polynomial chaos expansion models for the … · 1 polynomial chaos expansion models for the 2...

40
POLYNOMIAL CHAOS EXPANSION MODELS FOR THE 1 MONITORING OF STRUCTURES UNDER OPERATIONAL 2 VARIABILITY 3 Minas D. Spiridonakos 1 , Not a Member, ASCE, Eleni N. Chatzi, 1 Member, ASCE and Bruno Sudret, 1 Not a Member, ASCE 4 ABSTRACT 5 This study overviews the development and implementation of a stochastic framework 6 efficiently fusing operational response data with external influencing agents, namely environ- 7 mental conditions, for representing structural behavior in its complete operational spectrum. 8 The delivered methodology results in the formulation of a structural performance indicator 9 serving as a precursor of potential damage or deterioration. The proposed approach relies on 10 the combined use of (i) a polynomial chaos expansion method, able to account for structural 11 variations tied to the variation of the acting agents; and (ii) an independent component anal- 12 ysis algorithm, for extraction of salient behavioral features. The efficacy of the introduced 13 method is demonstrated on field data from actual large-scale systems in both operational 14 and damaged states. 15 Keywords: condition indicator, externally acting agents, environmental conditions, damage 16 detection, polynomial chaos expansion (PCE), independent component analysis (ICA). 17 INTRODUCTION 18 Vibration-based monitoring methods (Deraemaeker and Worden 2010; Ostachowicz and 19 uemes 2013) have long surfaced as a promising approach to non-destructive evaluation, 20 and a highly researched subdomain of the structural health monitoring (SHM) field. Recent 21 1 ETH Zurich, Institute of Structural Engineering, Department of Civil, Environmental and Geomatic Engineering, Stefano-Franscini-Platz 5, 8093 Zurich, Switzerland. E-mail: [email protected], [email protected], [email protected] 1

Upload: others

Post on 23-Oct-2020

22 views

Category:

Documents


0 download

TRANSCRIPT

  • POLYNOMIAL CHAOS EXPANSION MODELS FOR THE1

    MONITORING OF STRUCTURES UNDER OPERATIONAL2

    VARIABILITY3

    Minas D. Spiridonakos1, Not a Member, ASCE, Eleni N. Chatzi,1 Member, ASCE

    and Bruno Sudret,1 Not a Member, ASCE

    4

    ABSTRACT5

    This study overviews the development and implementation of a stochastic framework6

    efficiently fusing operational response data with external influencing agents, namely environ-7

    mental conditions, for representing structural behavior in its complete operational spectrum.8

    The delivered methodology results in the formulation of a structural performance indicator9

    serving as a precursor of potential damage or deterioration. The proposed approach relies on10

    the combined use of (i) a polynomial chaos expansion method, able to account for structural11

    variations tied to the variation of the acting agents; and (ii) an independent component anal-12

    ysis algorithm, for extraction of salient behavioral features. The efficacy of the introduced13

    method is demonstrated on field data from actual large-scale systems in both operational14

    and damaged states.15

    Keywords: condition indicator, externally acting agents, environmental conditions, damage16

    detection, polynomial chaos expansion (PCE), independent component analysis (ICA).17

    INTRODUCTION18

    Vibration-based monitoring methods (Deraemaeker and Worden 2010; Ostachowicz and19

    Güemes 2013) have long surfaced as a promising approach to non-destructive evaluation,20

    and a highly researched subdomain of the structural health monitoring (SHM) field. Recent21

    1ETH Zurich, Institute of Structural Engineering, Department of Civil, Environmental and Geomatic

    Engineering, Stefano-Franscini-Platz 5, 8093 Zurich, Switzerland. E-mail: [email protected],

    [email protected], [email protected]

    1

    mailto:[email protected]:[email protected]:[email protected]

  • advances in sensor technologies, with associated decrease in corresponding costs, have further22

    revealed the potential of these methods to permanently escort the structure throughout its23

    entire life-cycle, reporting significant information related to structural condition. Automa-24

    tion, ability to work on a system level, adaptability for various types of structures, on-line25

    or off-line implementation, are only few of the features contributing to the afore-mentioned26

    concept. However, in deciphering the extracted information a number of shortcomings need27

    be overcome. The latter pertains to the common lack of loading records, since the usual28

    source of excitation is ambient loading, and secondly the susceptibility of these structures to29

    a number of uncontrollable operational conditions, such as temperature, temperature gra-30

    dients, humidity, wind speed, traffic intensity loading and others (Sohn 2007; Peeters et al.31

    2001). Even though the former problem is adequately treated today by a number of accurate32

    and robust output-only (operational) identification methods (Peeters and De Roeck 2001;33

    Reynders 2012), the latter is still under research and in focus of a number of recent studies34

    (Cross et al. 2013; Magalhães et al. 2012; Kullaa 2011a).35

    The effect of operational, i.e., loading and environmental, conditions on the structural36

    dynamics has been examined in various studies over the last 20 years. For example, Far-37

    rar and coworkers (Farrar et al. 1997) report eigenfrequency differences for the Alamosa38

