You are on page 1of 9

Proceedings of the IMAC-XXVII

February 9-12, 2009 Orlando, Florida USA


2009 Society for Experimental Mechanics Inc.

Interactions of Modal Analysis and Wave Propagation Using


Timoshenko Beam Model
A. M. Iglesias NSWC Dahlgren
A. L. Wicks, Virginia Tech
T. Schwartz, NASA Goddard
Abstract
While modal analysis and wave
propagation concepts are both rooted in the
fundamentals of structural dynamics, these
two approaches to modeling the dynamic
response of beams to inputs are not often
considered in tandem. A differential equation
representation of a beam is developed using
Timoshenko Beam equation. Shear is
included as well as a spatial and temporal
varying forcing function. The intention of this
model is to study wave propagations compressive waves (P waves), shear
wave(S waves), bending waves and
Rayleigh waves - relative to the applied
moving forcing functions, but the first step of
the study is to validate the model and then
move to adding structural parameters (i.e.
foundation stiffness).
Validation of the discrete solution is
developed as the transient wave response
converges to the modal solution for a bandlimited spatially fixed forcing function.
Frequency response functions and the
resulting natural frequencies are extracted
from the transformed response functions.
Introduction
Wave propagation and modal
analysis are two different methods of
considering the dynamic response of a
structure. Modal analysis has gained a
prominence in the field and is applied to all
manner of structures. Modal testing of large
structures, such as spacecraft or buildings,
presents a challenge to the experimentalist.
Often when testing these large structures it
is difficult to extract modal information that
can be correlated with the predictions made
by analytical means, such as a finite
element model (FEM). The FEM in general
will yield real normal mode shapes, while the
mode shapes found experimentally can
often be complex in nature and poorly
defined. The root cause of these difficulties

is of interest.
While improved data
acquisition and parameter
extraction
techniques can improve results for some
structures, the more fundamental approach
taken here is to question whether the
assumptions made in the development of
modal theory are valid when considering the
response of some large structures or
structures with high damping.
The intent of this paper is to present
a dynamic model of the response of a finite
Timoshenko beam supported by fixed-fixed
boundary conditions and subject to an
impulse excitation.
The assumptions
regarding the separation of the temporal and
spatial response made in a modal approach
need not be made in the development of this
model. In this manner following an impulse
excitation the response can be observed as
a propagating waves and transitions to a
modal response.
In this paper we will first have a
discussion of the motivation for constructing
this model. Secondly, the development of
the model will be explained. Last is a
description of how the model was validated
and confirmed to adhere to known behavior
of vibrating structures.
Background
This
section
discusses
the
motivation to develop the model presented
in this paper, as well as a comparison of
how real normal modes are realized in
modal analysis and wave propagation.
Justification is also given for the use of the
Timoshenko beam model.
Motivation
The primary purpose of the model
presented in this paper is to explore if the
assumptions made in the development of
modal analysis are valid for very large
structures.
Experimental modal tests
performed on very large structures such as
spacecraft, buildings, and bridges have

proved to be a challenge.
Traditional
excitation techniques may not be able to
adequately excite the structure resulting in
poor coherence in the measurements.
Isolating large structures with either very stiff
or very soft interfaces, as is often done with
small test articles, is often impossible. As a
result, accurate representation of the
boundary conditions can be difficult.
Channel counts are often rather high on
tests of large structures making it
cumbersome to manage all the data. In the
case of these problems the difficulty is in the
test setup.
Some have made efforts to
improve techniques with the aim of
improving the results in these situations.
However, the goal of this effort is to consider
a more fundamental obstacle to modal
testing of large structures, the ability of the
structure to physically realize real normal
modes. If true, modal analysis would not be
an appropriate method of characterizing the
vibration of these structures and wave
propagation methods may be more suitable.
Real Normal Modes of Vibration
In order to consider if modal
analysis is appropriate for very large
structures, an understanding about real
normal modes must first be developed
relative to wave propagation theory and
modal analysis theory.
Modal analysis
theory will be considered first.
In this
discussion the modes of a beam in bending
are considered for simplicity, but the
principles extend to more complex
structures.
Interestingly enough, even though
the dynamic responses of structures consist
of the summation of many propagating
waves, modal theory can bypass all those
details. The modal solution is based on the
roots of the equations of motion. Using an
eigenproblem approach, as is commonly
done, the natural frequencies are found by
the square root of the eigenvalues, and the
mode shapes are scaled versions of the
eigenfunctions. These natural frequencies
and mode shapes are the basis of forming
the system response when using the modal
superposition method. The response is
composed of a weighted summation of the
mode shapes. The weighting functions are
a function of time. This weighted summation
is where the separation of variables is
employed as a part of the assumed solution

