Stochastic System Identification For Operational Modal Analysis: A Review
Stochastic System Identification For Operational Modal Analysis: A Review
Stochastic System Identification For Operational Modal Analysis: A Review
LMS International,
Interleuvenlaan 68,
B-3001 Leuven,
Belgium
e-mail: [email protected]
Guido De Roeck
Department of Civil Engineering,
K.U. Leuven,
Kasteelpark Arenberg 40,
B-3001 Leuven,
Belgium
e-mail: [email protected]
Introduction
(1)
n 2 n 2
where M ,C 2 ,KR
are the mass, damping and stiffness matrices; q(t)Rn 2 is the displacement vector at continuous time t.
A dot over a time function denotes the derivative with respect to
time. The vector f (t)Rn 2 is the excitation force. For systems
with distributed parameters e.g., civil engineering structures, Eq.
1 is obtained as the Finite Element FE approximation of the
system with only n 2 degrees of freedom DOFs left. Although the
nearly physical model 1 is a good representation of a vibrating
structure, it is not directly useful in an experimental modelling
context. First, it is not possible and also not necessary to measure all DOFs of the FE model. Second, this equation is in
continuous-time, whereas measurements are available as discrete
time samples. And finally, there is some noise modelling needed:
there may be other unknown excitation sources apart from f (t)
and measurement noise is always present in real life.
Stochastic State-Space Model. It can be shown that, by applying model reduction, sampling and modelling the noise, Eq. 1
can be converted to following discrete-time stochastic state-space
model see for instance 7,11 for a detailed derivation:
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
x k1 Ax k w k
(2)
y k Cx k k
where y k R is the sampled output vector the measurements;
x k Rn is the discrete state vector; w k Rn is the process noise,
typically due to disturbances and modelling inaccuracies, but here
also due to the unknown excitation of the structure; k Rl is the
measurement noise, typically due to sensor inaccuracy, but here
also due to the unknown excitation of the structure; k is the time
instant; l is the number of outputs; n is the system order (n
2n 2 ). The matrix ARnn is the state transition matrix that
completely characterizes the dynamics of the system by its eigenvalues; and CRln is the output matrix that specifies how the
internal states are transformed to the outside world. The noise
vectors are both unmeasurable vector signals assumed to be zero
mean, white and with covariance matrices:
l
Q
wp
w Tq Tq T
p
S
S
R
pq
(3)
i , i* i i j 1 i2 i
(5)
(6)
ARMA Model. The more classical system identification methods 10 identify models that do not contain the state. It can be
shown 12,13 that the following so-called ARMA model is
equivalent to the stochastic state-space model 2:
y k 1 y k1 p y k p e k 1 e k1 p e kp
p1
0
0
I
1
V
V
V d
V d
d
V dp2
V dp2
V dp1
V dp1
(8)
R 0 E y k y Tk
GE x k1 y Tk ,
(10)
S y s
i1
1
l T R u
s i i i
i1
1
l T
s * i i i
s j
(11)
white noise inputs; R u Rmm is the white noise input covariance matrix.
(7)
S y z C zIA 1 GR 0 G T z 1 IA T 1 C T ze j t (9)
nn
i e i t ;
Frequency-Domain Model. Many identification methods identify frequency-domain models from samples of the Fourier transform of the measurement signals. Frequency-domain models are
readily obtained by applying the z-transform to the discrete-time
models 2, 7. For instance, the output power spectrum matrix of
a state-space system can be written as 14:
(4)
The observed mode shapes V are the first l rows of the eigenvector matrix. As in the state-space case, the eigenfrequencies and
damping ratios can be computed from the discrete eigenvalues in
d ; see Eq. 5.
R i E y ki y Tk lim
N
1
N
N1
k0
y ki y Tk
(12)
where the second equation follows from the ergodicity assumption. Of course, in reality, a finite number N of data is available
and a covariance estimate is simply obtained by dropping the limit
in 12.
