You are on page 1of 16

ARTICLE IN PRESS

Mechanical Systems
and
Signal Processing
Mechanical Systems and Signal Processing 19 (2005) 865–880
www.elsevier.com/locate/jnlabr/ymssp

Structural damage diagnosis under varying environmental


conditions—part II: local PCA for non-linear cases
A.-M. Yan, G. Kerschen, P. De Boe, J.-C. Golinval
University of Liege, Department of Aerospace, Mechanics and Materials, LTAS-Vibrations and
Identification of Structures, Chemin des Chevreuils 1, B-4000 Liege, Belgium
Received 20 October 2003; accepted 1 December 2004
Available online 5 February 2005

Abstract

It is well known that changes in vibration features of structures due to damage may be masked by the
effects of environmental variations. This influence has to be eliminated in the structural health-monitoring
process, especially when a long-term in situ monitoring is expected. In the companion paper [1] a linear
method based on principal component analysis (PCA) has been proposed and has shown encouraging
results for linear or even weakly non-linear cases. The present paper concerns a further extension of the
proposed method to handle non-linear cases, which may be encountered in some complex structures. The
method involves a two-step procedure, namely a clustering of the data space into several regions and then
the application of PCA in each local region. The application of local PCA allows performing a piecewise
linearisation of the non-linear problem. A close look at the choice of the distortion function used in data
clustering leads to two new clustering strategies. Whereas the first strategy is specifically suitable for the
application treated in this paper, the second one is more general. The local PCA-based damage detection
method is applied for the structural health monitoring of a real bridge using vibration data measured in situ
over a one-year period.
r 2005 Elsevier Ltd. All rights reserved.

Keywords: Structural health monitoring; Effects of environmental variations; Local principal component analysis;
Bridge diagnosis

Corresponding author. Tel.:+32 4 366 91 77; fax: +32 4 366 4856.


E-mail address: jc.golinval@ulg.ac.be (J.-C. Golinval).

0888-3270/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ymssp.2004.12.003
ARTICLE IN PRESS

866 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

1. Introduction

The aim of vibration-based structural health-monitoring (SHM) methods is to detect


the appearance of damages by evaluating changes in the identified vibration characteristics.
However, environmental variations which unavoidably occur during the different time periods of
data acquisition, also cause changes in the vibration features. Therefore, it is of primary
importance to take into account the masking effect of environmental variations in the diagnostics.
In a companion paper [1], the authors have presented a linear approach based on principal
component analysis (PCA) of vibration features identified during the monitoring of the structure.
The advantage of this method lies in the fact that the environmental variables (such as
temperature, temperature gradients, wind, humidity, etc.) do not need to be measured and that
their quantitative effect on the features does not need to be known. The method has been
successfully applied to various numerical and experimental situations in linear or weakly non-
linear cases.
PCA may be viewed as an optimal linear tool in the sense that it allows to minimise the distance
in the least squares sense between the original signals and their dimension-reduced representation.
In structural dynamics, PCA has been extensively applied to measured vibration signals for
dimensionality studies [2], reduced-order modelling [3], modal analysis [4] and model updating of
non-linear systems [5], etc. In the companion paper [1], it has been applied to remove the influence
of the environmental effects from the vibration features.
Although PCA has also been used for the analysis of non-linear systems, it should be borne in
mind that it is in essence a linear tool. Accordingly, it may sometimes be too simple for dealing
with real-world data. One way for improving its capability to deal with non-linearity is to increase
the order of the PCA model (i.e., to take more principal components). As shown in the companion
paper [1], this may lead to better results in SHM with environmental effects. However, this
approach may not always be sufficient and it is contrary to the objective of dimension reduction.
To address this issue, global non-linear variants of PCA have been proposed in the literature.
Sohn et al. [6] used an auto-associative neural network proposed by Kramer [7] to train the data
measured in varying environmental conditions. Target outputs are simply inputs to the network
and the bottleneck layer of network captures the embedded environmental factors. The method
was used to analyse a simulated computer hard disk drive taking the coefficients of the transfer
function as vibration features. However, as pointed out by Kambhatla and Leen [8], for general
applications, training a network may be expensive and sometimes it may fail to obtain a better
result than ordinal PCA due to trapping in poor local optima. Therefore, experience and prudence
are necessary while using this method. Some other non-linear PCA alternatives, such as principal
curves [9], were also proposed in the literature.
In this paper, our attention is directed towards a vector quantisation technique used in
local PCA [8,10]. The method involves a two-step procedure, namely a clustering of the data
space into several regions and then the application of PCA in each local region. From a
numerical point of view, this leads to piecewise linearisation of a non-linear problem. Specially,
a close look at the choice of the distortion function for data clustering leads to two new data
clustering strategies. The methods are illustrated by using numerical examples. Finally, a real
application of long-term damage monitoring under varying environmental conditions is
presented.
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 867