and also where it clashes with the wave


propagation concept.
In the modal
formulation of the beam response when an
impact is imparted at one end of beam, each
mode that does not have a node at the
location impacted is instantaneously excited
throughout the entire mode shape.
Physically, it is not possible to produce that
instantaneous response, but if an infinite
number of modes are included in the
summation, the higher frequency modes can
cancel out the unwanted instantaneous
response. As the solution proceeds in time,
the high frequency modes doing the
canceling decay much faster than the lower
frequency modes and allow the low
frequency modes to dominate response
across the full mode shape. The behavior
defining real normal modes in a wave
propagation sense is more complicated
because the individual traveling waves and
each of their reflections must be considered.
Only very simple structures can be
reasonably considered using this method
such as the beam in bending considered
here.
The deflection response of the
beam is comprised of the summation of
traveling waves and the reflections produced
when these waves reach a boundary. The
summation of these waves produces
constructive and destructive interference.
Real normal mode shapes are realized
when the interference pattern is stationary.
The stationary wave, or standing waves
pattern, is the result of a condition termed
phase
coincidence.
An
excellent
explanation of phase coincidence is given in
[2]. In brief the phase coincident condition
can be described by the behavior of a
simple harmonic wave traveling in a beam of
finite length.
If the harmonic wave originates at
the left end of the beam it will propagate
along its length until it reaches the right end
and is reflected, as shown in figure 1a and
1b. The sum of these two waves generates
an interference pattern. Depending on the
end conditions of the beam and the
frequency of the wave, the pattern may be
stationary. Furthermore after the reflected
wave returns to the left end it is reflected
again. If this second reflection, which is
traveling in the same direction as the initial
wave, is such that the phase is identical to
the phase of the initial wave, the phase

Figure1: Harmonic waves traveling forward, (a) and (c), and


backward (b) from on end to t he other of a sim ply supported
beam which interfere to produce the pure standing-wave field
shown in (d). Figure taken from Ref [2].

coincident condition is achieved. The result


is purely constructive interference, where if
the beam is undamped the response will
increase unbounded as the reflections
continue to accumulate. The frequencies
where this occurs are the natural
frequencies of the beam.
For a more
detailed look at this phenomenon, including
the different reflection behaviors for different
end conditions and a more in depth
discussion of the interference patterns, see
[2].
Selecting a Beam Model
The Euler-Bernoulli beam model is
the most common, but for this study it has
some
deficiencies
that
make
the
Timoshenko beam model a more attractive
option. The Euler-Bernoulli beam model
neglects the effects of shear deformation
and rotational inertia. In many cases the
effect is negligible if only the first few modes
are of interest and the length of the beam is
significantly longer than the dimensions of
the cross-section. When this condition is not
satisfied, the wave length of the modes of
interest may be close to the dimensions of
the cross-section resulting in a significant
discrepancy in the response of the two
models.
The other major effect of the
assumptions made in the Euler-Bernoulli
beam model relates to the phase velocity of
traveling waves. As traveling waves of
shorter and shorter wavelength are

considered, the phase velocity of the wave


increases unbounded. In the result for an
infinitely short wavelength, a wave of
infinitely high frequency, the response of the
beam is instantaneous regardless of the
distance between the point of excitation and
response. This is physically impossible and
a flaw in the model. On the other hand, the
Timoshenko beam model has bounded
phase velocity for decreasing wavelengths,
preventing the unwanted instantaneous
response and matching extraordinarily well
with the exact theory of elasticity. Other
beam models include the shear beam model
and the Rayleigh beam model. These two
models both build from the Euler-Bernoulli,
the shear beam model adds only the shear
deformation effects, while the Rayleigh
beam model adds only the effect of
rotational inertia.
Model Development
In this section the model is
developed from the equations of motion for
a Timoshenko beam with fixed-fixed
boundary conditions. The two second order
partial differential equations are transformed
into four first order ordinary differential
equations when made discrete in the spatial
variable. The ODEs are solved using
MATLABs ode45 solver.
Equations of Motion
The two second order partial
differential equations of motion that govern
the vibration of a Timoshenko beam are
developed in many texts. The notion used
in [1] is used here.
(x, t ) 2 y ( x, t )
2 y ( x, t )
+ A
GA

= q ( x, t )
2
x
t 2
x

2 ( x, t )
2 (x, t )

y ( x, t )
GA
= I
( x, t ) + EI
2
x
t 2

where
x=
t=
y(x,t) =
(x,t) =
shear
G=
A=
E=
I=
=

position
time
transverse displacement
slope of the centroidal axis due to
shear modulus
cross-sectional area
elastic modulus
second moment of area
Timoshenko shear coefficient

