of the original plate in term of
area)

Structural Modification on a 2D Structure (Plate)

140

Figure 5.3 Calculated FRF of original plate at 271FZ271UZ

Figure 5.4 Calculated FRF of the original plate at 401FZ271UZ

Structural Modification on a 2D Structure (Plate)

141
One whole, and three partial, FE models are employed to obtain the delta dynamic
stiffness matrices| | B . The dimensions of partial FE models 1, 2 and 3 are 294 mm x
246 mm, 343 mm x 287 mm and 392 mm x 328 mm, corresponding to 9 times, 12.25
times and 16 times of the cut-out and 36%, 49% and 64% of the original plate in terms
of area respectively. Using the equations for structural modification with reduced DOFs,
as described in Chapter 3, the predicted FRFs are compared with the results from the
full harmonic analysis of the modified plate, as shown in Figure 5.5 to Figure 5.8.

Figure 5.5 Comparison of FRFs of original and modified plates at 271FZ271UZ

Structural Modification on a 2D Structure (Plate)

142

Figure 5.6 Comparison of FRFs of original and modified plates at 401FZ271UZ

Figure 5.7 Comparison of predicted FRFs from different-sized FE models at
271FZ271UZ

Structural Modification on a 2D Structure (Plate)

143

Figure 5.8 Comparison of predicted FRFs from different-sized FE models at
401FZ271UZ

From Figure 5.5 to Figure 5.8, it can be seen that virtually identical FRF curves are
obtained from structural modification of the modified plate irrespective of whether the
delta dynamic stiffness matrix | | B is obtained from the whole, or the partial, FE
models.
Figure 5.9, to Figure 5.12 show the distributions of the FRAC values of the FRFs
predicted from the whole FE model and the partial FE models 1, 2 and 3 respectively,
compared with the FRFs calculated from full harmonic analysis of the modified plate,
with only the transverse direction being taken into consideration, i.e., only the FRFs,
iFZjUZ(i = 81,82, , 432, j = 81, 82, , 432), are used in the calculations of the
FRAC values with the original frequency resolution of the FRFs which is 0.25 Hz.

Structural Modification on a 2D Structure (Plate)

144

Figure 5.9 FRAC values between predicted FRFs from whole FE model and FRFs
from full harmonic analysis for modified plate

Figure 5.10 FRAC values between predicted FRFs from partial FE model 1 and
FRFs from full harmonic analysis for modified plate
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

145

Figure 5.11 FRAC values between predicted FRFs from partial FE model 2 and
FRFs from full harmonic analysis for modified plate

Figure 5.12 FRAC values between predicted FRFs from partial FE model 3 and
FRFs from full harmonic analysis for modified plate
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

146
Table 5.2 shows the maximum, minimum and average FRAC values of the predicted
FRFs from different-sized FE models compared with the calculated FRFs of the
modified plate from full harmonic analysis. The results are in good agreement, with an
average FRAC value of 0.75 for all cases. The fact that the maximum, minimum and
average FRAC values are exactly the same also indicates that, in this full model case,
the whole and partial FE models generate the same delta dynamic stiffness matrices.
This result is the same as that for the beam discussed in Section 4.1.2.1.

Table 5.2 Maximum, minimum and average FRAC values from different-sized FE
models compared with FRFs from full harmonic analysis
FE models Maximum Minimum Average
Whole 1 0.51 0.75
Partial FE model 1 1 0.51 0.75
Partial FE model 2 1 0.51 0.75
Partial FE model 3 1 0.51 0.75

5.1.2 Reduced Model - Clamped Rectangular Plate with Condensation
In this second example, the same plates are subjected to structural modification with
reduced DOFs and condensation of the rotational DOFs. The geometry and material
properties of the original and modified plates are the same as in the previous numerical
example, as are the FE models.
By employing the reduced harmonic analysis in ANSYS V10, the reduced FRF matrix
of the original plate is obtained using information for only the transverse direction, Z,
within the frequency range of 0.25 Hz to 800 Hz and a 0.25 Hz frequency resolution. As
shown in Figure 5.13 and Figure 5.14, calculated FRFs from the reduced and full

Structural Modification on a 2D Structure (Plate)

147
harmonic analysis of the original plate at different locations (271FZ271UZ and
401FZ271UZ) are in excellent agreement.

Figure 5.13 Comparison of calculated FRFs of original plate from reduced and full
harmonic analysis at 271FZ271UZ

Structural Modification on a 2D Structure (Plate)

148

Figure 5.14 Comparison of calculated FRFs of original plate from reduced and full
harmonic analysis at 401FZ271UZ

The same procedure as was applied to the reduced beam model in Chapter 4 is adopted
here to obtain the delta dynamic stiffness matrices [ ]
c
B from four different-sized FE
models, i.e., one whole and three partial. By coupling the reduced FRF matrix of the
original plate and the four matrices [ ]
c
B , the predicted FRFs are compared with the
calculated results for the modified plate, in Figure 5.15 and Figure 5.16.

Structural Modification on a 2D Structure (Plate)

149

Figure 5.15 Comparison of FRFs of original and modified plates, using different
methods, at 271FZ271UZ

Figure 5.16 Comparison of predicted FRFs from different-sized FE models at
271FZ271UZ, by coupling reduced harmonic analysis results and Guyan reduced
delta stiffness matrices


Structural Modification on a 2D Structure (Plate)

150
From Figure 5.15 and Figure 5.16, it can be seen that the structural modification, using
the reduced harmonic analysis results and the Guyan reduced delta dynamic stiffness
matrices from the whole FE model and partial FE model 3, yields virtually identical
results to those calculated by the full harmonic analysis of the modified plate. The
predicted FRFs from the partial FE models 1 and 2 dont match the calculated results
very well but those from partial FE model 2 are slightly better than those from partial
FE model 1. This indicates that there is a minimum size requirement for maintaining
accuracy in the partial FE model used for reduced structural modification. These results
are quite different from those shown for beam simulation in Chapter 4 and Ref [23]
which confirm that the predicted FRFs from all the partial FE models are in great
agreement with the calculated results for the modified beam.
Figure 5.17, to Figure 5.20 show the distributions of the FRAC values between the
FRFs predicted from the whole FE model, the partial FE models 1, 2 and 3, and the
FRFs calculated from full harmonic analysis of the modified plate respectively, with a
4Hz frequency bandwidth.

Structural Modification on a 2D Structure (Plate)

151

Figure 5.17 FRAC values between predicted FRFs from whole FE model and FRFs
from full harmonic analysis of the modified plate

Figure 5.18 FRAC values between predicted FRFs from partial FE model 1 and
FRFs from full harmonic analysis of the modified plate
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

152

Figure 5.19 FRAC values between predicted FRFs from partial FE 2 model and
FRFs from full harmonic analysis of the modified plate

Figure 5.20 FRAC values between predicted FRFs from partial FE model 3 and
FRFs from full harmonic analysis of the modified plate
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

153

Table 5.3 lists the maximum, minimum and average FRAC values of the predicted
FRFs from different-sized FE models compared with the calculated FRFs of the
modified plate from full harmonic analysis, with a 4Hz frequency bandwidth. The
predicted FRFs from the full FE model and partial FE model 3 are in good agreement
with the calculated FRFs, having average FRAC values of 0.98 and 0.94 respectively.
The average FRAC values between the predicted FRFs from partial FE models 2 and 1,
and the calculated FRFs are only 0.32 and 0.06 respectively. The decreasing FRAC
values, from those for the whole FE model to those for the smallest partial FE model,
also indicate that there is a minimum size required for maintaining accuracy in the
partial FE model used in reduced structural modification.

Table 5.3 Maximum, minimum and average FRAC values from different-sized FE
models compared with FRFs from full harmonic analysis
FE models Maximum Minimum Average
Whole 1 0.96 0.98
Partial FE model 3 1 0.85 0.94
Partial FE model 2 1 0.01 0.32
Partial FE model 1 1 0.0001 0.06

5.1.3 Simulation for Experimental Procedure
In the above two numerical examples, the whole FRF matrix is calculated for the
original plate by applying to its FE model either a full or a reduced harmonic analysis.
As mentioned in Chapter 4, usually in practice, the elements of only one row or one
column of the whole FRF matrix array are measured in modal testing. The rest of the
elements can be obtained by using FRF synthesizing procedure after identifying the

Structural Modification on a 2D Structure (Plate)

154
modal parameters, such as the natural frequencies, damping ratios and mode shapes,
through FRF curve-fitting technology.
Here, the software STAR Modal version 5 is adopted to identify the modal parameters
of the original plate by taking one column of elements under force excitation at node
385, within the frequency range of 0 to 800Hz and a frequency resolution of 0.25Hz. As
shown in
Table 5.4, the calculated natural frequencies from ANSYS and the identified natural
frequencies from STAR Modal version 5 are virtually identical. Table 4.11 shows the
identified mode shapes of the original beam within the frequency range from 0 to 800
Hz from STAR Modal version 5.

Table 5.4 Comparison of calculated natural frequencies and identified natural
frequencies
Mode
Calculated natural
frequency (Hz)
Identified natural
frequency (Hz)
1 178.27 178.27
2 323.07 323.07
3 399.86 399.86
4 532.23 532.23
5 556.44 556.44
6 738.88 738.88
7 753.46 753.45

Table 5.5 The identified mode shapes of the original beam within frequency range
0 to 800 Hz from STAR Modal version 5
Modes Mode shapes Modes Mode shapes
1

2


Structural Modification on a 2D Structure (Plate)

155
3

4

5

6

7



As the FE model of the original plate has 441 nodes, with nodes 1 to 80 clamped, and
taking the reciprocity characteristic of the FRF matrix into consideration, the synthesis
of the whole FRF matrix still needs 361*(361 1) / 2 65341 + = operations in STAR
Modal version 5 which is time-consuming and inefficient. Hence, a FRF synthesizing
code has been written in MATLAB using the following equation [164, 165]

*
*
1
2 ( ) 2 ( )
r r
N
ij ij
ij
r
r r
A A
H
j s p j s p
=
(
= +
(

(

(5.1)
where
ij
H is the FRF between DOF i and DOF j;
r
ij
A is the rth complex residue term between DOF i and DOF j;
r r r
p j = + is the pole location in the s-place for the rth mode;
s j = , is the Laplace transfer parameter;

Structural Modification on a 2D Structure (Plate)

156
r
is the rth modal damping, in this case the damping is taken from the
measured case and proportional damping is assumed;
r
is the rth natural frequency;
superscript

denotes the complex conjugate; and


N is the number of modes included in the FRF synthesizing procedure.
Within the frequency range of 0 to 800Hz, there is a total of 7 modes identified for the
original plate. In order to reduce the residual effects of the higher order modes, another
two modal identifications were carried out within the frequency ranges of 0 to 1600Hz
and 0 to 3200Hz. There are 17 modes in the former frequency range and 38 in the latter.
The synthesized FRFs, with different numbers of identified modes, are compared with
the calculated FRFs from the full harmonic analysis in Figure 5.21 and Figure 5.22.
These figures clearly demonstrate the effects of the truncation error due to unknown
higher-order modes. It can also be seen that, below 400Hz, the synthesized FRFs in the
three cases are nearly identical to the calculated FRFs. This is because the truncation
error due to the unknown higher order modes has little effect on the lower order modes.
As the frequency increases, the discrepancy between the synthesized FRFs and the
calculated FRF also increases. The fewer the number of modes included in the
synthesizing procedure, the larger the discrepancy.
The FRAC value distribution plots obtained by comparing the synthesized FRFs with
the calculated FRFs are shown in Figure 5.23, Figure 5.24 and Figure 5.25. It should be
noted that, as only one column of elements from the FRF matrix is available, the FRAC
values are calculated from iFZjUZ, where i=385, j=81 to 441. Thus the FRAC value
plots are bar plots with the response nodes along the x axis.

Structural Modification on a 2D Structure (Plate)

157

Figure 5.21 Comparison of synthesized FRFs and calculated FRFs from full
harmonic analysis of original plate at 385FZ271UZ

Figure 5.22 Comparison of synthesized FRFs and calculated FRFs from full
harmonic analysis original plate at 385FZ385UZ

Structural Modification on a 2D Structure (Plate)

158

Figure 5.23 Distribution of FRAC values of synthesized FRFs with 7 modes
compared with calculated FRFs of original plate

Figure 5.24 Distribution of FRAC values of synthesized FRFs with 17 modes
compared with calculated FRFs of original plate

Structural Modification on a 2D Structure (Plate)

159

Figure 5.25 Distribution of FRAC values of synthesized FRFs with 38 modes
compared with calculated FRFs of original plate

Table 5.6 shows the maximum, minimum and average FRAC values of the synthesized
FRFs compared with the calculated FRFs from the full harmonic analysis of the original
plate.

Table 5.6 Maximum, minimum and average FRAC values of synthesized FRFs
compared with calculated FRFs from full harmonic analysis of original plate
FE models Maximum Minimum Average
Synthesizing with 7 modes 0.999996 0.996965 0.999825
Synthesizing with 17 modes 1 0.9988 1
Synthesizing with 38 modes 1 0.9990 1

Figure 5.23, Figure 5.24, Figure 5.25 and Table 5.6 show that the synthesized FRFs are
in good agreement with the calculated FRFs from the full harmonic analysis of the

Structural Modification on a 2D Structure (Plate)

160
original plate, even with only 7 modes. By coupling the synthesized FRF matrix
obtained with different numbers of identified modes and the Guyan reduced delta
dynamic stiffness matrix from the whole FE model, the FRFs of the modified plate are
predicted and shown in Figure 5.26 and Figure 5.27.

Figure 5.26 Comparison of predicted FRFs from structural modification and
calculated FRFs from full harmonic analysis of modified plate at 385FZ271UZ

Structural Modification on a 2D Structure (Plate)

161

Figure 5.27 Comparison of predicted FRFs from structural modification and
calculated FRFs from full harmonic analysis of modified plate at 385FZ385UZ

In Figure 5.26 and Figure 5.27, it is really surprising to observe that the greater the
number of modes included in the synthesis procedure for the original plate, the worse
are the predicted FRFs from the structural modification. This type of disagreement is
due to the effects of modal truncation errors in the FRF synthesizing procedure for the
original plate. These errors occur due to the limited number of modes identified and
taken into account in the FRF synthesis. They consist of the loss of contributions from
the out-of-band higher modes contributing to dynamic behaviour in the frequency range
of interest [167]. Therefore, in this section, a compensation method [167] will be
introduced.

