You are on page 1of 10

Measurement 102 (2017) 121–130

Contents lists available at ScienceDirect

Measurement
journal homepage: www.elsevier.com/locate/measurement

Comparison of experimental and operational modal analysis on a


laboratory test plate
Esben Orlowitz ⇑, Anders Brandt
Department of Technology and Innovation, University of Southern Denmark, Denmark

a r t i c l e i n f o a b s t r a c t

Article history: Operational modal analysis (OMA) is widely used whenever the dynamic characteristics of structures that
Received 10 November 2016 do not fit into a laboratory are desired. In addition, OMA offers a test of the structure under its real bound-
Received in revised form 27 January 2017 ary conditions which may sometimes be preferable for validation of numerical models. Theoretically, the
Accepted 1 February 2017
natural frequencies and damping ratios should be identically estimated by an OMA test and an experi-
Available online 4 February 2017
mental modal analysis (EMA) test. However it is still often reported that EMA tests are more reliable.
The present paper presents a thorough comparison of EMA and OMA tests of a Plexiglas plate. The exper-
Keywords:
iments were carefully designed, to ensure that the plate was tested under similar boundary conditions.
Operational modal analysis
Experimental modal analysis
Estimated modal parameters from the EMA test and OMA test are presented and compared, for the first
Damping estimation ten modes of the plate. It is found that natural frequencies are deviating by less than 0.3%, damping ratios
Experimental test by less than 7%, whereas cross-MAC values between the mode shapes of the two tests are found to be
above 0.99. The experimental test was conducted first by an EMA test, followed by the OMA test and
finally another EMA test was conducted in order to catch any time-variance. It is concluded that no sig-
nificant differences were found between modal parameters obtained by OMA and EMA.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction all inputs (excitation forces) are measured, which is practically


unfeasible for many structures.
As simulation of vibration responses based on numerical mod- The last couple of decades, OMA, by which only the structural
els are increasingly used to investigate dynamic behavior of struc- responses are used, has been widely applied and described in liter-
tures, experiments are increasingly required for validation or ature. OMA is attractive in many situations because it can be
correction of the numerical models. As an example, a dynamic applied to structures in operation, and does not require excitation,
model of a structure with incorrect damping assumptions could which is practical for many large structures, see for example
give misleading life time estimations [1]. Because of the complex- [25,19,24,4,27]. In addition, OMA can sometimes be preferable
ity of damping effects, it is usually not possible to give a proper for validation of numerical models because the effects of boundary
analytical estimate, hence experimental tests are needed. How- conditions are included in the results, as opposed to EMA where
ever, damping can be a notoriously difficult parameter to consis- the structure is usually tested under free-free conditions. OMA
tently estimate from experiments, and the accuracy of estimates methods are based on some assumptions about the nature of the
of damping from, particularly, operational modal analysis (OMA) loads exciting the structure; the loads being the result of random
is currently receiving a lot of research interest. processes, that may be colored as long as the poles of the force col-
Experimental modal analysis (EMA) has been used for decades oring are different from those of the structure. In addition the
to extract information about the structural damping. EMA is con- structure should be linear and time-invariant [2–4].
sidered reliable because it is based on input-output system identi- Theoretically, modal parameters should be identically esti-
fication, which allows validation e.g. of the estimated frequency mated via an OMA test and a classical EMA test [5]. However, it
response functions (FRFs) by coherence functions. However, the has still sometimes been reported that an EMA test is more reliable
strength of EMA is also limiting its applicability as it requires that because of the available information and controlled environment
[6]. Furthermore it has also been reported that the damping esti-
mates from OMA are over estimated, hence suggesting a critical
bias error [7,8].
⇑ Corresponding author.
E-mail address: esbenorlowitz@gmail.com (E. Orlowitz).

http://dx.doi.org/10.1016/j.measurement.2017.02.001
0263-2241/Ó 2017 Elsevier Ltd. All rights reserved.
122 E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130