The identification methods of Section 3 require a frequency
domain representation of the output signals. The frequencydomain representation of stochastic signals is provided by the
power spectrum S y Cll , defined as the discrete-time Fourier
transform of the covariance sequence:
S y e
j t
R k e j kt
(13)
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
(14)
(15)
where UCll is a complex unitary matrix containing the singular vectors as columns. The diagonal matrix Rll contains the
real positive singular values in descending order. This method
based upon the diagonalization of the spectral density matrix, as
it was called, was already used in the beginning of the eighties to
obtain the modes of a vibrating system subjected to natural excitation 20. Some years later, the method was also applied to Frequency Response Functions FRFs and became known as the
Complex Mode Indication Function CMIF. As suggested by the
name, the CMIF was originally intended as a tool to count the
number of modes that is present in measurement data. As a useful
byproduct, the CMIF also identifies the modal parameters from
FRFs 21. Recently, the spectrum-driven method received again
attention as an alternative for the PP method in civil engineering
applications 22. In this paper, the old method was given a new
name, the frequency-domain decomposition method.
Journal of Dynamic Systems, Measurement, and Control
Fig. 1 The complex mode indication function CMIF. The singular values of the spectrum matrix are plotted as a function of
the frequency. Around 2.4 Hz and 7 Hz, two singular values are
significant, indicating that there are two close modes.
(16)
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
(17)
where the first equality says that e k and y ki are uncorrelated; and
the second equality follows from the zero-mean property of the
noise sequence. If, on the contrary the residuals are correlated
with past data, they still contain useful but unmodelled information and the model is not ideal. The derivation of the IV method
starts by imposing conditions like 17 to the ARMA model 7 in
order to get rid of the right-hand side the MA part. The oldest
noise term is e k p ; so by post-multiplying the ARMA model by
T
y k
pi for i0 and by taking the expectation we obtain:
T
T
i0 : E y k y kpi
1 E y k1 y kpi
T
p E y kp y kpi
0
(18)
T
Because of stationarity and Eq. 12, we have: E y k y ki
i0 : R pi 1 R pi1 p R i 0
(19)
By replacing the output covariances by their estimates and writing down the equation for all available time lags i, the AR parameters 1 , . . . , p can be estimated by solving the resulting overdetermined set of equations in a least squares sense. Finally, the
eigenvalues and the observed mode shapes are obtained from the
eigenvalue decomposition of the companion matrix of the AR
coefficients; see Eq. 8.
As in the previous methods of Section 3, it is also in the case of
the IV method possible to reduce the dimensions of the involved
matrices and the related computational effort by making use of a
subset of reference sensors. Only the covariances between all outputs and this subset have to be computed. Details can be found in
7. This corresponds to classical modal analysis, where the impulse response matrices are rectangular matrices having l rows
i.e., the number of outputs and m columns i.e., the number of
inputs. In output-only cases, the impulse responses are substituted by output covariances and the inputs by the reference outputs; see also 8,29.
A typical problem of estimating a parametric model from data
is the determination of the model order. A pth order ARMA model
based on l outputs contains pl poles. Consequently, the expected number of poles covered by the data gives an indication
of the model order. This expected number can be based on physical insight or counted as two times the number of peaks in the
frequency-plot of a nonparametric spectrum estimate; see also the
PP method, Section 3.1. A more accurate model order estimate is
provided by the CMIF, a frequency-plot of the singular values of
a nonparametric spectrum estimate Section 3.2.
More formal procedures estimate models of different order and
compare these models according to a quality criterion such as
Akaikes Final Prediction Error or Rissanens Minimum Description Length criterion 10.
However, in modal analysis one is usually not interested in a
good model as such, but rather in the modal parameters extracted
from that model. Practical experience with parametric models in
modal analysis applications learned that it is better to over-specify
the model order and to eliminate spurious numerical poles afterwards, so that only true physical system poles are left. The famous
stabilization diagram 1,3 is a great tool to achieve this goal. The
poles corresponding to a certain model order are compared to the
poles of a one-order-lower model. If the eigenfrequency, the
damping ratio and the related mode shape or modal participation
factor differences are within preset limits, the pole is labeled as a
stable one. The spurious numerical poles will not stabilize at all
during this process and can be sorted out of the modal parameter
data set more easily. Such a stabilization diagram is represented in
Fig. 2.
