BLIND SOURCE SEPARATION TECHNIQUES ANOTHER WAY OF DOING OPERATIONAL MODAL ANALYSIS

Similar documents
EMD-BASED STOCHASTIC SUBSPACE IDENTIFICATION OF CIVIL ENGINEERING STRUCTURES UNDER OPERATIONAL CONDITIONS

Damping Estimation Using Free Decays and Ambient Vibration Tests Magalhães, Filipe; Brincker, Rune; Cunha, Álvaro

VARIANCE COMPUTATION OF MODAL PARAMETER ES- TIMATES FROM UPC SUBSPACE IDENTIFICATION

Automated Modal Parameter Estimation For Operational Modal Analysis of Large Systems

Comparison of output-only methods for condition monitoring of industrials systems

MODAL IDENTIFICATION AND DAMAGE DETECTION ON A CONCRETE HIGHWAY BRIDGE BY FREQUENCY DOMAIN DECOMPOSITION

Assessment of the Frequency Domain Decomposition Method: Comparison of Operational and Classical Modal Analysis Results

Modal parameter identification from output data only

Identification of modal parameters from ambient vibration data using eigensystem realization algorithm with correlation technique

CONTRIBUTION TO THE IDENTIFICATION OF THE DYNAMIC BEHAVIOUR OF FLOATING HARBOUR SYSTEMS USING FREQUENCY DOMAIN DECOMPOSITION

Mode Identifiability of a Multi-Span Cable-Stayed Bridge Utilizing Stochastic Subspace Identification

Subspace-based damage detection on steel frame structure under changing excitation

DEVELOPMENT AND USE OF OPERATIONAL MODAL ANALYSIS

Modal Based Fatigue Monitoring of Steel Structures

Smooth Orthogonal Decomposition for Modal Analysis of Randomly Excited Systems

COMPARISON OF MODE SHAPE VECTORS IN OPERATIONAL MODAL ANALYSIS DEALING WITH CLOSELY SPACED MODES.

Damage Localization under Ambient Vibration Using Changes in Flexibility

Semi-Blind approaches to source separation: introduction to the special session

Identification Techniques for Operational Modal Analysis An Overview and Practical Experiences

where A 2 IR m n is the mixing matrix, s(t) is the n-dimensional source vector (n» m), and v(t) is additive white noise that is statistically independ

Identification of Nonlinear Mechanical Systems: State of the Art and Recent Trends

Eliminating the Influence of Harmonic Components in Operational Modal Analysis

ORIENTED PCA AND BLIND SIGNAL SEPARATION

Tutorial on Nonlinear Modal Analysis of Mechanical Systems

Robust extraction of specific signals with temporal structure

Structural Dynamics Lecture 7. Outline of Lecture 7. Multi-Degree-of-Freedom Systems (cont.) System Reduction. Vibration due to Movable Supports.

Stochastic Dynamics of SDOF Systems (cont.).

An Indicator for Separation of Structural and Harmonic Modes in Output-Only Modal Testing Brincker, Rune; Andersen, P.; Møller, N.

Operational Modal Analysis of Rotating Machinery

Tensor approach for blind FIR channel identification using 4th-order cumulants

Vibration-Based Monitoring of Tensile Loads: System Development and Application

VIBRATION-BASED DAMAGE DETECTION UNDER CHANGING ENVIRONMENTAL CONDITIONS

Transactions on Modelling and Simulation vol 16, 1997 WIT Press, ISSN X

1 Introduction Blind source separation (BSS) is a fundamental problem which is encountered in a variety of signal processing problems where multiple s

IOMAC' May Guimarães - Portugal IMPACT-SYNCHRONOUS MODAL ANALYSIS (ISMA) AN ATTEMPT TO FIND AN ALTERNATIVE

Estimation of Unsteady Loading for Sting Mounted Wind Tunnel Models

Analytical solution of the blind source separation problem using derivatives

IOMAC'15 6 th International Operational Modal Analysis Conference

Joint input-response predictions in structural dynamics

Structural changes detection with use of operational spatial filter

An Iterative Blind Source Separation Method for Convolutive Mixtures of Images

Effects of wind and traffic excitation on the mode identifiability of a cable-stayed bridge

Nonlinear Normal Modes: Theoretical Curiosity or Practical Concept?