    Canyon Bridge of approximately 5% over a 24 hour period due to temperature and spatial39

    temperature gradients variation. Alampalli in his paper (Alampalli 2000) identifies natural40

    frequencies differences of up to 50% for an abandoned bridge in Claverack (NY) and at-41

    tributes this variation to the freezing of the bridge supports. Maximum differences varying42

    from 14 to 18% were also calculated for the first four natural frequencies of the Z24 bridge43

    monitored during a nine-months period for the SIMCES project (Peeters and De Roeck44

    2001). Catbas et al. (Catbas et al. 2008) found that the structural reliability of a long-span45

    truss bridge in USA was highly affected by the ambient temperature, while Cross et al.46

    (Cross et al. 2013) investigated the effect of temperature, traffic loading and wind speed on47

    the modal properties of the Tamar (UK) suspension bridge and identified differences up to48

    2

  • almost 5%. Finally, Yuen and Kuok focused on the description of the relationship between49

    natural frequencies and environmental factors (temperature and humidity) for tall buildings,50

    while also utilized one-year measurement of a 22-storey reinforced concrete building (East51

    Asia Hall) (Yuen and Kuok 2010).52

    The approaches that have been followed in the literature for incorporating or discarding53

    the operational variability from structural models may be primary classified into two cate-54

    gories: i) output-only methods which aim at discarding the influence of operational factors55

    solely from vibration response measurements and/or extracted features (e.g. modal prop-56

    erties) (Kullaa 2011b; Reynders et al. 2014), and ii) input-output methods which aim at57

    modelling the relationship between the measured vibration data and/or the extracted fea-58

    tures with respect to measured operational conditions (Peeters and De Roeck 2001; Yuen59

    and Kuok 2010). Borrowed from the machine learning literature, the terms unsupervised60

    and supervised learning, respectively, have also been used to describe the aforementioned ap-61

    proaches. In the former category methods such as the Principal Component Analysis (PCA)62

    (Magalhães et al. 2012) or its nonlinear ramifications (kernel PCA (Reynders and De Roeck63

    2015), Factor Analysis), neural networks, and others have been employed for solving the64

    unsupervised learning problem by searching and discarding patterns in multivariate output65

    features, which may reveal the influence of the unobserved input variables. The supervised66

    learning alternative on the other hand follows a somewhat different approach, which typically67

    involves formulation of a regression problem.68

    The present study follows the second approach (Spiridonakos and Chatzi 2014; Spiridon-69

    akos and Chatzi 2015), which delivers additional insight into the mechanism of variation of70

    structural properties (outputs) and its dependence upon influencing agents (inputs) (Wenzel71

    2009), which are often quantifiable via observations. A functional representation between in-72

    puts and outputs is delivered, tracking the behavior of the structure under study for a wide73

    range of operational conditions, even potentially unobserved ones through extrapolation.74

    Information on the inputs may be extracted by deploying a small number of sensors track-75

    3

  • ing externally acting agents, such as environmental (temperature, humidity, wave, wind) or76

    traffic/train crossing loads, along with the array of vibration sensors monitoring structural77

    response.78

    The methodological framework proposed herein relies on the expansion of appropriately79

    selected structural features, such as modal characteristics of the healthy structure estimated80

    by means of operational modal analysis methods, onto a polynomial chaos basis (PCE)81

    (Soize and Ghanem 2004; Berveiller and Lemaire 2006; Ghanem and D. 2003; Blatman and82

    Sudret 2010) for the projection of these estimates on the probability space of the measured83

    influencing agents. This process is carried out during a training phase resulting in a model,84

    which is able to predict the evolution of structural response based on information of the85

    input variables. Following the training phase, the PC-based estimated statistical properties,86

    e.g. the prediction error, may the serve as condition indicators for the monitored system87

    warning of irregular behavior or revealing evidence of a damaged/deteriorated system.88

    In order to illustrate the workings of the method, the proposed framework is applied on89

    a series of data from actual structural systems, including the Überführung Bärenbohlstrasse90

    bridge in Zürich, as well as the benchmark SHM problem of the Z24-bridge in Switzerland91

    (Peeters and De Roeck 2001; Reynders and De Roeck 2009). The results of these studies92

    demonstrate a tool, which may reliably be employed into practice for condition monitoring93

    of large-scale real world structures operating in a wide range of conditions.94

    The remainder of the paper is structured as follows: In Section 2 the SHM framework95

    based on PCE and ICA is introduced. The application case studies and the analysis results96

    are presented in Section 3. Finally, the conclusions of this study are summarized in Section97

    4.98

    STRUCTURAL IDENTIFICATION FRAMEWORK99

    The developed SHM framework aims at a global representation of the system, in the100

    sense that it may accurately describe the dynamic properties of the structure for a range of101

    operational conditions. The method employs conventional system identification methods in102

    4

  • order to extract statistical characteristics (features) of the healthy structure. A polynomial103

    chaos expansion (PCE) model (Berveiller and Lemaire 2006; Blatman and Sudret 2010)104

    is then adopted for the representation of the extracted features on the space spanned by105

    random variables describing the measured operational conditions. The extracted features106

    could be either modal properties of the structure or even the parameters of an identified107

    model.108

    A schematic diagram of the training phase of the proposed framework is illustrated in109