=
q(x,t) =

mass density per unit length


applied forcing function

Isolating the time derivatives in each, they


can be rewritten as,
2 y ( x, t ) q ( x, t ) G
=

t 2

( x, t ) 2 y ( x, t )

x 2
x
2
2 ( x, t ) GA y ( x, t )
E ( x, t )
=
( x, t ) +

2
2
I x
t
x

To break these two second order


equations into four first order equations with
respect to time, the variables z1, z2, z3, and
z4 are defined as,

z1 = y
z3 =

y
t

z4 =
t
z2 =

The four resulting equations are first order


with respect to time.
z1 (x, t )
= z 2 ( x, t )
t
z2 ( x, t ) q( x, t ) G z3 ( x, t ) 2 z1 ( x, t )

A
x
t
x 2

z3 ( x, t )
= z 4 ( x, t )
t
2
z4 ( x, t ) GA z1 ( x, t )
E z 3 ( x, t )
=
z 3 ( x, t ) +

2
I x
t
x

If the spatial variable, x, is taken at


discrete points the partial derivatives can be
approximated by the central difference
approximation. The discrete spatial variable
is defined as xi = ih where h is the spatial
step size and xi is defined over the
range 0 x L . The central difference
approximation of the first and second
derivatives with respect to x are given by,
z ( xi , t ) z (xi +1 , t ) z (xi 1 , t )
CDA
=
2h
x
2
z (xi , t ) z (xi +1 , t ) 2 z ( xi , t ) + z (xi 1 , t )
CDA
=
2
h2
x

Applying
the
discrete
spatial
variable and substituting the derivative

approximation, allows the four PDEs to be


written as four first order ODEs.
z1 ( xi , t )
= z 2 (xi , t )
t
z 2 (xi , t ) q(xi , t ) G z 3 ( xi+1 , t ) z3 (xi 1 , t ) z1 (xi +1 , t ) 2 z1 ( xi , t ) + z1 (xi 1 , t )
=

2h
h2
A

t

z3 (xi , t )
= z 4 ( xi , t )
t
z 4 (xi , t ) GA z1 (xi +1 , t ) z1 (xi 1 , t )
E z3 ( xi+1 , t ) 2 z3 ( xi , t ) + z3 (xi 1 , t )
=
z3 (xi , t ) +

h2
t
I
2h

Addition of Damping
Viscous damping can be added to
this set of equations by adding a term
proportional to the velocity of the deflection.
Returning to the first of the equations of
motion and adding the damping term yields,
(x, t ) 2 y ( x, t )
2 y ( x, t )
y ( x, t )
+ A
GA

+C
= q ( x, t )
2
x
t 2
t
x

where C is the damping coefficient. Now


making substitutions with the previously
defined z variables, the discrete form of the
equation can be written as,
z 2 ( xi , t ) q( xi , t ) G z3 ( xi +1 , t ) z3 ( xi 1 , t ) z1 ( xi+1 , t ) 2 z1 ( xi , t ) + z1 ( xi 1 , t ) C
=

z 2 ( xi , t )

2h
t
h2
A

A

To include damping in the solution this


equation replaces the second of the four first
order ODEs.
Boundary Conditions
The boundary conditions are given
as fixed-fixed which makes the deflection,
slope, slope due to shear, and curvature due
to shear all zero at both ends of the beam.
First consider deflection and slope at x = 0,

y ( x0 , t ) = 0

y (x0 , t )
=0
x

The boundary conditions are made discrete


in x just as was done in the equations of
motion. The derivative in x however cannot
be treated with the central difference
approximation because the term xi-1 does
not exist when i=0. Instead the forward
difference approximation is used.

y ( x0 , t ) y (x1 , t ) y ( x0 , t )
FDA
=0
=
h
x

This expression can be rearranged to show


that for the slope to equal zero the first and
second points of the beam will have equal
deflection. Since, the deflection at the first
point is already known to be zero,

y ( x1 , t ) = y ( x0 , t ) = 0
Similarly at the end where x = L, the
backwards difference approximation is used
because of the restriction of the rightward
terms

y (x L , t ) y ( xL1 , t ) y ( x, t )
=0
BDA
=
h
x
can be rewritten as

y ( x L 1 , t ) = y ( x, t ) = 0
The same procedure can be used for the
slope due to shear and curvature due to
shear to find