Little comparison between EMA and OMA has been published IES-Plate proposed in [12], although, for practical reasons, the plate
where exactly the same structure and experimental setup have thickness was chosen slightly different from the IES-Plate, by
been used for both cases. In [9] however, a comparison of EMA choosing 20 mm thickness, which is a standard thickness in
and OMA was presented for some case studies of structures with Europe.
simple to more complex geometries and it was shown that for sim- The measurement grid consisted of 35 out-of-plane DOFs dis-
ple geometries the differences of the extracted modes were small. tributed uniformly over the plate as shown in Fig. 2 and responses
However, difficulties to extract all modes from OMA tests were from all DOFs were measured simultaneously. The accelerometers
reported and in addition it was reported that choosing the loca- and cables significantly mass load the structure, and it is also likely
tions of the excitations in an OMA laboratory test is non-trivial. that there is additional damping added. This is, however, irrele-
For a simple geometry with random spatial excitation, problems vant, as the present experiments are designed so that it is the plate
with finding all modes from the OMA test were reported, and it together with the instrumentation that is investigated. The condi-
was found that the mode shapes from the two tests were not iden- tions for the EMA and OMA tests should therefore be equal, albeit
tical. The problem of undiscovered modes increased when the exci- not being those of the free-free plate.
tation was localized to a single point for the OMA test. The following data acquisition (DAQ) and sensor equipment
For a more complex structure also investigated in [9], a consid- were used:
erable number of modes were not well estimated or not estimated
at all. Damping ratios were not reported and mode shapes were in  3 National Instrument 4497, 16 channel, 24bit analog inputs
general showing poor similarity between the EMA and OMA tests. cards.
However, it should be noted that only a single reference DOF was  35 Dytran 3097A2 accelerometers, 100 mV/g, IEPE, 4.3 grams.
used for the calculation of spectra and the measurement time  1 Dytran 5800B4 impulse hammer, 10 mV/N, IEPE, with plastic
was not reported. Both the number of reference channels and the tip.
measurement time are crucial for OMA [6,10].  In-house software for DAQ control, based on MATLAB DAQ-
In [11] a comparison of EMA (using impact testing) and OMA toolbox.
was carried out, which showed large deviations in the estimated
damping ratios. However, the methods used for parameter estima- The plate was suspended by springs via a thin fishing line, see
tion for OMA and EMA were not the same and thus the applied set- Fig. 1, giving it relatively low rigid body modes (<3 Hz). For all tests
tings for estimation are not comparable. The authors of [11] a sampling frequency of 5 kHz was used.
pointed out that the difference in vibration levels between the For the EMA test, frequency response functions (FRFs) were
two tests could be the reason for the large deviations of the damp- measured by impact testing. FRFs relative to two different excita-
ing ratios. tion DOFs were measured, in order to have multi-reference data
In [8] a comparison of EMA and OMA was performed on the for the parameter extraction. DOFs 1 and 29 were thus excited,
same data set, acquired from excitation with two shakers and rov- and time sequences of 20 impacts with 20 s in-between each
ing accelerometers over a test structure. For OMA the measured impact were acquired for each reference point.
input was simply ignored, using only the responses. It was For the OMA test, 300 s of data were acquired from all DOFs
observed that the damping ratios estimated by OMA were higher simultaneously, corresponding to approx. 40000 periods of the
than the ones from EMA and the same was observed for the natural lowest natural frequency. This rather long measurement time
frequencies. However, in this study the dynamical characteristics was chosen to reduce possible effects on modal parameters by lim-
of the shakers seems too have been included in the OMA, but not ited measurement time, see e.g. [10,4]. The excitation was applied
in the EMA, results. For EMA the influence of the shakers is by gently tapping the tip of a pencil randomly around the plate.
excluded by using frequency responses relative to the inputs, but The idea of tapping rather than scratching was chosen in order to
this is not the case for OMA. Thus the boundary conditions were minimize the interaction with the plate as much as possible, as it
very different in the OMA test compared to the EMA test, which was initially experienced that scratching the plate influenced its
explains the higher damping in the OMA estimates. dynamics in a significant way, both on natural frequencies and
It is worth noticing that one of the attractive features of OMA is damping ratios.
that it indeed includes the boundary conditions. This means, how- Two impact tests were performed; one test was performed
ever, that in order to compare EMA and OMA, great care must be before and one after the OMA test, in order to be able to detect
taken in the design of the experiment to ensure that the boundary any time-variance. The impact test before the OMA test is referred
conditions are identical in both tests. to as EMA1 and the one after is referred to as EMA2 in the following.
The present paper presents an experimental study of EMA and All three tests were conducted consecutively within approximately
OMA tests on a Plexiglas plate where the EMA and OMA tests have one hour, under which no interaction with the experimental setup
been carefully designed so that they test the structure under very (except the excitation) was allowed. Between the end of the EMA1
similar boundary conditions. The plate, which is a simple geome- test and the start of the OMA test, 15 min elapsed, and from the
try, is designed to have closely spaced modes and is considered a end of the OMA test to the start of EMA2, 5 min elapsed, all due
good structure for comparison. Modal parameters are estimated to initial quality checks. Each of the impact tests took approxi-
and compared between the EMA and OMA tests. mately 15 min from start to end.
Damping ratio estimates are the primary focus in the present
work as they are known to be the most challenging, although both
natural frequencies and modes shapes are also compared for 3. Data processing and analysis
completeness.
3.1. Impact tests

