Вы находитесь на странице: 1из 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/297841095

PolyMAX modal parameter estimation: Challenging automotive and aerospace


applications

Article · January 2004

CITATION READS

1 221

4 authors, including:

Bart Peeters Herman Van der Auweraer


Siemens Industry Software Siemens Industry Software NV
242 PUBLICATIONS   5,021 CITATIONS    324 PUBLICATIONS   3,660 CITATIONS   

SEE PROFILE SEE PROFILE

Jan Leuridan
Siemens Industry Software (SISW) NV
56 PUBLICATIONS   1,203 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

DeMoTestEV – DEsign, Modelling and TESTing tools for Electrical Vehicles View project

Virtual Sensing View project

All content following this page was uploaded by Bart Peeters on 28 April 2017.

The user has requested enhancement of the downloaded file.


In Proceedings of VDI-Schwingungstagung 2004 - Modalanalyse und Identifikation, Wiesloch bei
Heidelberg, Germany, 25-26 May 2004.

PolyMAX Modal Parameter Estimation

Challenging automotive and aerospace applications

Dr. Ir. B. Peeters(1), Dr. Ir. H. Van der Auweraer(1), Dr. Ir. J. Leuridan(1),
Dr. Ing. T. Vasel(2), (1)LMS International, (2)LMS Deutschland

Abstract:
Recently, a new non-iterative frequency-domain parameter estimation method was
proposed. This so-called PolyMAX method is based on a least-squares approach and uses
multiple-input-multiple-output frequency response functions as primary data. The PolyMAX
polyreference frequency-domain method can be implemented in a very similar way as the
industry standard polyreference (time-domain) LSCE method: in a first step a stabilisation
diagram is constructed containing frequency, damping and participation information. Next,
the mode shapes are found in a second least-squares step, based on the user selection of
stable poles. One of the specific advantages of the technique lies in the very stable
identification of the system poles and participation factors as a function of the specified
system order, leading to easy-to-interpret stabilisation diagrams. This implies a potential for
automating the method and to apply it to “difficult” estimation cases such as high-order
and/or highly damped systems with large modal overlap. Several challenging real-life
automotive and aerospace case studies are discussed in this paper. The new PolyMAX
method will be compared with classical methods concerning stability, accuracy of the
estimated modal parameters and quality of the frequency response function synthesis.

1. Introduction
Experimental Modal Analysis (EMA) is currently one of the key technologies in structural
dynamics analysis. Based on the academic fundaments of system identification, it has
evolved to become a “standard” approach in mechanical product development. Essential in
this evolution is that modal analysis research has, from the start, taken the point of view of
industrial applicability, focusing on solving the specific problems related to testing and
modelling large industrial structures. The merit of each new method or new approach has
always been checked against the added value it brought in terms of helping the application
engineers to derive better models. The result is that EMA is now considered as a
“commodity” tool, continuously expanding its application base.
While in the past isolated structures with low damping were analysed, modal analysis is
nowadays also explored on complex structures with high damping such as trimmed car
bodies. Next to this, EMA has evolved to a standard tool for Finite Element Model updating
and is exploited in combination with numerical technologies for hybrid engineering. These
recent evolutions have highlighted the current limitations of the EMA process:
• The task of selecting the correct model order and discriminating between spurious and
structural system poles is quite complex, in particular in the case of high-order and/or
highly damped structures. This inevitably results in highly operator-dependant results,
and requires numerous iterations in the analysis procedure.
• Whereas the quality of the current modal parameter estimation technology is satisfying
for undamped or slightly damped structures, there is an increasing need for better modal
parameters estimations for highly damped structures.
• Instead of a variety of parameter estimation techniques, each optimised for a specific test
situation, there is a need for a single reliable and robust method that can be used in a
wide variety of applications.
As will be demonstrated in this paper, the PolyMAX modal parameter estimation method
provides an answer to these challenges. In next section, the theory behind the PolyMAX
method is briefly explained. Afterwards, the method will be applied to difficult industrial
cases: the identification of trimmed cars and the analysis of flight flutter test data.

