Sei sulla pagina 1di 13

HOW TO MATCH THEORETICAL AND EXPERIMENTAL BOUNDARY

CONDITIONS OF A CANTILEVER BEAM


Thiago Ritto 1* , Romulo Aguiar1 , Rubens Sampaio1 , and Edson Cataldo2
1

Mechanical Engineering Department


PUC-Rio
Rua Marques de Sao Vicente, 225, Gavea, CEP: 22453-900, RJ, Brazil
e-mail: thiagoritto@gmail.com, romulo@mec.puc-rio.br, rsampaio@mec.puc-rio.br
2

Applied Mathematics Department, Graduate Program in Telecommunications Engineering,


Graduate Program in Mechanical Engineering
UFF
Rua Mario Santos Braga, S/N, Centro, Niteroi, CEP: 24020-140, RJ, Brazil,
e-mail: ecataldo@im.uff.br

Keywords: boundary condition, Timoshenko beam, experimental modal analysis, finite elements
ABSTRACT
The purpose of this paper is to describe how to match experimental and theoretical frequencies
of a cantilever beam through a prescribed error tolerance. In some cases there is a clear difference between a clamped boundary condition in a real system or in an experiment (called in
this article as e-clamped condition) and the clamped condition as regularly modeled in theory
(the t-clamped condition). In such occasions the e-clamped (experimental) might not follow the
t-clamped (theoretical) condition. Actually, it hardly will. Depending on the e-clamped boundary condition the error between the natural frequencies of t-clamped and e-clamped might be
greater than 10%. This work presents a theoretical model that relaxes the boundary condition
of zero rotation. One parameter, a torsional stiffness at the fixed end of the beam, is used in
order to fit the natural frequencies of the numerical simulations with the experimental data.
The idea is to identify the torsional stiffness that minimizes the error between the numerical and
experimental frequencies. For the numerical simulations, the Finite Element Method (with Timoshenko beam element) is used, and for the experiments, multiple rubber layers are mounted
to the clamped side of the beam so that different e-clamped conditions can be analyzed and
reproduced if necessary. The numerical and the experimental results show excellent agreement.
1.

INTRODUCTION

In structural dynamics modeling, the clamped end is used as a boundary condition of a mount,
joint, etc. The theoretical clamped boundary condition implies that no movement in any direction nor rotation is allowed, mentioned in this work as t-clamped condition [1, 2]. However,
E110

in the field or in a laboratory experiment, this clamped condition, mentioned throughout this
work as e-clamped, might not respect the same restriction of no movement or rotation, as in the
t-clamped condition. Sometimes the mount has some flexibility and due to such characteristic
the entire numerical simulation might not have a good agrement with experimental data. Nevertheless, this small change in the flexibility of a clamped boundary condition may substantially
change the dynamic response of the entire system, such as natural frequencies and vibration
modes.
The present work proposes an alternative way to model the e-clamped boundary condition
of a cantilever beam, because a simple boundary condition might not be that simple [3]. One
parameter (torsional stiffness) is introduced in the clamped side in order to fit the numerical
simulations with the experiments. This work shows that such minor change in the boundary
condition is enough to represent the experimental data. In the context of Finite Element Model
Updating there are some studies of this kind as, for instance, [4, 5].
A similar idea was used by [6] in the impact of the temperature field on vibration of beams.
Due to the relatively high possible temperature variations in beams it was developed a new
model for the clamped and the simply supported beam. To obtain better analytical results in
comparison with experimental data, a spring was added and the comparison between the analytical model and experimental data showed good agreement.
To identify the optimal value of the torsional stiffness that minimizes the error between numerical and experimental results an optimization procedure is performed. The results achieved
herein might be used to change the dynamical characteristics of a structure, avoiding unwanted
frequencies. One can add some flexibility into a mount, for instance, and change the characteristic of the boundary condition in order to change the system natural frequencies. With such
action it is possible to change the dynamical characteristics avoiding undesirable frequencies
with a procedure that can be less expensive than change the system itself.
2.
2.1

EXPERIMENTAL APPARATUS AND RESULTS


Test rig apparatus and data acquisition methodology

The experimental apparatus attempts to obtain the frequency response of the cantilever beam
when the e-clamped condition is changed. The experimental apparatus is shown in Fig. 1.