Interesting to note and very relevant for civil engineering practice is that the IV method is robust against nonstationary inputs
i.e., white noise with time-varying covariances. This does not
only follow from practical experience but has also been theoretically proven in 32.
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 2 Stabilization diagram obtained with the IV method. The used symbols are: for
a stable pole; .v for a pole with stable frequency and vector; .d for a pole with stable
frequency and damping; .f for a pole with stable frequency and . for a new pole. Two
zooms are added that concentrate on the close modes around 2.4 and 7 Hz.
extended with the SVD to treat noisy data in 34 and 35. The
so-called Eigensystem Realization Algorithm ERA, developed
by Juang 36,37, is a modal analysis application of these deterministic realization algorithms. The stochastic output-only realization problem is solved in 12,38 40. Application of stochastic
realization to modal parameter estimation was reported by Benveniste and Fuchs 32. They also proved that their algorithm is
robust against nonstationary inputs e.g., a white noise sequence
with time-varying covariance.
Stochastic realization relies upon a fundamental property of stochastic state-space systems. It can be proven see for instance
41,42 that the output covariances R i 12 can be decomposed
as:
R i CA i1 G
(20)
T 1i
Ri
R i1
R i1
Ri
R 2i1
C
CA
CA i1
R 2i2
R1
R2
Ri
A i1 GAG G lnO i i
(21)
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 3 Stabilization diagram obtained with the SSI-COV method. By comparing this diagram with the IV diagram Fig. 2, it is clear that the IV method requires higher model
orders to find stable poles.
However, to obtain a good model for modal analysis applications, it is probably a better idea to construct a stabilization diagram, by identifying a whole set of models with different order.
The stabilization diagram was already introduced in Section 4.1.
In case of the SSI-COV method, an efficient construction of the
stabilization diagram is achieved by computing the SVD of the
covariance Toeplitz matrix only once. Models of different order
are then obtained by including a different number of singular values and vectors in the computation of O i and i . A stabilization
diagram obtained with the SSI-COV method is shown in Fig. 3.
5.1 Data-Driven Stochastic Subspace Identification SSIDATA. Recently, a lot of research effort in the system identification community was spent to subspace identification as evidenced by the book of Van Overschee and De Moor 42 and the
second edition of Ljungs book 10. Subspace methods identify
state-space models from input and output data by applying robust numerical techniques such as QR factorization, SVD and
least squares. As opposed to SSI-COV, the DATA-driven Stochastic Subspace Identification method SSI-DATA avoids the computation of covariances between the outputs. It is replaced by
projecting the row space of future outputs into the row space of
past outputs. In fact, the notions covariances and projections are
closely related. They both are aimed to cancel out the uncorrelated noise. The first SSI-DATA algorithms can be found in 44.
A general overview of data-driven subspace identification both
deterministic and stochastic is provided in 42.
Although somewhat more involved as compared to previous
methods, it is also possible with SSI-DATA to reduce the dimensions of the matrices and the memory requirements of the algorithm by introducing the idea of the reference sensors. This is
demonstrated in 7,11. It is beyond the scope of this paper to
explain the SSI-DATA method in detail. The interested reader is
referred to the above-cited literature.
664 Vol. 123, DECEMBER 2001
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
5.2
(22)
Experimental Comparison
Fig. 5 Eigenfrequency estimation results from 100 MonteCarlo simulations. The estimates are divided by the true values
a value of 1 on the graphs indicates a perfect estimate. These
relative frequencies are shown as dots. The scatter of this
quantity gives an idea about the variance of the estimate. The
average estimate is also shown as a dashed line. The deviation of this quantity from 1 full line corresponds to the bias of
the estimate. The rows show the 3 modes; the columns represent the results of 3 identification methods: PP, IV and SSIDATA.
sponses at 6 horizontal DOFs are simulated and afterwards contaminated by white measurement noise with N/S10% N/S is
the ratio of the rms values of the noise and output signal. The
noisy outputs are then fed to 5 system identification methods: PP
Section 3.1, CMIF Section 3.2, IV Section 4.1, SSI-COV
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Section 4.2 and SSI-DATA Section 5.1. More details about the
structure and the simulations can be found in 7.