Structural Modification on a 2D Structure (Plate)

162
5.1.3.1 Synthesis with compensation
For a system with n DOFs, the FRFs can be expressed using Equation (5.1). If N
denotes the number of identified modes, then Equation (5.1) can be re-written as

* *
* *
1 1
2 ( ) 2 ( ) 2 ( ) 2 ( )
r r r r
N n
ij ij ij ij
ij
r r N
r r r r
A A A A
H
j s p j s p j s p j s p
= = +
( (
= + + +
( (

( (


(5.2)
The first term on the right side in Equation(5.2), which is the same as the right hand side
of Equation (5.1), denotes the effects of the identified modes on the FRF
ij
H . The
second term represents the residual effects of the unknown modes of the structure under
investigation on the FRF
ij
H . If, in the second term, r is larger than N, then

*
( )
*
*
( )
2 ( ) 2 ( )
2 ( ( )) 2 ( ( ))
r r
ij ij r
r r
r r
ij ij
r r r r
A A
f s
j s p j s p
A A
j j j j j j
= +

= +
+
(5.3)
Taking only the first term on the right side of Equation (5.3) into consideration,

( )
( )
2 ( ( ))
r
ij r
L
r r
A
f
j j j


=
+
(5.4)
where
r r r
ij r i j
A c = (5.5)
where
r
ij
A is the rth modal constant;

r
i
is the displacement of the DOF i in the rth mode shape; and

r
j
is the displacement of the DOF j in the rth mode shape.
By substituting Equation (5.5) into Equation (5.4) and expanding it using the Taylor
series, we obtain

Structural Modification on a 2D Structure (Plate)

163

( )
2
' ''
' ''
2
0 0
( )
2( )
( 0)
(0) (0)( 0) (0) ...
2!
...
2( ) 2( ) 2( ) 2
r r
r i j r
L
r r
r r r r r r
r i j r i j r i j
r r r r r r
c
f
j
f f f
c c c
j j j


= =
=
+

= + + +
| | | |
= + + +
| |
| |
+ + +
\ . \ .

(5.6)
with
'
2
1
2 () 2 ( )
r r r r
r i j r i j
r r r r
c c
j j


| |
=
|
|
+ +
\ .
(5.7)

'' '
2 3
1 2
2 () 2 ( ) 2 ( )
r r r r r r
r i j r i j r i j
r r r r r r
c c c
j j j


| | | |
= =
| |
| |
+ + +
\ . \ .
(5.8)
If only the first and second order terms from the Taylor Series are taken into
consideration, Equation (5.6) can be approximated as

' ''
2
( )
0 0
2
2 3
( )
2( ) 2( ) 2( ) 2
2( ) 2( ) 2( )
r r r r r r
r i j r i j r i j r
L
r r r r r r
r r r r r r
r i j r i j r i j
r r r r r r
c c c
f
j j j
c c c
j j j








= =
| | | |
+ +
| |
| |
+ + +
\ . \ .
+ +
+ + +

(5.9)
Similarly, we can get

* * * *
( )
* * * * * *
2
2 3
( )
2 ( ( )) 2( )
2( ) 2( ) 2( )
r r r r
r i j r i j r
R
r r r r
r r r r r r
r i j r i j r i j
r r r r r r
c c
f
j j j j
c c c
j j j





= =
+ +
+ +
+ + +
(5.10)
The compensation term then becomes
*
( ) ( )
*
1 1
2
2 3
1
* * * * * *
2
( ) ( )
2 ( ) 2 ( )
2( ) 2( ) 2( )
2( ) 2( )
r r
n n
ij ij r r
L R
r N r N r r
r r r r r r
n
r i j r i j r i j
r N r r r r r r
r r r r r r
r i j r i j r i j
r r r r
A A
f f
j s p j s p
c c c
j j j
c c c
j j





= + = +
= +
(
( + = +
(


(

(
+ +
(
+ + +
(

+ + +
+ +

2
3
1
2( )
n
r N r r
j


= +
(
(
+
(

(5.11)
which can be simplified as

Structural Modification on a 2D Structure (Plate)

164

*
2
*
1
2 ( ) 2 ( )
r r
n
ij ij
ij ij ij
r N r r
A A
R S D
j s p j s p

= +
(
+ + +
(

(

(5.12)
where
* *
1
2( ) 2( )
r r r r
n
r i j r i j
ij
r N r r r r
c c
R
j j


= +
(
= +
(
+ +
(

is the static compensation coefficient;



* *
2 2
1
2( ) 2( )
r r r r
n
r i j r i j
ij
r N r r r r
c c
S
j j


= +
(
= +
(
+ +
(

is the first-order dynamic compensation


coefficient; and

* *
3 3
1
2( ) 2( )
r r r r
n
r i j r i j
ij
r N r r r r
c c
D
j j


= +
(
= +
(
+ +
(

is the second-order dynamic


compensation coefficient.
The static compensation coefficient,
ij
R , and the dynamic compensation coefficients,
ij
S
and
ij
D , are determined by calculating the differences between the measured FRFs and
the synthesized FRFs at three distinct frequencies,
low
,
middle
and
high
, the choice of
which will influence the results. As a rule of thumb [167, 168], one can select
low

below the first flexible mode in the frequency range of interest,
middle
at about 50% of
the frequency range of interest, and
high
at about 75% of the frequency range of
interest while none of their locations should be close to any peaks or troughs. If it is
possible to calculate more than three frequencies, the compensation coefficients could
be determined by means of the least-square fit. In this study,
low
was set at 50 Hz,
middle
was set at 390 Hz (to avoid the natural frequency at about 400 Hz) and
high
was
set at 600 Hz.
This compensation method is intended to approximate the contributions of all the modes
beyond the frequency range to the FRF matrix in the whole frequency range of interest.
Figure 5.28, Figure 5.29 and Figure 5.30 show the comparison of the synthesized FRFs

Structural Modification on a 2D Structure (Plate)

165
using 7 modes, 17 modes and 38 modes for synthesis respectively, with and without
compensation, and the calculated FRFs from the full harmonic analysis of the original
plate. Figure 5.31, to Figure 5.36 show the distributions of the FRAC values of the
synthesized FRFs, with both static and second-order dynamic compensation, compared
with the calculated FRFs at 385FZ385UZ. Nodes 385 and node 361 were the locations
where the accelerometers were placed in the experiment. These two locations were also
selected because their responses were reasonable for all the mode shapes within the
frequency range of interest.

Figure 5.28 Comparison of synthesized FRFs with 7 modes and calculated FRFs
from full harmonic analysis of the original plate at 385FZ385UZ

Structural Modification on a 2D Structure (Plate)

166

Figure 5.29 Comparison of synthesized FRFs with 17 modes and calculated FRFs
from full harmonic analysis of original plate at 385FZ385UZ

Figure 5.30 Comparison of synthesized FRFs with 38 modes and calculated FRFs
from full harmonic analysis of the original plate at 385FZ385UZ

Structural Modification on a 2D Structure (Plate)

167

Figure 5.31 Distribution of FRAC values of 7 modes synthesized FRFs with static
compensation compared with calculated FRFs of original plate

Figure 5.32 Distribution of FRAC values of 7 modes synthesized FRFs with
second-order compensation compared with calculated FRFs of the original plate

Structural Modification on a 2D Structure (Plate)

168

Figure 5.33 Distribution of FRAC values of 17 modes synthesized FRFs with static
compensation compared with calculated FRFs of the original plate

Figure 5.34 Distribution of FRAC values of 17 modes synthesized FRFs with
second-order compensation compared with calculated FRFs of the original plate

Structural Modification on a 2D Structure (Plate)

169

Figure 5.35 Distribution of FRAC values of 38 modes synthesized FRFs with static
compensation compared with calculated FRFs of the original plate

Figure 5.36 Distribution of FRAC values of 38 modes synthesized FRFs with
second-order compensation compared with calculated FRFs of the original plate


Structural Modification on a 2D Structure (Plate)

170
From Figure 5.28 to Figure 5.36, the FRFs synthesized with 17 modes with static and
second-order dynamic compensations appear to be accurate enough and will be used to
carry out structural modification without incurring substantial computational resources
as would be required for synthesis with 38 modes. By following the same procedure as
for the reduced model, the FRFs of the modified plate are predicted by coupling the
synthesized FRFs with the static and second-order dynamic compensations and the
Guyan reduced delta dynamic stiffness matrices. Figure 5.37 and Figure 5.38 show
comparisons of the predicted FRFs from structural modification of the whole FE model
and the calculated FRFs from the full harmonic analysis of the modified plate.

Figure 5.37 Comparison of predicted FRFs from structural modification, by
coupling 17 modes synthesized FRFs and Guyan reduced delta dynamic stiffness
matrix from whole FE model, and calculated FRFs from full harmonic analysis of
the modified plate at 385FZ271UZ

Structural Modification on a 2D Structure (Plate)

171

Figure 5.38 Comparison of predicted FRFs from structural modification, by
coupling 17 modes synthesized FRFs and Guyan reduced delta dynamic stiffness
matrix from whole FE model, and calculated FRFs from full harmonic analysis of
the modified plate at 385FZ385UZ

It can be seen from Figure 5.37 and Figure 5.38 that the agreement between the
predicted FRFs and the calculated FRFs has increased dramatically by structural
modification through the coupling of the synthesized FRFs with compensation and the
Guyan reduced delta dynamic stiffness matrix from the whole FE model, even with only
the static compensation.
Figure 5.39, to Figure 5.44 show comparisons of the predicted FRFs from structural
modification by coupling the 17 modes-synthesized FRFs and the Guyan reduced delta
dynamic stiffness matrix from partial FE models 1, 2 and 3 with the calculated FRFs of
the modified plate respectively.

Structural Modification on a 2D Structure (Plate)

172

Figure 5.39 Comparison of predicted FRFs from structural modification, by
coupling 17 modes-synthesized FRFs and Guyan reduced delta dynamic stiffness
matrix from partial FE model 1, and calculated FRFs from full harmonic analysis
of the modified plate at 385FZ271UZ

Figure 5.40 Comparison of predicted FRFs from structural modification, by
coupling 17 modes-synthesized FRFs and Guyan reduced delta dynamic stiffness

Structural Modification on a 2D Structure (Plate)

173
matrix from partial FE model 1, and calculated FRFs from full harmonic analysis
of the modified plate at 385FZ385UZ

Figure 5.41 Comparison of predicted FRFs from structural modification, by
coupling 17 modes-synthesized FRFs and Guyan reduced delta dynamic stiffness
matrix from partial FE model 2, and calculated FRFs from full harmonic analysis
of the modified plate at 385FZ271UZ


Structural Modification on a 2D Structure (Plate)

174

Figure 5.42 Comparison of predicted FRFs from structural modification, by
coupling 17 modes-synthesized FRFs and Guyan reduced delta dynamic stiffness
matrix from partial FE model 2, and calculated FRFs from full harmonic analysis
of the modified plate at 385FZ385UZ

Figure 5.43 Comparison of predicted FRFs from structural modification, by
coupling 17 modes-synthesized FRFs and Guyan reduced delta dynamic stiffness

Structural Modification on a 2D Structure (Plate)

175
matrix from partial FE model 3, and calculated FRFs from full harmonic analysis
of the modified plate at 385FZ271UZ

Figure 5.44 Comparison of predicted FRFs from structural modification, by
coupling 17 modes-synthesized FRFs and Guyan reduced delta dynamic stiffness
matrix from partial FE model 3, and calculated FRFs from full harmonic analysis
of the modified plate at 385FZ385UZ

In Figure 5.39 and Figure 5.40, regardless of the compensation method used, the
predicted results are not acceptable due to the limited size of partial FE model 1. In
Figure 5.41 and Figure 5.42, the introduced compensation method improves the
predicted results from partial FE model 2 to the same extent as that in section 5.1.2. In
Figure 5.43 and Figure 5.44, as expected, the predicted FRFs from structural
modification for partial FE model 3 are dramatically improved by the compensation
method.
Figure 5.45, to Figure 5.52 show the FRAC values between the predicted FRFs from the
whole FE model, the partial FE models 1, 2 and 3 and the calculated FRFs, with a 4 Hz
frequency band.

Structural Modification on a 2D Structure (Plate)

176
Table 5.7 shows the maximum, minimum and average FRAC values of the predicted
FRFs compared with the calculated FRFs from the full harmonic analysis of the
modified plate, with the same frequency bandwidth.

Figure 5.45 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and static compensation from whole FE model
compared with calculated FRFs of the modified plate with 4 Hz frequency
bandwidth.
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

177

Figure 5.46 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and second-order dynamic compensation from whole
FE model compared with calculated FRFs of the modified plate with 4 Hz
frequency bandwidth.

Figure 5.47 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and static compensation from partial model 1
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

178
compared with calculated FRFs of the modified plate with 4 Hz frequency
bandwidth.


Figure 5.48 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and second-order dynamic compensation from partial
model 1 compared with calculated FRFs of the modified plate with 4 Hz frequency
bandwidth.
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

179

Figure 5.49 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and static compensation from partial FE model 2
compared with calculated FRFs of the modified plate with 4 Hz frequency
bandwidth.

Figure 5.50 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and second-order dynamic compensation from partial
Node number
N
o
d
e

n
u
m
b
e
r

Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

180
FE model 2 compared with calculated FRFs of the modified plate with 4 Hz
frequency bandwidth.


Figure 5.51 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and static compensation from partial FE model 3
compared with calculated FRFs of the modified plate with 4 Hz frequency
bandwidth.
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

181

Figure 5.52 Distribution of FRAC values of predicted FRFs from structural
modification with 17 modes and second-order dynamic compensation from partial
FE model 3 compared with calculated FRFs of the modified plate with 4 Hz
frequency bandwidth.

Table 5.7 Maximum, minimum and average FRAC values of predicted FRFs
compared with calculated FRFs from full harmonic analysis of the modified plate
with 4 Hz frequency bandwidth.
FE models
Original FRF from static
compensation
Original FRF from second-order
dynamic compensation
Maximum Minimum Average Maximum Minimum Average
Whole FE model 1 0.8564 0.9634 1 0.9871 0.9954
Partial FE model 3 1 0.7675 0.9050 1 0.8817 0.9462
Partial FE model 2 1 0.0133 0.3132 1 0.0227 0.3163
Partial FE model 1 0.9998 0.0001 0.0537 0.9999 0.001 0.0580

It can be seen that the agreement between the predicted FRFs from the whole FE model
and partial FE model 3 and the calculated FRFs improves after the employment of the
synthesis procedure with compensation. For the whole FE model, the average FRAC
Node number
N
o
d
e

n
u
m
b
e
r


Structural Modification on a 2D Structure (Plate)

182
values are 0.963 and 0.995 for the static and second-order dynamic compensations
respectively. For partial FE model 3, the average FRAC values are 0.905 and 0.946 for
these compensations respectively.
Although the FRAC values from partial FE models 2 and 1 also improve after the
synthesis procedure with compensation is conducted, they are still unacceptable due
mainly to their limited sizes. The decreasing FRAC values, from the highest for the
whole FE model to the lowest for the smallest partial FE model, also indicates that, for
the partial FE models used for reduced structural modification, there is a minimum size
required in order to maintain accuracy.
5.2 STRUCTURAL MODIFICATION WITH EXPERIMENTAL DATA
In this section, structural modification is conducted with experimental data obtained
from an aluminum plate with all four edges clamped and subjected to modification. The
modification is applied in the form of a rectangular cut-out from the centre of the plate.
5.2.1 Modal Testing
Experimental modal testing is conducted on both the original and modified plates. Each
plate is screw-clamped between two steel frames while the bottom frame is welded to a
four-legged heavy supporting frame structure. This set-up is used to provide solid
support conditions for vibration in the transverse direction of the plate. The valid plate
area for measurement is 490410mm with 21 measurement points equally spaced along
both sides. The measurement points are arranged in a clockwise spiral manner with 1 at
the upper right corner and 441 at the center. The four edges of the valid area are then

Structural Modification on a 2D Structure (Plate)

183
clamped to the frame, as shown in Figure 5.53 and Figure 5.54, and the available total
measurement points are 361, from point 81 to point 441.
Modal testing with an impact hammer was conducted with a B&K type 8202 impact
hammer exciting every node in the transverse direction, Z. The response measurements
in the same direction were taken by B&K type 4374 accelerometers on points 361 and
385. The accelerometers were mounted on the plate using beeswax. The modal test was
conducted from 0 to 800 Hz with a resolution of 0.25 Hz. Force measurements were
obtained from a B&K type 8200 force transducer integrated on the impact hammer.