2. PolyMAX theory in brief


Basically, the PolyMAX method is a further evolution of the so-called least-squares complex
frequency-domain (LSCF) method. Originally, LSCF was introduced to find initial values for
the iterative maximum likelihood method [1]. The method estimates a so-called common-
denominator transfer function model [2]. Quickly it was found that these “initial values”
yielded already very accurate modal parameters with a very small computational effort
[1][3][4]. The most important advantage of the LSCF estimator over the available and widely
applied parameter estimation techniques [5] is the fact that very clear stabilisation diagrams
are obtained (Section 3). A thorough analysis of different variants of the common-
denominator LSCF method can be found in [4]. A complete background on frequency-
domain system identification can be found in [6].
It was found that the identified common-denominator model closely fitted the measured
frequency response function (FRF) data. However, when converting this model to a modal
model (5) by reducing the residues to a rank-one matrix using the singular value
decomposition (SVD), the quality of the fit decreased [3]. Another feature of the common-
denominator implementation is that the stabilisation diagram can only be constructed using
pole information (eigenfrequencies and damping ratios). Neither participation factors nor
mode shapes are available at first instance [7]. The theoretically associated drawback is that
closely spaced poles will erroneously show up as a single pole.
These two reasons provided the motivation for a polyreference version of the LSCF method,
using a so-called right matrix-fraction model. This method is called “PolyMAX”. In this
approach, also the participation factors are available when constructing the stabilisation
diagram. The main benefits PolyMAX are the facts that the SVD step to decompose the
residues can be avoided and that closely spaced poles can be separated. The method was
introduced in [7][8].

Data model
Time-domain methods, such as the polyreference LSCE method [5][9][10], typically require
impulse responses (obtained as the inverse Fourier transforms of the FRFs) as primary data.
Frequency-domain methods such as PolyMAX use measured Frequency Response
Functions (FRFs) as primary data. In PolyMAX, following so-called right matrix-fraction
model is assumed to represent the measured FRFs:
−1

[H (ω)] = ∑ z [βr ] .  ∑ z r [α r ]


p p
r
(1)
r =0  r =0 
where [H (ω)] ∈ l×m
is the matrix containing the FRFs between all m inputs and all l

outputs; [β r ]∈ l ×m
are the numerator matrix polynomial coefficients; [α r ]∈ m× m
are the
denominator matrix polynomial coefficients and p is the model order. Please note that a so-
called z -domain model (i.e. a frequency-domain model that is derived from a discrete-time
model) is used in (1), with:

jω∆t
z=e (2)
where ∆t is the sampling time. Equation (1) can be written down for all values ω of the
frequency axis of the FRF data. Basically, the unknown model coefficients [α r ], [β r ] are then
found as the least-squares solution of these equations (after linearisation). More details
about this procedure can be found in [7][8].
Poles and modal participation factors
Once the denominator coefficients [α r ] are determined, the poles and modal participation
factors are retrieved as the eigenvalues and eigenvectors of their companion matrix:

 0 I … 0 0 
 
 0 0 … 0 0 
 … … … … … .P = P Λ (3)
 
 0 0 … 0 I 
 [ ] [ ]
 − αT
0 − α 1T … − α Tp − 2 [ ] [
− α Tp −1 ]


mp×mp
The modal participation factors are the last m rows of P ∈ ; the matrix Λ ∈ mp× mp

λ i ∆t
contains the (discrete-time) poles e on its diagonal. They are related to the

eigenfrequencies ω i [rad/s] and damping ratios ξ i [-] as follows ( •* denotes complex

conjugate):

λ i , λ*i = −ξ i ω i ± j 1 − ξ i2 ω i (4)

This procedure is similar to what happens in the time-domain LSCE method and allows
constructing a stabilisation diagram for increasing model orders p and using stability criteria
for eigenfrequencies, damping ratios and modal participation factors. Examples of such
stabilisation diagrams will be given in Section 3.

Mode shapes
Although theoretically the mode shapes could be derived from the model coefficients
[α r ], [β r ] , we proceed in a different way. A so-called pole-residue model is considered:

[H (ω)] = ∑
n
{v i } < l iT >
+
{v }< l
*
i i
H
>

[LR ] + [UR] (5)
i =1 jω − λ i jω − λ *
i ω2

where n is the number of modes; • H denotes complex conjugate transpose of a matrix;

{v i }∈ l
are the mode shapes; < l iT >∈ m
are the modal participation factors and λ i are

the poles. [LR ], [UR ] ∈ l ×m


are respectively the lower and upper residuals modelling the
influence of the out-of-band modes in the considered frequency band. The interpretation of
the stabilisation diagram yields a set of poles λ i and corresponding participation factors

< l iT > . Since the mode shapes {v i } and the lower and upper residuals are the only
unknowns, they are readily obtained by solving (5) in a linear least-squares sense. This
second step is commonly called least-squares frequency-domain (LSFD) method [5][9]. The
same mode-shape estimation method is normally also used in conjunction with the time-
domain LSCE method.