2. Experimental setup The acquired time sequences from the impact test were pro-
cessed to obtain FRFs. FRFs and coherence functions were esti-
The Plexiglass (PMMA) plate that has been experimentally mated by averaging the spectra from two selected impacts of the
tested can be seen in Fig. 1. The dimensions of the plate are time sequences (the input and all the responses) using the H1 esti-
533 mm  321 mm  20 mm. The plate is similar to the so-called mator in the MATLAB toolbox ABRAVIBE [13]. The two impacts
E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130 123

Fig. 1. Experimental setup. Plexiglas plate suspended vertically in two springs with 35 accelerometers mounted. In the lower right corner the impact hammer and pencil used
for excitation in the EMA and OMA measurements, respectively, can be seen.

EMA2 test, respectively. For EMA1 the FRFs and coherence func-
tions between all response locations and the references (70 in
total) are shown in Fig. 4 and similar results were found for EMA2.
As can be seen in Fig. 4(b), the coherence functions are very
close to unity at frequencies above, say, 100 Hz, which is well
below the first natural frequency. Above this frequency, no signif-
icant dips in the coherence functions are present.
Reciprocity plots for EMA1 are shown in Fig. 5 and are also indi-
cating good measurements; the slight differences that can be
observed may be caused by impacts chosen not being exactly in
the point where the accelerometer in the excitation points are
located, or slight sensitivity differences between the accelerom-
eters used. Neither of these causes of the differences will affect the
pole estimates, but only the mode shapes, and the reciprocity is
Fig. 2. Measurement grid on the plate consisting of 35 measurement points with a
therefore acceptable in the present work where the primary
separation of 75  85 mm. The full dimensions of the plate are 533  321 mm and parameter of interest is damping. A similar examination of the
thickness 20 mm. reciprocity of the EMA2 test showed similar results. The reciprocity
plot is one of the validation tools that are available when input and
output data are measured in EMA applications.

producing a FRF estimate with optimum coherence were selected


by an automatic process as described in [14]. 3.2. OMA test
The blocksize for the estimation of FRFs was 64 k. This blocksize
was selected by a comparison of FRF estimates using increasing For the OMA test, correlation functions (CFs) were estimated as
blocksizes, and selecting the FRF estimate for which a further substitutes for impulse responses. A requirement for estimating
increase in blocksize does not have higher peaks of the FRFs. This CFs is that the responses are stationary. The reverse arrangements
procedure ensures that bias errors in the FRF estimates are negligi- test was therefore applied on the rms-levels of frames of the
ble. For noise suppression, a force window of 1% of the blocksize responses, to test for (wide sense) stationarity [15,14]. A signifi-
was applied to the force signal and an exponential window cance level of 2% and 100 frames were used as recommended in
decreasing to 0.001% at its end was applied to all signals. The [16], and all responses passed the reverse arrangements test, indi-
effects on the damping estimates due to the exponential window cating that the signals are likely to be stationary.
were removed after the parameter extraction as described in [14]. The rms-levels of the responses ranged from 113 to 339 mm/s2
The peak force levels of the force impacts were approximately with a median of 207 mm/s2. Maximum peak values for the
15 N. In Fig. 3, a comparison of two FRFs from two different pairs responses ranged from 6 to 26 m/s2 which is significantly lower
of impacts of different force levels are shown. The similarities than the first part of the transient responses from the impact tests,
between the FRFs indicate that linearity can be assumed within which ranged between 50 and 150 m/s2. It is not possible to com-
the current force range. pare the transient decaying responses of the impact tests directly
Linearity is a fundamental assumption for both EMA and OMA with the continuous responses from the OMA test, but as linearity
and as the excitations of the two tests are different in nature, has already been verified, as described above, this should not be an
any amplitude dependent damping would make the comparison issue.
of EMA and OMA invalid. In Fig. 6 two typical power spectral densities (PSDs) are shown
As stated earlier, two impact tests were conducted, one before from which it can be observed that the dynamic range of the mea-
the OMA test and one immediately after; the EMA1 test and surements is sufficient.
124 E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130