    Fig. 1.110

    Polynomial Chaos Expansion (PCE)111

    PCE is a tool employed for the propagation of uncertainty through dynamical systems,112

    due to uncertainty relating to the system parameters (Wiener 1938; Ghanem and Spanos113

    1997; Ghanem and D. 2003). The non-intrusive variant of the method is employed herein,114

    relying on expansion of the random output variables on polynomial chaos basis functions,115

    which are orthonormal with respect to the probability space of the system’s random inputs.116

    Only the fundamental concepts of the employed tool are provided herein for the purpose of117

    completeness. The interested reader is referred to (Blatman and Sudret 2010) for a deeper118

    insight into the details of the method.119

    Consider a system S comprising M random input parameters represented by indepen-120

    dent random variables Ξ1, . . . ,ΞM , e.g. temperatures measured at different locations of the121

    structure, gathered in a random vector Ξ with joint Probability Density Function (PDF)122

    pΞ(ξ). Since the input variables are herein linked to measured quantities, it is reasonable to123

    assume that information may be collected concerning their probability distribution.124

    The system output, e.g. natural frequency estimates, denoted by Y = S(Ξ) will also be125

    random. Assuming that Y has finite variance, it may be expressed as follows:126

    Y = S(Ξ) =∑

    d(j)∈NMθjφd(j)(Ξ) (1)127

    5

  • where θj are unknown deterministic coefficients of projection, d is the vector of multi-indices128

    describing the multivariate polynomial basis, and φd(j)(Ξ) are the polynomial basis (PC)129

    functions, which are orthonormal to pΞ(ξ). The basis functions φd(j)(Ξ) may be constructed130

    through tensor products of the corresponding univariate polynomials.131

    Each univariate probability density function may be associated with a well-known fam-132

    ily of orthogonal polynomials, serving for construction of the corresponding PC basis. For133

    instance, the normal distribution is associated with Hermite polynomials while the uni-134

    form distribution with Legendre (Table 1). A list of the most common probability density135

    functions along with the corresponding orthogonal polynomials and the relations for their136

    construction may be found in (Xiu and Karniadakis 2002).137

    Obviously, the resulting series must be truncated to a finite number of terms for purposes138

    of practicality, commonly resulting in selection of the multivariate polynomial basis with a139

    total maximum degree |dj| =∑M

    m=1 dj,m ≤ P for every j. In this case the dimensionality of140

    the functional subspace is equal to141

    p =(M + P )!

    M !P !(2)142

    whereM is the number of random variables and P the maximum basis degree. However, more143

    compact representations may be achieved by dropping the terms which do not significantly144

    contribute to the accuracy of the model, judged in terms of a predefined criterion (Blatman145

    and Sudret 2010; Blatman and Sudret 2011).146

    When truncating the infinite series of expansion of Eq. 1, the resulting PCE model is147

    fully parameterized in terms of a finite number of deterministic coefficients of projection θj.148

    The parameter vector θj may be estimated by solving Eq. 1 in a least squares sense. Toward149

    this end, the data of the input-output data have to be employed.150

    In visualizing the connection to the problem handled herein, consider the example a151

    structural system subjected to ambient excitation which is susceptible to variations of the152

    6

  • environmental conditions, which would herein serve as the system inputs. A given output153

    variable, such as the estimate of one of the system’s natural frequencies, could be expanded on154

    a PC basis constructed to be orthogonal to the experimentally estimated PDF of temperature155

    data. In this way, an indication is obtained regarding the sensitivity of the natural frequencies156

    of the structure to the manner in which stochasticity, relating to temperature, propagates157

    through the structural system.158

    At this point, it needs to be noted that the application of the PCE method on real data159

    is accompanied by a number of difficulties which have to be overcome. These are:160

    1. the constraint of independence between the input variables161

    2. the statistical characterization of the input variable data, that is the estimation of162

    the underlying PDFs163

    3. the selection of appropriate PC subspace164

    A first necessary assumption obviously lies in that the considered input variables need be165

    independent. However, operational condition data is normally acquired from a dense grid of166

    sensors measuring operational conditions is normally installed in large structures, and thus167

    a considerable percentile of the measured inputs is highly correlated. In order to extract168

    only a small number of independent latent input variables, the Independent Component169

    Analysis (ICA) algorithm is herein utilized in a first step for obtaining a set of independent170

    transformed input variables out of the original dependent set (Hyvärinen and Oja 2000).171

    The ICA algorithm is described in the next section.172

    The second point pertaining to the fitting of PDFs on the measured input variable data173

    may be addressed by various approaches, including i) approximation of the underlying PDFs174

    by well-known parametric PDFs, ii) transformation to uniformly distributed data via use of175

    nonparametric kernel estimates for the cumulative distribution function, and iii) implemen-176

    tation of a data-driven approach, namely arbitrary PC (aPC) (Oladyshkin and Nowak 2012).177

    Finally, to what the third point is concerned, the Least Angle Regression (LAR) method178