2. Local principal component analysis

Consider a set of n-dimensional data Y 2 <nN ; where N is, in a general case, the number
of sampling points. In this work, data is defined as the identified natural frequencies of a structure,
n denotes the number of vibrating modes and N corresponds to the number of modal
identifications. Let us partition the data space into q subregions Y ¼ (Y1,yYq). Each
subregion data set is described by a hyperplane, which results from the application of local
PCA. The key issue in this procedure is how to properly cluster the data regions. Following
the generalised Lloyd algorithm [11], the partition of q regions with centroid (also called
codebook) Ȳ ¼ ðȲ 1 ; . . . Ȳ q Þ satisfies the condition that the jth subregion data Yj consists of any
data y 2 <n that is closer to Ȳ j than to any other centroid. The distance may be measured
using the Euclidean distance or, more reasonably, the reconstruction error [8]. Since the
initially chosen codebook is not exactly a centroid in each formed region, an iterative process is
generally necessary. Such a partition in subregions may be performed several times with each
time a randomly chosen initial codebook in order to find the best partition. However, such a
data partition by Euclidean distance criterion [11], or by reconstruction error criterion [8] is
not always optimal; this will be discussed later in this paper. In the following sections, two
new data-partition strategies are proposed: the first one is appropriate for damage diagnosis
under changing environmental conditions; the second one is rather a general technique for
extensive applications.

2.1. Partition in hyper-sphere model based on data structure observation

The data partition proposed here is based on the observation of the n-dimensional data
structure. Let us recall that the data considered in this discussion are the identified natural
frequencies of n modes. Firstly, (n1) projections on two-dimensional subspaces (yi : y1 ), here
i ¼ 2; . . . n; are constructed in order to observe l possible non-linearity tuning points Lj 2 <n ;
j ¼ 1; . . . ; l: An example is shown in Fig. 1 in the case of n ¼ 2 and l ¼ 2: Since non-linearity in
each 2D diagram is not always similar, a synthetical (or average) assessment is sometimes
necessary for determining Lj : The subregions Yj are defined by the condition:
Yj ¼ fy : jjLj1 jjojjyjjojjLj jj; (1)
where j ¼ 1 . . . q; q ¼ l þ 1; q is the number of subregions; || || denotes Euclidean norm; by
definition, ||L0|| ¼ 0 and ||Lq|| ¼ N. In this way, the data are partitioned into q subregions in the
form of hypersphere. Obviously, in each subregion data may be better approximated by local
PCA. Note that the first principal component has been assumed dominant in data partitioning,
which is reasonable in most situations. It is also emphasised that the number of subregions should
be as few as possible in order that data in each region distribute along local principal directions in
a global sense. Taking too many subregions may result in over-fitting of noise, which will be
illustrated later by numerical examples.
Once the data partition has been performed, the singular value decomposition (SVD) may be
performed on the local covariance matrix:
ðYj  Ȳ j ÞðYj  Ȳ j ÞT ¼ Uj R2j UTj ; (2)
ARTICLE IN PRESS