(a)

(b)

Figure 1. Test rig: a) photo; b) scheme.


The experiment is composed by a simple cantilever beam, supported by a cast iron bench.
The beam is made of aluminum, with a 511 mm length and a retangular section of 30.7 by
3.04 mm. The mount is composed by steel couplings, which tries to guarantee the boundary
E110

condition of zero displacement and zero rotation by pressing the aluminum beam against both
couplings, and the mount is fixed at the cast iron rig, using 7/16 inch bolts for all couplings.
The frequency response is obtained [7] using an impulse modal hammer (Endevco Model
30927). The system response is captured by two mini-accelerometers (Endevco, Model 25B),
one positioned at the free end of the beam and the other at the middle. All signals (hammer and
accelerometers) are preconditioned before analyzed by the signal processor (HP 35650).

Figure 2. Mount photo.

The specification of all sensors and the modal hammer are shown in Tab. 1.

Table 1. Hammer and sensors specs.


Accelerometer 1 - 25B SN BL47
Sensitivity
Weight
Measure Range
Resonance frequency
Accelerometer 2 - 25B SN BL557
Sensitivity
Weight
Measure Range
Resonance frequency
Impulse Modal Hammer - 30927 SN 1660
Sensitivity
Max impulse
Range

4.7902
0.2 103
1000
50

mV /g
Kg
g
kHz

4.7707
0.2 103
1000
50

mV /g
Kg
g
kHz

99.7
1000
50

mV /lb
lbf
lb

The methodology applied here is to obtain the frequency response of the beam as the boundary condition at the clamped end is varied. This boundary condition is varied by applying
multiple rubber layers at the mount, as shown on the sketch of Fig. 3. The number of layers
vary from 0 (no rubber layer) to 10 rubber layers.

E110

(a)

(b)

Figure 3. Rubber layer: (a) Scheme; (b) Photo (1 layer).


For each boundary condition the frequency response of the cantilever beam is obtained using
the modal hammer. The modal test is ran two times, each time is an average of six distinct
impulses applied by the hammer in different points of the beam.
Finally the experimental results will be used to validate the mathematical modeling of the
cantilever beam, with the objective to predict the natural frequencies for a given number of
layers.
2.2

Experimental results

As mentioned before, for each boundary condition the modal tests are ran twice, with each test
an average of six impulses applied in different points of the beam. The natural frequencies are
obtained from the magnitude signal of both accelerometers and confirmed by the phase chart.
Figure 4 shows the frequency response of the cantilever beam with no rubber layers.

Figure 4. Frequency response, 0 rubber layers, data from accelerometer # 1, two modal tests.

E110

Figure 4 shows the response of the accelerometer #1 in both experimental data acquisitions.
All experimental data was acquired and analyzed from accelerometers #1 and #2, but once the
response of accelerometer #1 gives a better resolution than accelerometer #2, the results shown
in this article are from accelerometer #1. Figure 4 shows 6 major peaks, corresponding to the
first six natural frequencies, the first natural frequency at 9 Hz and the sixth natural frequency
around 790 Hz. It is also noticeable the presence of several minor sharp peaks in multiples of
60 Hz. Those peaks correspond to the local electric supply and its harmonics, and those peaks
should not be considered for the modal analysis.
Figure 5 shows the frequency response in four distinct boundary conditions: 1, 3, 6 and 10
rubber layers.

(a)

(b)

(c)

(d)

Figure 5: Frequency response: (a) 1 rubber layer; (b) 3 rubber layers; (c) 6 rubber layers; (d) 10
rubber layers.
From the charts of all boundary conditions it is possible to obtain the natural frequencies for
each case and compare each frequency of each particular vibration mode with the variation of
the boundary condition. This is shown in Tab. 2.