    7

  • proposed in (Blatman and Sudret 2011) is herein adopted for the selection of a sparse PC179

    basis.180

    Once these preliminary processing steps are performed, the training phase of the scheme181

    is carried out for extracting the PCE model. The schematic diagram of the training phase182

    of the proposed framework is illustrated in Fig. 1. Following this phase, the PCE-based esti-183

    mated statistical properties of the features may be further exploited for extracting condition184

    indicators as explained int he applications section.185

    Independent Component Analysis (ICA)186

    ICA is a source separation method which aims at estimating independent unobservable187

    (latent) variables which are mixed into observed variables (Hyvärinen and Oja 2000). The188

    key of the method lies in its search for non-Gaussian components, in contrast to principal189

    component analysis and other second order methods, which rely on the covariance matrix190

    of random variables. To this end, ICA combines a static linear mixture model with higher191

    order statistics in order to identify unobservable variables. More specifically, considering192

    n observed linear mixtures of n independent components {si, i = 1, . . . , n} the following193

    relation is assumed (Hyvärinen and Oja 2000):194

    xj = a(j,1)s1 + a(j,2)s2 + . . .+ a(j,n)sn, for j = 1, . . . , n (3)195

    or by using vector-matrix notation196

    x = As (4)197

    with x designating the random observed variables vector, while198

    A =

    a1,1 . . . a(1,n)

    .... . .

    ...

    an,1 . . . a(n,n)

    199denotes the mixing matrix, and s is the vector of unobserved independent variables. The200

    latter quantities are unknown and have to be estimated by the random vector x and the also201

    8

  • unknown mixing matrix A. The vector of the independent latent variables s may be readily202

    obtained via the inverse of the mixing matrix A, herein denoted as W = A−1:203

    s = Wx (5)204

    The ICA focuses on deriving matrix A. As already noted, the key in estimating the ICA205

    model is the non-Gaussianity of the unobserved sources. Therefore, ICA estimation algo-206

    rithms are based on nonlinear optimization methods, which aim at the maximization of the207

    non-Gaussianity of each one of the latent variables wTx, with w designating a column vector208

    of matrix W . The measure of non-Gaussianity may be based on kurtosis, negentropy, and209

    others (Hyvärinen and Oja 2000). The flowchart of a typical ICA algorithm implementation210

    is provided in Fig. 2.211

    Finally, it should be added that the characteristics of the ICA-based approach are par-212

    ticularly attractive within the context of the proposed SHM framework. The advantage is213

    twofold since PCE requires feeding of independent input variables and, additionally, only214

    a reduced number of latent variables may be identified depending on the results of the215

    eigenvalue decomposition of the ICA algorithm. Within such a context, ICA may serve for216

    critically reducing the large bulk of data acquired via SHM systems by extracting only a217

    reduced number of salient input and output features.218

    APPLICATION CASE STUDIES219

    Implementation on Systems under Operational Conditions220

    As part of a project funded by FEDRO on “Structural Identification for Condition As-221

    sessment of Swiss Bridges”, the Überführung Bärenbohlstrasse bridge has been chosen as a222

    suitable test case due to its dimensions, static system, and ease of access (Fig. 3). According223

    to existing documents, the bridge was designed in 1977 and construction was completed in224

    1980. This implies that the bridge is currently at 34 years of age, therefore not extremely225

    aged. The bridge was monitored from July 2013 till July 2014 through a permanent system226

    9

  • acquiring hourly ambient vibration response and environmental condition measurements.227

    More specifically, 18 acceleration sensors measuring ambient vibration responses (vertical228

    axis), and two environmental condition sensors for measuring air temperature and humidity229

    were installed. After the twelve-month period, more than 6500 datasets were collected. For230

    these datasets, the estimation of the modal properties of the structure based on the NExT-231

    ERA method was realized. The method’s hyper-parameters (state space model order equal232

    to 30; 80 Markov parameters for the construction of the Hankel matrix) were kept constant233

    for all datasets and they were selected based on initial modeling of a randomly selected234

    training set of vibration response data.235

    The first four natural frequencies estimated via the NExT-ERA method are selected as236

    the output variables to be represented by the proposed methodology. These estimates are237

    plotted versus time in Fig. 4. It may be observed that all four natural frequencies vary with238

    time in three different patterns: i) almost linear increase from July 2013 till end of November239

    2013, ii) high variability from end of November 2013 till end of December 2013, and iii) linear240

    decrease from January 2014 till July 2014 when they seem to actually return to their July241

    2013 values.242

    A substantial part of this variability may be attributed to the seasonal variation of the243

    environmental conditions. This dependency is clearly depicted in Fig. 5 where the natural244

    frequency estimates are plotted versus temperature and humidity. A bilinear temperature -245

    natural frequency relationship may be observed between negative and positive temperatures,246

    while a more complex relationship seems to characterize the dependency of the identified247

    natural frequencies and humidity.248

    In order to apply the PCE approach, the independence of the random input variables,249

    i.e., temperature and humidity, has to be checked. Looking at the scatter plot of Fig. 6a, it is250

    evident that these are highly correlated (correlation close to 0.75). Therefore, the ICA has to251

    be implemented before PCE is applied on the natural frequency estimates. The scatter plots252

    of the corresponding ICA-based estimates of the independent random (latent) variables are253