868 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

2
y
PC-I 2
PC-I 3

L2
+
YIII

× ×
×× ×
× × ×
× × ×
× × PC-I 1
× ×
×
×× +
YII

× ×
L1

YI

O 1
y

Fig. 1. Proposed data partition method 1 for local PCA: ‘o’ reference data; ‘  ’ damaged data (2D case, three
subregions and one principal component considered).

where Ȳ j is the mean co-ordinate or centroid in the jth subregion; U is an orthonormal matrix, the
columns of which form a hyperplane of principal components; R is a diagonal matrix of singular
values indicating the energies associated with the corresponding principal components. Taking the
first m columns of Uj, where m is the number of principal components corresponding to the
number of independent environmental effects to build the loading matrix Tj (see Ref. [1]), the
measured features in each subregion are projected into an environmental-factor characterised
space:
Xj ¼ Tj ðYj  Ȳ j Þ; (3)
where X 2 <mN is called the scores matrix. The important meaning of this matrix in the present
work is that feature variations due to environmental effects should occur in (or around) this X-
defined hyperplane and only damages could drive features to deviate away. Therefore, damages
may be detected by observing and comparing the feature deviations in potential damage state with
respect to the reference state.
The loss of information in the above projection can be assessed by re-mapping the projected
data back to the original space
^ j ¼ TT Xj ¼ TT T j ðYj  Ȳ j Þ þ Ȳ j :
Y (4)
j j
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 869

The residual error matrix is estimated as


^ j:
Ej ¼ Y j  Y (5)
The Novelty Index (NI) may be defined by the Euclidean norm:
NI Ek ¼ jjEj ðkÞjj (6)
or by the Mahalanobis norm
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
M
NI k ¼ ETj ðkÞR1 j Ej ðkÞ; (7)
where R ¼ ð1=N j ÞðYj  Ȳ j ÞðYj  Ȳ j ÞT ; k ¼ 1; . . . ; N j ; with N j the number of data in the jth region.
Assembling NI of all subregions forms a complete novelty index series. The partition of data
from the damaged structure is made in the same way as for the reference data. Finally, damage
detection is realised using a similar procedure as for ordinal PCA in linear or weakly non-linear
cases [1]. It should be pointed out that when using local PCA for damage detection, the reference
structure should be measured and identified in a full range of environmental conditions in order to
capture existing non-linear information of features for all possible environmental variations.

2.2. Partition based on a combined criterion of Euclidean and reconstruction norms

The method proposed in Section 2.1 is simple but it requires to define subjectively the non-
linearity turning points based on an observation of the data structure. For a more extensive
application of local PCA, such as data or model reduction, an automatic and general partition
method is certainly more appropriate. In what follows, two existing algorithms are first
introduced and assessed: the first one concerns a generalisation of Lloyd’s algorithm (Euclidean
partition) [11] and the second one is Kambhatla and Leen’s algorithm (Projection partition) [8].
Then an improvement is proposed.
Suppose the data are to be partitioned into q subregions. The generalised Lloyd algorithm
based on Lloyd’s optimality condition is the simplest way for achieving this goal.

1. Each region Yj consists of any data y that lies closer to a centroid Ȳ j than to any other centroid
Ȳ k for k ¼ 1; . . . ; q but kaj: Mathematically, one has
Yj ¼ fy : d E ðy; Ȳ j Þod E ðy; Ȳ i Þ 8jaig; (8)
E
where d ðy; Ȳ j Þ ¼ jjy  Ȳ j jj is a distortion function defined by the Euclidean distance || ||.
Thus, data points are assigned to their nearest subregion.
2. Centroid Ȳ j is placed at the mean of data in the jth region. Mathematically, it means
N
1 X j

Ȳ j ¼ y 8y 2 Yj : (9)
N j k¼1 k

The convergence is guaranteed by using an iterative process. However, this algorithm is not
optimal in the sense that the data partition is performed independently of the projection that
follows. In fact, it leads to uniform partition regardless of real non-linear distribution of data,
which is illustrated in Fig. 2a. Therefore, Kambhatla and Leen [8] proposed a modification of the
ARTICLE IN PRESS