E110

Table 2. Natural frequencies - changing boundary conditions.


natural frequency (Hz)
rubber layers
0
1
2
3
4
6
8
10

1st

2nd

3rd

4th

5th

6th

9.25
9.00
9.00
9.00
9.00
9.00
9.00
9.00

58.80
57.00
57.00
56.75
56.75
56.65
56.50
56.25

163.75
160.90
160.75
160.00
160.15
159.90
159.75
158.75

321.50
315.60
315.25
314.00
314.20
313.60
313.20
311.50

534.50
524.75
525.40
523.00
523.00
522.75
521.50
519.25

793.20
778.75
780.00
777.25
777.00
775.75
773.50
770.75

Observing the system frequency response and the data shown in Tab. 2, it can be noticed that
there is no variation of the first frequency with the presence of the rubber layers, but it is worth
it to remark that the accuracy of the experiment is 0.25Hz, what makes it difficult to distinguish
9.0Hz from 9.1Hz, for example.
The higher the number of rubber layers more flexible the boundary conditions becomes and
lower the natural frequencies, compared to the condition without rubber layers. A comparison
between the opposite sides of the experiment, i.e., the frequency response of the beam without
rubber layers on the clamped end and the system with 10 rubber layers is shown in Fig. 6.

(a)

(b)

Figure 6: Frequency response, boundary condition comparison. Thin line: 0 rubber layers;
thick line: 10 rubber layers. a) 0-200 Hz, b) 300-800Hz
3.
3.1

MODEL OF THE CLAMPED BOUNDARY CONDITION


The FE model and the clamped boundary condition

Consider the beam element shown in Fig. 7.

E110

Figure 7. Beam element.


In the Finite Element Method, the clamped boundary condition is usually modeled as u1a |x=0 =
u1v |x=0 = u1r |x=0 = 0, see Fig. 8(a). The model proposed in this paper relax the condition
u1r |x=0 = 0. Instead of imposing no rotational displacement, a torsional stiffness is incorporated to the system, Fig. 8(b).

(a)

(b)

Figure 8: a) Clamped boundary condition usually applied in Finite Element Method, u1a |x=0 =
u1v |x=0 = u1r |x=0 = 0; b) u1r |x=0 6= 0, torsional stiffness kt .
The Finite Element Method was used to discretize a cantilever beam, clamped in one end
and free in the other. The Timoshenko beam element used has three degrees of freedom (axial,
vertical and rotational) at each node (Fig. 7). The variational form of strain energy is given by:
H =

Z
0

[u0a (EAu0a ) + u00v (EIu00v ) + u0r (GIu0r )]dx .

(1)

Where 0 is the derivative with respect to x (d/dx). The degrees of freedom are: the axial displacement (ua ); the vertical displacement (uv ); and the rotation (ur ). E is the elastic modulus,
G is the shearing modulus, I is the inertia momentum, L is the beam length and A is the cross
section area. The virtual work done by inertial forces (T ) and damping forces (D) are the
following:

{T, D} =

Z
0

{ua (A
ua ) + uv (A
uv ) + ur (I ur ), ua cu a uv cu v ur cu r }dx . (2)

The damping forces are taking into account as a Rayleigh damping proportional to the mass,
where c is the damping constant. The element matrices are the following ones:

h
(e)

[M ] = A
420

140 0
0
156
0
22h
70 0
0
54
0
13h

0
22h
4h2
0
13h
3h2

170
0
0
140
0
0

0
54
13h
0
156
22h

0
13h
3h2
0
22h
4h2

[C] =

c
[M ]
A

E110

E
[K (e) ] = 1+m

A(1 + m)/l
0
0
A(1 + m)/l
0
0

0
12I/l3
6I/l2
0
12I/l3
6I/l2

0
12I/l3
6I/l2
0
12I/l3
6I/l2

0
6I/l2
2I(1 m/2)/l
0
6I/l2
4I(1 + m/4)/l

0
6I/l2
4I(1 + m/4)/l
0
6I/l2
2I(1 m/2)/l

A(1 + m)/l
0
0
A(1 + m)/l
0
0

12
EI
Where m =
, ks is the shear correction factor and l is the element size.
2
h
GAks
The essential boundary conditions at x = 0 are given by ua |x=0 = 0 and uv |x=0 = 0. The
system is discretized using the Finite Element Method, and after assembling the matrices, one
can write:




+ [K]u(t) = f(t) .
[M ]
u(t) + [C]u(t)

(3)

Where [M ], [C] and [K] are the mass, damping and stiffness matrices, which are real and
positive-definite. The external force is represented by vector f(t) = (f1a , f1v , f1r , f2a , f2v , f2r , . . . , fnr )
and the displacements (u1a , u1v , u1r , u2a , u2v , u2r , . . . , unr ) are the components of the vector
u(t).
The data used for the numerical simulation were: c = 10 N/(m.s); E = 2e11 Pa; = 7850
kg/m3 ; = 0.3; G = 0.5E/(1 + ); ks = 5/6; L = 511 mm, b = 30.7mm; h = 3.04mm.
Figure 9 shows the beam modeled.