(x0 , t ) = (x1 , t ) = (xL 1 , t ) = (xL , t ) = 0


These boundary conditions are included in
the MATLAB function created to be solved
by the ode45 solver.
Modeling the Impulse
One method of modeling an impulse
is to provide an initial velocity at a single
point to simulate a sharp blow with a very
stiff object. The result is the input of
vibration energy at all frequencies raising
concerns about aliasing in the response.
Because the signal is not digitized from
analog, a traditional anti-aliasing filter is not
acceptable. In order to eliminate aliasing,
the impulse to the beam will be provided by
the forcing function term q(x,t) that has
known frequency content. A rectangular
profile in the frequency domain given by

1 < c
rect (c ) =
0 > c
will ensure frequency content to be below
c. The inverse Fourier transform of the
rectangular function is the sinc function,

sinc(t ) =

sin (ct )
c t

Because the sinc function is centered about


t = 0, as shift is required to include the entire
pulse in the simulation at positive values of t.
To further reduce the amplitude of the tails
of the sinc function impulse a Gaussian
function is applied to the input. The addition
of the Gaussian window further emphasizes
the low frequency content. As a result the
effective cutoff frequency becomes even
lower than that specified by c. The extent
of this effect is briefly revisited in the model
validation section.
The variables c and shift represent
the cutoff frequency and the time shift to be
used in the simulation. The final input to the
function, input_index, defines the index of
the x vector that is the point of excitation.
Model Validation
In this section, several methods are
used to verify that the behavior of the model
is representative of the known behavior of a
vibrating beam. In the time domain the
beam response is animated, and the
responses of individual points are plotted
against time to confirm they are realistic
responses.
To observe the frequency
domain behavior, the auto-power spectra of
the impulse and individual responses are
considered as well as frequency response
functions (FRFs). Further observation is
made using a time-frequency analysis by
plotting the smoothed pseudo Wigner-Ville
To
distribution of individual points.
demonstrate the methods used to verify the
model, consideration will be given here to
one case study.
Definition of Case Study Parameters
L=1m
x = [0:1:100]
E = 200 GPa
shift = 4 ms
G = 10 GPa
wc = 5000 Hz
Base = 5 cm
input_index = 26
Height = 10 cm
= 7800 kg/m^3

appearance of a combination of the first and


second modes of vibration. It is also evident
that as time progresses, the amplitude of the
oscillations is decreasing.

(a)

(b)

(c)

Displacement

The Excitation Impulse in the Time-domain


The impulse in the time domain was
a shifted sinc function windowed with a
Gaussian distribution. Figure 2 shows the
resultant
impulse.
Three
primary
observations can be made from figure 2.
First, the shape of the resultant impulse is
indeed of the nature of a sinc function.
Secondly, the Gaussian window applied to
the sinc function has reduced the portion of
the impulse in the tails of the sinc function.
The result is a focusing of the impulse on a
shorter time span, making the impulse more
realistic. Thirdly, the 4ms shift employed is
sufficient to prevent any significant portion of
the impulse from occurring prior to t = 0. If a
portion of the impulse were to be cutoff at t =
0 the result would be a discontinuity that
would introduce unwanted high frequency
content and could potentially cause aliasing
problems.

(d)

(e)

(f)

(g)
beam length

Figure 3: Beam vertical displacement at (a)


3.6 ms, (b) 4.0 ms, (c) 4.4 ms, (d) 4.8 ms,
(e) 5.2 ms, (f) 5.6 ms, and (g) 6.0 ms,

Figure 2: The time-domain representation of


the excitation pulse with a shift of 4ms
Animation of the Response
The first method of validation is
simply an animation of the beams response.
While the animation itself cannot be shared
here, figure 3 shows some frames of the
time response along the length of the beam.
There are several important aspects of the
response to be noted, beginning with the
basic behavior of the beam. The profile of
the beam is smooth throughout and does
not appear to have any discontinuities. Also
the fixed boundary conditions are enforced
and the responses of the points adjacent to
the boundary are realistically constrained.
While difficult to fully demonstrate in
this format, an observer of the animation
familiar with modal analysis can see that as
time passes the response takes on the

Beam Response in the Time Domain


Figure 4a shows the response at the
point where x = 0.75 meters plotted versus
time. The appearance of the response
indicates that it is dominated by the
summation of some number of decaying
sinusoidal functions. The decaying property
confirms the realistic application of damping
to the system equations. The summation of
sinusoidal functions suggests that there is a
particular set of frequencies which continue
to respond significantly longer than the rest
of the observed bandwidth. This behavior is
indicative of the natural frequencies of a
structure. Figure 4 suggests that more than
one mode is excited at x = 0.75
In contrast, figure 4b taken at x =
0.5 appears to be dominated by a single
decaying sinusoid indicating a single mode
of vibration dominating the response. In this
beam example the second mode is
expected to have a node at x = 0.5, thus
there would be no response at that location
in the second mode and the first mode
would dominate the response if no higher
modes were excited as appears to be the

case in this example. The frequency domain


analysis will further investigate if these are
indeed resonant frequencies.

frequency increases. The peaks in the autopower spectra indicate possible resonant
modes of vibration. The relative phase
information found in the FRF is required to
confirm that these peaks are indeed
resonances.

