Вы находитесь на странице: 1из 10

Order-based resonance identification using operational PolyMAX

Karl Janssens(1), Zsolt Kollar(2), Bart Peeters(1), Steven Pauwels(1), Herman Van der Auweraer(1)
(1)

LMS International, Interleuvenlaan 68, B-3001 Leuven, Belgium, karl.janssens@lms.be (2) Budapest University of Technology and Economics, Budapest, Hungary

ABSTRACT This paper presents a new order-based method, which has been developed to automatically identify resonances from operational data in an engine run-up. The method considers the run-up as a multi-sine sweep excitation and combines advanced Order Tracking with Operational Modal Analysis to identify the resonances. A three-steps approach is used. In a first step, an automatic order detection algorithm is applied to identify the significant engine orders in the measured data. These significant orders are then accurately tracked in both amplitude and phase by using an advanced time-varying DFT Order Tracking method. Once this is achieved, the PolyMAX modal parameter estimation algorithm is applied to the tracked engine orders to identify the resonances. The method is illustrated and discussed using the measured in-vehicle sound data and tacho pulse signal of a 4cylinder car in run-up conditions. The obtained resonance identification results are compared with those of a more traditional, spectrum-based method for Operational Modal Analysis.

1 INTRODUCTION One of the main concerns in the development of new cars is the acoustic comfort for driver and passengers. Interior vehicle sound plays an important role in brand identification and product differentiation. There are various sound characteristics that contribute to a typical brand sound. Resonances, amplitude modulations, order nonlinearities, booming and masking effects are typical examples of important acoustic features. An important challenge of sound engineers consists in developing intelligent algorithms to automatically identify and objectively characterize the main acoustic features from an interior car sound. The use of such algorithms in conjunction with a Virtual Car Sound synthesis tool (e.g. [1][2]) will allow engineers and designers to evaluate the sound quality perception of various design alternatives in a shorter period of time without the need for extensive jury testing. This paper concentrates on the automatic identification of resonances from engine run-up data. Two methods are presented and compared. The first one is a classical, spectrum-based method for Operational Modal Analysis, that is suffering from end-of-order related peaks in the frequency spectrum. The second one is a new orderbased approach which overcomes the limitations of the classical method for the considered run-up application. This order-based method considers the engine run-up as a multi-sine sweep excitation and combines advanced Order Tracking with Operational Modal Analysis to identify the resonances.

2 SPECTRUM-BASED OPERATIONAL MODAL ANALYSIS The classical, spectrum-based method to identify resonances employs Operational Modal Analysis in a traditional way. Operational Modal Analysis is a technique to estimate the modal parameters of a system under unknown excitation [3][4]. Theoretically, it needs to be assumed that this unknown excitation is white noise. Indeed, if only

response information is available, no distinction can be made between spectral peaks due to true modes of the structure or spectral peaks due to harmonic components of the excitation. Practically, it has been shown that a flat excitation spectrum suffices and that, for instance, Operational Modal Analysis can also be applied to structural responses due to unmeasured impact or sine sweep excitation. This fact gave us the motivation for applying the method to engine run-up data. The excitation is considered here as a kind of sine sweep data, exciting a broad frequency band. 2.1 Data pre-processing The data consist of a tacho pulse signal and a microphone measurement in the cavity of a 4-cylinder passenger car. The data were measured in a run-up from 1200 to 6000 rpm in third gear. Figure 1 shows the sound signal, which has been band-pass filtered between 100 and 600 Hz, as the analysis is concentrated to this frequency band. Most modal parameter estimation methods do not directly use the raw measurements, but rely on reduced data such as cross correlations and cross spectra between signals measured simultaneously at different locations. The correlations Ri R l l between the measured signals y k R l are estimated as:

Ri =

1 N

N 1 k =0

y k +i y kT

(1)

where k is the sample index, N is the total number of samples and i the correlation sample index, also called the time lag. A high-speed FFT-based implementation exists to compute the correlations as in (1). See [5]. In the present application, only 1 microphone signal is considered. The auto-correlation sequence for time lags i = 0 8191 is shown in figure 2. For instance, correlation-driven Stochastic Subspace Identification uses these functions as primary data. See [3][4]. Frequency-domain methods require cross spectra as primary data. As a non-parametric spectrum estimate, the so-called weighted correlogram can be used. It is computed as the DFT of the weighted estimated correlation sequence (1):

S yy () =

k = L

wk Rk exp( jkt )