BLIND SEPARATION OF INSTANTANEOUS MIXTURES OF NON STATIONARY SOURCES

Lecture Notes 5: Multiresolution Analysis

Uncertainty Quantification for Stochastic Subspace Identification of Multi-Setup Measurements

Modal identification of heritage structures: Molise s bell towers

Tutorial on Blind Source Separation and Independent Component Analysis

HST.582J/6.555J/16.456J

Application of Classical and Output-Only Modal Analysis to a Laser Cutting Machine

Review of modal testing

A UNIFIED PRESENTATION OF BLIND SEPARATION METHODS FOR CONVOLUTIVE MIXTURES USING BLOCK-DIAGONALIZATION

Stochastic structural dynamic analysis with random damping parameters

University of Bristol - Explore Bristol Research. Publisher's PDF, also known as Version of record

MULTICHANNEL SIGNAL PROCESSING USING SPATIAL RANK COVARIANCE MATRICES

EXPERIMENTAL IDENTIFICATION OF A NON-LINEAR BEAM, CONDITIONED REVERSE PATH METHOD

Investigation of traffic-induced floor vibrations in a building

A priori verification of local FE model based force identification.

Input-Output Peak Picking Modal Identification & Output only Modal Identification and Damage Detection of Structures using

Operational Modal Analysis of the Braga Sports Stadium Suspended Roof

1330. Comparative study of model updating methods using frequency response function data

Modal identification of output-only systems using frequency domain decomposition

EXPERIMENTAL STUDY ON CONCRETE BOX GIRDER BRIDGE UNDER TRAFFIC INDUCED VIBRATION

Curve Fitting Analytical Mode Shapes to Experimental Data

OPERATIONAL MODAL ANALYSIS BY USING TRANSMISSIBILITY MEASUREMENTS WITH CHANGING DISTRIBUTED LOADS.

Operational modal analysis using forced excitation and input-output autoregressive coefficients

AN ALGORITHM FOR DAMAGE DETECTION AND LOCALIZATION USIBG OUTPUT-ONLY RESPONSE FOR CIVIL ENGINEERING STRUCTURES SUBJECTED TO SEISMIC EXCITATIONS

Dynamic damage identification using linear and nonlinear testing methods on a two-span prestressed concrete bridge

Modal Testing and System identification of a three story steel frame ArdalanSabamehr 1, Ashutosh Bagchi 2, Lucia Trica 3

SINGLE DEGREE OF FREEDOM SYSTEM IDENTIFICATION USING LEAST SQUARES, SUBSPACE AND ERA-OKID IDENTIFICATION ALGORITHMS

Investigation of Operational Modal Analysis Damping Estimates MASTER OF SCIENCE

IDENTIFICATION OF THE MODAL MASSES OF AN UAV STRUCTURE IN OPERATIONAL ENVIRONMENT

Dessi, D., D Orazio, D.

ANNEX A: ANALYSIS METHODOLOGIES

Improved PARAFAC based Blind MIMO System Estimation

Index 319. G Gaussian noise, Ground-loop current, 61

Improving the Accuracy of Dynamic Vibration Fatigue Simulation

Multi-Order Covariance Computation for Estimates in Stochastic Subspace Identification Using QR Decompositions

Sampling considerations for modal analysis with damping

TIME-DOMAIN OUTPUT ONLY MODAL PARAMETER EXTRACTION AND ITS APPLICATION

Design of Structures for Earthquake Resistance

HST.582J / 6.555J / J Biomedical Signal and Image Processing Spring 2007

ABSTRACT Modal parameters obtained from modal testing (such as modal vectors, natural frequencies, and damping ratios) have been used extensively in s

Damage Identification in Wind Turbine Blades

Improvement of Frequency Domain Output-Only Modal Identification from the Application of the Random Decrement Technique

An example of correlation matrix based mode shape expansion in OMA

Variance estimation of modal parameters from input/output covariance-driven subspace identification

NUMERICAL STUDY OF INTRINSIC FEATURES OF ISOLAS IN A 2-DOF NONLINEAR SYSTEM

Reduction in number of dofs

A METHOD OF ADAPTATION BETWEEN STEEPEST- DESCENT AND NEWTON S ALGORITHM FOR MULTI- CHANNEL ACTIVE CONTROL OF TONAL NOISE AND VIBRATION