Figure 5.53 Schematic of experimental set-up for original plate


Structural Modification on a 2D Structure (Plate)

184

Figure 5.54 Photograph of modified plate under modal testing

For the impact hammer test, averages of up to 10 measurements were taken at each
excitation point and the force exponential window was used to improve the signal-to-
noise ratio. Figure 5.55, Figure 5.56 and Figure 5.57 show the measured FRFs.

Figure 5.55 Measured FRF of original plate at the driving point (361FZ361UZ)
from impact hammer test

Structural Modification on a 2D Structure (Plate)

185

Figure 5.56 Measured FRF of original plate at (385FZ361UZ) from impact
hammer test

Figure 5.57 Measured FRF of original plate at driving point (385FZ385UZ) from
impact hammer test

Structural Modification on a 2D Structure (Plate)

186
The measured FRFs of the original plate are imported into STAR Modal version 5 for
identification of the modal parameters. Table 5.8 shows the comparison of the natural
frequencies of the original plate obtained by different methods (0-800Hz).

Table 5.8 Comparison of natural frequencies of the original plate obtained by
different methods (0-800Hz)
Mode Modal test(impact hammer) FEM

Measured at
Point 361
Frequency (Hz)
Measured at
Point 385
Frequency (Hz)
Frequency
(Hz)
Error
compared with
Measurement
at Point 361
(%)
Error
compared
with
Measurement
at Point 385
(%)
1 167.2434 167.2444 178.27 6.59 6.59
2 309.876 309.8847 323.07 4.26 4.25
3 368.553 368.5554 399.86 8.49 8.49
4 500.6339 500.6304 532.23 6.31 6.31
5 538.4232 538.4251 556.44 3.35 3.35
6 680.7468 680.7521 738.88 8.54 8.54
7 719.2799 719.2858 753.46 4.75 4.75

From Table 5.8, the error between the measured and calculated frequencies is too large
to be accepted. In order to determine the effective plate size under modal testing
because of less than perfect clamping boundary condition, an optimization procedure by
obtaining the minimum graphically was conducted by setting the measured frequencies
as the target in the following equation for the 4-sided clamped plate [169].

1 1
4 4 3
2 2
1 2 1 2 1 2 1 2
4 4 2 2 2
2 2 ( )
12 (1 )
ij
G G J J H H J J Eh
f
a b a b


( ( +
= + +
( (


(5.13)
where a is the length of plate, 490 690 a ;
b is the width of plate, 410 610 b ;
h is thickness of plate;

Structural Modification on a 2D Structure (Plate)

187
i is number of half-waves in mode shape along horizontal axis;
j is the number of half-waves in mode shape along vertical axis;
E is Youngs modulus;
is mass per unit area of plate ( *h for plate with material density, );
is Poissons ratio; and
G , H , J are the coefficients in approximate formulas for isotropic and
orthotropic rectangular plate, for clamped-clamped boundary condition, J H = .
After the optimization procedure, the effective size of the original plate was determined
to be 499.4 X 431.8mm, compared with its physical measurement values of 490
410mm. The Youngs modulus is 70.382GPa, as determined by a static bending
experiment on a strip of the plates material. The density is 2671 kg/m
3
, as determined
by the mass over the volume. As shown in Table 5.9, the natural frequencies for the first
7 modes calculated from the updated size and material properties of the plate agree with
those determined by modal testing to within 0.4%.

Table 5.9 Comparison of natural frequencies of original plate with updated size
and material properties (0-800Hz)
Mode Modal test(impact hammer) Prediction by updated FE Model

Measured at
Point 361
Frequency (Hz)
Measured at
Point 385
Frequency (Hz)
Frequency
(Hz)
Error
compared with
Measurement
at Point 361
(%)
Error
compared with
Measurement
at Point 385
(%)
1 167.24 167.24 166.94 -0.179 -0.179
2 309.88 309.88 309.54 -0.110 -0.110
3 368.55 368.56 368.81 0.071 0.068
4 500.63 500.63 499.19 -0.288 -0.288
5 538.42 538.43 538.10 -0.060 -0.061
6 680.75 680.75 679.13 -0.238 -0.238
7 719.28 719.29 716.50 -0.387 -0.388

Structural Modification on a 2D Structure (Plate)

188
5.2.2 Structural Modification
The updated FE model of the original plate is used to calculate the FRF matrix in the
transverse direction, Z. The Guyan reduced matrices are obtained from a whole FE
model and three partial FE models. The predicted FRFs are compared with the FRFs
experimentally measured at different locations (361FZ361UZ, 385FZ361UZ and
385FZ385UZ) on the modified structure in Figure 5.58, Figure 5.59 and Figure 5.60.

Figure 5.58 Comparison of predicted FRFs from different-sized FE models and
experimentally measured FRFs at 361FZ361UZ

Structural Modification on a 2D Structure (Plate)

189

Figure 5.59 Comparison of predicted FRFs from different-sized FE models and
experimentally measured FRFs at 385FZ361UZ

Figure 5.60 Comparison of predicted FRFs from different-sized FE models and
experimentally measured FRFs at 385FZ385UZ

Structural Modification on a 2D Structure (Plate)

190
From Figure 5.58 to Figure 5.60, it can be seen that, for the first 3 modes, the predicted
FRFs of the whole FE model and partial FE model 3 are in reasonably good agreement
with the measured FRFs of the modified plate. The predicted FRFs of partial FE model
1 are not as good due to its limited size. The predicted FRFs of partial FE model 2 are
slightly better than the results from partial FE model 1 but are still unacceptable. It is
noted that, for the higher-order modes, the discrepancy between the predictions and
measurements increases due to Guyan reduction procedure which is an inexact process
and will always produce frequencies different from those of the full model. As the mode
number increases, so does the discrepancy. This is mainly due to the fact that significant
inertia terms of the eliminated DOFs, which are critical to the accuracy of the higher
modes of the system, are discarded [126, 128].
5.3 ADDITIONAL PROCESS FOR DETERMINATION OF THE
MINIMUM AREA
In this section, an attempt was made to determine the minimum area for the partial
model to generate acceptable results from structural modification.
5.3.1 10 Elements per Side for the Original and Modified Plates
In this case, the original and modified plates are divided into 10 shell elements with 6
DOFs per node. For the original plate, the node sequence starts from 1 at the upper right
corner and, in a clockwise spiral manner, ends at 121 in the center. For the modified
plate, the node numbering follows the same sequence as the original plate but is from 1
to 120. The cut-out area is 2 elements per side in the center which is from nodes 113 to
121 with nodes 113 to 120 as the physical interface nodes. The clamped boundary

Structural Modification on a 2D Structure (Plate)

191
condition is set on nodes 1 to 40. By using the reduced harmonic analysis in ANSYS
V10, the FRF matrix of the original plate is obtained within the frequency range of 0.25
Hz to 800 Hz and a 0.25 Hz frequency resolution. One whole, and two partial, FE
models are employed to obtain the delta dynamic stiffness matrices| | B . The
dimensions of partial plates 1 and 2 are 294 mm x 246 mm and 392 mm x 328 mm,
corresponding to 9 times and 16 times of the cut-out and 36% and 64% of the original
plate in terms of area respectively.
Structural modification is conducted by coupling the calculated FRF matrix and delta
dynamic stiffness matrices
| | B from different-sized FE models. The FRAC value is
calculated with the original frequency resolution of the FRFs and the average value is
compared with the ratio of the partial model over the modifying area in Figure 5.61.
FRAC_mag in the figure stands for the FRAC value by only taking the magnitude of the
complex FRF into consideration.
0
0.2
0.4
0.6
0.8
1
1.2
0 5 10 15 20 25 30
the ratio of partial model over modifying part
F
R
A
C

v
a
l
u
e
FRAC_mag
FRAC

Figure 5.61 FRAC value vs the ratio of partial models over the modifying part


Structural Modification on a 2D Structure (Plate)

192
It can be seen that the predicted results from partial FE models 1 and 2 are not
acceptable, while the results from whole FE model is acceptable with its average FRAC
value close to 1. Though the partial model 2 here covers the same area as the partial
model 3 in Section 5.1.2, the predicted result is still not acceptable because of the rough
FE model.
5.3.2 20 Elements per Side for the Original and Modified Plates
In this case, the original and modified plates are divided into 20 shell elements with 6
DOFs per node as in Section 5.1, while the cut-outs are varied from 2 elements per side,
4 elements per side to 6 elements per side in the center, corresponding to 1%, 4% and
9% of the original plate respectively.
5.3.2.1 2 elements per side cut-out
One whole, and four partial, FE models are employed to obtain the delta dynamic
stiffness matrices
| | B .The dimensions of partial plates 1, 2, 3 and 4 are 294 mm x 246
mm, 343mm x 287mm, 392mm x 238mm, and 441 mm x 369 mm, corresponding to 36
times, 49 times, 64 times and 81 times of the cut-out and 36%, 49%, 64%and 81% of
the original plate in terms of area respectively. Structural modification is carried out for
this cut-out.
The FRAC value is calculated with 4 Hz bandwidth and the average value is compared
with the ratio of the partial model over the modifying area in Figure 5.62.
From Figure 5.62, it can be seen that the predicted results from partial FE models and
whole FE model are in great agreement with the calculated FRFs for the modified plate
with the average FRAC values large than 0.7.


Structural Modification on a 2D Structure (Plate)

193
0
0.2
0.4
0.6
0.8
1
1.2
1.00 21.00 41.00 61.00 81.00 101.00 121.00
The ratio of partial model over modifying part
F
R
A
C

v
a
l
u
e
FRAC_mag
FRAC

Figure 5.62 FRAC value vs the ratio of partial models over the modifying part for
the 2 elements per side cut-out

5.3.2.2 4 elements per side cut-out
One whole, and four partial, FE models are employed to obtain the delta dynamic
stiffness matrices
| | B . The dimensions of partial plates 1, 2, 3 and 4 are 294 mm x 246
mm, 343mm x 287mm, 392mm x 238mm, and 441 mm x 369 mm, corresponding to 9
times, 12.25 times, 16 times and 20.25 times of the cut-out and 36%, 49%, 64% and
81% of the original plate in terms of area respectively. Structural modification is carried
out for this cut-out.
The FRAC value is calculated with 4 Hz bandwidth and the average value is compared
with the ratio of the partial model over the modifying area in Figure 5.63.


Structural Modification on a 2D Structure (Plate)

194
0
0.2
0.4
0.6
0.8
1
1.2
1.0 6.0 11.0 16.0 21.0 26.0 31.0
The ratio of partial model over modifying part
F
R
A
C

v
a
l
u
e
FRAC_mag
FRAC

Figure 5.63 FRAC value vs the ratio of partial models over the modifying part for
the 4 elements per side cut-out

From Figure 5.63, it can be seen that the predicted results from partial FE model 3,
partial FE model 4 and whole FE model are in great agreement with the calculated FRFs
for the modified plate with the average FRAC values large than 0.9. The FRAC values
from partial FE model 1 and partial FE model 2 are both lower than 0.5.
5.3.2.3 6 elements per side cut-out
One whole, and three partial, FE models are employed to obtain the delta dynamic
stiffness matrices
| | B
. The dimensions of partial plates 1, 2, 3 and 4 are 294 mm x 246
mm, 343mm x 287mm, 392mm x 238mm, and 441 mm x 369 mm, corresponding to 4
times, 5.4 times, 7.1 times and 9 times of the cut-out and 36%, 49%, 64% and 81% of
the original plate in terms of area respectively. Structural modification is carried out for
this cut-out.

Structural Modification on a 2D Structure (Plate)

195
The FRAC value is calculated with 4 Hz bandwidth and the average value is compared
with the ratio of the partial model over the modifying area in Figure 5.64.

0
0.2
0.4
0.6
0.8
1
1.2
1.0 3.0 5.0 7.0 9.0 11.0 13.0
The ratio of partial model over modifying part
F
R
A
C

v
a
l
u
e
FRAC_mag
FRAC

Figure 5.64 FRAC value vs the ratio of partial models over the modifying part for
the 6 elements per side cut-out

From Figure 5.64, it can be seen that only the predicted results from partial FE model 4
and whole FE model are in great agreement with the calculated FRFs for the modified
plate with the average FRAC values large than 0.9. The FRAC values from partial FE
model 1, partial FE model 2 and partial FE model 3 are all lower than 0.5.
From Figure 5.62 to Figure 5.64, it can be seen that although the areas covered by the
partial FE model 1 to 4 are maintained at the same ratio compared with the area of the
original structure, the predicted results are varied for different size cut-outs. It is clear
that the minimum size of the partial FE model is not related to the percentage of the
original structure but to the modifying area. If we consider a FRAC value at 0.5 is
acceptable, we can determine the minimum size. For the 2 elements per side cut-out, the

Structural Modification on a 2D Structure (Plate)

196
minimum size can not be determined from Figure 5.62. For the 4 elements per side cut-
out, a minimum size at about 14 times that of the modifying area can be revealed from
Figure 5.63. For the 6 elements per side cut-out, a minimum size of about 8 times of the
modifying area can be revealed from Figure 5.64.
5.3.3 40 Elements per Side for the Original and Modified Plates
In this case, the original and modified plates are divided into 40 shell elements with 6
DOFs per node, while the cut-out is 8 elements per side in the center, corresponding to
4%of the original plate. The dimensions of partial plates 1, 2, 3 and 4 are 294 mm x 246
mm, 343mm x 287mm, 392mm x 238mm, and 441 mm x 369 mm, corresponding to 9
times, 12.25 times, 16 times and 20.25 times of the cut-out and 36%, 49%, 64% and
81% of the original plate in terms of area respectively. Structural modification is carried
out for this cut-out.
The FRAC value is calculated with 4 Hz bandwidth and the average value is compared
with the ratio of the partial model over the modifying area in Figure 5.65.
0
0.2
0.4
0.6
0.8
1
1.2
1.0 6.0 11.0 16.0 21.0 26.0 31.0
the ratio of partial model over modifying part
F
R
A
C

v
a
l
u
e
FRAC_mag
FRAC

Figure 5.65 FRAC value vs the ratio of partial models over the modifying part for
the 6 elements per side cut-out


Structural Modification on a 2D Structure (Plate)

197
From Figure 5.65, it can be seen that the predicted results from partial FE model 2,
partial FE model 3, partial FE model 4 and whole FE model are in great agreement with
the calculated FRFs for the modified plate with the average FRAC values large than 0.9.
The FRAC values from partial FE model 1 are still lower than 0.5. Comparing Figure
5.65 with Figure 5.63, it can be seen that although partial FE model 2 in both cases
cover the same area which is 12.25 times that of the modifying area, the agreement
between the predicted results has been improved dramatically by a finer FE mesh. In
this case, a minimum size of 11 times of the modifying area appears to be sufficient.
From the above comparison in Sections 5.3.1, 5.3.2 and 5.3.3, two conclusions can be
drawn: 1) the ratio of partial FE model over the modifying part is dependent on the
finite element mesh size; and 2) with a reasonably fine mesh (element size about 2.5%
of the size of the structure),the predicted FRF from structural modification seems to be
acceptable if the partial FE model is at least 11 times the size of the modifying part.
5.4 CONCLUSIONS
For a plate clamped on all four sides as an example of a 2-D structure, the effects of
distributed structural modifications (in the form of a cut-out) with reduced DOFs on its
dynamic responses have been determined efficiently and successfully. For example, for
the modified structure presented in Section 5.1.1, computation of the whole FRF matrix
within frequency range 0 to 800Hz with a 0.25Hz frequency resolution in ANSYS
would take a HP workstation with a core 2 duo E6300 processor about 36 hours to
finish while by using structural modification, it only takes the same computer less than
3 hours to finish. This is because in the full harmonic analysis in ANSYS, the matrix
inversion is involved whose order is (a+b), while in structural modification with