(2)

where L is the maximum number of time lags at which the correlations are estimated. This number is typically much smaller than the number of data samples to avoid the greater statistical variance associated with the higher lags of the correlation estimates. As the correlation samples at negative time lags ( k < 0 ) contain redundant information, it suffices to consider only the positive time lag when computing the spectra. This lead to so-called half spectra of which even the auto spectra have a phase different from zero:
+ S yy () = L w0 R 0 + wk R k exp( jkt ) 2 k =1

(3)

The advantage to estimate spectra based on correlation functions, is that the use of a Hanning window can be avoided. A Hanning window introduces a bias on the damping estimates. Instead, just like in impact testing, an exponential window can be applied to the correlation functions before computing the DFT. An exponential window reduces the effect of leakage and the influence of the higher time lags, which have a larger variance. Moreover, the application of an exponential window to correlations is compatible with the modal model and the pole estimates can be corrected. The weighted correlogram approach to estimate half spectra is illustrated in figure 2. An exponential window of 10% has been used and its effect on the correlation data and spectra is clearly visible.

2.30

-2.30

Pa Real

0.00

37.00

Figure 1: Engine run-up sound data.

10.0e-3

60.00

( Pa) 2 Real

40.00

180.00

-10.0e-3

-180.00

0.00

2.00

Phase
100.00

(Pa) 2 dB

Hz

600.00

Figure 2: (Left) Auto correlation before and after applying an exponential window of 10%. (Right) Corresponding half spectra.

2.2 Operational PolyMAX parameter estimation Once the (half) spectra are available, the operational PolyMAX parameter estimation method can be applied [6][7] [8]. Basically, the following modal model is estimated from the half spectra:
+ () = S yy i =1 n

{vi } < g i >


j i

{v }< g
* i

* i *i

>

(4)

where n is the number of modes, * is the complex conjugate of a matrix, {vi } C l are the mode shapes, < g i > C l are the so-called operational reference factors, which replace the modal participation factors in case of output-only data and i are the poles, which are occurring in complex-conjugated pairs and are related to the eigenfrequencies i and damping ratios i as follows:

i , *i = i i j 1 i2 i

(5)

In the example of this paper, only 1 sensor was analysed, so the mode shapes and operational reference factors are scalars and it is only interesting to consider the half poles. As an intermediate step in the analysis, a stabilization diagram is constructed by fitting models of increasing order (in this case ranging from 2 to 100) to the spectra (figure 3). The main asset of PolyMAX becomes immediately

clear when looking at the stabilization diagram: it is very clear and easy for the user to select the poles. Automatic pole selection is also very well possible. See [9]. Such an automatic modal analysis approach has been employed in this example. Once a set of poles is identified, it is possible to re-synthesize the data based on the right-hand side of (4) and after adding lower and upper residual terms. The original spectrum is compared to the synthesized one in figure 4. The good correspondence indicates that most relevant modes have been extracted from the data.

Figure 3: Operational PolyMAX stabilization diagram for engine run-up data.

60.00

A utoPow er MIC:1:S Sy nthes iz ed Cros s pow er MIC:1:S

40.00
180.00

-180.00

Phase

( Pa)2 dB
100.00

Hz

600.00

Figure 4: Measured (red/black) versus synthesized PolyMAX spectrum (green/grey).

2.3 Discussion of spectrum-based resonance identification results At first sight, the results of the spectrum-based approach seem to be very satisfactory. A very good agreement is obtained between measured and synthesized spectrum (figure 4). However, the results must be interpreted with care. Figure 5 shows the overall spectrum used to extract the modes as well as the rpm-frequency spectrogram.

A comparison of both graphs reveals that some of the peaks in the overall spectrum are originating from order components that suddenly stop at the maximum rpm. Cursors have been inserted at 150 Hz (i.e. order 1.5 at 6000 rpm) and 200 Hz (i.e. order 2 at 6000 rpm) in figure 5 to illustrate this. These peaks are identified as poles of the system. This is the weakness of the method: not only the real poles are identified, but also the end-of-order related poles which are physically not present in the system.

60.00

( Pa)2 dB

40.00

150

200

100.00

Hz

600.00

6000.00

80.00

Z-Axis: measured tracking

rpm

1200.00 100.00

150

200 Hz 600.00

20.00

Figure 5: Overall spectrum and rpm-frequency spectrogram of the engine run-up sound with indication of clear end-of-order related poles.

3 ORDER-BASED OPERATIONAL MODAL ANALYSIS The second method is a new developed order-based approach which does not suffer from end-of-order related peaks and only identifies the real poles of the system. This method considers the run-up as a multi-sine sweep excitation and combines advanced Order Tracking with Operational Modal Analysis to identify the resonances.

Pa2

dB

A three-steps approach is used. In a first step, an automatic order detection algorithm is employed to identify the significant engine orders in the engine run-up data. These significant orders are then accurately tracked in both amplitude and phase by using an advanced time-varying DFT Order Tracking method. Once this is achieved, the PolyMAX modal parameter estimation algorithm is applied to the extracted orders to identify the resonances. 3.1 Automatic order detection The automatic order detection algorithm is based on adaptive resampling. The measured tacho pulse train is used to convert the sound data from the time-domain to the angle-domain. Once this is achieved, a tracked spectral processing is done on the angle-domain data. From the obtained rpm-order spectrogram, a mean order spectrum is calculated. From this mean order spectrum, the significant orders can be automatically identified as those that are local maxima and clearly exceeding the noise floor. Figure 6 shows the mean order spectrum (up to order 10 in steps of 0.05) and automatic order detection results for the engine run-up data in figure 1. The detected significant orders (orders 0.4, 0.8, 1, 1.5, 2,) are marked with a blue star. These are local maxima that exceed the noise floor (yellow curve of local minima) with more than 3 dB (detection threshold).

Figure 6: The automatic order detection results for the run-up example in figure 1.

3.2 Time-varying DFT Order Tracking Once the signifiant orders are detected, an advanced time-varying DFT algorithm is applied to accurately track them in both amplitude and phase. The TVDFT transform [10] is a discrete Fourier transform with a kernel whose frequency is varying as a function of time as defined by the rpm of the engine. It can be written as follows:

am =

1 N

x(nt ) * cos(2 (om * t * rpm / 60)dt )


n =1 N
0

nt

(6.1)

1 bm = N

x(nt ) * sin(2 (om * t * rpm / 60)dt )


n =1
0

nt

(6.2)

where x(nt) are the data samples, N is the number of data samples within an analysis block, om is the order that is being analyzed and am and bm are the Fourier coefficients of the cosine and sine terms for om. The TVDFT Order Tracking method performs much better than the classical FFT-based approach, which gives unreliable phase-estimates.

One of the important parameters of the TVDFT algorithm is the block size N. Our implementation uses a default block size of 16 periods (order resolution of 1/16). For normal speed run-ups of 5 to 10 seconds, this suffices to adequately separate the half orders without suffering from order cross-talk effects. For slower run-ups, the number of periods can be increased to also separate and track more closely-spaced orders if needed. Another important parameter is the number of trigger pulses per revolution. For a normal speed engine run-up of 5 to 10 seconds, we are using a trigger signal (derived from the tacho pulse train) with 1 pulse per 2 revolutions. This is suffucient to accurately track the phase behaviour of the even, odd and half orders. For slower run-ups, a trigger signal with reduced number of pulses can be used (e.g 1 pulse per 10 revolutions). This then allows to also accurately track the phase behaviour of more exotic orders, for instance order 0.4 in the considered run-up example of figure 1. Such a trigger signal with reduced number of pulses is not recommended for faster run-ups, as we then obtain only a limited number of order tracking points for the complete rpm range, which is clearly not sufficient to adequately describe the amplitude and phase variations in function of rpm. As an example, figure 7 visualizes the order 0.4 and order 2 tracking results (amplitude and phase in function of rpm) for the considered engine run-up. For this slow run-up (from 1200 to 6000 rpm in 37s) we used an order resolution of 1/32 and a trigger signal with only 1 pulse per 10 revolutions.

Figure 7: The TVDFT Order Tracking results (amplitude and phase in function of rpm) for orders 0.4 and 2.

3.3 Operational Modal Analysis In the last step of the approach, the resonances are identified from the extracted orders with Operational Modal Analysis. The PolyMAX parameter estimation method is applied to the complex mean of the tracked orders in adjacent, partially overlapping frequency bands (100-220 Hz, 200-320 Hz, 300-420 Hz, 400-520 Hz, 500-620 Hz). The complex mean of orders within the different frequency bands is only calculated from those orders that fully lie within the bands. In this way, end-of-order dsicontinuities are avoided. An automatic pole selection method [9] is used to select the set of poles from the stabilization diagram. The good modal synthesis results in figure 8 indicate that the most relevant modes are extracted from the data.

0.025 0.020 0.015 0.010 0.005 0 200 100 0 -100 200 220 240 260 280 300 320 220 240 260 280 300 320

0.010 0.008 0.006 0.004 0.002 0 400 100 0 -100 400 420 440 460 480 500 520 420 440 460 480 500 520

Figure 8: The modal synthesis results for 2 frequency bands (synthesis in red, order data in blue).

3.4 Discussion of order-based resonance identification results Table 1 gives the order-based resonance identification results for the important modes with mean spectral energy above 50 dB. Figure 9 visualizes these modes on top of the rpm-frequency sound spectrogram. The clearly visible resonance areas around 250 Hz and 360 Hz are well detected.

It is also clear that the method did not suffer from end-of-order related problems. Poles of 150 Hz (frequency of order 1.5 at 6000 rpm) and 200 Hz (frequency of order 2 at 6000 rpm) were not identified this time. This is clearly the strenght of the order-based method compared to the classical, spectrum-based approach. Only physical poles of the system are identified.

Important modes Eigenfrequency [Hz] Damping ratio [%] 1 121.27 2.98 2 191.99 2.83 3 236.91 0.87 4 246.23 1.82 5 247.73 1.45 6 259.43 1.98 7 346.26 0.32 8 347.14 2.03 9 359.56 0.23 10 362.50 1.54 11 373.09 1.09
Table 1: Order-based resonance identification results.

Figure 9: Visualization of important modes on top of the rpm-frequency spectrogram.

4 CONCLUSIONS The presented order-based method for Operational Modal Analysis has proven to be a very promising technology to identify resonances from operational engine run-up data. The method performs much better than the classical, spectrum-based method, which is suffering from end-of-order related peaks in the spectrum. By applying the PolyMAX parameter estimation algorithm to tracked engine order data, the method overcomes these end-oforder related problems and only identifies these modes that are physically present in the system.

REFERENCES [1] K. JANSSENS, M. ADAMS, P. VAN DE PONSEELE AND L. VALLEJOS. The integration of experimental models in a real-time Virtual Car Sound engineering environment. In Proc. of ISMA 2002, Leuven, Belgium, 16-18 Sep., 2002. [2] K. JANSSENS, A. VECCHIO, P. MAS, H. VAN DER AUWERAER AND P. VAN DE PONSEELE. Sound quality evaluation of structural design modifications in a Virtual Car Sound environment. In Proc. of the Institute of Acoustics, 26(2), Southampton, UK, 29-31 Mar., 2004. [3] L. HERMANS AND H. VAN DER AUWERAER. Modal testing and analysis of structures under operational conditions: industrial applications. Mechanical Systems and Signal Processing, 13(2), 193-216, 1999. [4] B. PEETERS AND G. DE ROECK. Stochastic system identification for operational modal analysis: a review. ASME Journal of Dynamic Systems, Measurement, and Control, 123(4):659-667, 2001. [5] A.V. OPPENHEIM AND R.W. SCHAFER. Digital Signal Processing. Prentice-Hall, 1975. [6] B. PEETERS, F. DAMMEKENS, F. MAGALHES, H. VAN DER AUWERAER, E. CAETANO, AND . CUNHA. Multi-run Operational Modal Analysis of the Guadiana cable-stayed bridge. In Proc. of IMAC24, St. Louis (MO), USA, Jan.-Feb. 2006. [7] B. PEETERS, F. VANHOLLEBEKE, AND H. VAN DER AUWERAER. Operational PolyMAX for estimating the dynamic properties of a stadium structure during a football game. In Proc. of IMAC23, Orlando (FL), USA, Jan.-Feb. 2005. [8] B. PEETERS AND H. VAN DER AUWERAER. Recent developments in operational modal analysis. In Proc. of the 6th European Conf. on Structural Dynamics EURODYN 2005, Paris, France, 4-7 Sep. 2005. [9] J. LANSLOTS, B. RODIERS, AND B. PEETERS. Automated pole-selection: proof-of-concept & validation. In Proc. of ISMA 2004, Leuven, Belgium, Sep. 2004. [10] J.R. BLOUGH, D.L. BROWN AND H. VOLD. The time-variant discrete Fourier transform as an order tracking method. In Proc. of SAE Conference 1997, 1997.

Вам также может понравиться