2.0 Theory. 2.1 Ground Vibration Test

Influence of background noise on non-contact vibration measurements using particle velocity sensors

Natural frequencies and modal damping ratios identification of civil structures from ambient vibration data

Single Channel Signal Separation Using MAP-based Subspace Decomposition

Comparison of the Results Inferred from OMA and IEMA

Dynamics of Structures

IN-FLIGHT MODAL IDENTIFICATION OF AN AIRCRAFT USING OPERATIONAL MODAL ANALYSIS ABSTRACT

Full-Field Dynamic Stress/Strain from Limited Sets of Measured Data

Nonparametric identification of added masses in frequency domain: a numerical study

Transcription:

BLIND SOURCE SEPARATION TECHNIQUES ANOTHER WAY OF DOING OPERATIONAL MODAL ANALYSIS F. Poncelet, Aerospace and Mech. Eng. Dept., University of Liege, Belgium G. Kerschen, Aerospace and Mech. Eng. Dept., University of Liege, Belgium J.C. Golinval Aerospace and Mech. Eng. Dept., University of Liege, Belgium F. Marin V 2 i S.A. (http://www.v2i.be), Liege Belgium fponcelet, g.kerschen, jc.golinval@ulg.ac.be, f.marin@v2i.be Abstract In many practical applications, it is difficult or even impossible to measure the force applied to a structure. This is especially the case in civil engineering applications. Therefore, this paper proposes to explore the utility of blind source separation (BSS) techniques for operational modal analysis. The basic idea of BSS is to recover unobserved source signals from their observed mixtures. The feasibility and practicality of the proposed method are demonstrated using an experimental application. Introduction For modal analysis of large structures, it is unpractical and expensive to use artificial excitation (e.g., shakers). However, engineering structures are most often subject to ambient loads (e.g., traffic and wind) that can be exploited for modal parameter estimation. One difficulty is that the actual loading conditions cannot generally be measured, and output-only measurements are available. During the last few years, there have been several successful attempts to address this issue using operational modal analysis (OMA) techniques [-2]. Recently, signal processing techniques have been used to perform OMA through the estimation of the modal coordinates. For instance, Lardies et al. [3] exploit the wavelet transform to determine the response of each mode and to subsequently compute the modal parameters. In [4], output-only data are processed using the empirical mode decomposition (also known as Hilbert-Huang transform) to identify the different modal contributions. Digital band-pass filters are considered by Kim et al. [] for the same purpose. Although attractive in principle, these signal processing-based methods present several drawbacks such as edge effects and difficulty in identifying closely spaced modes. In this paper, we propose a new OMA method by borrowing one technique from the statistical literature. The technique, second-order blind identification (SOBI), decomposes measured signals in terms of elemental components. When SOBI is applied to the response of engineering structures, the elemental components are directly related to the modal coordinates, as shown in [6]. This is also the case for other blind source separation (BSS) techniques such as independent component analysis (ICA, [8]), but this is not discussed herein. The reader can refer to [7] for further detail.

2 Blind Source Separation (BSS) 2. Theoretical Background Recovering unobserved source signals from their observed mixtures is a generic problem in many domains and is referred to as blind source separation (BSS) in the literature. One well-known example is the cocktail-party problem, the objective of which is to retrieve the speech signals emitted by several persons speaking simultaneously in a room using only the signals recorded by a set of microphones located in the room. The simplest BSS model assumes the existence of n source signals s ( t),..., s ( t ) and the observation of as many mixtures x ( ),..., ( ) t xn t. Although convolutive and non-linear mixtures can be considered, we focus on linear and static mixtures for which BSS is well established. The noisy model can be expressed as or, in matrix form, n x ( t) = a s ( t) + σ ( t), i =,..., n i ij j i j= x() t = y() t + σ() t = As() t + σ() t where A is referred to as the mixing matrix, and σ() t is the noise vector corrupting the data. The basic idea of BSS is to recover the unobserved source signals s() t from their observed mixtures x() t. Blind means that very little, if anything, is known about the mixing matrix, and that fairly general assumptions are made about the source signals. BSS has two inherent indeterminacies, because the mixing matrix is only identifiable up to scaling and permutation of its rows. As a result, it is not possible to determine the order and the variances of the identified sources. Generally, they are scaled to have a unit variance. 2.2 Second-Order Blind Identification (SOBI) Most BSS approaches are based (explicitly or not) on a model in which the sources are independent and identically distributed variables. On the contrary, the objective of the SOBI algorithm is to take advantage, whenever possible, of the temporal structure of the sources for facilitating their separation [9]. SOBI is therefore an interesting technique for sources with different spectral contents, which is often the case in structural dynamics as explained in Section 3. For the sake of conciseness, a detailed description of the SOBI method is not carried out herein. The reader can refer to [6,9] for further detail. However, we mention that SOBI is based on the joint diagonalization of time-lagged covariance matrices R * ( τ) = E x ( t+ τ) x ( t) (3) These matrices are evaluated from the measured data, and the basic idea is to find a matrix U, which jointly diagonalizes all the covariance matrices. It can be proven that the unitary matrix U corresponds to the mixing matrix A. For illustration, SOBI is applied to a mixture of five sinusoids in Figure and provides an accurate identification of the sources. n () (2)