2
10

FRF [(m/s )/N]


1
10

2
0
10

-1
10
0 200 400 600 800 1000
Force spec. [Ns]

10 0

10 -2

10 -4
0 200 400 600 800 1000

Frequency [Hz]

Fig. 3. Comparison of frequency response function (upper) from different force levels, see the force spectra (lower).
FRF [(m/s 2 )/N]

0
10

0 200 400 600 800 1000


Frequency [Hz]
(a) Frequency response functions
1
Coherence

0.5

0
0 200 400 600 800 1000
Frequency [Hz]
(b) Coherence functions
Fig. 4. Frequency response functions (a) and coherence functions (b) from EMA1, the first impact test. In total 70 functions are plotted in both graphs, which shows that the
coherence functions are very good above 100 Hz, which is well below the first natural frequency. Thus the FRF estimates are very good in the frequency range of interest.

The CFs were computed as unbiased estimates by the peri- where m is the discrete lags corresponding to time m=f s , and L is the
odogram approach, see [17]. Thus, the discrete Fourier transforms length of the time sequences xðnÞ and yðnÞ sampled by time incre-
of zero padded sequences are computed, to avoid cyclic effects of ment Dt. Finally, the negative lags of the correlation functions are
the DFT. To obtain, for example, the cross-correlation between discarded.
two responses y1 and y2 , with the first signal as the reference,
the DFT of y2 is multiplied by the complex conjugate of the DFT 3.3. Parameter extraction
of y1 in the frequency domain, whereafter the inverse DFT of this
product produces the biased correlation function. The unbiased The estimated impulse response functions and correlation func-
estimate is then formed by proper weighting so that the correla- tions were processed by the multi-reference Ibrahim time domain
tion function between signal y1 ðtÞ and y2 ðtÞ becomes (MITD) method for modal parameter extraction [18]. The MITD
X
Dt Lm1 method is in principle equivalent to the more well known stochas-
R21 ðmÞ ¼ y ðnÞy1 ðn  mÞ ð1Þ tic subspace identification (SSI) method [19]. In the MITD method,
L  m n¼0 2
a block Hankel matrix of impulse responses or correlation func-
E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130 125

2
10

1
10

FRF [(m/s )/N]


0
2 10

-1
10

-2
10
In=1, Out=29
In=29, Out=1
-3
10
0 200 400 600 800 1000
Frequency [Hz]
Fig. 5. Reciprocity plot for the EMA1 test.

10 -3

DOF 1-1
Auto and Cross PSD [(m/s ) /Hz]

DOF 1-29
-4
10
2 2

10 -5

10 -6

-7
10

10 -8
0 200 400 600 800 1000
Frequency [Hz]
Fig. 6. Typical (auto and cross) power spectral densities for DOF 1–1 and 1–29. A blocksize of 2048 (sampling frequency 5k), Hanning window and an overlap of 50% has been
used, resulting in 1463 averages.

tions is constructed, which is decomposed by singular value The FRFs from the impact tests were converted to impulse
decomposition (SVD). From the SVD compressed block Hankel response functions (IRFs) before parameter estimation by the MITD
matrix, an eigenvalue problem can be established for different sys- method. The first 5 samples of the IRFs were neglected as these
tem order (number of poles), depending on the number of singular samples are contaminated by noise on the responses [20]. The mul-
values used. A detailed description of the method can be found in tivariate mode indicator function [21] of the EMA1 data are shown
[18]. in Fig. 7 together with the stabilization diagram from the MITD
The modal parameter estimates (poles and eigenvectors) from method showing the estimated poles at increasing model orders
different levels of compression by the SVD, depending on the sys- up to 100, a similar stabilization diagram was obtained for EMA2.
tem order, were plotted in a stabilization diagram, sometimes As can be observed from the mode indicator function there are
referred to as a consistency diagram, as a tool to help finding the two closely spaced modes around 140 Hz, which is the reason for
reliable physical modes. The natural frequencies and modal damp- using two references in the modal analysis.
ing ratios were then calculated from the stable poles selected from For both EMA and OMA, 200 samples of the impulse response/-
the stabilization diagram. The mode shapes were calculated from correlation functions were used for the estimation of modes, corre-
the eigenvectors corresponding to the selected poles by expansion sponding to 40 ms, or approximately 5.5 periods of the lowest
via the left singular vectors from the SVD compression, producing natural frequency of the plate. This amount was found to give
unscaled mode shapes. the best stabilization of the first ten estimated modes.
126 E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130

80

70

60

Number of Poles
50

40

30

20

10

0
0 200 400 600 800 1000
Frequency [Hz]

Fig. 7. Stabilization diagram of impact test data from EMA1 with overlaid multivariate mode indicator function. Crosses and circles indicate stable and unstable poles,
respectively.