Figure 9. Beam modeled.


3.2

Natural frequencies

To calculate the natural frequencies without damping the following eigenvalue problem must
be solved:
(i2 [M ] + [K])i = 0 ,

(4)

where i is the i-th natural frequency, [M ] and [K] are the mass and stiffness matrices and
i is the vibration mode.
E110

To determine the number of elements needed, the error of the sixth natural frequency is
calculated (because the higher frequency obtained through experiment was the sixth). It was
necessary 32 elements to have an error lower than 0.1%.
The i-th natural frequency with damping is calculated in the following way:
q

di = i 1 i2 ,

(5)

where i is the damping rate of the i-th mode, which can be calculated using the normal
modes () and the damping matrix ([C]):
i =
4.

Ti [C]i
.
2i

(6)

IDENTIFICATION PROCEDURE

Since the torsional stiffness is not known, its value must be identified [8], this is an inverse
problem [9]. The cost function chosen to be minimized was the following:
6
X
(di (kt ) expi )2

,
(7)
2
expi
where expi is the i-th natural frequency obtained through experiment. The function fminsearch of MATLAB was used to solve the optimization problem. Table 3 shows the torsional
stiffness obtained for different number of layers.
min
kt

i=1

Table 3. Torsional stiffness results


N0 of Layers 0
kt (kN/m)
2.8
5.

1
1.2

2
1.2

3
1.1

4
1.1

6
1.1

8
1.0

10
0.9

COMPARISON BETWEEN NUMERICAL AND EXPERIMENTAL RESULTS

The first numerical simulation performed considered the t-clamped condition, u1a |x=0 = u1v |x=0
= u1r |x=0 = 0. The error between this result and the experiment with no rubber layers is shown
in Tab. 4, second column.
Even without the presence of rubber layers, the error between the simulation (t-clamped) and
the experiment is higher than 2.6%. Nevertheless, if the boundary condition is modeled with
a torsional spring, kt = 2.8e3N/m, the error drops considerably, as shown on Tab. 4, third
column.
Table 4. No rubber layer. Percent error for kt = 2.8e3N/m
0 layer
clamped condition
natural freq
error(%)
st
1
2.65
nd
2
2.40
3rd
1.73
th
4
1.50
5th
0.89
th
6
1.51

kt = 2.8e3N/m
error(%)
0.65
0.46
0.14
0.31
0.86
0.19
E110

Relaxing the condition ur = 0 and using a single extra parameter, kt , it is possible to obtain
excellent approximations. Either in an experiment or in any engineering application, it is not
simple to precisely assure the clamped boundary condition, even in this controlled test, in which
is noted that the clamped condition is not ideal. However, by simply adding a torsional stiffness,
the results become satisfactory. Mounting a rubber layer on the clamped side it is expected that
mount flexibility, therefore a smaller kt will be needed to match simulation with experimental
results.
The error between the numerical and experimental results is synthesized on Tab. 5. Depending on the number of rubber layers, a different kt needs to be fitted, so the best match between
the simulation and the experiment is achieved.

Table 5. Error between numerical and experimental results


Layers
kt (kN/m)
1st
2nd
3rd
4th
5th
6th

0
2.8
0.65
0.46
0.14
0.31
0.86
0.19

1
1.2
0.80
0.07
0.49
0.36
0.74
0.08

2
1.2
0.80
0.07
0.39
0.25
0.87
0.08

Error(%)
3
1.1
0.39
0.16
0.22
0.11
0.63
0.07

4
1.1
0.39
0.16
0.32
0.17
0.63
0.10

6
1.0
0.26
0.23
0.26
0.06
0.66
0.20

8
1.0
0.06
0.33
0.30
0.06
0.52
0.40

10
0.9
0.49
0.13
0.69
0.38
0.80
0.16

In order to attempt to establish a relation between the torsional stiffness kt and the number of
rubber layers at the mount, Fig. 13 shows how kt decreases with the number of layers.