(a)

(a)

(b)
Figure 4: The deflection at (a) x = 0.75m,
and (b) x = 0.5m from time 0 to 100 ms.
Auto-Power Spectra and Frequency
Response Functions
To further validate both the impulse
and responses, the frequency domain is
also considered. The auto-power spectra
and FRFs were calculated. Figure 5(a)
shows the calculated auto-power spectra of
the excitation impulse. The plot shows the
result of the combined sinc function and
Gaussian distribution used to specify the
input to the beam.
As the frequency
increases the energy rolls off dramatically
with the -3dB point being lower than 500 Hz.
The auto-power spectra of the
response at x = 0.5 and x = 0.75 are shown
in figure 5(b). Much like the auto-power
spectrum of the impulse, the auto-power
spectra of the responses rolls off as the

(b)
Figure 5: The auto-power spectrum of (a)
the excitation impulse and (b) the beam
deflection at x = 0.5 and x = 0.75.
The FRFs plotted in figure 6 show
the telltale phase shifting associated with
resonant frequencies and confirms that the
model does indeed demonstrate the
resonant behavior that was sought. Also of
note are the differences between the FRF at
x = 0.5 and x = 0.75. Because the node of
the second mode exists at the midpoint of
the beam, the second peak is suppressed in
the FRF of the response at x = 0.5.

Figure: 6: Frequency response functions


(FRF) of the deflection response at x = 0.5
and x = 0.75 with the excitation impulse
used as the reference.

(a)

The Smoothed Pseudo Wigner-Ville


Distribution
Time-frequency techniques allow for
further
insight
into
how
frequency
characteristics change with time. Figure7
shows the smoothed pseudo Wigner-Ville
distribution of the response at x = 0.75
calculated using the time-frequency toolbox,
Mathworks.
Focusing on the well
represented first two modes several
observations can be made that confirm the
behavior of the model is consistent with the
behavior of real beams. First, at the time of
the impulse excitation the energy is
distributed across a number of frequencies,
but is quickly focused into the frequencies
representing the modes of vibration. This
behavior represents the transition from a
wave propagation dominated response to a
modal response. This is a very important
phenomenon to verify as it will be of key
interest when larger beams are considered.
Another important observation is the
manner in which the modes are decaying.
The decay itself was earlier observed in the
time domain, but in the time-frequency
representation it can also be observed that
the second mode is decaying faster than the
first mode. The behavior is also exhibited in
real beams, as the higher number modes
undergo more oscillations and generally
damp out quicker than their lower number
mode counterparts.

(b)
Figure 7: The smoothed pseudo WignerVille distribution of the deflection response
at (a) x = 0.75 and (b) x = 0.5
The smoothed pseudo Wiger-Ville
distribution of the response at x = 0.5 is
shown in figure 7(b). The observations
made of the response at x = 0.75 are again
made with the exception of the second
mode. As noted in the FRF and auto-power
spectrum of the response at x = 0.5, the
second mode is not evident at this location
because it lies on the node of the mode.
Conclusions
In this paper a model for the
response of a finite Timoshenko beam
supported
by
fixed-fixed
boundary
conditions and subject to an impulsive
excitation has been presented. The model
has been verified to behave like vibrating
beams are known to behave through time,
frequency, and time-frequency domain
observations. This model will be scaled to

calculate responses of long beams and does


not rely on the calculation of the modal
matrix or the method of separation of
variables. The model demonstrates the
ability to model traveling waves and the
modal response of the beam. As such this
model will allow further study into how the
assumptions made by modal analysis
techniques may or may not be valid in the
testing of large structures.
In order to improve this model and
make it closer to a real structure, we will add
the foundation stiffness and dispersion
phenomena that can affect the wave
propagation, resonance frequencies and the
time it takes to converge to the modal state.
This model will put us in the position to
analyze and understand wave propagation
interacting with a moving load.
References
1. Graff K, 1991, Wave Motion in Elastic
st
Solids (1 ed), New York: Dover
Publications, 1991
2. Fahy F, Gardonio P, 2007, Sound and
Structural Vibration: Radiation,
Transmission and Response (2nd ed),
American Press, New York.

You might also like