In Fig. 8 the stabilization diagram from the OMA test is shown cies for all three experiments are presented in Table 1 and, corre-
together with an averaged PSD function as a simple mode indicator spondingly, damping ratios are presented in Table 2. The auto-
function. MAC matrix of the averaged mode shapes are shown in Fig. 9 for
the impact test EMA1 and the OMA test; a similar auto-MAC matrix
4. Experimental results was obtained for EMA2 but is not shown.
From Table 1 it can be observed that the natural frequencies
Modal parameters were extracted for the first 10 modes decrease slightly between the first test (EMA1) and the last
(<1000 Hz) by the method described in Section 3, in the following (EMA2). A comparison of the differences of the results of the impact
manner. All modes fulfilling the stabilization criteria of deviations tests with respect to the OMA test are presented in Table 3 and
of 0.1% for natural frequencies, 5% for damping ratios and having graphically presented in Fig. 10. It can be observed that the rate
modal assurance criterion (MAC) values exceeding 0.9 were of decrease in natural frequencies is smaller between the last
extracted from the stabilization diagrams. Also the mode shapes two tests (OMA and EMA2) than between the first two tests
of the selected modes were averaged. This resulted in at least 26 (EMA1 and OMA).
pole estimates of each mode from different model orders, which This could suggest that the decrease is due to some environ-
were averaged and the corresponding scatter calculated (calcu- mental change in the surroundings, possibly temperature increas-
lated as the normalized random error er , i.e., the standard deviation ing in the laboratory during the measurements, causing the
divided by the average value). The resulting mean natural frequen- stiffness of the plate to decrease. It should also be noted, however,

80

70

60
Number of Poles

50

40

30

20

10

0
0 200 400 600 800 1000

Frequency [Hz]

Fig. 8. Stabilization diagram from the OMA test. Crosses and circles indicate stable and unstable modes, respectively, and the background curve is the power spectral density
of the response of DOF 1.
E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130 127

Table 1
Estimated (averaged) natural frequencies, f n , and corresponding normalized random errors, er , for all three experimental tests. The random errors are calculated based on at least
26 extracted modes.

EMA1 OMA EMA2

f n [Hz] er ðf n Þ [%] f n [Hz] er ðf n Þ [%] f n [Hz] er ðf n Þ [%]


Mode 1 142.0 0.008 141.9 0.019 141.8 0.015
Mode 2 142.8 0.004 142.6 0.011 142.5 0.003
Mode 3 323.6 0.055 323.3 0.005 323.1 0.031
Mode 4 393.4 0.007 393.0 0.001 392.7 0.009
Mode 5 409.1 0.008 408.8 0.003 408.5 0.002
Mode 6 504.5 0.021 503.8 0.005 503.7 0.017
Mode 7 584.2 0.038 583.6 0.007 583.3 0.040
Mode 8 713.1 0.010 711.3 0.003 712.2 0.004
Mode 9 800.9 0.004 798.8 0.002 799.9 0.008
Mode 10 941.6 0.014 940.3 0.010 940.5 0.032

Table 2
Estimated (averaged) damping ratios, fn , and corresponding normalized random errors, er , for all three experimental tests. The random errors are calculated based on at least 26
extracted modes.

EMA1 OMA EMA2


fn er ðfn Þ [%] fn er ðfn Þ [%] fn er ðfn Þ [%]
Mode 1 3.28 0.242 3.25 0.529 3.34 2.448
Mode 2 3.09 0.279 3.06 0.297 3.13 0.370
Mode 3 2.68 2.114 2.68 0.206 2.70 1.134
Mode 4 2.56 0.305 2.54 0.151 2.61 0.632
Mode 5 2.47 0.443 2.45 0.252 2.50 0.322
Mode 6 2.48 0.420 2.42 0.277 2.52 0.290
Mode 7 2.46 2.210 2.38 0.441 2.53 2.522
Mode 8 2.38 0.560 2.41 0.085 2.42 0.578
Mode 9 2.27 0.367 2.28 0.373 2.33 0.442
Mode 10 2.44 0.793 2.57 0.544 2.53 0.841

1 1

1 0.8 1 0.8

0.5 0.6 0.5 0.6

0.4 0.4
0 0
12 12
34 0.2 34 0.2
10 10
56
7 89 56
7 89
78 56 78 56
9 34 9 34
10 12 10 12
Mode # Mode #

(a) EMA1 (b) OMA

Fig. 9. Auto MAC matrices from the averaged mode shapes of the EMA1 and OMA tests. Similar results were obtained for EMA2.