x 2 x -.. 2 2. 3 s s 2 -.. 2 2. 3 x x 4 x 3 -.. 2 2. 3 -.. 2 2. 3 -.. 2 2. 3 s 3 s 4 s -.. 2 2. 3-2.. 2 2. 3-2 2.. 2 2. 3 -.. 2 2. 3 Time (s) -2.. 2 2. 3 Time (s) Fig. : Mixture of five sinusoids (left plots) and sources identified using SOBI (right plots) 3 Modal Coordinates: Virtual Sources Consider a mechanical system governed by the equations of motion Mx () t + Cx () t + Kx() t = f () t (4) where M, C and K are the mass, damping and stiffness matrices, respectively. The system can be seen as a dynamic mixture of sources, because its response x() t is the convolution product of the impulse response function h () t and the external force vector f () t. However, the application of BSS techniques to mechanical systems is difficult, because the problem of a convolutive mixture of sources is not yet completely solved. Without loss of generality, the conservative system is considered now. Because the normal modes provide a complete set for the expansion of an arbitrary vector, the response of system (4) may be expressed through modal expansion m x() t = n () t η () t = N η() t () i= () i where η () t are the modal coordinates, that is the amplitude modulations of the corresponding i n() i normal modes, and m is the number of degrees of freedom of the system. Expression () shows that when expanding the system response in terms of the vibration modes, the modal coordinates may act as virtual sources regardless of the number and type of the physical excitation forces [6-7]. In addition, the time response can be interpreted as a static mixture of these virtual sources, which renders the application of BSS techniques possible. Finally, the modal coordinates are monochromatic (i.e., sources with different spectral contents) for the free response and mostly monochromatic for the random response, which is exactly what SOBI requires. In summary, the key idea is to interpret the modal coordinates of a dynamic system as virtual sources with different spectral contents. In this context, SOBI provides a one-to-one mapping i

between the mixing matrix and the vibration modes of the structure, which forms the basis of a truly simple modal analysis procedure:. Perform experimental measurements of the response of the tested system to obtain time series at different sensing positions. 2. Apply SOBI to the measured time series to estimate the mixing matrix and the sources. 3. The mode shapes are contained in the mixing matrix A. 4. It is then straightforward to identify the natural frequencies and damping ratios of the corresponding vibration modes. In the free response case, this is carried out by fitting the time series of the sources s() t with exponentially damped harmonic functions 2 ( ) Yexp( ξωt)cos ξ ωt+ α (6) where ω and ξ are the natural frequency and damping ratio of the considered mode, respectively. The amplitude Y and the phase α are constants depending on the initial conditions. In the case of random excitation, the same procedure can be applied, but the identified (random) sources are first transformed into free decaying responses using NExT (Natural Excitation Technique) algorithm []. An interesting feature of the proposed method is that it does not require the measurement of the applied force and can perform OMA. This is particularly convenient in practical applications for which the external force cannot be measured (e.g., vibrations of a bridge due to traffic and wind). 4 Experimental Demonstration To support the previous theoretical findings, the proposed OMA technique was applied to the response of the truss structure depicted in Figure 2. For the free response, a hammer provided a short impulse to the system. For the random response, the structure was mounted on a 26kN electrodynamic shaker, as shown in Figure 2. 6 accelerometers were distributed on the structure (two at each corner, eight on each storey), measuring its response in a horizontal plane. The results were also compared with another OMA technique, the (covariance-driven) stochastic subspace identification (SSI) technique [,]. 4. Free Response The SOBI and SSI methods were first applied to the free response of the truss structure. The first 6 samples of the measured time series were taken into account, the sampling frequency being set to 2 Hz. For SSI, 2 block rows and columns were selected in the Hankel matrix. To build the time-lagged covariance matrices in SOBI (cf. Equation 3), 2 time lags τ were uniformly distributed between.2 and. s, which covers the frequency range of interest. Because there are 6 measurement locations, a total of 6 virtual sources can be considered. The identification of reliable virtual sources is greatly facilitated by computing the error realized during the fitting of the time series of the sources with exponentially damped harmonic functions. For instance, Figure 3 depicts the measured and fitted signals for two different sources. Clearly, the measured source in the top plot of Figure 3 can be considered as an actual source (fitting error=.%), whereas this is likely not the case for the source in the bottom plot (fitting error=69%). Figure 4 shows the fitting error of the 6 virtual sources. sources have a fitting error below 7% and can be safely retained. The sum of their participation in the system response is above 97.7%.