870 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

PII
y2 y2

A A
+ PI
Y2

A’ A’
y y
+
Y1

(a) y1 (b) y1

Fig. 2. Data partition for local PCA with two different distortion functions, according to distance: (a) from y to
centroid Ȳ j (Euclidean partition); (b) from y to principal direction (projection partition); Note: region-I data ‘J’; region
II data ‘K’.

above method, replacing the Euclidean distance by the projection distance (or called
reconstruction error) in Lloyd’s optimal conditions, as illustrated in Fig. 2(b). In this case, the
first Lloyd condition becomes
Yj ¼ fy : d P ðy; Pj Þod P ðy; Pi Þ 8jaig; (10)
P
where distortion function d ðy; Pj Þ represents the distance between data point y and the subspace
spanned by the principal directions P of the jth region. Since the membership of subregions may
change during the partition procedure, an iterative process is also needed to arrive at convergence.
Fig. 2 shows the difference between the Euclidean distance and the projection distance that are
used as the distortion function in the partition criterion, where 2D data with one principal
component are considered as an illustrating example. The Euclidean criterion has the advantage
of leading to a connected partition for each subregion. However, it always leads to uniform
partition regardless of real non-linear distribution. This represents a disadvantage specially when
the number of regions is small (e.g. in Fig 2a, the data were not ideally separated into two groups
with line AA0 as separating boundary). On the other hand, the projection partition method takes
reconstruction error as the distortion function. Although this method improved data partition in
many cases as shown in Ref. [8], it is not always ideal. As shown in Fig. 2b, data departing in far
distance from each other may be inappropriately partitioned into a subregion just because their
local principal directions are close by accident. The reason behind this behaviour is that the data
partition process may be infected by noise.
In order to avoid the difficulties encountered with these two algorithms, a modification is
proposed in this paper. The idea is to take advantages of both approaches but to avoid their
drawbacks. With this aim in mind, we use a characteristic distance dC in Lloyd’s optimal
conditions instead of the Euclidean or projection distances:
Yj ¼ fy : d C C
j od i 8jaig (11)
with
dC E P
j ¼ ad ðy; Ȳ j Þ þ ð1  aÞd ðy; Pj Þ; (12)
where a is a coefficient between 0 and 1. When taking a ¼ 0 or 1, one of the previous two
approaches is recovered. Considering that d E is generally larger in magnitude than d P ; one takes
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 871

generally a small value of ao0:4: A simple rule is that the fewer the number of subregions, the
smaller the value of a: More generally, an algorithm may be suggested to choose the smallest value
of a which ensures that regions are connected sets. This leads to a satisfactory data partition in
most situations. Indeed, as feature variations in real world are always continuous and smooth,
data in a region should then be connected. Numerical examples are given in the next section.

3. Applications

3.1. Numerical examples of data partition

A simple numerical example is first considered here to illustrate the proposed data partition
approaches. The data are bi-linearly distributed in a 2D space and are corrupted by noise, as
shown in Fig. 3. It is obvious that, with the first proposed method (Section 2.1), the data may be
easily partitioned into two regions by choosing a turning point at co-ordinates (x ¼ 1:13;
y ¼ 2:07), and each subregion may be well approached by a local PCA model. Now, our concern

Fig. 3. Comparison of different data partition approaches: (a) Euclidean partition; (b) projection partition; (c)
combined partition; and (d) too many subregions lead to over-fitting of noise.
ARTICLE IN PRESS