3. Challenging modal parameter estimation cases


Full vehicle
A typical example of a challenging modal analysis application is the identification of a
trimmed car. The trim material turns a nicely resonating car body into a highly damped
system with large modal overlap. In the present example, data from a Porsche 911 Targa
Carrera 4 was used. The accelerations of the fully equipped car were measured at 154
degrees of freedom (DOF), while 4 shakers were simultaneously exciting the structure. More
details about the test can be found in [11]. Interesting about the data set are, next to the high
damping, the large number of references. This allows to investigate whether the property of
very clear stabilisation diagrams present in the LSCF method [1][3][4] carries over to the
polyreference case, which is the PolyMAX method. The data were analysed in a frequency
range from 3.5 to 30 Hz using three techniques:
• The frequency domain direct parameter identification (FDPI) technique, which has
traditionally been used to analyse data from highly damped structures [5][9][12].
• The least squares complex exponential (LSCE) method, which is the “industry-standard”
[5][9][10].
• The new PolyMAX method [7][8].
Figure 1 shows the stabilisation diagrams for the 3 methods. FDPI yields a clearer diagram
than LSCE, confirming the common assumption that FDPI is the preferred method in case of
high damping [5][9]. However, both methods are clearly outperformed by the PolyMAX
method: especially at lower frequencies it is much clearer than the FDPI diagram and also a
larger number of stable poles is found. As shown in [11], more stable FDPI poles can be
found when splitting the considered frequency band in smaller sub-bands. For the subset of
PolyMAX poles that have an FDPI counterpart, Table 1 shows that the resulting estimations
for frequency and damping are very close. Also the mode shapes are very similar, which is
evidenced by the MAC values represented in Figure 2. The modal assurance criterion (MAC)
is defined as the squared correlation coefficient between two modal vectors v i , v j :

2
viH v j
MACij = (6)
viH vi v Hj v j
Figure 1: Porsche full vehicle stabilisation diagrams: (Top) FDPI; (Middle) LSCE; (Bottom)
PolyMAX.
Table 1: Eigenfrequencies and damping
ratios from FDPI and PolyMAX.
FDPI method PolyMAX
f [Hz] ξ [%] f [Hz] ξ [%]
3.96 5.4
4.24 9.6
4.81 11.6
6.02 4.1 6.02 4.2
8.58 6.4 8.57 6.5
14.56 6.1 14.59 5.8
15.74 6.3
17.05 5.8
17.99 5.5 18.25 4.9
20.91 2.7
21.78 2.9 21.81 2.7
22.58 3.4 Figure 2: MAC values assessing the
23.98 0.9
correlation between FDPI and
25.12 2.4
25.23 3.1 corresponding PolyMAX mode
26.05 2.1 shapes.
27.03 5.6 27.06 5.2

The excellent identification results obtained with the PolyMAX method are confirmed by
comparing the measured FRFs with the FRFs that can be synthesised from the modal
parameters based on (5). Such a comparison is made in Figure 3. The correspondence
between two FRFs can also be quantified in single numbers [9]. The correlation C H oi and the

normalised error E H oi between a measured Hˆ oi (ω k ) and synthesised FRF H oi (ωk ,θ) are

respectively defined as:

Nf 2
Nf
∑ Hˆ
2
(ωk ) H (ωk , θ) ∑ Hˆ
*
oi oi oi (ωk ) − H oi (ωk , θ)
k =1
C H oi = , EH oi = k =1
(7)
 Nf
 Nf
 Nf
 ∑ Hˆ oi (ωk ) Hˆ oi* (ωk )   ∑ H oi (ωk , θ) H oi* (ωk , θ)  ∑H
2
(ωk , θ)
 k =1   k =1  oi
   k =1

Overall, an impressive 93% correlation between measured and synthesised FRFs was
achieved.
7.00e-3 7.00e-3

( g/N)
( g/N)