346 Hz.3.2. -. -.2 346 Hz.3.2. -. -.2.2.4.6.8.2.4.6.8 Excitation direction Fig. 2: Experimental fixture mounted on a 26kN electrodynamic shaker 439 Hz... -. -.....2 Measured source 439 Hz... -. -.....2 Reconstructed source Fig. 3: Measured and reconstructed sources (free response) 9 8 7 Fitting error 6 4 3 2 Threshold 346 3 366 334 397 374 8 76 38 26 388 24 439 42 Source frequency (Hz) Fig. 4: Fitting error of the 6 virtual sources (free response, SOBI) The identification results are listed in Table, which shows that the truss structure possesses closely spaced modes. For SSI, an interval rather than a well-definite value is given for the damping ratios, because their values change with the selected model order. The MAC matrix between the modes identified using SSI and SOBI is shown in Figure. One can see a complete correspondence of the results obtained with both methods, which confirms that an accurate and consistent identification is carried out using SOBI.

Frequency (Hz) (SOBI) Frequency (Hz) (SSI) Damping ratio (%) (SOBI) Damping ratio (%) (SSI) 7.94 7.82.2 [..2].37.99.37 [.4.6] 3.7 3.76.2 [.2.28] 8.6 8.69.8 [.2.28] 26.3 26.48.8 [..] 334.24 334.32. [.2.] 34.7 34.76.4 [.4.] 36.79 36.8. [..6] 374.34 374.4. [..3] 38. 38.4.6 [.2.4] 396.9 396.8.8 [.7.] Table : Identified natural frequencies and damping ratios (free response) SOBI - Free Response 396.9 38. 374.3 36.7 34.7 334.2 26.3 8 3.7..99...99.99....9.8.7.6..4.3.3 7.9...2. 7.8.9 3.7 8.6 26.4 334.3 34.7 36.8 374.4 38.4 396.8 SSI - Free Response Fig. : MAC matrix between the modes identified using SOBI and SSI (free response) 4.2 Random Response The SOBI and SSI methods were then applied to the random response of the truss structure. The random force imparted to the truss structure by the electrodynamic shaker was not measured. 6 samples of the measured time series were taken into account, the sampling frequency being set to 2 Hz. For SSI, 2 block rows and columns were selected in the Hankel matrix. For SOBI, 2 time lags τ were uniformly distributed between.2 and. s. Figure 6 shows the fitting error of the 6 virtual sources. 4 sources have a fitting error below 8%, and the two remaining sources have an error above 6%. Source selection is therefore trivial. The

identification results are listed in Table 2. The MAC matrix between the modes identified using SSI and SOBI is shown in Figure 7. Apart from the first mode, there is an excellent agreement between the results obtained with both methods. If the results obtained using SSI in the free response case are taken as a reference, one can compare how well SOBI and SSI perform in the random case. Table 3 shows that they perform equally well. Another finding is that none of the methods seems able to accurately estimate the mode at 7 Hz using the random response. This is because the structural deformation of this mode takes place in a plane orthogonal to the excitation direction. The mode has therefore a very low participation in the system response. 9 8 7 Fitting error 6 4 3 2 346 34 366 8 27 397 37 333 39 24 368 38 7 4 24 Source frequency (Hz) Fig. 6: Fitting error of the 6 virtual sources (random response, SOBI) SOBI - Random Response 397 39 38 37 368 366 346 333 27 24 8 34 7.9.9...86...98.96.93.99.98.97.94.9.8.7.6..4.3.2. 7 34 8 24 27 333 346 366 369 374 378 39 397 SSI - Random Response Fig. 7: MAC matrix between the modes identified using SOBI and SSI (random response)