    10

  • shown in Fig. 6b, with the variables indicating a correlation close to 0. It should be noted254

    that these transformed variables no longer correspond to temperature and humidity.255

    The identified modal frequencies are then expanded onto the probability space of the256

    latent input variables through PCE. Toward this end, the input set is transformed into uni-257

    formly distributed variables via use of their non-parametrically estimated cumulative distri-258

    bution functions. Consequently, a PC basis consisting of multivariate Legendre polynomials259

    is calculated.260

    The set of natural frequency estimates is divided into an training (estimation) and a261

    validation set. The training set comprises the first 1250 values corresponding to the first262

    five months of the monitoring period till early December 2013, while the rest of the values263

    are used as validation set, that is 1114 values. The estimation set values are expanded on a264

    PC basis consisting of multivariate Legendre polynomials of maximum order five, that is a265

    total number of 21 basis functions, while it was observed that increasing the maximum order266

    further did not significantly increased the accuracy of expansion.267

    It is observed that the PCE-model estimates are capable of reliably reproducing the268

    estimation set values, while more importantly very good accuracy is achieved for values269

    lying in the validation set, including the more challenging winter months of the monitored270

    interval (Fig. 7).271

    Nonetheless, damage detection still comprises a difficult decision, which has to be based272

    on multivariate statistics relating to the PCE expansion errors for the four natural frequency273

    estimates. In order to simplify this problem, the ICA algorithm is utilized for estimating274

    a reduced number of independent random variables from the estimated natural frequencies275

    (output variables). As a result, a single damage (or condition) indicator is extracted relying276

    on a univariate statistical hypothesis testing for the characterization of the “health” state277

    of the structure. By implementing the ICA once again, this time retaining a single latent278

    output variable, the results shown in Fig. 8 are attained, corresponding to a one-dimensional279

    damage indicator. Based on the statistical characterization of the PCE model prediction280

    11

  • errors, confidence intervals may be calculated with respect to this simplified metric, and281

    used for damage detection as described before.282

    The strength of the methodology presented herein therefore lies in that it is able to283

    account for the influence of exogenous factors, such as environmental or other operational284

    conditions, upon structural behavior without use of a detailed numerical structural model.285

    Instead, this is a tool easily generalizable for various types of systems, without necessitating286

    a-priori knowledge on the structure itself, but merely a training interval during which the287

    structure is monitored. Hence, such a process is only implementable within a long-term288

    monitoring perspective.289

    Implementation on Systems under Damaged Conditions290

    The workings of the method will be now demonstrated via implementation on an ac-291

    tually damaged case, namely the well known case study of the Z24 bridge (Switzerland),292

    which was monitored for nine months and finally artificially damaged for the purposes of293

    the BRITE-EURAM project SIMCES. The bridge, dated from 1963, was overpassing the294

    A1 highway (Zurich-Bern), connecting Utzemstorf and Koppigen and its demolition was295

    planned after the monitoring period since a new railway adjacent to the highway required a296

    new bridge with a larger side span (Reynders and De Roeck 2009). The bridge was moni-297

    tored from November 1977 till August 1998 through a permanent system acquiring hourly298

    ambient vibration response and environmental condition measurements. More specifically,299

    16 acceleration sensors measuring ambient vibration responses (13 of them are shown in Fig.300

    9 while three more were installed on one of the piers), while 49 environmental condition301

    sensors were installed for measuring air temperature, wind characteristics, humidity, bridge302

    expansion, soil temperatures at the boundaries and bridge concrete temperatures (Reynders303

    and De Roeck 2009).304

    After the nine-month period, progressive damages were artificially induced within a pe-305

    riod of a (August 10 - September 11, 1998). The monitoring system was still in operation306

    while during night vibration tests based on hammer, dropping mass device and shakers were307

    12

  • performed. The specific types of induced damages are summarized in Table 2, while more308

    details on the damage scenarios and description of the simulation of the real damage cause309

    may be found in (Reynders and De Roeck 2009).310

    Vibration and environmental data311

    In this study, the temperatures measured at six locations at the center of the mid-312

    dle span along with the air temperature are used (Fig. 10), while the first four natural313

    frequencies estimated through Stochastic Subspace Identification (SSI) and an automated314

    operational analysis method are considered as the output of the structural system (Reyn-315

    ders and De Roeck 2009). The initial data set of 5652 values, corresponding to the hourly316

    obtained data of the monitoring system, are initially truncated to a 3932 long dataset for317

    which estimates of all four natural frequencies are available (Peeters and De Roeck 2001).318

    The dependency of natural frequency estimates on the temperature measured at the cen-319

    ter of the web is shown in Fig. 12. A bilinear temperature/natural frequency relationship320

    may be observed between negative and positive temperatures, while values depicting abnor-321

    mal behavior, that may be attributed to damage, are clearly indicated only for the second322

    mode.323

    Even though the list of input variables affecting the structural behavior of Z24 is not324

    limited to this set of temperatures, it is assumed that they may be adequate for describing a325

    large percentage of the variability of the estimated modal properties (Peeters and De Roeck326