Figure 10. Layer versus kt .


Except for the first point, which is the torsional stiffness for the boundary condition of no
rubber layers, the relationship between the torsional stiffness and the number of rubber layers
seems linear. Figure 11 shows the linear relationship between the number of layers and kt .

E110

Figure 11. Linear approximation for the relation layer kt .


With the relationship between the number of layers and kt in hands it is possible to simulate
a case for nlay = 20, for instance. Figure 12 shows a FRF for nlay = 0 and nlay = 20. It is
evident that the higher the number of rubber layers, the smaller the natural frequencies will be.

Figure 12. FRF, no layer versus twenty layers.


Figure 13 shows an extrapolation of the result for a large frequency band (0-2500 Hz), using
nlay = 0 and nlay = 1.

E110

Figure 13. FRF, from 0 to 2500 Hz.


To investigate a little more the proposed modeling, three different rubber materials were
tested: hard rubber, soft rubber, foam rubber. All the three cases were performed with one
layer. For the foam rubber the layer was mounted only in one surface of the beam. The results
are in Tab. 6.

Table 6: Percent error between the numerical and the experimental results for different materials
Natural Freq.
kt (kN/m)
1st
2nd
3rd
4th
5th
6th

Error(%)
hard rubber
1.3
1.30
0.15
0.48
0.47
0.88
0.65

soft rubber
1.1
0.52
0.36
0.44
0.34
0.56
0.21

foam rubber
1.1
0.64
0.29
0.66
0.27
0.31
0.29

The only error above 1% was in the first natural frequency of the hard rubber. But one
knows that the precision of the experiment is 0.25Hz, so if 1 equals 9.05, for instance (instead
of 9.00Hz), the error would be 0.08%. These results are again very motivating. The proposed
model works well even for different materials.
6.

CONCLUDING REMARKS

An alternative to model the clamped boundary condition of a cantilever beam was proposed
in this work. Only one parameter (torsional stiffness) was needed to well fit the numerical
results with the experimental ones. It is very difficult to find a perfect clamped condition in
experiments or in an engineering application, while relaxing the condition ur |x=0 = 0 (no
rotational displacement) and using a torsional spring (kt ) instead, seems to be a good idea to
model the clamped boundary condition of a cantilever beam. The numerical results shown
E110

very good agreement with the experimental results (errors lower than 1%). The relationship
between the number of rubber layers and kt is almost linear for the rubber used in the experience
herein. In this way, it is possible to use these results to change the dynamical characteristics of
a structure, avoiding unwanted frequencies.
ACKNOWLEDGMENTS
The authors acknowledge the financial support of CNPQ, CAPES, and FAPERJ.

REFERENCES
[1] J. N. Reddy. An Introduction to the Finite Element Method. McGraw-Hill, 2005.
[2] T. J. R. Hughes. The Finite Element Method - Linear Static and Dynamic Finite Element
Analysis. Prentice-Hall, Inc., Englewood Cliff, New Jersey, 1997.
[3] G.S. Aglietti and P.R. Cunnigham. Is a simple support really that simple? Journal of Sound
and VIbration, 257(2):321335, 2002.
[4] H. Berger, L. Barthe, and R. Ohayon. Parametric updating of a finite element model from
experimental modal characteristics. Mechanical Systems and Signal Processing, 4(3):233
242, 1990.
[5] Q. W. Zhang, C. C. Chang, and T. Y. P. Chang. Finite element model updating for structures
with parametric constraints. Earthquake Engineering and Structural Dynamics, 29(7):927
944, 2000.
[6] Jurij Avsec and Maks Oblak. Thermal vibrational analisys for simply supported beam and
clamped beam. Journal of Sound and Vibration, 2007.
[7] D. J. Ewins. Modal Testing: Theory and Practice. John Wiley & Sons Inc., 1984.
[8] G. Kerschen, K. Worden, A. Vakakis, and J. C. Golinval. Past, present and future of nonlinear system identification in structural dynamics. Mechanical Systems and Signal Processing, 20:505592, 2006.
[9] A. J. Silva Neto and F. D. Moura Neto. Problemas Inversos - Conceitos Fundamentais e
Aplicaes. Editora da Universidade do Estado do Rio de Janeiro, Brasil, 2005.

E110

Potrebbero piacerti anche