872 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

is to compare the second proposed approach (Section 2.2) with the two existing ones by vector-
quantisation technique. It may be observed that the data partitioning results either by the
Euclidean partition approach, Fig. 3a, or by the projection partition approach, Fig 3b, are not
ideal. As stated before, the former results in a uniform partition and the latter may lead to
undesirable disjointed partition. The proposed method leads to an improvement, as shown in Fig.
3c. A practical concern, which was not yet addressed, is how many subregions should be chosen.
If the minimisation of the reconstruction errors is taken as the criterion, it seems better to increase
the number of subregions. But from a practical point of view, the conclusion is just opposite. The
reason is that adopting too many subregions leads to an increase in computational and storage
costs, and to over-fitting of the effect of noise. This is shown in Fig. 3d, where the local principal
directions may deviate from the local trajectory. Therefore, it is recommended to use as few
subregions as possible. For this purpose, inspection of the data structure is helpful. In the case of
multiple-dimensional data, the inspection may be performed by use of several separated 2D
diagrams, as illustrated in the next application examples.

3.2. A simulated bridge model

The second example concerns the 3-span bridge model described in the companion paper [1].
The bridge is made of two materials and is subjected to multiple-site damages. This example has
been studied in detail by use of the PCA-based damage detection method proposed in Ref. [1] in
the case of a weakly non-linear behaviour in terms of the environmental variables. In the present
work, the proposed local PCA method is used to detect damage in the case of a non-linear
behaviour of the vibration features. Non-linearity results from the two following factors: (1) two
types of materials showing different temperature dependencies are used, (2) the structure is subject
to varying temperature gradients. Modal identification is supposed to be performed for each
0.5 1C interval of temperature (at reference point). Fig. 4a gives the evolution of the normalised
natural frequencies f i (i ¼ 2; . . . ; 6) in function of f1 . The separated 2D diagrams of Figs. 6b–f
allow observing the data structure when 1% of noise is added to the simulated data.
Based on these diagrams, the data sets are partitioned into two subregions by choosing a
turning point at L1 ¼ 1:01ff̄ i ; i ¼ 1:::6g; where f̄ i is the mean of the ith natural frequency. Note
that the turning points could be different from one mode to another. Figs. 5 and 6 compare the
results of PCA and local PCA analyses for damage detection and false-positive verification. It
may be observed that the introduced damages are well detected by both PCA and local PCA
analyses, although the non-linearity introduces a little distortion in the reference data when using
PCA. However, in the case of false-positive verification, a false alarm may be issued by PCA
analysis as shown in Fig. 6a, whereas correct monitoring is given by local PCA analysis (see Fig.
6b). Sometimes, this difficulty may be overcome by choosing a higher order of PCA model, as
presented in Ref. [1]. In the present work, local PCA with two subregions gives better results.

3.3. One-year monitoring of the Z24-bridge

This in situ experiment, conducted in the framework of the BRITE-EURAM programme CT96
0277 SIMCES, is reported in detail in Ref. [12]. The Z24-bridge is located in Switzerland over-
passing the national highway A1 between Bern and Zurich. It is a classical post-tensioned
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 873

Fig. 4. Data partitions for local PCA based on a 2D-data observation. (a) All feature data without noises and (b–f) 2D
data with noise, the first PC of subregion data is projected onto each 2D plane.

concrete box girder bridge with a main span of 30 m and 2 side-spans of 14 m (Fig. 7). The bridge
was monitored during 305 days from the 11th of November 1997 until the 11th of September
1998. A total of 49 sensors were chosen to capture various environmental parameters
(temperatures, wind characteristics, humidity, etc.). During the monitoring period, the bridge
was subjected to a series of realistic damage cases. Vibration measurements were gathered using
accelerometers generally every hour, and modal parameters were identified automatically and
examined by Peeters et al. [13,14]. A close look at the identified vibration features (Fig. 8) reveals
ARTICLE IN PRESS

874 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

Fig. 5. Damage detection by (a) PCA and (b) local PCA (two subregions). Reference set (full T- range), damaged set
(45 to 25 1C)

Fig. 6. False verification (without damage) by (a) PCA and (b) local PCA (two subregions). set I (full T- range), set II
(45 to 25 1C).