Structural Modification on a 2D Structure (Plate)

198
reduced DOFs the order of matrix inversion is only b, where a, b holds the same
meaning as in Section 3.1.2. It should be noted that the modification in this case, a cut-
out, is easy to model almost perfectly, as compared to the modifications to the beam,
where an additional component is attached by welding, with consequent uncertainties in
joint stiffness.
In the full model, identical FRF curves for the modified structure can be obtained by
coupling the FRFs of the original structure from the full harmonic analysis and the
matrices [ ] B from differen FE models with different size without reduction.
By using the numerically determined FRFs of the original structure from the reduced
harmonic analysis and the Guyan reduced matrices [ ] B , good agreement of the FRF
curves for the modified structure can be obtained from the whole FE model and partial
FE models 2 (49% of the plate and 12.25 times of the cut-out, in terms of area) and 3
(64% of the plate and 16 times of the cut-out, in terms of area). The predicted FRFs
dont match the calculated FRFs very well for the smallest partial model 1 (39% of the
plate and 9 times of the cut-out, in terms of area). A conclusion can be drawn that a
partial FE model should cover a certain area. From the additional study in section 5.3, it
has been found that the ratio of the partial FE model over the modifying part is
dependent on the finite element mesh size and with a reasonable fine mesh (element size
about 2.5% of the size of the structure), if the partial FE model is at least 11 times the
size of the modifying part, the predicted FRFs from structural modification seems to be
acceptable.
After running the numerical simulation for the experimental procedure, it was
unexpected to see that the greater the number of modes included in the synthesis
procedure, the worse are the predicted FRFs due to the truncation errors of the unknown

Structural Modification on a 2D Structure (Plate)

199
modes. After introducing the compensation procedure, the predicted FRFs from the
structural modification of the modified structure behave in the same manner as they do
in the reduced model.
The measured FRFs of the original structure were used to update the actual size of the
plate in the experiment because of less than perfect clamping boundary condition. By
coupling the FRFs of the updated plate from reduced harmonic analysis and the Guyan
reduced matrices [ ] B of different-sized FE models, good agreement is obtained
between the predicted FRFs from the whole FE model and partial FE model 3 (64% of
the original plate and 16 times of the cut-out) and the measured FRFs of the modified
structure for the first 3 modes below 400 Hz. The agreement is not as good at higher
frequencies for two reasons: (1) the discrepancy between the FE model of the modified
plate and the physical modified plate; and (2) Guyan reduction procedure. The predicted
FRFs from the partial FE models 1 and 2 do not match the measured FRFs very well for
the same reasons, in addition to the limited sizes of the models.
200
Chapter 6 STRUCTURAL MODIFICATION OF A
3D STRUCTURE (FRAME)

In this chapter, Structural Modification is applied to a free-free or a clamp supported 3D
box frame model which is used to validate the Structural Modification Method on 3D
frame structures with additional DOFs. In total, four cases are presented - three
numerical simulations and one experimental validation.
6.1 NUMERICAL SIMULATIONS
6.1.1 Full Model Clamped Supported 3D Box Frame without
Condensation
This numerical example is a 3D box frame made from steel with an aluminum plate
attached to one side to represent structural modification with additional DOFs. The
original 3D box frame model and the modified version are shown in Figure 6.1 and
Figure 6.2 respectively. The nominal material properties of the steel are: Youngs
modulus=210 GPa; Poissons ratio=0.30; and density=7850 kg/m
3
. The material
properties of the aluminum are: Youngs modulus=70 GPa; Poissons ratio=0.33; and
density=2800 kg/m
3
. The dimensions of the original 3D box frame are 300 mm x 200
mm x 100mm, constructed from rectangular steel bars with cross section of 10 mm x 10

Structural Modification on a 3D Structure (Frame)

201
mm for bars along the length, 10 mm x 15mm along the width and 10 mm x 25 mm
along the height. The dimensions of the aluminum plate are 200 mm x 100 mm x 6mm.

Figure 6.1 FE model of original 3D box frame


Figure 6.2 FE model of modified 3D box frame with aluminum plate attached


Structural Modification on a 3D Structure (Frame)

202
In each of these FE models, the 3D box frame is divided into 12, 8 and 4 beam elements
along the length, width and height respectively, with 6 DOFs per node.
Figure 6.3 shows the node numbers of the original 3D box frame. Totally there are 92
nodes, 96 beam elements. The plate is divided into 8*4 shell elements with 6 DOFs per
node. Totally there are 113 nodes, 96 beam elements and 32 shell elements in the
modified structure. The box frame is supported by constraining all DOFs of the nodes
along one edge nodes 1, 29, 30, 31 and 32. By employing the full harmonic analysis,
the FRF matrix of the original 3D box frame is obtained within the frequency range of
0.25 Hz to 800Hz and a 0.25 Hz frequency resolution. Figure 6.4, Figure 6.5 and Figure
6.6 show the calculated FRFs of the original 3D box frame at different locations
(13FZ13UZ, 47FX47UX and 75FY75UY). These three locations were selected because
of their responses by superimposing the mode shapes within the frequency range of
interest.


Structural Modification on a 3D Structure (Frame)

203

Figure 6.3 Node numbers of the original 3D box frame

Figure 6.4 Calculated FRFs of original 3D box frame at 13FZ13UZ

Structural Modification on a 3D Structure (Frame)

204

Figure 6.5 Calculated FRFs of original 3D box frame at 47FX47UX

Figure 6.6 Calculated FRFs of original 3D box frame at 75FY75UY


Structural Modification on a 3D Structure (Frame)

205
One whole and three partial FE models are employed to obtain the matrices | |
B
. The
partial FE models 1, 2 and 3 of the original 3D box frame are shown in Figure 6.7
Figure 6.8 and Figure 6.9 respectively. The partial FE model 1 spans one sixth of the
volume of the original box frame. The partial FE model 2 spans one third of the volume
of the original box frame. The partial FE model 3 spans half of the volume of the
original box frame.

Figure 6.7 Partial FE model 1 of original 3D box frame


Structural Modification on a 3D Structure (Frame)

206

Figure 6.8 Partial FE model 2 of original 3D box frame


Figure 6.9 Partial FE model 3 of original 3D box frame

Structural Modification on a 3D Structure (Frame)

207
Using the equations for structural modification with additional DOFs, as developed in
Chapter 3, the predicted FRFs of the whole FE model are compared with both the
calculated results from the full harmonic analysis of the modified structure and the
FRFs of the original structure in Figure 6.10, Figure 6.11 and Figure 6.12.

Figure 6.10 Comparison of FRFs of original and modified 3D frames at 13FZ13UZ

Structural Modification on a 3D Structure (Frame)

208

Figure 6.11 Comparison of FRFs of original and modified 3D frames at 47FX47UX

Figure 6.12 Comparison of FRFs of original and modified 3D frames at 75FY75UY


Structural Modification on a 3D Structure (Frame)

209
It can be seen from Figure 6.10 to Figure 6.12 that structural modification of the whole
FE model yields virtually identical FRFs to the calculated FRFs from full harmonic
analysis of the modified structure.
In Figure 6.13, Figure 6.14 and Figure 6.15, the predicted FRFs of the whole and partial
FE models are compared with the calculated results from full harmonic analysis of the
modified structure at various locations (13FZ13UZ, 47FX47UX and 75FY75UY).

Figure 6.13 Comparison of predicted FRFs of different-sized FE models at
13FZ13UZ

Structural Modification on a 3D Structure (Frame)

210

Figure 6.14 Comparison of predicted FRFs of different-sized FE models at
47FX47UX

Figure 6.15 Comparison of predicted FRFs of different-sized FE models at
75FY75UY

Structural Modification on a 3D Structure (Frame)

211
It can be seen from Figure 6.10 to Figure 6.15 that the FRFs obtained from structural
modification are virtually identical to those calculated from a full harmonic analysis of
the modified 3D frame irrespective of whether the matrix [ ] B is obtained from the
whole model or a partial FE model.
6.1.2 Reduced Model - Clamped Supported 3D Box Frame with
Condensation

By using the reduced harmonic analysis, the reduced FRF matrix of the original 3D
frame contains information in only the translational directions X, Y and Z, within the
frequency range of 0.25 Hz to 800 Hz and a 0.25 Hz frequency resolution. The same
procedure as was used for the reduced beam model in Chapter 4 is adopted to obtain the
matrices
| |
c
B of the four different-sized FE models, that is, one whole and three partial
FE models. As shown in Figure 6.16, Figure 6.17 and Figure 6.18, the calculated FRFs
from the reduced harmonic analysis and those from the full harmonic analysis of the
original 3D frame at different locations (13FZ13UZ, 47FX47UX and 75FY75UY) are
in reasonably good agreement in the frequency range below 400 Hz. However, there are
discrepancies in the frequency range higher than 400 Hz for the magnitudes and phases,
though the peaks and troughs align with each other.

Structural Modification on a 3D Structure (Frame)

212

Figure 6.16 Comparison of calculated FRFs of original 3D frame from reduced
harmonic analysis and from full harmonic analysis at 13FZ13UZ

Figure 6.17 Comparison of calculated FRFs of original 3D frame from reduced
harmonic analysis and from full harmonic analysis at 47FX47UX

Structural Modification on a 3D Structure (Frame)

213

Figure 6.18 Comparison of calculated FRFs of original 3D frame from reduced
harmonic analysis and from full harmonic analysis at 75FY75UY

Figure 6.19, Figure 6.20 and Figure 6.21 show the predicted FRFs obtained by coupling
the reduced FRF matrix of the original 3D frame and the Guyan reduced delta dynamic
stiffness matrix of the whole FE model compared with the FRFs from the reduced
harmonic analysis of the modified 3D frame.

Structural Modification on a 3D Structure (Frame)

214

Figure 6.19 Comparison of FRFs of original and modified 3D frames obtained by
different methods at 13FZ13UZ

Figure 6.20 Comparison of FRFs of original and modified 3D frames obtained by
different methods at 47FX47UX

Structural Modification on a 3D Structure (Frame)

215

Figure 6.21 Comparison of FRFs of original and modified 3D frames obtained by
different methods at 75FY75UY

From Figure 6.19 to Figure 6.21, it can be seen that the predicted FRFs obtained from
structural modification by coupling the reduced FRF matrix and the Guyan reduced
delta dynamic stiffness matrix of the whole FE model are virtually identical to the
calculated FRFs obtained from the reduced harmonic analysis of the modified 3D frame.
Figure 6.22, Figure 6.23 and Figure 6.24 show comparisons of the predicted FRFs of
different-sized FE models and the calculated FRFs of the modified 3D frame.

Structural Modification on a 3D Structure (Frame)

216

Figure 6.22 Comparison of predicted FRFs of different-sized FE models at
47FX47UX obtained by coupling reduced FRF matrix and Guyan reduced delta
dynamic stiffness matrix

Figure 6.23 Comparison of predicted FRFs of different-sized FE models at
75FY75UY obtained by coupling reduced FRF matrix and Guyan reduced delta
dynamic stiffness matrix

Structural Modification on a 3D Structure (Frame)

217

Figure 6.24 Comparison of predicted FRFs of different-sized FE models at
13FZ13UZ obtained by coupling reduced FRF matrix and Guyan reduced delta
dynamic stiffness matrix

From Figure 6.22 to Figure 6.24, it can be seen that the predicted results from structural
modification using the reduced harmonic analysis results and the Guyan reduced delta
dynamic stiffness matrices of the whole FE model, partial FE model 3 and partial FE
model 2 are virtually identical and are in good agreement with those calculated by the
reduced harmonic analysis of the modified 3D frame.
However, the predicted FRFs of partial FE model 1 dont match the calculated results
very well. This indicates that there is a minimum size required to maintain accuracy
when a partial FE model is used for reduced structural modification. In this case, it
appears that the minimum size for a partial model would be close to that for partial FE
model 2 which is one third of the volume of the original 3D box frame.

Structural Modification on a 3D Structure (Frame)

218
These results differ from those obtained from the beam simulation for 1-D structure in
Chapter 4 and Ref [23] in which the predicted FRFs of all the partial FE models are in
great agreement with the calculated results for the modified beam. Nonetheless, they
match the finding in Chapter 5 that a minimum area is required to achieve acceptable
results from structural modification using a partial FE model for a 2-D structure.
Figure 6.25, to Figure 6.28 show the distributions of the FRAC values between the
FRFs predicted by the whole, and the three partial, FE models and the FRFs calculated
from the reduced harmonic analysis of the modified 3D frame respectively, with a 4Hz
frequency bandwidth. The numbers on the x-axis and y-axis of these figures correspond
to the FRAC values in the x,y and z degrees of freedom for each node according to
Table 6.1.

Table 6.1 Corresponding number sequence to the node DOFS
Node 2 Node 3 Node 27 Node 28
x y z x y z x y z x y z x y z x y z
1 2 3 4 5 6 76 77 78 79 80 81
Node 33 Node 34 Node 112 Node 113
x y z x y z x y z x y z x y z x y z
82 83 84 85 86 87 319 320 321 322 323 324
Note: All DOFs of the nodes 1, 29, 30, 31 and 32 are constrained.

Structural Modification on a 3D Structure (Frame)

219

Figure 6.25 FRAC values between predicted FRFs of whole FE model and FRFs
from reduced harmonic analysis of modified 3D frame

Figure 6.26 FRAC values between predicted FRFs of partial FE model 1 and FRFs
from reduced harmonic analysis of modified 3D frame
DOF number
D
O
F

n
u
m
b
e
r

1
1
DOF number
D
O
F

n
u
m
b
e
r

1
1

Structural Modification on a 3D Structure (Frame)

220

Figure 6.27 FRAC values between predicted FRFs of partial FE model 2 and FRFs
from reduced harmonic analysis of modified 3D frame

Figure 6.28 FRAC values between predicted FRFs of partial FE model 3 and FRFs
from reduced harmonic analysis of modified 3D frame
DOF number
D
O
F

n
u
m
b
e
r

1
1
DOF number
D
O
F

n
u
m
b
e
r

1
1

Structural Modification on a 3D Structure (Frame)

221
Table 6.2 shows the maximum, minimum and average FRAC values of the predicted
FRFs of different-sized FE models compared with the calculated FRFs of the modified
3D frame from the reduced harmonic analysis, with a 4Hz frequency bandwidth.