Table 3
Difference of the results of the impact tests (EMA1 and EMA2) with respect to the OMA test.

Difference of: EMA1 wrt. OMA EMA2 wrt. OMA

f n [%] fn [%] Cross-MAC f n [%] fn [%] Cross-MAC

Mode 1 0.101 1.089 0.999 0.059 2.903 0.999


Mode 2 0.136 0.847 0.999 0.043 2.176 0.999
Mode 3 0.119 0.058 1.000 0.047 0.696 1.000
Mode 4 0.079 0.708 0.997 0.082 2.742 0.997
Mode 5 0.071 0.933 1.000 0.078 2.061 1.000
Mode 6 0.138 2.588 1.000 0.012 4.390 1.000
Mode 7 0.105 3.144 1.000 0.054 6.286 1.000
Mode 8 0.248 1.275 0.994 0.117 0.500 0.994
Mode 9 0.274 0.312 0.992 0.148 2.259 0.992
Mode 10 0.136 5.075 0.997 0.020 1.674 0.997

that the changes are very small; less than 0.3%. As stated in Sec- the smaller decrease of natural frequencies between the latter
tion 2, the time elapsed between EMA1 and OMA was larger than tests. A divergence from this can be observed for modes 8 and 9
the time between OMA and EMA2 which is in compliance with from the OMA test, see Table 1, that cannot be explained only by
128 E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130

Frequency wrt. OMA

Relative deviation [%]


0.2

-0.2 EMA1
EMA2
-0.4
1 2 3 4 5 6 7 8 9 10
Mode #
Damping wrt. OMA
Relative deviation [%]

10

-10
1 2 3 4 5 6 7 8 9 10
Mode #

Fig. 10. Relative deviations of EMA1 and EMA2 results with respect to the OMA results.

the observed scatter. No obvious reason has been found for this averaged together. For example, if the estimated damping is plot-
divergence, but all errors are small. ted as a function of model order, as presented for modes 3 and
The damping estimates, see Table 3 and Fig. 10, show larger 10 in Fig. 11, it can be observed that the averaging over increasing
deviations between the three tests than for the natural frequencies, model order influences the results in some cases, e.g. for mode 3 as
which is expected, since the inaccuracy of damping estimates are seen in Fig. 11(a). For mode 10, however, as shown in Fig. 11(b), the
well known to be larger than of frequency estimates. The results observed deviation is more consistent as a function of model order,
of the two EMA tests are, however, in better agreement with each but as stated earlier there is a difference in the OMA and EMA tests
other than with the OMA results, although the estimated damping due to the excitation which could influence the observed deviation
ratios are found consistently larger for EMA2 compared to EMA1. between tests.
No consistency is found in the comparison of the damping esti- In addition to the influence of the model order at which the
mates from the OMA test and the EMA tests and again from a prac- modes are estimated, as discussed above, other parameters in the
tical point of view the observed errors are small. estimation process may also be of importance. To make the com-
The observed errors are influenced by the choice of evaluation parison between the OMA and EMA results as independent as pos-
of the results in the present work, where all stabilized modes are sible of any such differences, the analysis was made in this study

0.029
Damping ratio [%]

0.028

0.027

0.026
10 20 30 40 50 60 70 80
Model order
(a) Mode3
0.026
Damping ratio [%]

0.025

0.024
40 50 60 70 80
Model order
(b) Mode10
Fig. 11. Damping ratio estimates depending on model order in the estimation for mode 3 and mode 10. Comparing the estimates from for OMA ‘+’ and EMA1 ‘’.
E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130 129