Log
Log

FRF moto:9:+Z/karo:25:+Z FRF moto:9:+Z/karo:25:+Z


20.0e-6 20.0e-6
180.00 3.50 Linear 30.00 180.00 3.50 Linear 30.00

Phase
Phase

Hz Hz

°
°

FRF moto:9:+Z/karo:25:+Z FRF moto:9:+Z/karo:25:+Z


-180.00 -180.00
3.50 Hz 30.00 3.50 Hz 30.00

7.00e-3 7.00e-3
( g/N)

( g/N)
Log

Log

FRF moto:9:+Z/moto:2:+Z FRF moto:9:+Z/moto:15:+Z


20.0e-6 20.0e-6
180.00 3.50 Linear 30.00 180.00 3.50 Linear 30.00
Phase

Phase

Hz Hz
°

FRF moto:9:+Z/moto:2:+Z FRF moto:9:+Z/moto:15:+Z


-180.00 -180.00
3.50 Hz 30.00 3.50 Hz 30.00

Figure 3: Comparison of the measured FRFs (grey/green) with FRFs synthesised from the
identified modal model (black/red). The FRFs between the 4 inputs and 1 typical
output are shown.

Partially trimmed car


Although the FDPI method performed satisfactory in the Porsche example discussed in
previous subsection, this is not always the case. In present subsection, data from a partially
trimmed car is used. The modal damping ratios are a bit lower (2–3%) than in the case of the
fully equipped Porsche. The considered frequency range is 35–75 Hz. There were more than
500 output DOFs and 2 inputs. Figure 4 compares the FDPI and the PolyMAX stabilisation
diagram. The FDPI stabilisation diagram is hardly to be interpreted at all and, again,
PolyMAX is very successful in finding the system poles.
Figure 4: Partially trimmed car stabilisation diagrams: (Top) FDPI; (Bottom) PolyMAX.

Noisy in-flight data


Cases exist where the FRF data are highly contaminated by noise. Flight flutter testing is
such a case. In the example considered here, both wing tips of a plane are excited during the
flight with so-called rotating vanes. These vanes generate a sine sweep through the
frequency range of interest. The forces are measured by strain gauges. Next to these
measurable forces, also turbulences are exciting the plane resulting in rather noisy FRFs.
Figure 5 shows some typical multiple coherences and the corresponding FRFs. They clearly
show the noisy character of the data. During the flight, accelerations were measured at 9
locations, while both wing tips were excited (2 inputs). The data were analysed using both
the LSCE and the PolyMAX method. Figure 6 shows both stabilisation diagrams. Also in this
example, PolyMAX has an advantage over the LSCE method in the sense that it is much
easier to select the true system poles.

1.00 300e-3
( (m/s2)/N )
Log
Amplitude
/

200e-6
180.00 4.00 Linear 20.00
Phase

Hz
°

0.00 -180.00
Hz Hz
Figure 5: Flight flutter test data. (Left) multiple coherences of a sensor at the wing tip close to
the excitation (black/red) and a sensor at the back of the plane (grey/green); (Right)
corresponding FRFs. The frequency axis is blind for confidentiality reasons.
Figure 6: Flight flutter test data stabilisation diagrams: (Top) LSCE; (Bottom) PolyMAX.

When looking at the synthesised FRFs (Figure 7), it appears that PolyMAX is able to
accurately model the data. Obviously this becomes more difficult as the FRF becomes
noisier (Figure 7 – Right). In this context, it is interesting to mention some recent
developments in which the turbulence excitation of the airplane is not just considered as
noise, but as useful input to the system. Very promising results are obtained with these so-
called combined deterministic-stochastic identification methods [13][14].
30.0e-3 0.03
( g/N)

( g/N)
Log

Log
20.0e-6 0.00
180.00 4.00 Linear 20.00 180.00 4.00 Linear 20.00
Phase

Phase
Hz Hz
°

°
-180.00 -180.00
Hz Hz

Figure 7: Comparison of the measured FRFs (green/grey) with FRFs synthesised from the
identified modal model (black/red). (Left) Sensor at the wing tip; (Right) Sensor at the
back of the plane.

4. Conclusions
The proposed PolyMAX modal parameter estimation method offers very good prospects in
view of the current EMA challenges. The method in general yields very clean stabilisation
diagrams, easing dramatically the problem of selecting the model order and the best
structural system poles. The analysis results compare favourably with current best-of-class
commercially available methods, without increasing the computational effort. Its very good
identification behaviour for noisy data sets as well as for high order, highly damped
structures allows to position PolyMAX as a new “standard” in modal parameter estimation. It
is commercially available in the LMS Test.Lab Structural Testing software [15].