Table 6.2 Maximum, minimum and average FRAC values of different-sized FE
models compared with FRFs from reduced harmonic analysis
FE models Maximum Minimum Average
Whole 1 0.7602 0.9067
Partial FE model 3 1 7.5764e-6 0.6781
Partial FE model 2 1 1.2193e-4 0.6464
Partial FE model 1 1 7.2696e-7 0.1744

From Figure 6.25 to Figure 6.28 and Table 6.2, it can be seen that the agreement
between the predicted FRFs from the whole FE model, partial FE model 3 and 2 and the
calculated FRFs are good, with average FRAC values of 091, 0.68 and 0.65 respectively;
however, the average FRAC value between the predicted FRFs from partial FE model 1
and the calculated FRFs is only 0.17.
The variation of FRAC values, from the highest for the whole FE model to the lowest
for the smallest partial FE model, also indicate that there is a minimum size required to
maintain the accuracy of the partial FE model used for reduced structural modification.
It appears that the minimum size could be close to partial FE model 2, corresponding to
one third of the volume of the original and modified frame.
6.1.3 Simulation for Experimental Procedure
As in the previous two chapters, a simulation for the experimental procedure is
presented. Only one column of the whole FRF matrix array is taken into consideration,
that is, the elements under the force excitation at node 26 of the whole FRF matrix,

Structural Modification on a 3D Structure (Frame)

222
within the frequency range of 0 to 800Hz with a frequency resolution of 0.25Hz. The
software STAR Modal version 5 is employed to identify the modal parameters of the
original 3D frame.
The rest of the elements of the whole FRF matrix are obtained by using FRF-synthesis
procedure (in section 4.2.3) after the identification of the modal parameters, such as
natural frequencies, damping ratios and mode shapes, using global polynomial FRF
curve-fitting technique,. Table 6.3 shows the comparison of the calculated natural
frequencies from ANSYS and the identified natural frequencies from STAR Modal
version 5 within the frequency range of 0 to 800Hz.
Table 6.3 illustrates that the identified natural frequencies from STAR Modal version 5
are in excellent agreement with the calculated natural frequencies from ANSYS for the
original 3D frame.

Table 6.3 Comparison of calculated and identified natural frequencies within 0
800 Hz
Mode Calculated natural frequency (Hz) Identified natural frequency (Hz)
1 53.53 53.53
2 61.98 61.98
3 103.42 103.42
4 152.04 152.04
5 193.07 193.06
6 476.80 476.81
7 483.13 483.15
8 495.25 495.25
9 504.15 504.15
10 545.38 545.38
11 561.66 561.65
12 582.16 582.16
13 616.01 616.02
14 648.43 648.44
15 751.71 751.72

Structural Modification on a 3D Structure (Frame)

223
Next, the FRF-synthesis MATLAB code (see section 4.2.3) is used here to obtain the
missing elements of the full FRF matrix of the original 3D frame. Within the frequency
range of 0 to 800 Hz, there are a total of 15 modes for the original 3D frame. In order to
reduce the residual effects of the unknown higher order modes, another modal
identification is carried out within the frequency range of 0 to 1600 Hz. There are 25
modes in this frequency range. Comparisons of the synthesized FRFs, with different
numbers of identified modes, and the calculated FRFs from the reduced harmonic
analysis at three different locations (13FZ13UZ, 47FX47UX and 75FY75UY) are
shown in Figure 6.29, Figure 6.30 and Figure 6.31.

Figure 6.29 Comparison of synthesized FRFs and calculated FRFs from reduced
harmonic analysis of original 3D frame at 13FZ13UZ

Structural Modification on a 3D Structure (Frame)

224

Figure 6.30 Comparison of synthesized FRFs and calculated FRFs from reduced
harmonic analysis of original 3D frame at 47FX47UX

Figure 6.31 Comparison of synthesized FRFs and calculated FRFs from reduced
harmonic analysis of original 3D frame at 75FY75UY

Structural Modification on a 3D Structure (Frame)

225
Figure 6.30 and Figure 6.31 clearly demonstrate the effects of the truncation error due to
the unknown higher-order modes. But only a minor truncation error phenomenon, above
the 400Hz frequency, can be observed in Figure 6.29.
By coupling the synthesized FRF matrix obtained using different numbers of identified
modes and the Guyan reduced delta dynamic stiffness matrix of the whole FE model,
the FRFs of the modified 3D frame are predicted and shown in Figure 6.32, Figure 6.33
and Figure 6.34 for 3 different locations ((13FZ13UZ, 47FX47UX and 75FY75UY)

Figure 6.32 Comparison of predicted FRFs from structural modification with
whole FE model and calculated FRFs from reduced harmonic analysis of modified
3D frame at 13FZ13UZ

Structural Modification on a 3D Structure (Frame)

226

Figure 6.33 Comparison of predicted FRFs from structural modification with
whole FE model and calculated FRFs from reduced harmonic analysis of modified
3D frame at 47FX47UX

Figure 6.34 Comparison of predicted FRFs from structural modification with
whole FE model and calculated FRFs from reduced harmonic analysis of modified
3D frame at 75FY75UY

Structural Modification on a 3D Structure (Frame)

227
From Figure 6.32 to Figure 6.34, it can be seen that the predicted FRFs from structural
modification using the synthesized original FRFs with 25 modes are in good agreement
with the FRFs from the reduced harmonic analysis of the modified 3D frame. The
predicted FRFs from structural modification using the synthesized original FRFs with
15 modes are not as good compared with the FRFs from the reduced harmonic analysis
of the modified 3D frame. This is due to model truncation errors in the synthesizing
procedure.
Figure 6.35, Figure 6.36 and Figure 6.37 show comparisons of the predicted FRFs from
structural modification, obtained by coupling the synthesized FRFs with 25 modes and
the Guyan reduced delta dynamic stiffness matrices of the whole and partial FE models,
and the calculated FRFs from the reduced harmonic analysis of the modified 3D frame
for the same 3 locations (13FZ13UZ, 47FX47UX and 75FY75UY).

Figure 6.35 Comparison of predicted FRFs from structural modification using
synthesized original FRFs with 25 modes and calculated FRFs of modified 3D
frame at 13FZ13UZ

Structural Modification on a 3D Structure (Frame)

228

Figure 6.36 Comparison of predicted FRFs from structural modification using
synthesized original FRFs with 25 modes and calculated FRFs of modified 3D
frame at 47FX47UX

Figure 6.37 Comparison of predicted FRFs from structural modification using
synthesized original FRFs with 25 modes and calculated FRFs of modified 3D
frame at 75FY75UY

Structural Modification on a 3D Structure (Frame)

229
It can be seen from Figure 6.35 to Figure 6.37 that the predicted FRFs from structural
modification obtained by coupling the synthesized original FRFs with 25 modes and the
Guyan reduced delta dynamic stiffness matrices of the whole FE model, partial FE
model 2 and partial model 3 are in good agreement with the calculated FRFs from the
reduced harmonic analysis for the modified 3D box frame. The predicted FRFs of
partial FE model 1 are not as good which is similar to results as discussed in section
6.1.2.
Figure 6.38, to Figure 6.41 show the distributions of the FRAC values between the
predicted FRFs and the calculated FRFs of the whole FE model, partial FE model 1,
partial FE model 2 and partial FE model 3 respectively, with a 4Hz frequency
bandwidth. Table 6.4 shows the maximum, minimum and average FRAC values, with a
4Hz frequency band.

Figure 6.38 FRAC values between predicted FRFs of whole FE model and
calculated FRFs from reduced harmonic analysis of modified 3D frame
DOF number
D
O
F

n
u
m
b
e
r

1
1

Structural Modification on a 3D Structure (Frame)

230

Figure 6.39 FRAC values between predicted FRFs of partial FE model 1 and
calculated FRFs from reduced harmonic analysis of modified 3D frame

Figure 6.40 FRAC values between predicted FRFs of partial FE model 2 and
calculated FRFs from reduced harmonic analysis of modified 3D frame
DOF number
D
O
F

n
u
m
b
e
r

1
1
DOF number
D
O
F

n
u
m
b
e
r

1
1

Structural Modification on a 3D Structure (Frame)

231

Figure 6.41 FRAC values between predicted FRFs of partial FE model 3 and
calculated FRFs from reduced harmonic analysis of modified 3D frame

Table 6.4 Maximum, minimum and average FRAC values of different-sized FE
models compared with FRFs from reduced harmonic analysis
FE models Maximum Minimum Average
Whole 0.997 0.009 0.6908
Partial FE model 3 0.9895 3.3048e-5 0.5290
Partial FE model 2 0.9850 1.2762e-4 0.4982
Partial FE model 1 0.9602 3.0472e-6 0.3670

It can be seen that the agreement between the predicted FRFs of the whole FE model
and partial FE model 3 and the calculated FRFs are acceptable with the average FRAC
value of over 0.5. Although the average FRAC value for partial FE models 2 is only a
little bit lower than 0.5, it is still considered as unacceptable. The average FRAC value
for partial FE models 1 is only 0.367 which is unacceptable. This is due mainly to the
small size used in these two partial FE models. The reduction of FRAC values, from
DOF number
D
O
F

n
u
m
b
e
r

1
1

Structural Modification on a 3D Structure (Frame)

232
those for the whole FE model to those for the partial FE model, also indicates that there
is a minimum size required to maintain accuracy for the partial FE model used for
reduced structural modification. In the previou numerical example, Section 6.1.2, the
minimum size was revealed at one third of the volume of the original structure. But in
this case study, it reveals that one third of the volume of the original structure can only
produce close to acceptable results. So the minimum size should be slightly big than one
third of the volume of the original structure. The low minimum value is due to the
longitudinal vibration mode is taken into consideration; for example, for partial FE
model 1, the minimum FRAC value happens at 86FZ37UZ which is involved in the
longitudinal vibration mode.
6.2 STRUCTURAL MODIFICATION WITH EXPERIMENTAL DATA
In this section, structural modification is conducted with experimental data obtained
from a 3D frame with free-free boundary conditions and subjected to modification in
the form of a plate being attached to one side of the 3D frame. The geometry of the 3D
frame is different from that of the 3D frame used in the above numerical examples due
to the difficulty in modeling the actual 3D frame with the neutral planes of the beams
not joint at the location with beam elements in FE model. It is constructed using steel
bars with a crosssection of 10 mm X 10 mm. The dimensions of the original 3D frame
are illustrated in Figure 6.42. Figure 6.43 and Figure 6.44 show the experiment set-up
and the photo of experiment.

Structural Modification on a 3D Structure (Frame)

233

Figure 6.42 Dimensions of 3D frame under testing (All dimensions in mm)

6.2.1 Modal Testing
Experimental modal testing was conducted on both the original and modified 3D frames.
The 3D frame was hung at its four corners by cotton thread. This set-up is used to
simulate free-free boundary conditions for vibration, as illustrated in Figure 6.43 and
Figure 6.44. The measurement points are arranged as shown in Figure 6.45. Modal
testing, using a B&K type 8202 impact hammer with a force transducer B&K type 8200
integrated to excite every node in the transverse directions, was conducted.
Response measurements in the transverse directions Y and Z were taken using B&K
type 4374 accelerometers on node 38. The accelerometers were mounted on the 3D
frame by beeswax. The modal test was performed in the range of 0 to 800Hz and a
resolution of 0.25 Hz.

Structural Modification on a 3D Structure (Frame)

234

Figure 6.43 Schematic of experimental set-up for original 3D frame

Structural Modification on a 3D Structure (Frame)

235

(a) Original 3D frame under impact hammer model testing

Structural Modification on a 3D Structure (Frame)

236

(b) Modified 3D frame

Figure 6.44 Photo of modal testing of 3D frame


Figure 6.45 Illustration of measurement points

Structural Modification on a 3D Structure (Frame)

237
For the impact hammer test, averages of up to 10 measurements were taken at each
excitation point and a force exponential window was used to improve the signal-to-
noise ratio Figure 6.46, Figure 6.47 and Figure 6.48 show the measured FRFs at
38FY38UY 38FZ38UZ, and 38FZ49UY. These three locations were selected because
their responses were reasonable for all the modes within the frequency range of interest.

Figure 6.46 Measured FRFs of original 3D box frame at driving point (38FY38UY)
from impact hammer test

Structural Modification on a 3D Structure (Frame)

238

Figure 6.47 Measured FRFs of original 3D box frame at driving point (38FZ38UZ)
from impact hammer test

Figure 6.48 Measured FRFs of original 3D box frame at random point
(38FZ49UY) from impact hammer test

Structural Modification on a 3D Structure (Frame)

239
The measured FRFs of the original 3D frame are imported into STAR Modal version 5
for identifying the modal parameters. Table 6.5 shows the comparison of the natural
frequencies of the original 3D box frame obtained by experimental modal testing and
FE Analysis (0-800Hz). First, textbook values for steel were used as the material
properties in the FEM calculation, that is, Youngs modulus=210 GPa, density=7850
kg/m
3
and Poissons ratio =0.3 in the FE analysis.

Table 6.5 Comparison of natural frequencies of original 3D frame obtained by
different methods (0-800Hz)
Mode Modal test(impact hammer) FEM

Measured
at point
38Y
Measured
at point
38Z
Average
Before model
updating
After model
updating
Frequency
(Hz)
Frequency
(Hz)
Frequency
(Hz)
Frequency
(Hz)
Error
(%)
Frequency
(Hz)
Error
(%)
1 186.60 186.59 186.60 187.73 0.61 186.76 0.09
2 / 220.21 220.21 219.27 -0.43 218.14 -0.94
3 / 363.51 363.51 366.98 0.95 365.09 0.43
4 401.77 / 401.77 402.00 0.06 399.93 -0.46
5 408.61 408.8 408.7 407.94 -0.19 405.83 -0.70
6 / 444.34 444.34 448.19 0.87 445.88 0.35
7 449.46 / 449.46 452.73 0.73 450.40 0.21
8 470.66 470.51 470.59 475.10 0.96 472.65 0.44
9 550.46 550.46 550.46 551.41 0.17 548.56 -0.35
10 / 605.88 605.88 607.47 0.26 604.33 -0.26
11 648.33 648.13 648.23 650.17 0.30 646.81 -0.22
12 653.31 / 653.31 656.24 0.45 652.86 -0.07
13 731.75 731.75 731.75 732.08 0.05 728.31 -0.47
* / indicates difficulty in identifying the mode

From the above table, we can see that, due to the discrepancies, such as the difference of
the material properties between the physical frame and the FE model and the joints in
the physical frame not being modelled in the FE model, the calculated natural

Structural Modification on a 3D Structure (Frame)

240
frequencies do not match the measured frequencies well. In order to determine the right
material parameters, the FE model of the original 3D frame is updated using FEMTools
[162, 163]. After model updating, the material parameters are Youngs modulus=207.97
GPa, density=7855 kg/m
3
and Poissons ratio=0.3.
6.2.2 Structural Modification
As described in section 4.2.4.2, the updated FE model of the original 3D frame is used
to calculate the FRF matrix in the translational directions. The Guyan reduced matrices
are obtained by a whole FE model and three partial FE models. Comparisons of the
predicted FRFs and the FRFs measured at three different locations (38FY38UY,
38FZ38UZ and 38FZ49UY) on the modified structure are shown in Figure 6.49, Figure
6.50 and Figure 6.51.