with the same parameter estimation algorithm, which should min- Comparison of the results from the OMA test with the EMA tests
imize any differences in parameter estimates due to uncertainties. showed overall very good agreement. The natural frequencies
It should further be considered that the relatively long measure- found by the OMA test were consistently lower than the ones of
ment time used for the present OMA analysis, results in small ran- the first EMA test, with deviations less than 0.3%, and deviated
dom errors in the correlation functions estimates, which in turn from the later EMA test with less than 0.2%. All cross-MAC values
results in less uncertainties in the modal parameter estimates. between the OMA test and the EMA tests were above 0.99, thus
The impulse responses used for the EMA parameter estimation showing close similarity of the mode shapes. The estimated damp-
have similarly small random errors due to the high coherence func- ing ratios between the EMA tests and the OMA test were less than
tion values. A thorough study of the influence of various parameter 7% and typically around 2%. The estimated damping ratios from the
estimation choices can be found in e.g. [6,26]. OMA test were both higher and lower than the EMA estimates,
hence no systematic error was observed.
5. Discussion The results of the present work strongly indicate that there is no
significant difference in estimated modal parameters obtained by
From Table 2 it is clear that the differences in damping esti- OMA and EMA provided that the conditions of the tests are compa-
mates between the three tests are insignificant, which is in agree- rable. We would like to emphasize, however, that OMA tests in
ment with the findings of [6,22]. From the tests in Tables 1 and 2, it most engineering applications are not comparable to typical EMA
can also be observed that the random errors for both natural fre- tests. During a typical OMA test, the structure has different bound-
quencies and damping ratios are comparable between the OMA ary conditions than the typical free-free conditions of an EMA test.
test and the EMA tests.Thus, the results support the hypothesis Therefore, it can be expected that OMA results in many (or even
that damping ratios estimated by OMA are as reliable as those esti- most) engineering applications will show higher damping values
mated by EMA. than a free-free EMA test. But this result is not due to differences
The possibility of different response levels and non-linear between the signal processing and modal parameter extraction,
damping could be a reason for the small deviations in the damping but to the actual boundary conditions of the tests.
estimates from the experimental EMA and OMA tests, but it was
shown in Section 3.1 that, at least within a given force range of
References
impacts, no indications of non-linear behavior was observed. This
strongly suggests that the small differences found here between [1] A. Linderholt, Y. Chen, E. Orlowitz, A. Brandt, A study of the coupling between
EMA and OMA damping estimates can be explained by the slight test data accuracy and life prediction, in: Proceedings of Conference on Noise
and Vibration Engineering, Leuven, 2014, pp. 415–428.
differences in boundary conditions between the tests, as the tap-
[2] S.R. Ibrahim, R. Brincker, J.C. Asmussen, Modal parameter identification from
ping of the plate during the OMA test is certainly not identical to responses of general unknown random inputs, in: Proceedings of XIV
the impact used for the EMA test. International Modal Analysis Conference, Dearborn, Michigan, 1996, pp.
The setup used in the present work where impact testing was 446–452.
[3] W. Weijtjens, J. Lataire, C. Devriendt, P. Guillaume, Dealing with periodical
used has the advantage that no interaction with the plate is done, loads and harmonics in operational modal analysis using time-varying
except the excitation. In further work acoustic excitation could transmissibility functions, Mech. Syst. Signal Process. 49 (12) (2014) 154–164.
perhaps be an opportunity to decrease the interaction for OMA [4] R. Brincker, C. Ventura, Introduction to Operational Modal Analysis, Wiley,
2015.
excitation in the laboratory, see e.g. [23]. However, the best excita- [5] G. James, T. Carne, J. Lauffer, The natural excitation technique (next) for modal
tion for OMA is a load distributed over the structure, which is elab- parameter extraction from operating structures, Int. J. Anal. Exp. Modal Anal.
orate to implement for acoustic excitation. 10 (1995) 260–277.
[6] F. Magalhães, .A. Cunha, E. Caetano, R. Brincker, Damping estimation using free
Finally, it is noted that the cross-MAC values in Table 3 show decays and ambient vibration tests, Mech. Syst. Signal Process. 24 (5) (2010)
that the mode shapes of the OMA and EMA tests, are very similar. 1274–1290, Special Issue: Operational Modal Analysis.
Thus, for the present experiments where the boundary conditions [7] P. Avitabile, Modal space: someone told me that operating modal analysis
produces better results and that damping is much more realistic, Exp. Tech. 30
were very similar between the tests, there is no indication that the (2006) 25–26.
OMA and EMA methods would result in any different modal [8] T. Lauwagie, R.V. Assche, J.V. der Straeten, W. Heylen, A comparison of
parameters. experimental, operational, and combined experimental-operational parameter
estimation techniques, in: Proceedings of Conference on Noise and Vibration
Engineering, Leuven, 2006.
6. Conclusions [9] L. Thibault, T. Marinone, P. Avitabile, C. Van Karsen, Comparison of modal
parameters estimated from operational and experimental modal analysis
The present paper presents a comparison of an Experimental approaches, in: Proceedings of International Modal analysis XXX, Jacksonville,
Florida, 2012.
Modal Analysis (EMA) test and an Operational Modal Analysis [10] M. Ozbek, D.J. Rixen, A new analysis methodology for estimating the
(OMA) test on a Plexiglas plate. The experimental setup was care- eigenfrequencies of systems with high modal damping, J. Sound Vib. 361
fully designed so that both tests would test the plate under very (2016) 290–306.
[11] D.F. Giraldo, W. Song, S.J. Dyke, J.M. Caicedo, Modal identification through
similar boundary conditions, in order to make a proper comparison ambient vibration: comparative study, J. Eng. Mech. 135 (2009) 759–770.
of modal parameters extracted by EMA and OMA possible. The [12] D. Smallwood, D. Gregory, A rectangular plate is proposed as an ies modal test
EMA test applied in the present work was based on impact testing structure, in: Proceedings of International Modal analysis IV, Los Angeles,
California, 1986.
and in the OMA test the excitation was applied by gently tapping [13] A. Brandt, ABRAVIBE – a MATLAB toolbox for noise and vibration analysis and
the tip of a pencil randomly around the plate. teaching, Department of Technology and Innovation, University of Southern
Two impact tests were performed, one test was performed Denmark, 2011, July http://www.abravibe.com.
[14] A. Brandt, Noise and Vibration Analysis – Signal Analysis and Experimental
before and one after the OMA test, to be able to detect any time- Procedures, Wiley, 2011.
variance. In addition, the three tests were conducted within a nar- [15] J. Bendat, A.G. Piersol, Random Data: Analysis and Measurement Procedures,
row time frame, to ensure no significant environmental changes Wiley Interscience, 1986.
[16] H. Himelblau, J.W. Wise, A.G. Piersol, M.R. Grundvig, Handbook for Dynamic
would occur. The two EMA tests, before and after the OMA test,
Data Acquisition and Analysis – IES Recommended Practices 012.1, Institute of
gave similar results with deviations between the modal parame- Environmental Sciences, 1994.
ters obtained from the two tests of less than 0.2% for natural fre- [17] E. Orlowitz, A. Brandt, Influence of correlation estimation methods on damping
quencies, less than 4% for the damping ratios of the first ten estimates, in: Proceedings of 5th International Operational Modal Analysis
Conference, Guimaraes, Portugal, 2013.
modes, and cross-MAC values between the mode shapes larger [18] R. Allemang, D. Brown, Experimental modal analysis and dynamic component
than 0.98. synthesis – vol 3: Modal parameter estimation, Tech. rep., USAF, 1987.
130 E. Orlowitz, A. Brandt / Measurement 102 (2017) 121–130