    2001).327

    Results328

    As a first step, and before expanding the natural frequency estimates to the PC functional329

    basis orthogonal to the PDFs of the random input variables, i.e. recorded temperatures, the330

    independency of the latter has to be checked. Looking at the scatter plots of Fig. 11,331

    drawn for each pair of measured temperatures, it is evident that they are highly correlated332

    (correlation larger than 0.95 for almost all pairs), while the smallest correlation is 0.89.333

    Therefore, ICA has to be implemented before PCE is applied on the natural frequency334

    13

  • estimates. The scatter plots of the corresponding ICA-based estimates of the independent335

    random (latent) variables, are shown in Fig. 13, with most of the variables indicating a336

    correlation smaller than 0.1, with the larger correlation being 0.13. It should be noted that337

    these transformed variables do not correspond to temperature anymore.338

    The identified modal frequencies are then expanded as a function of the random latent339

    input variables through PCE. To this end, the input variables are transformed into uniformly340

    distributed variables by using their non-parametrically estimated cumulative distribution341

    functions and PC basis consisting of multivariate Legendre polynomials is finally constructed.342

    The set of natural frequency estimates is in this case as well divided into a training and343

    validation set. The training set comprises 1500 values randomly selected from the first eight344

    months of the monitoring period (one month prior to the first damage), while the remaining345

    values are employed as a validation set. The latter includes the complete set recorded during346

    the last month of the monitoring period, plus the complete set of values corresponding to347

    the period of induced damages, i.e., 2744 values.348

    The values of the estimation set are expanded onto a PC basis consisting of multivariate349

    Legendre polynomials of maximum order three, that is a total number of 120 basis functions,350

    while it is observed that increasing the maximum order further does not significantly increase351

    the accuracy of the expansion. The standard deviations of the expansion errors, computed352

    on the validation set, are summarized in Table 3, with the maximum being approximately353

    equal to 0.13 Hz.354

    It may be observed that the PCE-model estimates are capable of reliably reproducing355

    the estimation set values, while more importantly a very good accuracy is also achieved for356

    values belonging to the validation set for the interval prior to damage (Fig. 14). Naturally,357

    predictions deviate for the period after the first damage is induced (August 9, 1998) which358

    correctly translates to damage indication.359

    In once again extracting a reduces order condition indicator, the ICA algorithm is uti-360

    lized for estimating a reduced number of independent salient features from both the vector361

    14

  • of temperature measurements (inputs) and the estimated natural frequencies (outputs). By362

    inspecting the eigenvalues of the covariance matrices (second step of the ICA; Fig. 2) it is363

    obvious that a single input latent variable is enough to describe the variance of the tempera-364

    ture measurements in a percentage equal to 95% (Fig. 15a). At the same time, by applying365

    the ICA on the vector of natural frequencies, and inspecting the eigenvalues of the corre-366

    sponding covariance matrix (Fig. 15b), a single output latent variable is retained, which is367

    able to account for almost 90% of the output variance.368

    The PCE of the output to the input (latent variables), with just 4 basis functions (uni-369

    variate Legendre polynomial of orders 0 to 3), renders the results demonstrated in Fig. 16,370

    for which an abnormal trend is noted during the damage period clearly signaling the damage371

    initiation and progression. As a next step to this framework, damage localization is fore-372

    seen, which may be explored via exploring spatial correlations in a grid of sensors. Kriging373

    methods could assist to such an end, and are left for a subsequent exploration.374

    Finalizing, it must be noted that in previous work, related to the same benchamark375

    problem (Peeters and De Roeck 2001), an ARX model was used to describe the temperature-376

    modal frequency relationship. Similar damage detection results were obtained, while the377

    description of nonlinearity was alleviated via use of the temperature range above zero.378

    CONCLUSIONS379

    Environmental condition data which are nowadays available in modern SHM systems380

    of large-scale civil engineering structures should be accounted for in favor of identifying381

    comprehensive dynamic models. For this reason, the present study proposes an identifica-382

    tion framework which utilizes polynomial chaos expansions for the accurate description of383

    the propagation of stochasticity stemming from environmental factors through the struc-384

    tural response. In order to compress/reduce the amount of recorded environmental data385

    and simultaneously identify a small number of statistically independent input variables, the386

    independent component analysis method is proposed as a suitable tool. In addition, its appli-387

    cability for extraction of simplified, or low dimensional damage indicators is also examined.388

    15

  • The implementation of the method on two field case studies demonstrated the effectiveness389

    and applicability of the proposed framework toward both aforementioned directions.390

    ACKNOWLEDGMENTS391

    The authors would like to gratefully acknowledge Prof. E. Reynders from the Department392

    of Civil Engineering, KU Leuven, for providing the SIMCES project data, serving as the393

    second test case discussed herein. This research was supported by the Federal Roads Office394