Figure 6.49 Comparison of predicted FRFs of different-sized FE models with
experimental measured FRFs at 38FY38UY

Structural Modification on a 3D Structure (Frame)

241

Figure 6.50 Comparison of predicted FRFs of different-sized FE models with
experimental measured FRFs at 38FZ38UZ

Figure 6.51 Comparison of predicted FRFs of different-sized FE models and
measured FRFs at 38FZ49UY

Structural Modification on a 3D Structure (Frame)

242
It can be observed that the predicted FRFs of the whole and partial FE models are in
good agreement with the measured FRFs of the modified structure below approximately
500Hz. Discrepancies between the predictions and measurements increase for the
higher-order modes. This is due to the discrepancies between the FE models and the
physical frames, such as the welding joints in the physical frame not being modelled in
the FE model, plate mounted on the side of the physical frame not being modelled
exactly due to the limitation of beam and plate elements in ANSYS (see Figure 6.2 and
Figure 6.44 (b)), and Guyan reduction used to obtain the delta dynamic stiffness
matrices.
6.3 CONCLUSIONS
The effects of distributed structural modification on dynamic responses have been
determined efficiently and successfully for a 3D box frame structure with additional
DOFs. For example, for the modified structure presented in Section 6.1.1, computation
of the whole FRF matrix within the frequency range 0 to 800Hz with a 0.25Hz
frequency resolution in ANSYS would take a HP workstation with a Core 2 Duo E6300
processor about 26 hours to finish while by using structural modification, it only takes
the same computer less than 2 hours to finish. This is because in the full harmonic
analysis in ANSYS, the matrix inversion is involved whose order is (a+b+c), while in
structural modification the order of matrix inversion is only b and c, where a, b and c
holds the same meaning as in Section 3.1.3.
By coupling the numerically determined FRFs of the original structure from full
harmonic analysis and the matrices [ ] B from different-sized FE models, without
condensation, identical FRF curves for the modified structure can be obtained

Structural Modification on a 3D Structure (Frame)

243
irrespective of whether the delta dynamic stiffness matrix [ ] B is obtained from either
a whole or a partial FE model.
By coupling the numerically determined FRFs of the original structure from reduced
harmonic analysis and Guyan reduced matrices [ ] B , reasonably good agreement of the
FRF curves for the modified structure can be obtained from the whole FE model and
partial FE models 3 and 2 corresponding to half and one third of the volume of the
original frame respectively. However, the predicted FRFs of the partial FE model 1
which is one sixth of the volume of the original frame do not match the calculated
results very well. This indicates that there is a minimum size required to maintain
accuracy for the partial FE model used for reduced structural modification. In this case,
it appears that the minimum size could be close to partial FE model 2 which is one third
of the volume of the original frame.
These results are different from those for beam simulation in Chapter 4 and Ref [23] in
which the predicted FRFs of all the partial FE models are in good agreement with the
calculated results for the modified beam. But they are broadly in line with the finding in
Chapter 5 that a minimum area is required to achieve acceptable results from structural
modification using a partial FE model.
By running a numerical simulation for the experimental procedure, excellent agreement
of the FRF curves for the modified structure were obtained when the delta dynamic
stiffness matrix [ ] B is obtained from the whole FE model and the partial FE model 3.
Although the predicted FRFs of partial FE model 2 are in good agreement with the
calculated FRFs at several points, the FRAC value calculations show that its predictions
are not quite acceptable. The predicted FRFs of partial FE model 1 do not match the
calculated results very well which is also confirmed by the FRAC value calculations.

Structural Modification on a 3D Structure (Frame)

244
FEM update technology was used to update the original structure from the measured
FRFs. By coupling the FRFs of the updated original structure from reduced harmonic
analysis and the Guyan reduced matrices
[ ] B
from different-sized FE models, good
agreement is obtained between the predicted and the measured FRFs of the modified
structure at frequencies below approximately 500 Hz. Agreement is not as good at
higher frequencies due to the discrepancies between the FE models of the modified 3D
frame, such as the welding joints in the physical frame not being modelled in the FE
model, plate mounted on the side of the physical frame not being modelled exactly and
Guyan reduction.
245
Chapter 7 CONCLUSIONS AND
RECOMMENDATIONS

7.1 GENERAL
In this thesis, the methodology used for efficiently assessing the effects of distributed
structural modification on the dynamic response of structures has been derived and
validated for three different cases: with no change in DOFs; with reduced DOFs; and
with additional DOFs.
The main results achieved and conclusions drawn from this research are summarised
below.
7.2 CONCLUSIONS
7.2.1 Structural Modification with No Change in DOFs on 1-D
Structure (Beam)
The effects of distributed structural modification on the dynamic responses of a beam as
an example of a 1-D structure with no change in DOFs have been determined
successfully.

Conclusions and Recommendations

246
In the full model, by coupling the numerically determined FRFs of the original structure
from full harmonic analysis and matrices,
[ ] B
obtained from different-sized FE
models without condensation, virtually identical FRF curves for the modified structure
can be obtained.
In the reduced model, by coupling the FRFs of the original structure from reduced
harmonic analysis and Guyan-reduced matrices,
[ ] B
, good agreement of the FRF
curves for the modified structure can be obtained from whole FE model, partial FE
model 3 (corresponding to 90% of the original beam) and partial FE model 2
(corresponding to 80% of the original beam). The predicted results from partial FE
model 1 (corresponding to 70% of the original beam) did not agree with the calculated
FRFs well. A further study by doubling the elements in this partial model reveals that
the main reason for the poor prediction is due to the limited number of elements in the
partial model.
7.2.2 Structural Modification with Additional DOFs on 1D Structure
(Beam)
The effects of distributed structural modification with additional DOFs on the dynamic
response of a beam as a 1-D structure have been determined efficiently and successfully.
Also the structural dynamic modification method with additional DOFs for coupling the
experimentally measured FRFs and the numerical models has been validated.
From the full model case, it can be concluded that structural modification can produce
virtually identical FRF curves for the modified structure by coupling the FRFs of the
original structure from full harmonic analysis and matrices
[ ] B
, from different-sized
FE models without condensation

Conclusions and Recommendations

247
By coupling the FRFs of the original structure from reduced harmonic analysis and
Guyan-reduced matrices,
[ ] B
, good agreement of the FRF curves for the modified
structure can be obtained irrespective of whether the delta dynamic stiffness matrix,
[ ] B
, is obtained from a whole or a partial FE model.
Good agreement of the magnitude of the FRF curves for the modified structure is
obtained from numerical simulations for the experimental procedure, regardless of
whether the delta dynamic stiffness matrix,
[ ] B
, is obtained from a whole or a partial
FE model. The phase of the FRF curves from structural modification does not match the
calculated FRFs for the modified beam well because the phase discrepancies in the
synthesized FRFs are inherited and magnified by the matrix manipulation involved in
the structural modification.
By using the experimentally determined FRFs of the original structure and the Guyan
reduced matrices,
[ ] B
, of different-sized FE models, good agreement is obtained
between the predicted and measured FRFs of the modified structure at frequencies of up
to 300Hz in this beam case. The agreement is not ideal at higher frequencies for three
reasons: the discrepancy between the FE models of the beams and the physical beams
such as welding joints between the small beam and the modifying beam not being
modelled in the FE simulation; Guyan reduction; and the mass loading effects
introduced by the shaker test. The phase of the FRF curves from structural modification
does not match the measured FRFs for the modified beam because the phase
discrepancies between the synthesized FRFs and the measured FRFs for the original
structure are inherited and magnified by the matrix manipulation involved in the
structural modification.

Conclusions and Recommendations

248
Due to the non-zero elements in the delta dynamic stiffness matrix not associated with
the physical interface DOFs after Guyan reduction, the definition of the interface DOFs
has been extended from physical to numerical. In other words, after Guyan reduction,
all the DOFs in the partial FE model except the passenger DOFs should be treated as the
interface DOFs.
7.2.3 Structural Modification with Reduced 2D Structure (Plate)
A plate clamped on all four sides is chosen as an example of a 2-D structure, and the
effects of distributed structural modifications (in the form of a cut-out) with reduced
DOFs on its dynamic responses have been determined efficiently and successfully
Identical FRF curves for the modified structure can be obtained in the full model by
coupling the numerically determined FRFs of the original structure from the full
harmonic analysis and matrices
[ ] B
from different-sized FE models without
condensation.
By coupling the FRFs of the original structure from the reduced harmonic analysis and
Guyan Reduction, good agreement of the FRF curves for the modified structure can be
obtained from the whole FE model and partial FE models 2 (49% of the plate and 12.25
times the size of the cut-out) and 3 (64% of the plate and 16 times the size of the cut-
out). The predicted FRFs do not match the calculated FRFs very well for the smallest
partial model 1 (39% of the plate and 9 times the size of the cut-out). A conclusion can
be drawn that a partial FE model should cover a minimum area. The additional study in
section 5.3 discloses that the ratio of the partial FE model over the modifying part is
dependent on the finite element mesh size and with a reasonable fine mesh (element size
about 2.5% of the size of the structure), if the partial FE model is at least 11 times the

Conclusions and Recommendations

249
size of the modifying part, the predicted FRFs from structural modification is likely to
be acceptable.
The numerical simulation of the experimental procedure displayed an unusual result in
that the greater the number of modes included in the synthesis procedure, the worse are
the predicted FRFs were. This is attributed to the truncation errors of the unknown
modes. After introducing the compensation procedure, the predicted FRFs from the
structural modification of the modified structure are similar to those of the reduced
model.
The measured FRFs of the original structure were used to update the actual size of the
plate in the experiment because of less than perfect clamping boundary condition. By
coupling the FRFs of the updated plate from reduced harmonic analysis and the Guyan
reduced matrices
[ ] B
from different-sized FE models, good agreement is obtained
between the predicted FRFs from the whole FE model and partial FE model 3 (64% of
the original plate and 16 times the size of the cut-out) and the measured FRFs of the
modified structure for the first 3 modes below 400 Hz. The agreement is not as good at
higher frequencies for two reasons: (1) the discrepancy between the FE model of the
modified plate and the actual modified physical plate; and (2) Guyan reduction. The
predicted FRFs from the partial FE models 1 and 2 do not match the measured FRFs
very well for the same reasons, in addition to the limited sizes of the models.
7.2.4 Structural Modification with Additional DOFs on 3D Structure
(Frame)
The effects of distributed structural modification on dynamic responses have been
determined efficiently and successfully for a 3D box frame structure with additional
DOFs. Also, structural modification has been validated for coupling two types of

Conclusions and Recommendations

250
structures, beam and plate. This illustrated the usefulness of the structural modification
approach for real life or complex problems.
In the full model, structural modification yields identical FRF curves for the modified
structure by coupling the FRFs of the original structure from full harmonic analysis and
the matrices
[ ] B
from different-sized FE models, corresponding to one sixth, one third
and half of the volume of the original frame respectively, without condensation.
By coupling the FRFs of the original structure from reduced harmonic analysis and
Guyan reduced matrices
[ ] B
, reasonably good agreement of the FRF curves for the
modified structure can be obtained from the whole FE model and partial FE models 3
and 2 corresponding to half and one third of the volume of the original frame
respectively. However, the predicted FRFs of the partial FE model 1 corresponding to
one sixth of the volume of the original frame do not match the calculated results very
well. This indicates that there is a minimum size required to maintain accuracy for the
partial FE model used for reduced structural modification. In this case, it appears that
the minimum size could be close to that of partial FE model 2 which is one third of the
volume of the original frame.
The numerical simulation for the experimental procedure discloses that excellent
agreement of the FRF curves for the modified structure were obtained when the delta
dynamic stiffness matrix
[ ] B
is obtained from the whole FE model and the partial FE
model 3. Although the predicted FRFs of partial FE model 2 are in good agreement with
the calculated FRFs at several points, the FRAC value calculations show that its
predictions are not quite acceptable. The predicted FRFs of partial FE model 1 do not
match the calculated results very well which is also confirmed by the FRAC value
calculations mainly due to its limited size.

Conclusions and Recommendations

251
FEM update technology was used to update the original structure from the measured
FRFs. By coupling the FRFs of the updated original structure from reduced harmonic
analysis and the Guyan reduced matrices
[ ] B
from different-sized FE models, good
agreement is obtained between the predicted and the measured FRFs of the modified
structure at frequencies below approximately 500 Hz. Agreement is not as good at
higher frequencies due to the discrepancies between the FE models of the numerical
models of the 3D frame and the physical 3D frames, such as the welding joints in the
physical frame not being modelled in the FE model, plate mounted on the side of the
physical frame not being modelled exactly, and Guyan reduction.
7.3 RECOMMENDATIONS FOR FUTURE WORK
Further numerical and experimental studies are necessary to extend the outcomes from
this study. These include:
Application of dynamic condensation in the structural modification to
overcome the shortcoming of Guyan reduction
Improve numerical model to model more exactly the physical model
More numerical studies to determine the minimum size for the partial FE
model precisely
In this study, the average FRAC value greater than or lower than 0.5
between the predicted FRFs and the calculated FRFs are used to determine
whether the prediction from structural modification is acceptable. More
numerical studies would be required to determine whether a FRAC value of
0.5 is acceptable for predictions in practice.
252
PUBLICATIONS ARISING FROM THIS THESIS
WORK
Journals
1. Hang, H., Shankar, K. and Lai, J.C.S., Prediction of the effects on dynamic response
due to distributed structural modification with additional degrees of freedom.
Mechanical Systems and Signal Processing, 2008. 22(8): p. 1809-1825
2. Hang, H., Shankar, K. and Lai, J.C.S., Effects of Distributed Structural Dynamic
Modification with Reduced Degrees of Freedom. Mechanical Systems and Signal
Processing, 2009, 23(7): p. 2154-2177
3. Hang, H., Shankar, K. and Lai, J.C.S., Effects of Distributed Structural Dynamic
Modification with Additional Degrees of Freedom on 3D Structure. Submitted to
Mechanical Systems and Signal Processing, 2009.

Conference proceedings
1. Huajiang Hang & Krishna Shankar & Joseph Lai, Prediction of Dynamic Response
Due To Distributed Structural Modifications, 19th Australian Conference on the
Mechanics of Structures and Materials, Christchurch, New Zealand 2006
2. Huajiang Hang & Krishna Shankar & Joseph Lai, Distributed Structural Dynamics
Modifications without Rotational Degrees of Freedom, 25th International Modal
Analysis Conference, Orlando, Florida USA 2007

253
REFERENCE

1. Sestieri, A. and W. D'Ambrogio, A Modification Method for Vibration Control
of Structures. Mechanical Systems and Signal Processing, 1989. 3(3): p. 229-
253.
2. D'Ambrogio, W., Some Remarks about Structural Modifications Involving
Additional Degrees of Freedom. Mechanical Systems and Signal Processing,
1990. 4(1): p. 95-99.
3. Tahtali, M., Vibration Analysis of Damped Structures and Structural Reanalysis
Using a New Structural Modification Method, in Mechanical Engineering 1992,
Middle East Technical University: Ankara, Turkey p. 78.
4. zgven , H.N., Structural Modifications Using Frequency Response Functions.
Mechanical Systems and Signal Processing, 1990. 4(1): p. 53-63.
5. D'Ambrogio, W. and A. Sestieri, Coupling Theoretical Data and Translational
FRFs to Perform Distributed Structural Modification. Mechanical Systems and
Signal Processing, 2001. 15(1): p. 157-172.
6. Braun, S. and Y.M. Ram. On structural modification in truncated systems. in
Proceedings of the 5th International Modal Analysis Conference. 1987. London,
U. K.
7. Elliott, K.B. and L.D. Mitchell. The Effect of Modal Truncation on Modal
Modification. in Proceedings of the 5th International Modal Analysis
Conference. 1987. London, U. K.