Figures 13, which were already introduced previously, contain
intermediate identification results of CMIF, IV and SSI-COV. As
explained in Sections 4.1 and 4.2, no order selection criteria as
such is applied. The stable poles are selected from the stabilization
diagrams and they dont have to originate from one model. The
modal parameter estimation results for the first three modes are
represented in Figs. 57. In our discussion of the CMIF method,
we did not include an alternative frequency or damping estimation
procedures as compared to the PP method. The only difference is
that the CMIF can detect closely spaced modes and finds the
eigenfrequencies in a more objective way. Therefore the CMIF
frequencies and damping ratios are not presented in the figures.
The results of SSI-COV and SSI-DATA are so close to each other,
that only SSI-DATA is presented.
The eigenfrequency estimates of the PP method can only take
the discrete values determined by the frequency resolution of the
spectrum Fig. 5. All methods yield unbiased eigenfrequency estimates. Although still small, the standard deviation of the PP
estimates is three times higher than for the other methods.
When looking at the damping estimates Fig. 6, the high bias
of the PP damping estimates is striking. It is rather a coincidence
that mode 2 and 3 have unbiased damping estimates, as the
situation changes when choosing different options for the
non-parametric spectrum estimate resolution, window, overlap,
averages.
Regarding the mode shape estimates Fig. 7, the IV estimates
for the first mode are too bad to fit into the scales. Also the
average correlation of the PP estimates of the third mode could
not be represented. The subspace methods clearly outperform the
others.
Conclusions
References
1 Heylen, W., Lammens, S., and Sas, P., 1995, Modal Analysis Theory and
Testing, Department of Mechanical Engineering, Katholieke Universiteit Leuven, Leuven, Belgium.
2 Maia, N. M. M., Silva, J. M. M., He, J., Lieven, N. A. J., Lin, R. M., Skingle,
G. W., To, W.-M., and Urgueira, A. P. V., 1997, Theoretical and Experimental
Modal Analysis, Research Studies Press, Taunton, Somerset, UK.
3 Allemang, R. J., 1999, Vibrations: Experimental Modal Analysis, Course
Notes, Seventh Edition, Structural Dynamics Research Laboratory, University
of Cincinnati, OH. http://www.sdrl.uc.edu/course info.html.
4 Ewins, D. J., 2000, Modal Testing: Theory, Practice and Application, Research
Studies Press, Baldock, Hertfordshire, UK.
5 Abdelghani, M., Verhaegen, M., Van Overschee, P., and De Moor, B., 1998,
Comparison study of subspace identification methods applied to flexible
structures, Mech. Syst. Signal Process., 12, No. 5, pp. 679 692.
6 Petsounis, K. A., and Fassois, S. D., 2001, Critical comparison and assessment of parametric time domain methods for the identification of vibrating
structures. Mech. Syst. Signal Process., accepted for publication.
7 Peeters, B., 2000, System Identification and Damage Detection in Civil Engineering, PhD thesis, Department of Civil Engineering, Katholieke Universiteit Leuven, Belgium, http://www.bwk.kuleuven.ac.be/bwm, Dec.
8 Hermans, L., and Van der Auweraer, H., 1999, Modal testing and analysis of
structures under operational conditions: industrial applications, Mech. Syst.
Signal Process., 13, No. 2, pp. 193216.
9 Peeters, B., and De Roeck, G., 2001, One-year monitoring of the Z24-Bridge:
environmental effects versus damage events, Earthquake Eng. Struct. Dyn.,
30, No. 2, 149171.
10 Ljung, L., 1999, System Identification: Theory for the User, Second Edition,
Prentice-Hall, Upper Saddle River, NJ.
11 Peeters, B., and De Roeck, G., 1999, Reference-based stochastic subspace
identification for output-only modal analysis, Mech. Syst. Signal Process., 13,
No. 6, pp. 855 878.