    (FEDRO) of Switzerland under Project #AGB 2012/015.395

    16

  • APPENDIX I. REFERENCES396

    Alampalli, S. (2000). “Effects of testing, analysis, damage, and environment on modal pa-397

    rameters.” Mechanical Systems and Signal Processing, 14(1), 63–74.398

    Berveiller, M., S. B. and Lemaire, M. (2006). “Stochastic finite elements: a non intrusive399

    approach by regression.” Eur. J. Comput. Mech., 15(1-3), 81–92.400

    Blatman, G. and Sudret, B. (2010). “An adaptive algorithm to build up sparse polynomial401

    chaos expansions for stochastic finite element analysis.” Probabilistic Engineering Mechan-402

    ics, 25(2), 183–197.403

    Blatman, G. and Sudret, B. (2011). “Adaptive sparse polynomial chaos expansion based on404

    least angle regression.” Journal of Computational Physics, 230(6), 2345–2367.405

    Catbas, F. N., Susoy, M., and Frangopol, D. M. (2008). “Structural health monitoring and406

    reliability estimation: Long span truss bridge application with environmental monitoring407

    data.” Engineering Structures, 30(9), 2347–2359.408

    Cross, E., Koo, K., Brownjohn, J., and Worden, K. (2013). “Long-term monitoring and data409

    analysis of the tamar bridge.” Mechanical Systems and Signal Processing, 35(1-2), 16–34.410

    A. Deraemaeker and K. Worden, eds. (2010). New Trends in Vibration Based Structural411

    Health Monitoring, Vol. 520 of CISM Courses and Lectures. Springer Vienna.412

    Farrar, C. R., Doebling, S. W., Cornwell, P. J., and Straser, E. G.413

    (1997). “Variability of modal parameters measured on the alam-414

    osa canyon bridge.” 15th International Modal Analysis Conference415

    (IMAC XV), .417

    Ghanem, R. and Spanos, P. (1997). “Spectral techniques for stochastic finite elements.”418

    Archives of Computational Methods in Engineering, 4(1), 63–100.419

    Ghanem, R. G. and D., S. P. (2003). Stochastic Finite Elements: A Spectral Approach.420

    Courier Corporation.421

    Hyvärinen, A. and Oja, E. (2000). “Independent component analysis: algorithms and appli-422

    17

  • cations.” Neural Networks, 13(4-5), 411–430.423

    Kullaa, J. (2011a). “Distinguishing between sensor fault, structural damage, and environ-424

    mental or operational effects in structural health monitoring.” Mechanical Systems and425

    Signal Processing, 25(8), 2976–2989.426

    Kullaa, J. (2011b). Vibration-Based Structural Health Monitoring Under Variable Environ-427

    mental or Operational Conditions, Vol. 520 of CISM Courses and Lectures. Springer, Chap-428

    ter 4, 107–181.429

    Magalhães, F., Cunha, A., and Caetano, E. (2012). “Vibration based structural health moni-430

    toring of an arch bridge: From automated oma to damage detection.” Mechanical Systems431

    and Signal Processing, 28, 212–228.432

    Oladyshkin, S. and Nowak, W. (2012). “Data-driven uncertainty quantification using the433

    arbitrary polynomial chaos expansion.” Reliability Engineering & System Safety, 106, 179–434

    190.435

    W. Ostachowicz and J. Güemes, eds. (2013). New Trends in Structural Health Monitoring,436

    Vol. 542 of CISM International Centre for Mechanical Sciences. Springer Vienna.437

    Peeters, B. and De Roeck, G. (2001). “One-year monitoring of the z24-bridge: environmental438

    effects versus damage events.” Earthquake Engineering and Structural Dynamics, 30(2),439

    149–171.440

    Peeters, B., Maeck, J., and De Roeck, G. (2001). “Vibration-based damage detection in civil441

    engineering: excitation sources and temperature effects.” Smart Materials and Structures,442

    10(3), 518–527.443

    Reynders, E. (2012). “System identification methods for (operational) modal analysis: Re-444

    view and comparison.” Archives of Computational Methods in Engineering, 19(1), 51–124.445

    Reynders, E. and De Roeck, G. (2009). “Continuous vibration monitoring and progres-446

    sive damage testing on the z24 bridge.” Encyclopedia of Structural Health Monitoring, C.447

    Boller, F.-K. Chang, and Y. Fujino, eds., John Wiley & Sons, Chapter 126, 1–10.448

    Reynders, E. and De Roeck, G. (2015). “Vibration-based damage identification: The z24449

    18

  • bridge benchmark.” Encyclopedia of Earthquake Engineering, M. Beer, I. A. Kougioumt-450

    zoglou, E. Patelli, and I. S.-K. Au, eds., Springer Berlin Heidelberg, 1–8.451

    Reynders, E., Wursten, G., and De Roeck, G. (2014). “Output-only structural health moni-452

    toring in changing environmental conditions by means of nonlinear system identification.”453

    Structural Health Monitoring, 13(1), 82–93.454

    Sohn, H. (2007). “Effects of environmental and operational variability on structural health455

    monitoring.” Philosophical Transactions of the Royal Society A, 365(1851), 539–560.456

    Soize, C. and Ghanem, R. (2004). “Physical systems with random uncertainties: Chaos rep-457

    resentations with arbitrary probability measure.” SIAM Journal on Scientific Computing,458