that the natural frequencies of the first four modes vary significantly with temperature. It was
found that the asphalt layer on the bridge surface plays a different role during warm and cold
periods, which in turn causes non-linearity. The natural frequencies indicate a bilinear relation in
function of the temperature, with a turning point at 0 1C [15].
In a previous work by Peeters et al., ARX models were established for the first four modes in
order to obtain a good correlation between the environmental parameters (i.e. the temperatures at
selected locations) and the natural frequencies of each mode. These models were then used to
analyse the different damage scenarios. A detailed comparison of this method with the present one
is presented in Ref. [16].
Using the PCA-based damage detection method proposed in the present paper, the
problem becomes much simpler: only vibration features (e.g. natural frequencies) are
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 875

Utzenstorf Koppigen

2.70 14.00 30.00 14.00 2.70


1.10

To Bern 4.50 To Zurich

ch
ba
otz
erh

Highway A1
Ob

North
Zurich
8.60

South Bern

Fig. 7. Z-24 bridge longitudinal section and top view [12].

Fig. 8. Time-variations of the natural frequencies of the first 4 modes (From November 11, 1997 to September 11,
1998).

needed to directly determine if the structure is damaged. Numerical tests show that the
choice of two principal components in the analysis (i.e. a model order m ¼ 2) gives the best
results.
Firstly, the results of the health monitoring and damage detection of the Z24-bridge, obtained
using linear PCA, are shown in Fig. 9. The results are quite satisfactory as the introduced various
damages are well detected. They are in good agreement with the results obtained by Peeters et al.
[13,14]. However, an abnormality appears around day 80–90, during which very cold weather
caused freezing of the asphalt layer on the bridge surface. It resulted in a strong non-linear
behaviour of the structure. To avoid this difficulty, data in the cold period (lower than 0 1C) were
ARTICLE IN PRESS

876 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

Fig. 9. (a) PCA for damage detection of Z24-bridge and (b) close-up of (a).

not considered in the analysis performed by Peeters. To improve the present detection, two data
partition techniques as proposed in Section 2, are used to partition the features into two
subregions as shown in Fig. 10. Thus local PCA may be performed in each subregion. Note that
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 877

Fig. 10. Features partitioned into two subregions (region I: ‘+’ or ‘  ’, II: ‘o’ ) by proposed two data partition
techniques: (a–c) by the first approach (turning point 1.005fi); (d–f) by second approach.

the two partitions lead to similar results. For this reason, only the results corresponding to the
first data-partition technique are given in Fig. 11. The comparison of the results obtained by using
local PCA (Fig. 11) with those obtained by linear PCA (Fig. 9) indicates two improvements: (1)
the abnormality of days 80–90 disappears as the non-linearity is taken into account by the local
PCA-based method; (2) the NI values increase slightly for the damage states (days 269–305), with
respect to those in the reference state; it means that the local PCA-based method is more sensitive
to the damage.
ARTICLE IN PRESS

878 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

Fig. 11. Local PCA for damage detection of Z24-bridge (two subregions) and (b) close-up of (a).

4. Conclusions

The PCA-based damage detection under varying environmental conditions, which was
developed in the companion paper [1], has been extended to non-linear cases. The proposed
ARTICLE IN PRESS

A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880 879

local PCA-based damage detection method assumes that the non-linearity present in structural
vibration features due to environmental factors may be modelled by local PCA when feature data
are well partitioned into several subregions. Therefore, the method involves a two-step procedure,
namely a clustering of the data space into several subregions and the application of PCA in each
local region. Two efficient data partition methods have been developed in this paper; the second
one is more general in its formulation and is less application-dependent. The local PCA models
are built in each partitioned data subregion, using the vibration features measured in reference
state under varying environmental conditions. Then, if no structural damage occurs afterwards,
the residual error of reconstruction of the input data remains small. On the contrary, if damage
occurs, the residual error increases significantly. The application of the proposed method to
simulated data, and especially to measured data obtained on a bridge during a one-year
monitoring, highlights its efficiency and suitability for structural health monitoring under varying
environmental conditions.

Acknowledgements

This work is supported by a grant from the Walloon government of Belgium as part of the
Research Convention No. 9613419. G. Kerschen is supported by a grant from the Belgian
National Fund for Scientific Research (FNRS). The authors would also like to acknowledge
Dr. Peeters (LMS International, Belgium) and Prof. De Roeck (K.U. Leuven, Belgium) for
providing the experimental data on the Z24-bridge.

References

[1] A.-M. Yan, G. Kerschen, P. De Boe, J.-C. Golinval, Structural damage diagnosis under changing environmental
conditions—Part I : linear analysis, Mechanical Systems and Signal Processing, 2003, submitted for publication.
[2] J.P. Cusumano, B.Y. Bai, Period-infinity periodic motions, chaos and spatial coherence in a 10-degree-of freedom
impact oscillator, Chaos and Solitons and Fractals 3 (1993) 515–535.
[3] M.F.A. Azeez, A.F. Vakakis, Proper orthogonal decomposition of a class of vibroimpact oscillations, Journal of
Sound and Vibration 240 (2001) 859–889.
[4] B.F. Feeny, R. Kappagantu, On the physical interpretation of the proper orthogonal modes in vibrations, Journal
of Sound and Vibration 211 (1998) 607–616.
[5] V. Lenaerts, G. Kerschen, J.-C. Golinval, Identification of a continuous structures with a geometrical non-
linearity, Part II: proper orthogonal decomposition, Journal of Sound and vibration 262 (2003) 907–919.
[6] H. Sohn, K. Worden, C.F. Farrar, Novelty detection under changing environmental conditions, SPIE’s Eigth
Annual International Symposium on Smart Structures and Materials, Newport Beach, CA (LA-UR-01-1894),
2001.
[7] M.A. Kramer, Nonlinear principal component analysis using autoassociative neural networks, AIChE Journal 37
(2) (1991) 233–243.
[8] N. Kambhatla, T.K. Leen, Dimension reduction by local PCA, Neural Computation 9 (1997) 1–17.
[9] T. Hastie, W. Stuetzle, Principal curves, Journal of the American Statistical Association 84 (1989) 502–516.
[10] G. Kerschen, J.-C. Golinval, Non-linear generalization of principal component analysis: from a global to a local
approach, Journal of Sound and Vibration 254 (5) (2002) 867–876.
[11] A. Gersho, R.M. Gray, Vector Quatization and Signal Compression, Kluwer Academic Publishers, Norwell, 1992.
[12] C. Kramer, C.A.M. De Smet, G. De Roeck, Z24-bridge damage detection tests, Proceedings of IMAC 17,
Kissimmee, FL, USA February 1999, pp. 1023–1029.
ARTICLE IN PRESS

880 A.-M. Yan et al. / Mechanical Systems and Signal Processing 19 (2005) 865–880

[13] B. Peeters, G. De Roeck, One year monitoring of the Z24-bridge: environmental influences versus damage events,
Proceedings of IMAC 18, the International Modal Analysis Conference, San Antonio, TX, USA, February 2000,
pp. 1570–1576.
[14] B. Peeters, J. Maeck, G. De Roeck, Dynamic monitoring of the Z24-Bridge: separating temperature effects from
damage, Proceedings of the European COST F3 Conference on System Identification and Structural Health
Monitoring, Madrid, Spain, June 2000, pp. 377–386.
[15] B. Peeters, System identification and damage detection in civil engineering. Ph.D. Thesis, Katholieke Universiteit
Leuven, 2000.
[16] A.M. Yan, J.C. Golinval, B. Peeters, G. De Roeck, A comparative study on damage detection of Z24-Bridge: one-
year monitoring with varying environmental conditions, Proceedings of Second E.W. on Structural Health
Monitoring 2004, July 7–9, 2004, Munich, Germany, pp. 791–799.

You might also like