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
31 Allemang, R. J., Brown, D. L., and Fladung, W. A., 1994, Modal parameter
estimation: a unified matrix polynomial approach, Proceedings of IMAC 12,
the International Modal Analysis Conference, pp. 501514, Honolulu, HI.
32 Benveniste, A., and Fuchs, J.-J., 1985, Single sample modal identification of
a nonstationary stochastic process, IEEE Trans. Autom. Control, AC-30, No.
1, pp. 66 74.
33 Ho, B. L., and Kalman, R. E., 1966, Effective construction of linear statevariable models from input/output data, Regelungstechnik, 14, pp. 545548.
34 Zeiger, H. P., and McEwen, A. J., 1974, Approximate linear realization of
given dimension via Hos algorithm, IEEE Trans. Autom. Control, AC-19,
No. 2, pp. 53.
35 Kung, S. Y., 1974, A new identification and model reduction algorithm via
singular value decomposition. Procedings of the 12th Asilomar Conference
on Circuits, Systems and Computers, pp. 705714, Asilomar, CA, USA, Nov.
36 Juang, J.-N., and Pappa, R. S., 1985, An eigensystem realization algorithm
for modal parameter identification and model reduction, J. Guid. Control
Dyn.
37 Juang, J.-N., 1994, Applied System Identification, Prentice Hall, Englewood
Cliffs, NJ.
38 Desai, U. B., Pal, D., and Kirkpatrick, R. D., 1985, A realization approach to
stochastic model reduction, Int. J. Control 42, pp. 821 839.
39 Aoki M., 1987, State Space Modelling of Time Series, Springer-Verlag, Berlin,
Germany.
40 Arun, K. S., and Kung, S. Y., 1990, Balanced approximation of stochastic
systems, SIAM J. Matrix Anal. Appl., 11, pp. 42 68.
41 Akaike, H., 1974, Stochastic theory of minimal realization, IEEE Trans.
Autom. Control, 19, pp. 667 674.
42 Van Overschee, P., and De Moor, B. 1996, Subspace Identification for Linear
Systems: Theory-Implementation-Applications, Kluwer Academic Publishers,
Dordrecht, The Netherlands.
43 Camba-Mendez, G., and Kapetanios, G., 2001, Testing the rank of the Hankel covariance matrix: a statistical approach, IEEE Trans. Autom. Control,
AC-46, No. 2, pp. 325330.
44 Van Overschee, P., and De Moor, B., 1993, Subspace algorithm for the stochastic identification problem, Automatica, 29, No. 3, pp. 649 660.
45 Viberg, M., 1995, Subspace-based methods for the identification of linear
time-invariant systems, Automatica, 31, No. 12, pp. 18351851.
46 Kirkegaard P. H., and Andersen, P., 1997, State space identification of civil
engineering structures from output measurements, Proceedings of IMAC 15,
The International Modal Analysis Conference, pp. 889 895, Orlando, FL.
47 Peeters, B., De Roeck, G., and Andersen, P., 1999, Stochastic system identification: uncertainty of the estimated modal parameters, Proceedings of
IMAC 17, The International Modal Analysis Conference, pp. 231237, Kissimmee, FL, Feb.
48 Piombo, B., Ciorcelli, E., Garibaldi, L., and Fasana, A., 1993, Structures
identification using ARMAV models, Proceedings of IMAC 11, the International Modal Analysis Conference, pp. 588 592, Orlando, FL.
49 Andersen, P., 1997, Identification of Civil Engineering Structures Using Vector ARMA Models, PhD thesis, Department of Building Technology and
Structural Engineering, Aalborg University, Denmark, May.
50 Pandit, S. M., 1991, Modal and Spectrum Analysis: Data Dependent Systems
in State Space, Wiley, New York.
51 De, Roeck G., Claesen, W., and Van den Broeck, P., 1995, DDS-methodology
applied to parameter identification of civil engineering structures, Vibration
and Noise 95, pp. 341353, Venice, Italy, Apr.
Downloaded 17 Feb 2013 to 152.3.102.242. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm