20
ARTICLE IN PRESS Neurocomputing ( ) www.elsevier.com/locate/neucom Blind separation of rotating machine sources: bilinear forms and convolutive mixtures Alexander Ypma a; b; , Amir Leshem c; d , Robert P.W. Duin a a Pattern recognition group, Department of Applied Physics, TU Delft, Lorentzweg 1, 2628 CJ Delft, The Netherlands b SNN, Geert Grooteplein 21, 6525 EZ Nijmegen, The Netherlands c Metalink Broadband Access, Yakum Business Park, 60972, Israel d Circuits and Systems group, Department of Electrical Engineering, TU Delft, Mekelweg 4, 2628 CD Delft, The Netherlands Abstract We propose the use of blind source separation (BSS) for separation of a machine signature from distorted measurements. Based on an analysis of the mixing processes relevant for machine source separation, we indicate that instantaneous mixing may hold in acoustic monitoring. We then present a bilinear forms-based approach to instantaneous source separation. For simulated acoustic mixing, we show that this method may give rise to a more robust separation. For vibrational monitoring, a convolutive mixture model is more appropriate. The demixing algo- rithm by Nguyen Thi–Jutten allows for separation of the contributions of two coupled machines, both in an experimental setup and in a real-world situation. We conclude that BSS is a feasible approach for blind separation of distorted rotating machine sources. c 2002 Elsevier Science B.V. All rights reserved. Keywords: Blind source separation; Machine vibration; Bilinear forms; Convolutive mixture; Condition monitoring 1. Introduction Many machines contain rotating elements, which causes that specic frequency com- ponents are present in the machine vibration spectrum. Deviations from normal behav- ior show up as distinct dierences with respect to the normal spectrum (the machine Corresponding author. E-mail addresses: [email protected] (A. Ypma), [email protected] (A. Leshem). 0925-2312/02/$ - see front matter c 2002 Elsevier Science B.V. All rights reserved. PII: S0925-2312(02)00524-6

Blind separation of rotating machine sources: bilinear forms and ...leshema/papers/neurocomputing2002.pdf · In Section 2,we analyze the mixing processes relevant for blind source

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

ARTICLE IN PRESS

Neurocomputing ( ) –www.elsevier.com/locate/neucom

Blind separation of rotating machine sources:bilinear forms and convolutive mixtures

Alexander Ypmaa;b;∗, Amir Leshemc;d , Robert P.W. Duina

aPattern recognition group, Department of Applied Physics, TU Delft, Lorentzweg 1,2628 CJ Delft, The Netherlands

bSNN, Geert Grooteplein 21, 6525 EZ Nijmegen, The NetherlandscMetalink Broadband Access, Yakum Business Park, 60972, Israel

dCircuits and Systems group, Department of Electrical Engineering, TU Delft, Mekelweg 4,2628 CD Delft, The Netherlands

Abstract

We propose the use of blind source separation (BSS) for separation of a machine signaturefrom distorted measurements. Based on an analysis of the mixing processes relevant for machinesource separation, we indicate that instantaneous mixing may hold in acoustic monitoring. Wethen present a bilinear forms-based approach to instantaneous source separation. For simulatedacoustic mixing, we show that this method may give rise to a more robust separation. Forvibrational monitoring, a convolutive mixture model is more appropriate. The demixing algo-rithm by Nguyen Thi–Jutten allows for separation of the contributions of two coupled machines,both in an experimental setup and in a real-world situation. We conclude that BSS is afeasible approach for blind separation of distorted rotating machine sources. c© 2002 ElsevierScience B.V. All rights reserved.

Keywords: Blind source separation; Machine vibration; Bilinear forms; Convolutive mixture; Conditionmonitoring

1. Introduction

Many machines contain rotating elements, which causes that speci3c frequency com-ponents are present in the machine vibration spectrum. Deviations from normal behav-ior show up as distinct di5erences with respect to the normal spectrum (the machine

∗ Corresponding author.E-mail addresses: [email protected] (A. Ypma), [email protected] (A. Leshem).

0925-2312/02/$ - see front matter c© 2002 Elsevier Science B.V. All rights reserved.PII: S 0 9 2 5 -2312(02)00524 -6

2 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

;ngerprint). This is the basis for many techniques for machine diagnostics and con-dition monitoring [22]. In this paper, we investigate to what extent the contribu-tion of a machine source to a multichannel acoustic or vibration measurement canbe separated when using only general assumptions on the source signals and themixing process. Fault-related machine vibration is usually corrupted with structuralmachine vibration and noise from interfering machinery. In the presence of nearbynon-isolated machines (e.g. in ship engine rooms), severe interference may be presentin the measurements. Many well-known techniques from signal processing have beenapplied to interference removal [20]. They either rely on detailed knowledge aboutthe type of source that is to be identi3ed (e.g. an impulsive excitation, as with bear-ing failures) or on the identi3cation of the transmission characteristic (in a modalanalysis). In contrast, we take a blind approach, where we aim at separation ofmachine sources using only sensor measurements and general characteristics of thesupposed source signals. In an acoustical measurement setup, the machine sourcepropagates through the air and mixes on the sensor array with other sound sources.We explain the di5erence between both types of mixing in the nextsection.

In previous work [10,16] separation of the underlying sources in vibro-acousticalsystems was studied using a singular value decomposition of the cross-spectral ma-trix or using modal and eigenvector beamformers [17]. Gelle [13] concluded thatthe instantaneous mixing model did not hold for vibration measurements from themachine under consideration. Supposedly, this is related to the mechanical proper-ties (e.g. rigidity) and size of the structure under investigation. Subband 3ltering ofthe measurements may give rise to approximately instantaneous mixing. We proposedin earlier work to separate machine vibration sources using independent componentanalysis (ICA) [33] for cases where the instantaneous mixing assumption holds. Weused the MDL-based ICA algorithm [25] for this task; this algorithm is related tothe second-order blind identi3cation (SOBI) method by Belouchrani et al. [2], butdi5ers in spirit (it is based on the MDL-principle) and in the cost function that isminimized.

In Section 2, we analyze the mixing processes relevant for blind source sepa-ration (BSS) in a rotating machine monitoring context. In Section 3, we describethe blind separation problem and the algorithms we used in our later experiments.An important family of separation algorithms is the orthogonal source separationalgorithms that 3rst prewhiten the signals and then estimate the missingunitary matrix through joint diagonalization of a set of matrices. The bilinear formsapproach that we propose in this paper uni3es several orthogonal separation algo-rithms and allows for combinations of these algorithms. We then describe the NguyenThi–Jutten [23] algorithm for demixing of convolutive mixtures. In Section 4, wedemonstrate that the bilinear forms framework allows for combination of several sep-aration criteria, where the demixing may be more robust than with conventionalalgorithms. In Section 5 the BSS approach is validated in three di5erent setups withrotating machines, exhibiting both acoustic and vibrational mixing. We show thatthe relevant machine contribution can be separated out of the measurementsblindly.

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 3

2. Mixing of rotating machine sources

Nearly every machine will emit a certain amount of noise and vibration: machinesare often not perfectly balanced, contact between moving elements in the machine maynot be ideal, manufacturing imperfections will be present or vibration may occur as aconsequence of the operation [11]. Vibration sources in machines are usually eventsthat generate forces and motion during machine operation. These include [20]: shaftimbalance, impacts (e.g. due to bearing faults), Guctuating forces during gear mesh(as a result of small variations during machine operation or variations in local contactsti5ness), electromagnetic forces and pressure- and Gow-related sources. As a result,(series of) harmonic components will dominate the spectrum of the measurements.Particular contributions to machine vibration are caused by nonstationary or nonlinearphenomena, which can be characterized with time–frequency methods [19,24] or withmethods for cyclostationary signal analysis [4].

2.1. Acoustic mixing

In acoustic monitoring, the mixing process can sometimes be modelled as instanta-neous [28]. A narrowband acoustic signal s(t) that enters a sensor array from the far3eld will experience a time delay � when travelling from one sensor to the next, de-pending on the direction of arrival � and the interelement spacing � (in wavelengths).If this delay is small compared to the inverse of the bandwidth W of the signal, i.e.

W��1 (1)

the delayed signal can be modelled as a phase-shifted version of the “3rst” signal,where the phase shift � equals

� = ej2L sin �: (2)

If we have a sensor array with equispaced omnidirectional sensors at the locations�i; i = 1; : : : ; M (in wavelengths), the M -dimensional measurement vector x(t) can bemodelled as

x(t) = [e j 1 ; e j 2 ; : : : ; e j M ]Ta(�)s(t) = a(�)s(t); (3)

where the steering vector a(�) is expressed in terms of i(�),

i(�) = 2�i sin � (4)

and the sensor gain pattern a(�). If one considers the case where N independent(narrowband) sources at di5erent locations are impinging on the array simultaneously,a linear mixing matrix between sources and sensors can be postulated:

x(t) = As(t); A = [a(�1) : : : a(�N )]; (5)

where the mixing vectors ai are attenuated or ampli3ed steering vectors (assuming thatthe multipath propagation model holds). In practice, the narrowbandness assumption

4 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

has to be veri3ed. A useful heuristic is to check whether the product of source band-width W and maximum inter-element delay � is suMciently small. The planar waveassumption (i.e. the source is in the far 3eld) may not be valid if the sensor array isclose to a source. Moreover, the environment may cause reGections and scattering ofthe signal into multipath rays. For measurements in open or suburban terrain, the angleand delay spread due to these e5ects is expected to be small [28]. If one of the aboveassumptions does not hold, a convolutive mixture model should be used. The maindi5erence between (informed) beamforming techniques and methods for blind beam-forming and source separation is the criterion that is used for separation: beamforminguses the directivity of the sources and assumes that disturbances make nonzero angleswith the relevant sources. Blind source separation on the other hand assumes statisticalindependence or spatiotemporal uncorrelatedness among the source signals and doesnot exploit the parametric structure of the mixing matrix, only its linearity.

2.2. Mixing in vibrating structures

Vibration in machines takes the form of compressional (longitudinal) and bendingwaves mainly. Regardless the frequency, compressional waves travel at the same speed,whereas bending waves with di5erent frequencies travel at di5erent speeds. The latterphenomenon is called dispersion [20]; dispersion changes the waveform. Moreover, amachine will usually exhibit reverberation e5ects: multiple reGections and propagationalong di5erent paths in the structure. Dispersion and reverberation can severely dis-tort waveforms indicative of faults. The degree of coherence between two vibrationmeasurements at di5erent positions on a machine will vary from completely coherent(when the vibration is dominated by rigid body motion or when strong reverberationis present), to completely incoherent. In between those two extremes, the situationarises where both “spatial diversity and redundancy” exists between a set of machinevibration sensors; this is the case we focus on. Note that already in 1987, Lyon [20]recommended that adaptive systems should be developed that learn machine transferfunctions from arrays of accelerometers.

In damaged rotating machines, a characteristic signal that is indicative of wear orfailure will be repeated during each machine revolution and will be 3ltered at themachine casing with the local impulse response. The transmission path between asensor and a vibration source is hence modelled as a linear 3lter hij(t) [8,20]. For thegeneral situation with multiple sensors and multiple vibration sources (both internal tothe machine and external interferers), we model the response at position j as

xj(t) =Nf∑i=1

hFij(t)?Fi(t) +

Ni∑i=1

hIij(t)?Ii(t) + hS

j (t)?m(t) + nj(t); (6)

where ? denotes the convolution operation. The above expression is close to the expres-sion given in [8]. It is a convolutive mixture of Nf fault-related vibration sources Fi(t),vibration from Ni interfering machinery components Ii(t) and 3ltered modal machineresponse m(t) due to all remaining excitation sources arising with normal machineoperation and structural imperfections. We introduced a “structural 3lter” hS

j (t) that

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 5

Fig. 1. (a) (Left): Experimental setup with coupled pumps; (b) (right): contributions from various internaland external sources are measured at the machine casing.

accounts for the “attenuation” of a mode due to sensor measurement direction andposition (e.g. when putting a sensor on the position of a nodal line). By making theattenuation of certain modes explicit, we emphasize that the position of a sensor canhighly inGuence the actual measured response. We also include a set of interferingsources into the model. It is not obvious that these interferers (possibly from othermachines or remote machine components) appear as an additive contribution. In Section5.2, we investigate the suitability of this model in a setup with two coupled pumps, seeFig. 1(a). Depending on the purpose of the monitoring setup, a source can be designatedas a fault source or as an interference source. The ambient and sensor noise are givenby the nj(t) term. The various components of machine vibration measurements aredisplayed schematically in Fig. 1(b).

3. Methods for blind source separation

We give an overview of blind source separation in the next section, introduce thebilinear forms framework for instantaneous source separation and describe a methodfor convolutive demixing (simultaneous demixing and equalization).

3.1. Instantaneous mixing

Consider the case where one has access to multiple realizations x(t) of a mixtureof independent signals s(t). Assuming a linear mixing matrix A and allowing additivenoise n(t) leads to the model

x(t) = As(t) + n(t); (7)

where boldface variables denote random vectors with zero mean and 3nite covariance.Source vector s(t) is assumed to have statistically independent or spatiotemporallyuncorrelated components. In a real-world source separation problem, the vectors x(t)and s(t) live in Rn and Rm, respectively (or in Cn and Cm for certain man made oranalytic real-world signals) and the number of sources is assumed to be less than or

6 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

equal to the number of measurements (m6 n). The term n(t) represents white Gaussiannoise with covariance matrix �2I . The blind source separation problem (BSS) is now:given T realizations of x(t), estimate both mixing matrix A and the independent sourcess(t). Unless the noise covariance can be estimated, s(t) cannot be recovered exactly.Approximate reconstruction can be done with independent component analysis. Hereone recovers the sources s(t) with signals y(t)

y(t) = Wx(t) = s(t) (8)

such that y is a random vector whose components maximize a contrast function,chosen to be maximal when the components of the vector under investigation are sta-tistically independent. One can only recover the original sources up to scaling andpermutation, i.e. W = �PA−1, where � and P are a scaling and a permutation matrix,respectively. This means that the proper amplitudes of the signal cannot be retrievedwithout any prior knowledge on the sources. The contrast function is usually basedon the higher-order statistics (HOS) of the (empirical) distribution of the reconstructedsources [6,9,14]. If the sources do not show signi3cant skewness or kurtosis (non-gaussianity is not present or cannot be determined accurately due to low SNRs), thepresence of temporal coherence in the sources can be used to separate the sources whenthe sources have di5erent frequency spectra [2]. Second-order statistics are consideredmore robust than higher-order statistics [26,27].

3.1.1. BSS with bilinear formsA bilinear form g is a function g(·; ·) of two signals that expresses the degree of

similarity between the signals in terms of a generalized cross-correlation.Let U and V be vector subspaces over C.

De�nition 3.1. A function g(u; v) :U × V → C is called a bilinear form if it satis3esthe following properties:

(1) g(u1 + u2; v) = g(u1; v) + g(u2; v) for all u1; u2 ∈U; v∈V .(2) g(u; v1 + v2) = g(u; v1) + g(u; v2) for all u∈U; v1; v2 ∈V .(3) g(�u; v) = �g(u; v) for all u∈U; v∈V and �∈C.(4) g(u; �v) = �∗g(u; v) for all u∈U; v∈V and �∈C.

An example is the inner product bilinear form, denoted by g0(x1; x2)

g0(x1; x2) = limN→∞

1N

N∑t=0

x1(t)∗x2(t); (9)

where the ∗ operator denotes complex conjugation. We will describe other bilinearforms in the sequel. In order to be suitable for source separation, a bilinear form gmust satisfy

g(yk ; yl) = ck'kl; 16 k; l6m; (10)

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 7

where the yi; i = 1; : : : ; m, are components of the estimated source vector y. We havenow dropped the time index t. Moreover, with 'kl we denote the Kronecker deltaand ck is nonzero for all k. Both sources and mixtures are now considered to becomplex-valued random variables. When applied to a multidimensional vector y, abilinear form gives rise to a generalized covariance matrix:

g(y; y) = Rgyy =

g(y1; y1) · · · g(y1; ym)

......

g(ym; y1) · · · g(ym; ym)

: (11)

This matrix will be diagonal if the reconstructions yi are proper estimates of thesources. Bilinearity implies that

g(∑

�kxk ;∑

)lxl)

= �HRgxx) (12)

for a linear mixture x with components xi, i = 1; : : : ; n. In this formula, Rgxx is de3ned

analogous to Eq. (11), and parameter vectors � and ) are written as � = [�1; : : : ; �n]T

and ) = [)1; : : : ; )n]T.Bilinear forms can be used as criteria for source separation. A proper demixing

matrix W should separate the sources, which results in generalized covariance matricesthat are diagonal for all chosen bilinear forms. It is known that in the noiseless case,prewhitening of the data leaves a unitary transformation Q to be estimated in order toretrieve the demixing matrix W [2]. We need at least one more bilinear form (otherthan the inner product) to identify this unitary transformation. The separation algorithmnow consists of two steps:

Prewhitening: Let Vs = span{s1; : : : ; sm}. Find a basis z1; : : : ; zm of Vs such thatg0(zk ; zl) = 'kl. The z1; : : : ; zm components are called prewhitened measurements, sinceno cross-correlation exists between these components.Estimate Q: The remaining unitary transformation Q is found, such that

Q[z1; : : : ; zm]T = [s1; : : : ; sm]T.

The 3rst step can be performed using principal component analysis. The second stepis implemented through the use of joint diagonalization. To solve for the unitary factorwe compute Rgl

zz for l = 1; : : : ; M , and search for a unitary matrix Q such that QRglzzQ

H

are simultaneously (approximately) diagonal. This can either be implemented using theJADE algorithm [7] or using the joint Schur decomposition [29]. In JADE, a set ofconsecutive Jacobi rotations is performed in order to diagonalize the set of matricesRglzz approximately; because of 3nite sample size e5ects and noise these generalized

covariance matrices are never exactly diagonalizable. After the JADE operation, eachof the matrices individually need not be diagonalized exactly, but the criterion function

C(R; Q) =M∑l=1

o�(QRglzzQ

H ) (13)

8 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

is minimized. Here, the function o�(·) of a matrix is the sum of the squared non-diagonal elements, and R is the set of matrices to be (approximately) diagonalized.The consequence of prewhitening is that in3nite weight is put to the data correlationmatrix, which may lead to estimation errors in cases with small sample sizes or noise.Recently, an attempt has been made to approximately diagonalize a set of matrices withan arbitrary (not necessarily unitary) matrix Q [30]. In this paper we will not considerthis approach, although it would 3t well into our framework. The previous formulationallows the incorporation of several existing algorithms for source separation, dependingon the choice of bilinear form. We will mention three examples, which are used forsource separation in our experiments. More information is given in [18,31].

Temporal correlation: We obtain the SOBI algorithm [2] by choosing

gki(x1; x2) = limN→∞

1N

N∑t=0

x1(t)∗x2(t − ki): (14)

Time–frequency distributions: We obtain Belouchrani’s time–frequency algorithm[3] by choosing

gt;f(x1; x2) =∑m;l∈Z

’(m; l)x1(t + m + l)∗x2(t + m− l)e−4jfl (15)

using time–frequency atoms (t; f) and smoothing kernel /(m; l).Cyclostationarity: We note that the solution to the generalized eigenvalue problem

that arises with the phase-SCORE algorithm [1] is equivalent to the solution obtainedwith joint diagonalization of the bilinear forms

g�;k(x1; x2) = limN→∞

1N

N∑t=0

x1(t)∗x2(t − k)e−2j�t : (16)

This allows for an extension of the phase-SCORE algorithm: in the case of multiplefeatures (�i; ki), all we need to do is to jointly diagonalize the corresponding bilinearforms.

We note that also subband 3ltering approaches to blind source separation (e.g. [15])may be uni3ed in the bilinear forms framework [31].

3.2. Convolutive mixing

We have seen in Section 2 that sources in vibrating structures will both be 3lteredand mixed. The convolutive mixture model is then more appropriate:

xi(t) =n∑

j=1

{N∑

k=0

Aji(k)sj(t − k)

}; i = 1; : : : ; m; (17)

where A is now a matrix of FIR-3lters Aji and the sources and mixtures are againconsidered real-valued random variables. An additive noise term can easily be included

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 9

in the model. The aim is to 3nd the matrix of FIR-3lters that jointly demixes themeasurements (to undo the spatial mixing operation) and equalizes the sources (toundo the temporal 3ltering operation performed by the transmitting medium). Thisproblem can be solved in both the time domain [23,26] and the frequency domain[12,21]. The frequency domain methods often use the fact that a convolutive mixturecorresponds to an instantaneous mixture in each (suMciently narrow) frequency band[21].

3.2.1. Algorithm by Nguyen Thi and JuttenAn example of the time-domain approach to convolutive demixing is the algorithm

by Nguyen Thi and Jutten [23]. In this algorithm, a convolutive mixture of two sourcess1; s2 measured with two sensors x1; x2 (also referred to as a 2×2 convolutive mixture)is modelled in the Z-domain by

x1(z) = A11(z) · s1(z) + A21(z) · s2(z);

x2(z) = A12(z) · s1(z) + A22(z) · s2(z); (18)

where the 3lters Aji(z) are linear and causal FIR 3lters of order M :

Aji(z) =M−1∑k=0

aji(k) · z−k : (19)

The model is simpli3ed by assuming that each sensor is close to one source, whichleads to 3lters A11(z) = A22(z) = 1. The aim is now to estimate the inverse 3ltermatrix

B(z) =

[1 B21(z)

B12(z) 1

](20)

such that

B(z)A(z) =

[H11(z) 0

0 H22(z)

]: (21)

Note that the solution can be determined blindly only up to a permutation and a3ltering, i.e. a matrix B(z) that results in a product matrix B(z)A(z) with zeros andFIR 3lters Hii interchanged with respect to the above matrix also separates the sources,while any remaining FIR 3lter Hii suMces. The inverse matrix to be identi3ed consistsof FIR 3lters only and the inverse “close sensor” 3lters are constrained to unity. Thealgorithm performs an adaptive cancellation of higher-order cross-statistics, e.g. by3nding the zeros of E[s2

i (t)sj(t − k)] with the update rule

cij(t + 1; k) = cij(t; k) + 1s3i (t)sj(t − k); (22)

where cij(t; k) is the estimated demixing 3lter coeMcient at tap k and time t (using arecursive demixing architecture [23]) and 1 is the (positive) step size. It was shown in[5] that convolutively mixed harmonic signals may be separated on the basis of theirhigher-order statistics.

10 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

4. Combining di�erent bilinear forms

In the bilinear forms framework we can include several second-order criteria simulta-neously for source separation. We investigate the merits of this approach with simulateddirectional narrowband sources that exhibit both temporal coherence and cyclostation-arity. We generated two cyclostationary sources with temporal correlations, i.e. twocomplex modulated analytic AR(1)-sources with random initial phase. These sourceswere mixed such that the mixing simulates impinging on an array with 5 sensors atdirections of arrival of 12 and 13◦s. The length of the signals was 512 samples. Apartfrom direction of arrival, the source signals only di5er with respect to their modulationfrequency f1 = f0 + Lf. We chose f0 to be equal to 0:3fs, where the samplingfrequency fs was de3ned to be 1. The maximal frequency shift Lf was 0:1fs. Wecompare the results obtained with SOBI, SCORE and a combination of both methodsif we vary the frequency shift Lf. We repeat 250 runs of the source separation al-gorithm (SOBI, SCORE or their combination), and the median SINRs (in dB) over250 runs, averaged for both sources, are computed. We chose the set of delays as{1; : : : ; 25}, i.e. in both SOBI and SCORE we used 25 generalized covariance matri-ces (and 50 matrices in the combination algorithm). In Fig. 2(a) the result is plottedfor source SNR −5 dB. It can be seen that for small spectral di5erence only SCORE(denoted as cyclo in the 3gure) will be able to do the separation. For large spectraldi5erence (Lf¿0:01) there is suMcient information in the lagged covariance matrices,which results in better performance of SOBI. The combination algorithm (denoted ascyclosobi) follows the better performance of both methods: for small spectral shift itbehaves like SCORE, whereas for larger shifts it approaches SOBI. We noticed a largevariation in 250 repetitions of an experiment. In order to quantify this, we determinedthe cumulative density of the averaged SINR for both sources over 250 repetitions. InFig. 2(b) the empirical cumulative distribution for a small and a large spectral shift, i.e.

0 0.02 0.04 0.06 0.08 0.11

2

3

4

5

6

7

8

delta f (in f0)

aver

aged

med

ian

SIN

R (

dB)

SINR vs. spectral difference; noiselevel = _5 dB

cyclo sobi cyclosobi

_2 0 2 4 6 8 100

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

cdf’s at noiselevel _5 dB; ∆_ f1=0.001, ∆_ f2=0.1

SINR (dB)

prob

abili

ty d

ensi

ty

cy1 so1 socy1cy2 so2 socy2

Fig. 2. (a) (Left): Average median SNR vs. Lf for SOBI, SCORE (cyclo) and their combination (cyclosobi),SNR=−5 dB SNR, delays={1 : 25}; (b) (right): empirical cumulative distribution functions for small (dotted,denoted by cy1, etc.) and large (solid, denoted by cy2, etc.) frequency shift.

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 11

Lf = {0:001; 0:1}, is shown for SOBI, SCORE and their combination. For small shiftthe combination behaves like SCORE, whereas for large shift it behaves like SOBI. Inboth cases it follows the maximum performance. From several experiments we noticedthat the bene3t of combining SCORE with SOBI depends on a proper choice of timelags [31,32]. In [34] it was shown that for SOBI this choice may be related to thesubbands where the sources are spectrally di5ering, which may be used in the futureas a criterion for choosing the set of time lags. We conclude that with suitable choicesfor the set of time delays, combination of SCORE and SOBI for separation of sourcesthat exhibit both types of structure allows for approximation of the individually bestperformances in cases where the dominant signal structure (cyclostationarity or temporalcoherence) is not known a priori.

5. Separation of machine noise

5.1. Separation of acoustical sources

In an outdoor acoustical measurement setup in a suburb of a Dutch town, noise froma passing car was measured on a logarithmically spaced 5-microphone array. Severedisturbances were present like: a rotating machine running stationary around 60 Hz thatwas positioned at approximately 15 m from the 3rst sensor in the array at an angle of45◦; an airplane crossing the scene at high altitude; wind bursts from vegetation in thevicinity of sensors 4 and 5 in the array. The measurements were lowpass-3ltered, so thesources remaining in the measurements were wind, car and machine. The inter-elementspacing was such that beamforming techniques could not be used for source separation(severe spatial aliasing was present). Moreover, the fact that vegetation and car noisewere impinging on the sensor with similar directions of arrival was expected to prohibitthe use of beamforming for source separation. In the measurement setup, no signi3cantreGections from the environment or from the ground were expected. A calculation ofthe maximum W� product for the sources and the inter-sensor delays, assuming a maxi-mum frequency of interest of 75 Hz and a maximum direction of arrival of 45◦ revealedthat the instantaneous mixing assumption held approximately for the 3rst 4 sensors inthe array (W�≈0:7). A 3rst attempt at source separation using delay-and-sum beam-forming (we time aligned the signals with respect to the most dominant source usingcross correlation between the sensors) corroborated our expectation that beamformingis not suitable for this setup. In the spectra corresponding to the aligned measurements,mixing can still be observed, Fig. 3(a). We attempted further separation by using blindinstantaneous source separation with bilinear forms.

The time–frequency plot of a typical measurement is shown in Fig. 4(a). The rotatingmachine (stationary component around 60 Hz) was not visible in any time–frequencyplot of the measurements. After prewhitening we obtained spectra that contained severalcomponents where machine and car were severely mixed. Since one of the sources ismoving (the car), we can take the nonstationarity of the signal due to this source intoaccount. We applied the time–frequency algorithm of Eq. (15) with kernel /(m; l) = 1(so we used a Wigner distribution) to the measurements, using the 20 matrices that

12 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

0 20 40 60 80 100 120 140 160 180 2000

5

10x 10

5

0 20 40 60 80 100 120 140 160 180 2000

5

10x10 5

0 20 40 60 80 100 120 140 160 180 2000

1

2x106

0 20 40 60 80 100 120 140 160 180 2000

5

10x105

0 20 40 60 80 100 120 140 160 180 2000

1

2x106

File 43 _ measured

0 20 40 60 80 100 120 140 160 180 2000

500

0 20 40 60 80 100 120 140 160 180 2000

500

0 20 40 60 80 100 120 140 160 180 2000

500

1000

0 20 40 60 80 100 120 140 160 180 2000

200

400

0 20 40 60 80 100 120 140 160 180 2000

500

1000

File 43 _ separated

Fig. 3. (a) (Left): Spectra of outdoor acoustical measurements, after time-alignment; (b) (right): separatedcomponents using time–frequency BSS.

1 2 3 4 5 6 7 80

10

20

30

40

50

60

70

80

90

wv, lin. scale, contour, Threshold=0%

Time [s]

Fre

quen

cy [H

z]

1 2 3 4 5 6 7 80

10

20

30

40

50

60

70

80

90

wv, lin. scale, contour, Threshold=0%

Time [s]

Fre

quen

cy [H

z]

Fig. 4. Time–frequency plot of ((a) left) a typical measurement: and ((b) right) a component consisting ofrotating machine plus wind.

represent the most energetic tf-atoms. The signals were downsampled to approximately200 Hz and the time–frequency atoms had a width of 50 time samples and approx-imately 1:5 Hz. This resulted in the spectra of Fig. 3(b). The rotating machine isseparated into the 3fth component: harmonic peaks at 30 and 60 Hz are clearly dis-cernible. Its time–frequency signature is plotted in Fig. 3(b). The second componentis predominantly a wind component and the 3rst component contains mainly the carsignature, see Figs. 3(b) and 5(a). The component due to sensor 5 seems to be ignoredin the separation procedure: comparison of the tf-plots corresponding to component 5in Fig. 3(a) and component 2 in Fig. 3(b) revealed nearly identical tf-signatures. Thiscan be explained by the fact that sources registered on this sensor cannot be consid-ered as phase-shifted versions of registrations at other sensors. However, care shouldbe taken that the size of the signal subspace is estimated correctly. In our experimentwe assumed 5 spatially coherent sources, where there are only two present (along

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 13

1 2 3 4 5 6 7 80

10

20

30

40

50

60

70

80

90

Time [s]

Fre

quen

cy [H

z]

1 2 3 4 5 6 7 80

10

20

30

40

50

60

70

80

90

Time [s]

Fre

quen

cy [H

z]

wv, lin. scale, contour, Threshold=0% wv, lin. scale, contour, Threshold=0%

Fig. 5. Time–frequency plot of 2 independent components in acoustic measurements: (a) (left): car signature;(b) (right): spurious incomplete car signature.

with the colored noise due to wind and vegetation). The time–frequency signature cor-responding to component 4 is shown in Fig. 5(b). Since we forced a solution into5 components with distinct time–frequency behavior (according to the most energetictime–frequency descriptors), di5erent temporal parts of the car signature are repeated inspuriously reconstructed components 3 and 4. The important point to note is, however,that the rotating machine component (the main spatially coherent interference source)could be separated largely from the car noise. We remark that similar decompositionswere obtained with di5erent choices for temporal bandwidth (10 and 500 time samples,respectively) and number of included tf-atoms (100) in the time–frequency algorithm.Application of conventional separation algorithms like SOBI and JADE also led tocomparable results.

5.2. Demixing of two connected pumps

In a laboratory test bench, a small water pump was attached to a larger submersiblepump, see Fig. 1(a). The small pump runs at 100 Hz. The running speed of the largerpump is 1500 RPM at a nominal driving frequency (of the frequency converter drivingthe machine) of 50 Hz. Due to slip and load di5ering from the nominal load, this resultsin an e5ective fundamental frequency around 28 Hz at a driving frequency of 60 Hz.In the spectrum of the submersible pump, harmonics of this fundamental frequencyare clearly visible. In Fig. 6(a) the uppermost spectrum corresponds to a measurementon the small pump, where the small pump is the only operating machine. The bottomtwo spectra are measurements at two di5erent positions on the large pump, whereonly the large pump was running. Here, channel 1 denotes measurements from thesensor on the small pump, channels 2–4 represent orthogonal measurement directionsfrom a triaxial sensor that is mounted on the middle of the large pump and channel 5is from a sensor at the bottom (the vane) of the large pump. When both pumps arerunning simultaneously, the measured spectra on the small machine and on the positioncorresponding to channel 3 on the large machine are shown in Fig. 6(b). Since thereare only two sources present, we could apply the Nguyen Thi–Jutten algorithm to pairsof mixed sources. Each time, the mixture signal from the small pump was paired with

14 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

0 20 40 60 80 100 120 140 1600

0.5

1

1.5

Spectrum of individual sources at 60 Hz, channels 1, 3 and 5

0 20 40 60 80 100 120 140 1600

2

4

0 20 40 60 80 100 120 140 1600

1

2

0 20 40 60 80 100 120 140 1600

2

4

6

8

10

0 20 40 60 80 100 120 140 1600

1

2

3

4

Spectrum of measurements at 60 Hz, channels 1 and 3

Fig. 6. (a) (Left): Measurements at driving frequency 60 Hz while one pump is running, on small pump(top) and two positions (middle: channel 3; bottom: channel 5) on large pump; (b) (right): measurementswhile both pumps are running simultaneously, on small pump (top) and on large pump (bottom, channel 3).

0 20 40 60 80 100 120 140 1600

1

2

3

4

5

0 20 40 60 80 100 120 140 1600

0.5

1

1.5

2

2.5

Spectrum of unmixed sources at 60 Hz, channels 1 and 2, order 14

0 20 40 60 80 100 120 140 1600

1

2

3

4

5

0 20 40 60 80 100 120 140 1600

1

2

3

4

5

6

Spectrum of unmixed sources at 60 Hz, channels 1 and 3

Fig. 7. Reconstruction of small pump (top 3gures) and machine signature (bottom 3gures) on two di5erentpositions on large pump: (a) (left): channel 2; (b) (right): channel 3.

a mixture signal from the large pump. The demixing results using a 3lter order of 14and large pump channels 2 and 3 are shown in Figs. 7(a) and (b), respectively. It canbe observed that the harmonic series due to the large pump is separated from the smallpump contribution. We repeated the experiments on measurements with the large pumprunning at 23 Hz (due to a driving speed of 50 Hz, slip and non-nominal load). Themeasured spectra when both machines are running are shown in Fig. 8(a). Note that1st, 3rd and 4th harmonics of running speed are clearly discernible at both machines.The separated signatures using a 3lter order of 14 are shown in Fig. 8(b). Thesedecompositions are representative of most decompositions we obtained using di5erentsensor combinations. Sometimes there remains a small amount of cross-talk in thereconstruction obtained with channel 5, which is possibly due to the fact that channel

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 15

0 20 40 60 80 100 120 140 1600

1

2

3

4

5

6

0 20 40 60 80 100 120 140 1600

0.5

1

1.5

2

Spectrum of measurements at 50 Hz, channels 1 and 5

0 20 40 60 80 100 120 140 1600

1

2

3

4

5

6

0 20 40 60 80 100 120 140 1600

0.5

1

1.5

Spectrum of unmixed sources at 50 Hz, channels 1 and 5

Fig. 8. (a) (Left): Measurements at driving frequency 50 Hz, while both machines are running simultaneously,on small pump (top) and on large pump (bottom, channel 5); (b) (right): reconstruction of small pump (top)and machine signature (bottom, channel 5).

5 is more remote to the oilpump than channels 2 and 3. In general, it appears that themain bene3t is in demixing the big pump signature from the smaller pump (gains upto 2 dB can be obtained); demixing of the (usually small) contribution of the smallpump to the big pump can lead to a somewhat distorted big pump signature, whichmay o5set the demixing gain (in terms of SINR). We also noticed that instantaneousdemixing leads to very unsatisfactory results for this setup.

5.3. Vibrational artifact removal

Vibration was measured on a pumpset in pumping station “Buma” at Lemmer (TheNetherlands), where a gearbox fault (severe pitting in a gearwheel) was present [31].Mounted onto the gearbox casing was a small oil pump. We measured the vibration onthe gearbox casing and on the oilpump casing, when only the oilpump was running.In Fig. 9(a) the corresponding vibration spectra are shown. Channel 3 (top subplot)is measured on the gearbox, channel 7 (bottom subplot) is measured on the oilpump.Measurement channels 1 and 3 correspond to a sensor position near the driving shaftof the gearbox, channels 4 and 5 correspond to positions at the top and bottom of thegearbox casing. The oilpump predominantly emits vibration at multiples of 25 Hz. Theoilpump contribution is observable on the large machine casing as well, though smallin amplitude. If both machines are running simultaneously, we measure the spectra ofFig. 9(b) on several positions at the large machine (top three plots) and at the oilpump(bottom plot).

5.3.1. Instantaneous separationThe signals were downsampled to approximately 400 Hz and an inspection of the

eigenvalues of the data covariance matrix revealed that the signal subspace contained2 to 3 components, see Fig. 10(a). After performing time–frequency ICA on the mul-

16 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

0 50 100 150 200 250 300 350 4000

2

4

6

8

0 50 100 150 200 250 300 350 4000

100

200

300

400

500

600

Spectrum of oilpump, channels 3 and 7

0 50 100 150 200 250 300 350 4000

100

200

0 50 100 150 200 250 300 350 4000

500

1000

0 50 100 150 200 250 300 350 4000

100

200

0 50 100 150 200 250 300 350 4000

500

1000

Machine vibration measurements_measured

Fig. 9. (a) (Left): Vibration spectrum due to oilpump, measured at gearbox (top, channel 3) and oilpump(bottom, channel 7); (b) (right): vibration with both machines running simultaneously, measured at gearbox(top three plots, channels 1, 3, and 5) and oilpump (bottom, channel 7).

0 50 100 150 200 250 300 350 4000

200

400

600

0 50 100 150 200 250 300 350 4000

200

400

600

0 50 100 150 200 250 300 350 4000

200

400

Machine vibration measurements _ prewhitened

0 50 100 150 200 250 300 350 4000

200

400

600

800

0 50 100 150 200 250 300 350 4000

100

200

300

400

0 50 100 150 200 250 300 350 4000

500

1000

Machine vibration measurements _ separated

Fig. 10. (a) (Left): Spectra of prewhitened vibration measurements; (b) (right): separated components usinginstantaneous tf-separation algorithm.

tichannel measurement (spectral resolution in algorithm was 1024 bins and 50 mostenergetic atoms were used), the decomposition shown in Fig. 10(b) was obtained. The3rst component consists mainly of the gearmesh-related component at 285 Hz (whichwas high due to the gearbox fault), whereas the second component contains predom-inantly structural vibration of the oil pump. The third component contains mainly thefundamental gearmesh frequency (216 Hz). This component was measured on the oilpump as well, see Fig. 9(b), 4th channel, but suppressed in the “cleansed” oil pumpmeasurement, see Fig. 10(b), 2nd channel. Prewhitening is clearly insuMcient for sourceseparation, see Fig. 10(a), since the peaks at 216, 285 and around 100 Hz are present

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 17

0 50 100 150 200 250 300 350 4000

100

200

300

400

500

0 50 100 150 200 250 300 350 4000

100

200

300

400

500

Spectrum of Buma measurements, channels 3 and 7

0 50 100 150 200 250 300 350 4000

100

200

300

400

500

0 50 100 150 200 250 300 350 4000

50

100

150

200

250

300

Spectrum of Buma reconstructions, channels 3 and 7, order = 22

Fig. 11. (a) (Left): Measurements while both machines are running simultaneously, on gearbox casing (top)and oilpump (bottom); b. (right): reconstructed gearbox signature (top) and oilpump signature (bottom).

in all three components. We noticed that proper choice of frequency band and timeshift was critical for the performance of the algorithm: bad choices did not allow forsigni3cant improvement over PCA. In this decomposition, the relevant gearmesh-relatedfrequencies are removed from the oilpump signature. However, both frequencies dueto gearmesh are distributed over two di5erent components, which may indicate thatsigni3cant time delays are present between di5erent registrations of the gearmesh com-ponents on the sensor array. This suggests that for the mechanical structure at handthe mixing process should be modelled as convolutive.

5.3.2. Convolutive separationWe now assumed two underlying sources and applied the Nguyen Thi–Jutten al-

gorithm to pairs of mixtures, analogous to the laboratory experiment with the twoconnected pumps. In Fig. 11(a) the mutual cross-talk on both machines is again vis-ible from the spectra measured at two of the positions shown in Fig. 9, where bothmachines were running simultaneously. A convolutive demixing with 3lter order 22yielded the reconstructions shown in Fig. 11(b). The oilpump contribution to the largemachine is suppressed slightly; however, the removal of the large peaks at 216 and285 Hz (related to gearmesh) from the oilpump signature is much more signi3cant.Note that the small peak around 210 Hz in the oilpump reconstruction should not beconfused with the gearmesh component at 216 Hz. The demixing results depend on aproper choice of the 3lter order. Demixing with a 3lter order of 45 gives comparableresults. Demixing with much smaller or much larger 3lter order yields unsatisfactoryor unstable results for the 2 × 2 convolutive demixing case, which gives an indicationof the time constants in the mixing process of the system. The (spurious) separationof the 216 and the 285 Hz components in the 3 × 3 (3 sources, three sensors) instan-taneous demixing case of the previous section indicates that a proper estimation of thenumber of underlying sources and use of the proper mixing model is important forproper blind separation of machine signatures.

18 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

6. Conclusion

When monitoring rotating machines, it is valuable to determine a machine signaturethat is free from interfering noise due to neighboring machines or due to the envi-ronment. This will allow a focus on the relevant spectral signature without spectralmasking e5ects caused by interferers (that may cause false alarms). We use the spatialdiversity and redundancy in a multichannel measurement of machine noise to sepa-rate the machine signature blindly from disturbing noise sources in the scene. Basedon an analysis of the mixing process that takes place in mechanical structures and insound propagation, we indicate that both instantaneous and convolutive mixtures maybe obtained in machine monitoring. We showed that using the bilinear forms frame-work it is possible to include several types of signal characteristics simultaneously forseparation of instantaneously mixed sources. This can be bene3cial when the sourcesexhibit several types of structure and it is a priori not clear under which conditionseach criterion can be used for separation. Speci3cally, it allows the use of particu-lar rotating machine-related signal structure like cyclostationarity and time–frequencystructure. For simulated directional sources, we showed that combination of di5erentsignal characteristics may give rise to a more robust separation procedure.

In three case studies we showed experimentally that the relevant machine compo-nent can be separated blindly out of a set of acoustical or vibration measurements. Inthe former case, a (stationary) rotating machine source could be separated from thecontribution due to a (nonstationary) car using the time–frequency algorithm, whichcan be seen as an instance of the bilinear forms framework. In the latter (convolu-tive mixing) case, demixing results improve signi3cantly when a convolutive demixingalgorithm is employed. Both observations are in accordance with our prior analysisof the mixing processes in rotating machine applications. We conclude that BSS is afeasible approach for blind separation of distorted rotating machine sources.

Acknowledgements

This work is partially supported by the Dutch Technology Foundation STW, ProjectNo. DTN-44.3584 “Machine diagnostics by neural networks”, by TNO-TPD b.v. (Delft)and the Dutch Ministry of Housing, Spatial planning and Environment (Depart. ofVROM). We like to thank Ronald Ligteringen (TU Delft), TechnoFysica b.v. (Baren-drecht) and Landustrie b.v. (Sneek) for assistance with measurements, and WillemKeijzer (RND Mechanical Engineering b.v. Delft) and Guillaume Gelle (ReimsUniversity, France) for fruitful discussions. Moreover, we thank the anonymous re-viewers for many useful remarks.

References

[1] B.G. Agee, S.V. Schell, W.A. Gardner, Spectral self-coherence restoral: a new approach to blind adaptivesignal extraction using antenna arrays, Proc. IEEE 78 (4) (1990) 753–767.

ARTICLE IN PRESSA. Ypma et al. / Neurocomputing ( ) – 19

[2] A. Belouchrani, K. Abed-Meraim, J.-F. Cardoso, E. Moulines, A blind source separation technique usingsecond-order statistics, IEEE Trans. Signal Process. 45 (2) (1997) 434–444.

[3] A. Belouchrani, M. Amin, Blind source separation using time-frequency distributions: algorithm andasymptotic performance, in: Proceedings of the ICASSP97, Vol. 5, Munchen, 1997, pp. 3469–3472.

[4] C. Capdessus, M. Sidahmed, J.L. Lacoume, Cyclostationary processes: application in gear faults earlydiagnosis, Mech. Systems Signal Process. 14 (3) (2000) 371–385.

[5] V. Capdevielle, Ch. ServiUere, J.-L. Lacoume, Blind separation of wideband sources: application torotating machine signals, in: Proceedings of the EUSIPCO ’96, 1996, pp. 2085–2088.

[6] J.F. Cardoso, High-order contrast for ICA, Neural Comput. 11 (1) (1999) 157–192.[7] J.F. Cardoso, A. Souloumiac, Blind beamforming for non-gaussian signals, IEE Proc. F 140 (6) (1993)

362–370.[8] C. Cempel, Vibroacoustic Condition Monitoring, Ellis Horwood, Chichester, UK, 1991.[9] P. Comon, Independent component analysis—a new concept? Signal Process. 36 (3) (1994) 287–314.

[10] J.M. Danthez, R. Aquilina, Separation of broadband sources concept of the Labrador software, Mech.Systems Signal Process. 11 (1) (1997) 91–106.

[11] I.K. Epps, An investigation into vibrations excited by discrete faults in rolling element bearings, Ph.D.Thesis, University of Canterbury, New Zealand, 1991.

[12] P. Fabry, Ch. ServiUere, Blind separation of noisy harmonic signals using only second-order statistics,in: Proceedings of the ICECS’99, Cypress, September 1999.

[13] G. Gelle, Les statistiques d’ordre superieur appliquees a la detection et a la separation de sources,Utilisation en analyse vibratoire et acoustique, Ph.D. Thesis, Universite de Reims Champagne-Ardenne,1998.

[14] A. HyvXarinen, E. Oja, A fast 3xed-point algorithm for independent component analysis, Neural Comput.9 (7) (1997) 1483–1492.

[15] B.-U. Koehler, R. Orglmeister, ICA of electroencephalographic data using wavelet decomposition, in:Proceedings of the VIII Mediterranian Conference on MBEC, Cyprus, 1998.

[16] M.S. Kompella, P. Davies, R.J. Bernhard, D.A. U5ord, A technique to determine the number ofincoherent sources contributing to the response of a system, Mech. Systems Signal Process. 8 (4)(1994) 363–380.

[17] W.A. Kuperman, G. Turek, Matched 3eld acoustics, Mech. Systems Signal Process. 11 (1) (1997)141–148.

[18] A. Leshem, Source separation using bilinear forms, in: Proceedings of the Eighth InternationalConference on Higher-Order Statistical Signal Processing, Aussois, France, 1999.

[19] P.J. Loughlin, G.D. Bernard, Cohen-Posch (positive) time-frequency distributions and their applicationto machine vibration analysis, Mech. Systems Signal Process. 11 (4) (1997) 561–576.

[20] R.H. Lyon, Machinery Noise and Diagnostics, Butterworths, Boston, 1987.[21] C. Mejuto, J.C. Principe, A second-order method for blind source separation of convolutive mixtures,

in: Proceedings of ICA’99, 1999.[22] J.S. Mitchell, An Introduction to Machinery Analysis and Monitoring, 2nd Edition, PennWell Publishing

Co., Tulsa, USA, 1993.[23] H.L. Nguyen Thi, Ch. Jutten, Blind sources separation for convolutive mixtures, Signal Process. 45 (2)

(1995) 209–229.[24] H. Oehlmann, Analyse temps-frYequence de signaux vibratoires de boi tes de vitesses, Ph.D. Thesis,

UniversitYe H. PoincarYe, C. de Recherche en Automatique, Nancy, 1996.[25] P. Pajunen, Blind source separation using algorithmic information theory, Neurocomputing 22 (1998)

35–48.[26] L. Parra, C. Spence, Convolutive blind separation of nonstationary sources, IEEE Trans. Speech Audio

Process. 8 (3) (2000) 320–327.[27] I. Schiessl, M. Stetter, J.E.W. Mayhew, S. Askew, N. McLoughlin, J.B. Levitt, J.S. Lund, K. Obermayer,

Blind separation of spatial signal patterns from optical imaging records, in: Proceedings of the ICA’99,1999, pp. 179–184.

[28] A.J. van der Veen, Algebraic methods for deterministic blind beamforming, Proc. IEEE 86 (10) (1998)1987–2008.

20 A. Ypma et al. / Neurocomputing ( ) –

ARTICLE IN PRESS

[29] A.-J. van der Veen, A. Paulraj, An analytical constant modulus algorithm, IEEE Trans. Signal Process.44 (5) (1996) 1136–1155.

[30] A. Yeredor, Approximate joint diagonalization using non-orthogonal matrices, in: Proceedings of theICA2000, Helsinki, 2000, pp. 33–38.

[31] A. Ypma, Learning methods for machine vibration analysis and health monitoring, Ph.D. Thesis, PatternRecognition Group, TU Delft, The Netherlands, 2001. Download: http://www.mbfys.kun.nl/∼ypma/papers/list of papers.html.

[32] A. Ypma, A. Leshem, Blind separation of machine vibration with bilinear forms, in: Proceedings ofICA-2000, Helsinki, June 2000, pp. 405–410.

[33] A. Ypma, P. Pajunen, Rotating machine vibration analysis with second-order independent componentanalysis, in: Proceedings of the ICA’99, Aussois, January 1999, pp. 37–42.

[34] A. Ziehe, G. Nolte, G. Curio, K.-R. MXuller, Optimal 3ltering algorithms for source separation, in:Proceedings of the ICA2000, Helsinki, 2000, pp. 127–132.

Alexander Ypma was born in The Netherlands in 1971. He received the B.Sc. in computer science fromTwente University in 1990 and the M.Sc. in computer science from Groningen University in 1995. In1996, Ypma started his Ph.D. research at the Pattern Recognition Group of Delft University of Technologyunder supervision of Dr. R.P.W. Duin, on the topic of “Machine diagnostics by neural networks”. He hassuccessfully defended his Ph.D. thesis at Delft UT in 2001. Currently, he is working as a postdoc at theFoundation for Neural Networks SNN at Nijmegen University. Ypma’s scienti3c interests are in the 3eldsof learning methods, Independent Component Analysis and dynamical systems. He is also interested inapplications like (medical and industrial) health monitoring, sound=speech=music recognition and recognitionof brain signals EEG, MEG.

Amir Leshem received the B.Sc. (cum laude) in m athematics and physics, the M.Sc. (cum laude) in math-ematics, and the Ph.D. in mathematics all from the Hebrew University, Jerusalem, Israel, in 1986,1990 and1998 respectively.From 1984 to 1991 he served in the IDF. From 1990 to 1997 he was a researcher withRAFAEL signal processing center. From 1992 to 1997 he was also a teaching assistant at the Institute ofMathematics, Hebrew University, Jerusalem. From 1998–2000 he was with Faculty of Information Tech-nology and Systems, Delft University of Technology, The Netherlands, as a postdoctoral fellow workingon algorithms for the reduction of electromagnetic interference in radio-astronomical observations and sig-nal processing for communication. Since September 2000 he is with Metalink Broadband Access where heis director of advanced technology. He is responsible for the research and development of new modemtechnologies. He is also a visiting researcher at Delft University of Technology.His main research interestsinclude transmission over copper lines, array and statistical signal processing, radio-astronomical imagingmethods, set theory, logic and foundations of mathematics.

Robert P.W. Duin studied applied physics at Delft University of Technology in the Netherlands. In 1978 hereceived the Ph.D. degree for a thesis on the accuracy of statistical pattern recognizers. In his research heincluded various aspects of the automatic interpretation of measurements, learning systems and classi3ers.Between 1980 and 1990 he studied and developed hardware architectures and software con3gurations forinteractive image analysis.At this moment he is an associate professor of the Faculty of Applied Sciencesof Delft University of Technology. His present research interest is in the design and evaluation of learningalgorithms for pattern recognition applications. This includes in particular neural network classi3ers, sup-port vector classi3ers and classi3er combining strategies. Recently, he started to study alternative objectrepresentations for classi3cation and became thereby interested in the use of relational methods for patternrecognition.