[19] B. Peeters, G. De Roeck, Reference-based stochastic subspace identification for 29th, Los Angeles, California, Springer, New York, New York, NY, 2011, pp.
output-only modal analysis, Mech. Syst. Signal Process. 13 (1999) 855–878. 359–374.
[20] H. Vold, K. Napolitano, D. Hensley, M. Richardson, Aliasing in modal parameter [24] C. Rainieri, G. Fabbrocino, Operational Modal Analysis of Civil Engineering
estimation, SOUND&VIBRATION, 2008. Structures, Springer, 2014.
[21] R. Williams, J. Crowley, H. Vold, The multivariate mode indicator function in [25] G.H. James, T.G. Carne, J.P. Lauffer, A.R. Nord, Modal testing using natural
modal analysis, in: Proceedings of the III International Modal Analysis excitation, in: Proceedings of International Modal analysis X, San Diego, 1992.
Conference, Orlando, 1985. [26] C. Rainieri, G. Fabbrocino, Development and validation of an automated
[22] E. Orlowitz, A. Brandt, Producing simulated time data for operational modal operational modal analysis algorithm for vibration-based monitoring and
analysis, in: Proceedings of International Modal analysis XXXIII, Orlando, tensile load estimation, Mech. Syst. Signal Process. (2015) 512–534.
Florida, 2015. [27] C. Rainieri, A. Dey, G. Fabbrocinoa, F.S. Magistrisa, Interpretation of the
[23] Y.F. Xu, W.D. Zhu, Operational modal analysis of a rectangular plate using experimentally measured dynamic response of an embedded retaining wall by
noncontact acoustic excitation, in: Proceedings of International Modal analysis finite element models, Measurement (2016).

You might also like