Reference

254
8. Skingle, G.W., Structural Dynamic Modification Using Experimental Data, in
Imperial College of Science, Technology and Medicine. 1989, University of
London: London, UK. p. 259.
9. Kyprianou, A., J.E. Mottershead, and H. Ouyang, Structural modification. Part
2: assignment of natural frequencies and antiresonances by an added beam.
Journal of Sound and Vibration, 2005. 284(1-2): p. 267-281.
10. Arora, J.S., Survey of Structural Reanalysis Techniques. Journal of Structural
Division ASCE, 1976. 102(Apr): p. 783-802.
11. Wang, B., A. Palazzolo, and W. Pilkey, Re-Analysis, Modal Synthesis and
Dynamic Design. Computational Mechanics, ASME, 1981.
12. Snyder, V.W., Structural Modification and Modal Analysis -- A Survey.
International Journal of Analytical and Experimental Modal Analysis, 1986. 1: p.
45-52.
13. Avitabile, P., Twenty Years of Structural Dynamic Modification - A Review.
Sound and Vibration, 2003. 37(1): p. 14-27.
14. Weissenburger, J.T., Effect of Local Modifications on the Vibration
Characteristics of Linear Systems. Journal of Applied Mechanics, Transactions
of the ASME, 1968. 35(June): p. 327-332.
15. Pomazal, R.J. and V.W. Snyder, Local Modifications of Damped Linear Systems.
AIAA Journal, 1971. 9(11): p. 2216-2221.
16. Hallquist, J. and V. Snyder, Synthesis of Two Discrete Vibratory Systems Using
Eigenvalue Modification. AIAA Journal, 1973. 11(2): p. 247-249.

Reference

255
17. Hirai, I., T. Yoshimura, and K. Takamura, Short Communications on a Direct
Eigenvalue Analysis for Locally Method Structures. International Journal for
Numerical Methods in Engineering, 1973. 6(3): p. 441-456.
18. Wang, B.P., G. Clark, and F.H. Chu. Structural Dynamic Modification Using
Modal Analysis Data. in Proceedings of the 3rd International Modal Analysis
Conference. 1985. Orlando, Florida, USA.
19. Wallack, P., P. Skoog, and M. Richardson. Simultaneous Structural Dynamics
Modification in Proceedings of the 6th International Modal Analysis Conference.
1988. Orlando, FL. U. S. A.
20. D'Ambrogio, W. Consistent modelling of continuous structural dynamic
modifications. in Proceedings of the 9th International Modal Analysis
Conference. 1991. Florence,Italy.
21. Hang, H., K. Shankar, and J.C.S. Lai, Prediction of Dynamic Response Due To
Distributed Structural Modifications, in ACMSM19. 2006, Taylor & Francis:
Christchurch, New Zealand.
22. Hang, H., K. Shankar, and J.C.S. Lai, Distributed Structural Dynamics
Modifications without Rotational Degrees of Freedom, in Proceedings of the
25th International Modal Analysis Conference. 2007: Orlando, FL., U. S. A.
23. Hang, H., K. Shankar, and J. Lai, Prediction of the effects on dynamic response
due to distributed structural modification with additional degrees of freedom.
Mechanical Systems and Signal Processing, 2008. 22(8): p. 1809-1825.
24. Hang, H., K. Shankar, and J. Lai, Effects of Distributed Structural Dynamic
Modification with Reduced Degrees of Freedom. Mechanical Systems and
Signal Processing, 2009. 23(7): p. 2154-2177.

Reference

256
25. Venkayya, V.B., Structural Optimization: A Review and Some
Recommendations. International Journal for Numerical Methods in Engineering,
1978. 13: p. 203-228.
26. Done, G.T.S. and M.A.Y. Rangacharyulu, Use of Optimization in Helicopter
Vibration Control by Structural Modification. Journal of Sound and Vibration,
1981. 74(4): p. 507-518.
27. Wang, B.P., et al., Structural Modification to Achieve Antiresonance in
Helicopters. Journal of Aircraft, 1982. 19(6): p. 499-504.
28. Wang, B.P., F.H. Chu, and C. Trundle. Reanalysis Technique Used to Improve
Local Uncertainties in Modal Analysis. in Proceedings of the 3rd International
Modal Analysis Conference. 1985. Orlando, Florida, U.S.A.
29. Kuang, J.-H. and L.-S. Chen. A Sensitivity Analysis on the Eigen Solutions of
Linearly Damped System. in Proceedings of the 6th International Modal
Analysis Conference. 1988. Orlando, FL. U. S. A.
30. Zhang, Q., et al. Prediction of Mass Modification for Desired Natural
Frequencies. in Proceedings of the 6th International Modal Analysis Conference.
1988. Orlando, Florida, USA.
31. Lisowski, W., W. Bochniak, and T. Uhl. Experimental Modal Analysis and
Structural Modification of the SW3 Helicopter. in Proceedings of ISMA21:
International Conference on Noise and Vibration Engineering. 1996. Leuven,
Belgium.
32. Li, T. and J. He, Local Structural Modification Using Mass and Stiffness
Changes. Engineering Structures, 1999. 21: p. 1028-1037.

Reference

257
33. Park, Y.H. and Y.S. Park, Structure Optimization to Enhance Its Natural
Frequencies Based on Measured Frequency Response Functions. Journal of
Sound and Vibration, 2000. 229(5): p. 1235-1255.
34. Park, Y.H. and Y.S. Park, Structural modification based on measured frequency
response functions: an exact eigenproperties reallocation. Journal of Sound and
Vibration, 2000. 237(3): p. 411-426.
35. Mottershead, J.E., C. Mares, and M.I. Friswell, An inverse method for the
assignment of vibration nodes. Mechanical Systems and Signal Processing, 2001.
15(1): p. 87-100.
36. Kyprianou, A., J.E. Mottershead, and H. Ouyang, Assignment of natural
frequencies by an added mass and one or more springs. Mechanical Systems
and Signal Processing, 2004. 18(2): p. 263-289.
37. Buchberger, B. Groebner Bases: A Short Introduction for Systems Theorists. in
Proceeding of EUROCAST 2001. Canary Islands, Spain.
38. Mottershead, J.E., A. Kyprianou, and H. Ouyang, Structural modification. Part
1: rotational receptances. Journal of Sound and Vibration, 2005. 284(1-2): p.
249-265.
39. Olsson, P. and P. Lidstrm, Inverse structural modification using constraints.
Journal of Sound and Vibration, 2007. 303(3-5): p. 767-779.
40. Stanimirovic, P. and M. Stankovic, Determinants of Rectangular Matrices and
Moore-Penrose Inverse. Novi Sad Journal of Mathematics, 1997. 27(1): p. 53-69.
41. Fleury, C., First and Second Order Convex Approximation Strategies in
Structural Optimization. Structural Optimization, 1989. 1(1): p. 3-10.

Reference

258
42. Svanberg, K., The Method of Moving Asymptotes - A New Method for Structural
Optimization. International Journal for Numerical Methods in Engineering, 1987.
24: p. 359-373.
43. Schittkowski, K., NLPQL: A FORTRAN subroutine solving constrained
nonlinear programming problems. Annals of Operations Research, 1986. 5(2): p.
485-500.
44. Zhang, W. and C. Fleury, A Modification of Convex Approximation Methods for
Structural Optimization. Computers and Structures, 1997. 64(1-4): p. 89-95.
45. Fox, R.L. and M.P. Kapoor, Rates of Change of Eigenvalues and Eigenvectors.
AIAA Journal, 1968. 6: p. 2426-2429.
46. Sharp, R.S. and P.C. Brooks, Sensitivities of Frequency Response Functions of
Linear Dynamic Systems to Variations in Design Parameter Values. Journal of
Sound and Vibration, 1988. 126: p. 167-172.
47. Chen, T.Y., Structural modification with frequency response constraints for
undamped MDOF systems. Computer and Structures, 1990. 36(6): p. 1013-1018.
48. To, W.M. and D.J. Ewins, Structural Modification Analysis Using Rayleigh
Quotient Iteration. International Journal of Mechanical Sciences, 1990. 32(3): p.
169-189.
49. Johnson, C. and V. Synder. Simultaneous vs. Sequential Modifications. in
Proceedings of the 7th International Modal Analysis Conference. 1989. Las
Vegas, Nevada, U.S.A.
50. Avitabile, P. Twenty Years of Structural Dynamic Modification - A Review. in
Proceedings of the 20th International Modal Analysis Conference. 2002. Los
Angeles, California, U. S. A.

Reference

259
51. Simpson, A. and B. Tabarrok, On Kron's Eigenvalue Procedure and Related
Methods of Frequency Analysis. Journal of Mechanics and Applied Mathematics,
1968. 21(1): p. 1-39.
52. Eormenti, D. and S. welaratna, Structural Dynamics Modification -- An
Extension to Modal Analysis. SAE paper, 1981. No.81-1043.
53. Vilmann, C.R. and V.W. Snyder, Line Modification of Continuous Vibratory
Systems. ASCE Journal of the Stuctural Division, 1978. 104(12): p. 1819-1826.
54. Wang, B.P. The limitation of local structural dynamic modification. in
Proceedings of the 3rd International Modal Analysis Conference. 1985. Orlando,
FL. U. S. A.
55. Avitabile, P., J.C. O'Callahan, and J. Milani. Comparison of Complex and
Proportional Mode Structural Dynamic Modification Techniques. in
Proceedings of the 6th International Modal Analysis Conference. 1988. Orlando,
FL. U. S. A.
56. Liu, J.L., Exact solution of nonlinear hysteretic responses using complex mode
superposition method and its application to base-isolated structures. Journal of
engineering mechanics, 2005. 131(3): p. 282-289.
57. Smith, M.J. and S.G. Hutton, A perturbation method for inverse frequency
modification of discrete undamped systems. Journal of Applied Mechanics
(ASME), 1994. 61: p. 887-892.
58. Ravi, S.S.A., T.K. Kundra, and B.C. Nakra, Single step eigen perturbation
method for structural dynamic modification. Mechanics Research
Communications, 1995. 22(4): p. 363-369.

Reference

260
59. Chen, S.-H. and Y.-L. Liu. Substructure analysis of complex structure with weak
connections using matrix perturbation. in Proceedings of the 3rd International
Modal Analysis Conference. 1985. Orlando, FL. U. S. A.
60. Ravi, S.S.A., T.K. Kundra, and B.C. Nakra, Eigenvalue Reanalysis of Sandwich
Beams With Viscoelastic Core Using Perturbation Method. International Journal
of Analytical and Experimental Modal Analysis, 1994. 9(3): p. 203-217.
61. Benfratello, S. and G. Muscolino, A perturbation approach for the response of
dynamically modified structural systems. Computers and Structures, 1998. 68: p.
101-112.
62. Ewins, D.J., Modal Testing: Theory and Practice, ed. J.B. Roberts. 1986:
Research Studies Press LTD. for Brel & Kjr.
63. Deel, J.C. and Y.W. Luk. Modal Testing Considerations for Structural
Modification Applications. in Proceedings of the 3rd International Modal
Analysis Conference. 1985. Orlando, Florida, U. S. A.
64. Sestieri, A. and W. D'Ambrogio. Why Be Modal: i.e. How to Avoid the Use of
Modes in the Modification of Vibrating Systems. in Proceedings of the 6th
International Modal Analysis Conference. 1988. Orlando, FL. U. S. A.
65. Hurty, W.C., Dynamic Analysis of Structural Systems Using Component Modes.
AIAA Journal, 1965. 3(4): p. 678-685.
66. Hintz, R.M., Analytical Methods in Component Modal Synthesis. AIAA Journal,
1975. 13(8): p. 1007-1016.
67. Craig, R.R., ed. A review of time-domain and frequency-domain component
modes synthesis methods. Combined Experimental/Analytical Modeling of

Reference

261
Dynamic Structural Systems Using Substructure Synthesis, ed. D.R. Martinez
and A.K. Miller. 1985.
68. Hwang, W.S. and D.H. Lee, Substructure analysis of complex systems using
rigid body information of components. Proceedings of the Institution of
Mechanical Engineers, Part D: Journal of Automobile Engineering, 2002.
216(10): p. 811-817.
69. Ram, Y.M. and J.E. Mottershead, Receptance method in active vibration control.
AIAA Journal, 2007. 45(3): p. 562 -567.
70. Herbert, M.R. and D.W. Kientzy, Applications of Structural Dynamics
Modification. SAE paper No. 80-1125, 1980.
71. Eastep, F., N. Khot, and R. Grandhi, Improving the active vibrational control of
large space structures through structural modifications. Acta Astronautica,
1987. 15(6/7): p. 383-389.
72. Thomson, W.T., Theory of Vibration with Applications. 2nd ed. 1981: Prentice-
Hall International Inc.
73. Meirovitch, L., Elements of Vibration Analysis. 2nd ed. 1986: McGraw-Hill
Book Company.
74. Gorman, D.J., Vibration Analysis of Plates by the Superposition Method. Series
on Stability, Vibration and Control of Systems, ed. A. Guran and D.J. Inman.
1999, Singapore, New Jersey, London, Hong Kong: Word Scientific.
75. Dossing, O., Structural Testing, Part I: Mechanical Mobility Measurements.
1988: Brel & Kjr.
76. Dossing, O., Structural Testing, Part II: Modal Analysis and Simulation. 1988:
Brel & Kjr.

Reference

262
77. He, J. and Z. Fu, Modal Analysis. 2001, Oxford, England: Butterworth-
Heinemann.
78. Heylen, W., S. Lammens, and P. Sas, Modal Analysis Theory and Testing. 1997:
Katholieke Universiteit leuven, Belgium.
79. Maia, N.M.M. and J.M.M. Silva, Theoretical and Experimental Modal Analysis.
1998, Baldock, England: Research Studies Press Ltd.
80. Stein, P.K. Experimental error? No! Experimenter's error! A measurement
engineer's view of experimental modal analysis. in Proceedings of the 3rd
International Modal Analysis Conference. 1985. Orlando, FL. U. S. A.
81. Brinkman, B.A. and D.J. Macioce. Understanding modal parameter terminology
and mode shape scaling. in Proceedings of the 3rd International Modal Analysis
Conference. 1985. Orlando, FL. U. S. A.
82. Avitabile, P., J.C. OCallahan, and J. Milani. Model correlation and
orthogonality criteria. in Proceedings of the 6th International Modal Analysis
Conference. 1988. Orlando, FL. U. S. A. .
83. Hutton, D.V., Fundamentals of Finite Element Analysis. 2004: The
McGrawHill Companies.
84. Moaveni, S., Finite Element Analysis Theory and Application with ANSYS. 1999,
Upper Saddle River, New Jersey: Prentice Hass.
85. Zienkiewicz, O.C. and R.L. Taylor, The Finite Element Method Volume 1 - The
Basis. 2000, Oxford, England: Butterworth-Heinemann.
86. Zienkiewicz, O.C. and R.L. Taylor, The Finite Element Method Volume 2 - Solid
Mechanics. 2000, Oxford, England: Butterworth-Heinemann.

Reference

263
87. Zienkiewicz, O.C. and R.L. Taylor, The Finite Element Method Volume 3 -
Fluid Dynamics. 2000, Oxford, England: Butterworth-Heinemann.
88. Trethewey, M.W., et al. Measurement and Application of Rotational Degrees of
Freedom in Structural Vibration. in Proceedings of the International
Conference on Noise and Vibration Control. 1993. St. Petersburg, Russia.
89. Moreno, D., et al., Modal vibration analysis of a metal plate by using a laser
vibrometer and the POD method. Journal of Optics A: Pure and Applied Optics,
2005. 7(6): p. 356-363.
90. Trethewey, W.M., H.J. Sommer, and J.A. Cafeo, A dual beam laser vibrometer
for measurement of dynamic structural rotations and displacements. Journal of
Sound and Vibration, 1993. 164(1): p. 67-84.
91. Bokelberg, E.H., H.J. Sommer, and W.M. Trethewey, A six-degree-of-freedom
laser vibrometer, part I: theoretical development. Journal of Sound and
Vibration, 1994. 178(5): p. 643-654.
92. Bokelberg, E.H., H.J. Sommer, and W.M. Trethewey, A six-degree-of-freedom
laser vibrometer, part II: experimental validation. Journal of Sound and
Vibration, 1994. 178(5): p. 655-667.
93. D'Ambrogio, W. and A. Sestieri. Models for accounting/eliminating rotational
DOFs in distributed structural modification. in Proceedings of the 18th
International Modal Analysis Conference. 2000. San Antonio, U.S.A.
94. Smith, E.J., Measurement of the total structural mobility matrix. Shock and
Vibration Bulletin, 1969. 40(7): p. 51-84.