5. Acknowledgements
The discussed results have been obtained in the context of the EUREKA project E!2419
FLITE (Flight Test Easy) which is conducted in cooperation with K.U.Leuven (B),
V.U.Brussels (B), INRIA (F), Sopemea (F), Airbus (F), Dassault (F), AGH (PL), ILOT (PL),
PZL Mielec (PL).
6. References
[1] GUILLAUME P., P. VERBOVEN AND S. VANLANDUIT. Frequency-domain maximum
likelihood identification of modal parameters with confidence intervals. In Proceedings
of ISMA 23, the International Conference on Noise and Vibration Engineering, Leuven,
Belgium, 16–18 September 1998.
[2] GUILLAUME P., R. PINTELON AND J. SCHOUKENS. Parametric identification of
multivariable systems in the frequency domain - a survey. In Proceedings of ISMA 21,
the International Conference on Noise and Vibration Engineering, 1069–1082, Leuven,
Belgium, 18–20 September 1996.
[3] VAN DER AUWERAER H., P. GUILLAUME, P. VERBOVEN AND S. VANLANDUIT. Application of
a fast-stabilizing frequency domain parameter estimation method. ASME Journal of
Dynamic Systems, Measurement, and Control, 123(4), 651–658, 2001.
[4] VERBOVEN, P. Frequency domain system identification for modal analysis. PhD Thesis,
Vrije Universiteit Brussel, Belgium, 2002.
[5] HEYLEN W., S. LAMMENS AND P. SAS. Modal Analysis Theory and Testing. Department
of Mechanical Engineering, Katholieke Universiteit Leuven, Leuven, Belgium, 1995.
[6] PINTELON R. AND J. SCHOUKENS. System Identification: a Frequency Domain Approach.
IEEE Press, New York, USA, 2001.
[7] GUILLAUME P., P. VERBOVEN, S. VANLANDUIT, H. VAN DER AUWERAER AND B. PEETERS.
A poly-reference implementation of the least-squares complex frequency-domain
estimator. In Proceedings of IMAC 21, the International Modal Analysis Conference,
Kissimmee (FL), USA, February 2003.
[8] PEETERS B., P. GUILLAUME, H. VAN DER AUWERAER, B. CAUBERGHE, P. VERBOVEN AND
J. LEURIDAN Automotive and aerospace applications of the PolyMAX modal parameter
estimation method. In Proceedings of IMAC 22, the International Modal Analysis
Conference, Dearborn (MI), USA, January 2004.
[9] LMS INTERNATIONAL. The LMS Theory and Background Book, Leuven, Belgium,
www.lmsintl.com, 2000.
[10] BROWN D.L., R.J. ALLEMANG, R. ZIMMERMAN AND M. MERGEAY. Parameter estimation
techniques for modal analysis. Society of Automotive Engineers, Paper No. 790221,
1979.
[11] VAN DER AUWERAER H., C. LIEFOOGHE, K. WYCKAERT AND J. DEBILLE. Comparative
study of excitation and parameter estimation techniques on a fully equipped car. In
Proceedings of IMAC 11, the International Modal Analysis Conference, 627–633,
Kissimmee (FL), USA, 1–4 February 1993.
[12] LEMBREGTS F., R. SNOEYS AND J. LEURIDAN. Application and evaluation of multiple input
modal parameter estimation. International Journal of Analytical and Experimental Modal
Analysis, 2(1), 19–31, 1987.
[13] CAUBERGHE B., P. GUILLAUME, P. VERBOVEN AND E. PARLOO. Identification of modal
parameters including unmeasured forces and transient effects. Journal of Sound and
Vibration, 256(3), 609–625, 2003.
[14] GOETHALS I. AND B. DE MOOR. Subspace identification combined with new mode
selection techniques for modal analysis of an airplane. In Proceedings of the 13th IFAC
Symposium on System Identification (SYSID 2003), Rotterdam, The Netherlands,
September 2003.
[15] LMS INTERNATIONAL. LMS Test.Lab – Structural Testing Rev 4B, Leuven, Belgium,
www.lmsintl.com, 2003.

View publication stats

Вам также может понравиться