Freq. (Hz-SOBI) Freq. (Hz-SSI) Damp. (%-SOBI) Damp. (%-SSI) 74.7 74.68 2. [.7 2.].6.28 2.3 [. 2.] 33.9 33.77.8 [.6.8] 8.87 8.98.23 [.2.3] 24.29 24.38.6 [..] 27.47 27.47. [.9.] 333.2 333.34.2 [..] 34.64 34..9 [..2] 36.6 36.76.2 [.7.] 368.9 369.3.33 [..3] 374.34 374.69.6 [.2.4] 38.6 378.34.7 [..7] 39.8 39.9.33 [.4.] 396.83 397.2.7 [..2] Table 2: Identified natural frequencies and damping ratios (random response) Frequency MAC(SOBI random, SSI free) MAC(SSI random, SSI free) 7.64.63.99.96 34.. 8.99.99 24.78.94 27.99.99 333.. 346.93.98 366.92.92 374.9.9 379.94.94 39.8.7 397.9.97 Table 3: Correlation of the modes identified using the random response (SOBI and SSI) with those identified using the free response (SSI)

Conclusion In this paper, a new operational modal analysis method is introduced by extracting modal coordinates from structural responses through second-order blind identification. The experimental application shows that the method holds promise for identification of mechanical systems: A truly simple identification scheme is proposed, because the straightforward application of SOBI to the measured data yields the modal parameters. A seemingly robust criterion has been developed for the selection of reliable sources. The use of stabilization charts, which always require a great deal of expertise, is therefore avoided. In addition, the selection of a model order, a common issue for conventional modal analysis techniques such as SSI, is not necessary. Compared to SSI, the computation load is very reduced, which makes the method a potential candidate for online modal analysis. A possible limitation of the method is that sensors should always be chosen in number greater or equal to the number of active modes. This will be addressed in subsequent studies. 6 Acknowledgments The author F. Poncelet is supported by a grant from the Walloon Government (Contract No. 68 VAMOSNL). The author G. Kerschen is supported by a grant from the Belgian National Fund for Scientific Research (FNRS). 7 References [] [2] [3] [4] [] Peeters, B, De Roeck, G., Stochastic system identification for OMA: A review, Journal of Dynamic Systems Measurement and Control 23, 69-667, 2 Brincker, R, Moller. N, Operational modal analysis - a new technique to explore, Sound and Vibration 4, -6, 26. Lardies, J., Minh Nghi, T.A., Berthillier, M., Modal parameter estimation using the wavelet transform, Archive of Applied Mechanics 73, 78-733, 24 Yang, J.N. et al., Identification of natural frequencies and dampings of in situ tall buildings, Journal of Engineering Mechanics 3, 7-77, 24 Kim, B.H., Stubbs, N., Park, T., A new method to extract modal parameters using output-only responses, Journal of Sound and Vibration 282, 2-23, 2. [6] Poncelet, F., Kerschen, G., Golinval, J.C., Verhelst, D., Output-only modal analysis using BSS techniques, Mechanical Systems and Signal Processing, in press. [7] Kerschen, G., Poncelet, F., Golinval, J.C., Physical interpretation of independent component analysis in structural dynamics, Mechanical Systems and Signal Processing 2, 6-7, 27 [8] Comon, P., ICA: a new concept?, Signal Processing 36, 287 34, 994. [9] Belouchrani, A., Abed-Meraim, K., Cardoso, J.F., Molines, E., A BSS technique using 2nd-order statistics, IEEE Trans. on Signal Processing 4, 434 44, 997. [] James, G.H., Carne, T.G., Lauffer, J.P., The natural excitation technique for modal parameter extraction from operating wind turbines. SAND92-666, UC-26, 993. [] Van Overschee, P., De Moor, B., Subspace Identification for Linear Systems: Theory, Implementation, Applications, Kluwer Academic Publishers, 996.