Reference

264
95. Ewins, D.J. and M.G. Sainsbury, Mobility measurements for the vibration
analysis of connected structures. Shock and Vibration Bulletin, 1972. 42(1): p.
105-122.
96. Ewins, D.J. and P.T. Gleeson, Experimental determination of multi directional
mobility data for beams. Shock and Vibration Bulletin, 1975. 45(5): p. 153-173.
97. Cheng, L. and Y.C. Qu, Rotational compliance measurements of a flexible plane
structure using an attached beam-like tip, part 1: analysis and numerical
simulation. Journal of Vibration and Acousics, Transactions of the ASME, 1997.
119(4): p. 596-602.
98. Qu, Y.C., L. Cheng, and D. Rancourt, Rotational compliance measurements of a
flexible plane structure using an attached beam-like tip, part 2: experimental
study. Journal of Vibration and Acousics, Transactions of the ASME, 1997. 119:
p. 603-608.
99. Ewins, D.J., On predicting point mobility plots from measurement of other
mobility parameters. Journal of Sound and Vibration 1980. 70: p. 69-75.
100. Yasuda, C., et al. An estimation method for rotational degrees of freedom using
a mass additive technique. in Proceedings of the 2nd International Modal
Analysis Conference. 1984. Orlando, Florida, U. S. A.
101. Kanda, H., et al. Structural dynamic modifications using mass additive
technique. in Proceedings of the 4th International Modal Analysis Conference.
1986. Los Angeles, California, U. S. A.
102. Petersson, B., On the use of giant magnetostrictive devices for moment
excitation. Journal of Sound and Vibration, 1987. 116(1): p. 191-194.

Reference

265
103. Gibbs, B.M. and B. Petersson, Moment excitation and mobility measurement in
studies of structure-borne sound emission. Acoustics Bulletin, 1993. 18(3): p.
19-21.
104. EU Contract BRPR-CT97-540, P.N.B.-. Quantitative treatment and testing of
rotational degrees of freedom (QUATTRO): guidelines for experimental
techniques.
105. Su, J. and B.M. Gibbs, Measurement of point moment mobility in the presence of
non-zero cross mobility. Applied Acoustics, 1998. 54(1): p. 9-26.
106. Sanderson, M.A. and C.R. Fred, Direct measurement of moment mobility, part
I: a theoretical study. Journal of Sound and Vibration, 1995. 179(4): p. 669-684.
107. Sanderson, M.A., Direct measurement of moment mobility, part II: an
experimental study. Journal of Sound and Vibration, 1995. 179(4): p. 685-696.
108. Stanbridge, A.B. and D.J. Ewins, Modal testing using a scanning laser Doppler
vibrometer. Mechanical Systems and Signal Processing, 1999. 13(2): p. 255-270.
109. Duarte, M.L.M. and D.J. Ewins, Rotational degrees of freedom for structural
coupling analysis via finite-difference technique with residual compensation.
Mechanical Systems and Signal Processing, 2000. 14(2): p. 205227.
110. Yoshimura, T. and N. Hosoya. FRF estimation on rotational DOFs by rigid
block attachment method. in Proceedings of ISMA25: International Conference
on Noise and Vibration Engineering. 2000.
111. Trethewey, W.M. and H.J. Sommer. Measurement of rotational DOF frequency
response functions with pure moment excitation. in Proceeding of the 3rd
International Conference on Structural Dynamics Modeling Test, Analysis,
Correlation and Validation. 2002. Madeira, Portugal.

Reference

266
112. Dong, J. and K.G. McConnell. Extracting multi-directional FRF matrices with
Instrument Cluster. in Proceedings of the 20th International Modal Analysis
Conference. 2002. Los Angeles, California, U. S. A.
113. OCallahan, J.C., I.W. Lieu, and C.M. Chou. Determination of rotational
degrees of freedom for moment transfers in structural modifications. in
Proceedings of the 3rd International Modal Analysis Conference. 1985. Orlando,
Florida, U. S. A.
114. OCallahan, J.C., et al. An efficient method of determining rotational degrees of
freedom from analytical and experimental modal data. in Proceedings of the 4th
International Modal Analysis Conference. 1986. Los Angeles, California, U. S.
A.
115. Avitabile, P., et al. Expansion of rotational degrees of freedom for structural
dynamic modifications. in Proceedings of the 5th International Modal Analysis
Conference. 1987. London, England.
116. Mitchell-Dignan, M. and G.C. Pardon. The Estimation of Rotational Degrees-of-
Freedom Using Shape Functions. in Proceedings of the 6th International Modal
Analysis Conference. 1988. Orlando, FL. U. S. A.
117. Silva, J.M.M., N.M.M. Maia, and A.M.R. Ribeiro. Some Applications Of
Coupling/Uncoupling Techniques In Structural Dynamics - Part 1: Solving The
Mass Cancellation Problem. in Proceedings of the 15th International Modal
Analysis Conference. 1997. Orlando, FL. U.S.A.
118. Maia, N.M.M., J.M.M. Silva, and A.M.R. Ribeiro. Some Applications Of
Coupling/Uncoupling Techniques In Structural Dynamics - Part 2: Generation
Of The Whole FRF Matrix From Measurements On A Single Column - The Mass

Reference

267
Uncoupling Method (MUM). in Proceedings of the 15th International Modal
Analysis Conference. 1997. Orlando, FL. U. S. A.
119. Maia, N.M.M., J.M.M. Silva, and A.M.R. Ribeiro. Some Applications Of
Coupling/Uncoupling Techniques In Structural Dynamics - Part 3: Estimation
Of Rotational Frequency-Response Functions Using MUM. in Proceedings of
the 15th International Modal Analysis Conference. 1997. Orlando, FL. U. S. A.
120. Silva, J.M.M., N.M.M. Maia, and A.M.R. Ribeiro. Estimation of Frequency
Response Functions involving RDOFs Using an Uncoupling Technique. in
International Conference on Applications of Modal Analysis99. 1999. Gold
Coast, Queensland, Australia.
121. Silva, J.M.M., N.M.M. Maia, and A.M.R. Ribeiro. An Indirect Method for the
Estimation of FRFs Involving Rotational DOFs. in Proceedings of ISMA25:
International Conference on Noise and Vibration Engineering. 2000. Leuven,
Belgium.
122. Avitabile, P. and J.C. OCallahan, Frequency response function expansion for
unmeasured translation and rotation DOFs for impedance modelling
applications. Mechanical Systems and Signal Processing, 2003. 17(4): p. 723
747.
123. Lin, R. and Y. Xia, A New Eigensolution of Structures via Dynamic
Condensation. Journal of Sound and Vibration, 2003. 266: p. 93-106.
124. Guyan, R.J., Reduction of Stiffness and Mass Matrices. AIAA Journal, 1965.
3(Feb): p. 380.
125. Irons, B.M., Structural Eigenvalue Problems: Elimination of Unwanted
Variables. AIAA Journal, 1965. 3(May): p. 961-962.

Reference

268
126. Bouhaddi, N. and R. Fillod, A method for selecting master DOF in dynamic
substructuring using the Guyan condensation method. Computers and Structures,
1992. 45(5): p. 941-946.
127. Bouhaddi, N. and R. Fillod, Model Reduction by a Simplified Variant of
Dynammic Condensation. Journal of Sound and Vibration, 1996. 191(2): p. 233-
250.
128. Penny, J.E.T., M.I. Friswell, and S.D. Garvey, Automatic choice of measurement
locations for dynamic testing. AIAA Journal, 1994. 32(2): p. 407-414.
129. Paz, M., Dynamic condensation. AIAA Journal, 1984. 22(5): p. 724-727.
130. Paz, M., Modified dynamic condensation method. Journal of Structural
Engineering, ASCE, 1989. 115(1): p. 23-238.
131. Suarez, L.E. and M.P. Singh, Dynamic Condensation Method for Structural
Eigenvalue Analysis. AIAA Journal, 1992. 30(4): p. 1046-1054.
132. Zhang, N., Dynamic Condensation of Mass and Stiffness Matrices. Journal of
Sound and Vibration, 1995. 188(4): p. 601-615.
133. Qu, Z. and Z. Fu, An Iterative Method for Dynamic Condensation of Structural
Matrices. Mechanical Systems and Signal Processing, 2000. 14(4): p. 667-678.
134. Qu, Z. and W. Chang, Dynamic condensation method for viscously damped
vibration systems in engineering. Engineering Structures, 2000. 23: p. 1426-
1432.
135. Qu, Z.-Q. and R.P. Selvam, Insight into the dynamic condensation technique of
non-classically damped models. Journal of Sound and Vibration, 2004. 272: p.
581-606.

Reference

269
136. Qu, Z.-Q., R.P. Selvam, and Y. Jung, Model Condensation for Non-Classically
Damped Systems Part II: Iterative Schemes for Dynamic Condensation
Mechanical Systems and Signal Processing, 2003. 17(5): p. 1017-1032.
137. Kim, K.-O., Dynamic Condensation for Structural Redesign. AIAA Journal,
1985. 23(11): p. 1830-1831.
138. Kim, K.-O., Hybrid Dynamic Condensation for Eigenproblems. Computer and
Structures, 1995. 56(1): p. 105-112.
139. Kim, K.-o. and W.J. Anderson, Generalized Dynamic Reduction in Finite
Element Dynamic Optimization. AIAA Journal, 1984. 22(11): p. 1616-1617.
140. Takewaki, I. and K. Uetani, Efficient redesign of damped large structural
systems via domain decomposition with exact dynamic condensation. Computer
Methods in Applied Mechanics and Engineering, 1999. 178: p. 367-382.
141. Bouhaddi, N. and R. Fillod, Substructuring using a linearized dynamic
condensation method. Computer and Structures, 1992. 45: p. 679-683.
142. Bouhaddi, N. and R. Fillod, Substructuring by a Two Level Dynamic
Condensation Method. Computer and Structures 1996. 60(3): p. 403-409.
143. Qiu, J.-B., Z.-G. Ying, and F.W. Williams, Exact Modal Synthesis Techniques
Using Residual Constraint Modes. International Journal for Numerical Methods
in Engineering, 1997. 40: p. 2475-2492.
144. Leung, A.Y.-T., An Accurate Method of Dynamic Condensation in Structural
Analysis. International Journal for Numerical Methods in Engineering, 1978.
12(11): p. 1705-1715.
145. Flippen, L.D.J., Polynomial-basis model reduction. Mathematical and computer
modelling, 1992. 16(12): p. 121-132.

Reference

270
146. Flippen, L.D.J., Fundamental law preservation in model reduction.
Mathematical and computer modelling, 1992. 16(11): p. 171-182.
147. Flippen, L.D.J., Abstract Zwanzig Model Reduction Theory with Application to
Discretized Linear Systems. Mathematical and computer modelling, 1992.
16(10): p. 121-134.
148. Flippen, L.D.J., A theory of condensation model reduction. Computers &
mathematics with applications, 1994. 27(2): p. 9-40.
149. Flippen, L.D.J., Constitutive-Operator Smoothing by Condensation. Computers
& mathematics with applications, 1994. 27(6): p. 5-18.
150. Flippen, L.D.J., Interpolation-Based Condensation of Algebraic Semi-Discrete
Models with Frequency Response Application. Computers & mathematics with
applications, 1995. 29(9): p. 39-52.
151. Dyka, C.T., R.P. Ingel, and L.D. Flippen, A New Approach to Dynamic
Condensation for FEM. Computer and Structures, 1996. 61(4): p. 763-773.
152. O'Callahan, J.C., P. Avitabile, and R. Riemer. System equivalent reduction
expansion process (SEREP). in Proceedings of the 7th International Modal
Analysis Conference. 1989. Las Vegas, Nevada, U.S.A.
153. Sastry, C.V.S., et al., An iterative system equivalent reduction expansion process
for extraction of high frequency response from reduced order finite element
model. Computer Methods in Applied Mechanics and Engineering, 2003. 192: p.
1821-1840.
154. OCallahan, J.C. A procedure for an Improved Reduced System (IRS). in
Proceedings of the 7th International Modal Analysis Conference. 1989. Las
Vegas, Nevada, U. S. A.

Reference

271
155. Friswell, M.I., S.D. Garvey, and J.E.T. Penny, Model Reduction Using Dynamic
and Iterated IRS Techniques. Journal of Sound and Vibration, 1995. 186(2): p.
311-323.
156. Friswell, M.I., S.D. Garvey, and J.E.T. Penny, The Convergence of the Iterated
IRS Method. Journal of Sound and Vibration, 1998. 211(1): p. 123-132.
157. Friswell, M.I., J.E.T. Penny, and S.D. Garvey, The Application of the IRS and
Balanced Realization Methods to Obtain Reduced Models of Structure with
Local Non-Linearities. Journal of Sound and Vibration, 1996. 196(4): p. 453-
468.
158. Kim, K.-O. and M.-K. Kang, Convergence Acceleration of Iterative Modal
Reduction Methods,AIAA Journal, 2001. 39(1): p. 134-140.
159. Xia, Y. and R. Lin, Improvement on the iterated IRS method for structural
eigensolutions. Journal of Sound and Vibration, 2004. 270: p. 713-727.
160. Jones, R. Structural modification using modal and frequency domain techniques.
in The 10th International Seminar on Modal Analysis. 1985. Leuven, Belgium.
161. Allemang, R.J. The modal assurance criterion (MAC): twenty years of use and
abuse. in Proceedings of the 20th International Modal Analysis Conference.
2002. Los Angeles, California, U. S. A.
162. Dynamic-Design-Solutions, FEMtools Users Guide Version 3.0. 2003.
163. Dynamic-Design-Solutions, FEMtools Examples Manual Version 3.0. 2003.
164. SPECTRAL-DYNAMICS, The STAR System Manuals, User Manual 3405-0113.
1994.
165. SPECTRAL-DYNAMICS, The STAR System Manuals, Reference Manual 3405-
0114. 1994.

Reference

272
166. Avitabile, P., Modal Space - In Our Own Little World: I heard someone say
"Pete doesn't do windows", what's the scoop? SEM Experimental Techniques,
1999(October).
167. Cuppens, K., P. Sas, and L. Hermans. Evaluation of the FRF Based
Substructuring and Modal Synthesis Technique Applied to Vehicle FE Data. in
Proceedings of ISMA25: International Conference on Noise and Vibration
Engineering. 2000. Leuven, Belgium.
168. Hermans, L., et al. Estimation and use of residual modes in modal coupling
calculations: A case study. in Proceedings of the 18th International Modal
Analysis Conference. 2000. San Antonio, Texas, U. S. A.
169. Blevins, R.D., Formulas for natural frequency and mode shape. 1979, New
York: Van Nostrand Reinhold Company.

You might also like