    26(2), 395–410.459

    Spiridonakos, M. and Chatzi, E. (2014). “Polynomial chaos expansion models for shm under460

    environmental variability.” Proceedings of Eurodyn 2014, 9th International Conference on461

    Structural Dynamics, June 30 July 2, 2014, Porto, Portugal.462

    Spiridonakos, M. and Chatzi, E. (2015). “Stochastic structural identification from vibrational463

    and environmental data.” In: Beer M., Patelli E., Kougioumtzoglou I., Au I. (Ed.) En-464

    cyclopedia of Earthquake Engineering: SpringerReference (www.springerreference.com).465

    Springer-Verlag Berlin Heidelberg, 1–16.466

    Wenzel, H. (2009). “The influence of environmental factors.” Encyclopedia of Structural467

    Health Monitoring, Y. F. C. Boller, F.-K. Chang, ed., John Wiley & Sons, Ltd, Chapter468

    120, 1–9.469

    Wiener, N. (1938). “The homogeneous chaos.” American Journal of Mathematics, 60(4), pp.470

    897–936.471

    Xiu, D. and Karniadakis, G. E. (2002). “The wiener-askey polynomial chaos for stochastic472

    differential equations.” SIAM Journal on Scientific Computing, 24(2), 619–644.473

    Yuen, K. and Kuok, S. (2010). “Ambient interference in long-term monitoring of buildings.”474

    Engineering Structures, 32(8), 2379–2386.475

    19

  • List of Tables476

    20

  • TABLE 1. Types of Wiener-Askey polynomial chaos and their underlying random vari-ables.

    PDF Support Polynomials

    Normal (Gaussian) (−∞,∞) HermiteGamma [0,∞) LaguerreBeta [0, 1] Jacobi

    Uniform [−1, 1] Legendre

    21

  • TABLE 2. Types of artificial damages induced on bridge Z24 (Reynders and De Roeck2009).

    # Date (1998) Damage type

    1 4.8 First reference measurement

    2 9.8 Second reference measurement

    3 10.8 Lowering of pier, 20 mm

    4 12.8 Lowering of pier, 40 mm

    5 17.8 Lowering of pier, 80 mm

    6 18.8 Lowering of pier, 95 mm

    7 19.8 Tilt of foundation

    8 20.8 Third reference measurement

    9 25.8 Spalling of concrete, 24 m2

    10 26.8 Spalling of concrete, 12 m2

    11 27.8 Landslide at abutment

    12 31.8 Failure of concrete hinge

    13 2.9 Failure of anchor heads I

    14 3.9 Failure of anchor heads II

    15 7.9 Rupture of tendons I

    16 8.9 Rupture of tendons II

    17 9.9 Rupture of tendons III

    22

  • TABLE 3. Errors of the polynomial chaos expansion .

    Natural freq. (Hz) std(PCE error) (Hz)

    ω1 0.0266

    ω2 0.0526

    ω3 0.1001

    ω4 0.1346

    23

  • List of Figures477

    24

  • FIG. 1. Schematic diagram of the proposed identification method.

    25

  • FIG. 2. Flowchart of the ICA algorithm.

    26

  • FIG. 3. Test-case: Überführung Bärenbohlstrasse bridge.

    27

  • FIG. 4. The first four identified natural frequencies of the bridge for the variousdatasets of the long-term monitoring system.

    28

  • FIG. 5. The first four identified natural frequencies versus (a) temperature; (b) hu-midity.

    29

  • FIG. 6. Histograms of (a) the measured temperature and humidity; (b) the latentvariables occurring after application of the ICA, along with corresponding scatter plots.

    30

  • FIG. 7. PCE natural frequency estimates for both the estimation and validation setcontrasted to the true NExT-ERA based estimates.

    31

  • FIG. 8. Extracted feature variable compared to the PCE modeling estimates (top) andthe extracted damage index based on the PCE prediction errors for the ÜberführungBärenbohlstrasse.

    32

  • FIG. 9. The Z24-Bridge: longitudinal section and top view (Reynders and De Roeck2009).

    33

  • FIG. 10. Cross-section of the Z24-Bridge and location of the thermocouples (Reyndersand De Roeck 2009).

    34

  • FIG. 11. The first four natural frequency estimates obtained through the SSI methodplotted against the temperature measured at the center of the web.

    35

  • FIG. 12. Histograms of the measured temperatures (upper row and left column) alongwith the scatter plots and the estimated correlation values.

    36

  • FIG. 13. Histograms of the ICA-based estimated independent random variables (upperrow and left column) along with the scatter plots and the estimated correlation values.

    37

  • FIG. 14. PCE natural frequency estimates for both the training and validation setcontrasted to the true SSI-based estimates.

    38

  • FIG. 15. Normalized eigenvalues of a) the temperature vector and b) natural frequencyestimates vector.

    39

  • FIG. 16. PCE estimates for both the estimation and validation set based on a singleinput and a single output variable.

    40

    Polynomial Chaos Expansion (PCE)Independent Component Analysis (ICA)Implementation on Systems under Operational ConditionsImplementation on Systems under Damaged ConditionsVibration and environmental dataResults