Line 2: Line 2:
  
  
<div id="_Hlk479174183" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
<div id="_GoBack" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
<big>'''Based Time-Scale Analysis Modal Identification using Mobile Sensors'''</big></div>
+
<span style="text-align: center; font-size: 75%;">'''Modal Identification with Mobile Sensors Using Cohen’s Class Time-Frequency Distributions'''</span></div>
 +
-->
  
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
'''Abstract:''' '' Dynamic characterization of structures from field measurements is useful for different purposes (e.g. retrofit validation, model updating, structural health monitoring, etc.). The identification of high spatial density mode shapes has been recently a challenge tackled using mobile sensors. These sensors travel over the structure and continuously acquire vibration data that is used to identify modal coordinates with a higher spatial density than can generally be obtained using a limited number of stationary sensors. The recorded signal from a mobile sensor is non-stationary, thus, it has significant variations in its spectral content over time, requiring a suitable processing to extract the properties not only for the time but also for the frequency domain. In this paper, Cohen’s class Time-Frequency Distributions (TFD) are proposed for the output-only dynamics identification of structures based on non-stationary signals recorded with mobile sensors. Identification is achieved through cross-time-frequency estimators using Smoothed Pseudo-Wigner-Ville (SPWVD) distribution. Results from numerical simulations using a simply supported beam subject to ambient vibration are shown and the sensibility of the proposed identification to the presence of measurement noise is evaluated. Numerical results show that use of the cross-time-frequency estimators is effective in extracting modal properties of the structures and filtering noise.''
<span style="text-align: center; font-size: 75%;">''MARIO A. MARMOLEJO C.''</span></div>
+
  
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
'''Keywords:''' ''Modal Identification, Mobile Sensors, Non-Stationary Signals, Time-Frequency Distributions''
<span style="text-align: center; font-size: 75%;">''Assistant Professor, Universidad del Quindio, Armenia, Colombia''</span></div>
+
  
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
==1. Introduction==
<span style="text-align: center; font-size: 75%;">''M.Sc, Ph.D Candidate, Universidad del Valle, Cali, Colombia,  ''[mailto:mario.marmolejo@correounivalle.edu.co ''mario.marmolejo@correounivalle.edu.co'']</span></div>
+
  
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Structural Health Monitoring (SHM) Systems usually play a crucial duty in early failure or damage detection for construction, maintenance or catastrophe management. Civil infrastructure and many other sectors can greatly benefit from systems that monitor behaviour under normal operating conditions,  without the need of interrupting regular day-to-day activity (Brownjohn, 2007; Karbhari et al., 2009).
<span style="text-align: center; font-size: 75%;">''Calle 13 100-00 Cali, Colombia. mobile: +57-321-706-5859''</span></div>
+
  
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
The future trend in instrumentation for SHM is the use of smart mobile sensing along with wireless intelligent networks, as they not only provide similar information to conventional cabled systems, but can also be installed at a much lower cost and process data autonomously using their embedded microprocessors and software (Lynch, 2006; Spencer et al., 2004). Modal shapes are traditionally identified with stationary sensors fixed at locations that provide representative structural responses, however, these schemes contain restricted spatial information (Figure 1). If more spatial and temporal data is obtained from a vibrating structure, the structure’s response can be evaluated more accurately, and hence the advantage of implementing the structure with as many stationary sensors as possible during the acquisition of data. However, despite the improved accuracy in estimations, these or dense sensor array approaches can be impractical for many reasons including the cost of sensors and setup time, memory constrains, network reliability, power requirements, and physical limitations due to structure geometries (Magalhães et al., 2008, 2009; Matarazzo T. J. and Pakzad, S. N. 2013). Mobile sensing is a novel alternative in Structural Health Monitoring (SHM). The objective is to solve failings from fixed sensor array setups. Few sensors are usually used to collect dense spatial information in mobile sensing setups. Mobile sensing provides several advantages over schemes using stationary sensors, but the main advantage is that a single mobile sensor can be used to record signals continuously along the structure. Modal coordinates are identified with a higher spatial density than using stationary sensors because the mobile sensor continuously records vibration data as it travels along trajectories on the structure (Figure 2). Mobile sensing is not only easier to implement, but also more cost effective when compared to dense stationary sensor arrays (Marulanda et al., 2016; Marulanda, J., 2014; Matarazzo T. J.  and Pakzad S. N., 2013).
<span style="text-align: center; font-size: 75%;">'' ''</span></div>
+
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
<span style="text-align: center; font-size: 75%;">''JOHANNIO MARULANDA''</span></div>
+
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image2.png|600px]] </span></div>
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
<span style="text-align: center; font-size: 75%;">''Ph.D, Associate Professor, Universidad del Valle, Cali, Colombia,  ''[mailto:johannio.marulanda@correounivalle.edu.co ''johannio.marulanda@correounivalle.edu.co'']</span></div>
+
<span style="text-align: center; font-size: 75%;">Figure 1. Spatial resolution of mode shapes: stationary sensors</span></div>
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
<span style="text-align: center; font-size: 75%;">''PETER THOMSON''</span></div>
+
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image3.png|600px]] </span></div>
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
<span style="text-align: center; font-size: 75%;">''Ph.D, Professor, Universidad del Valle, Cali, Colombia, ''[mailto:peter.thomson@correounivalle.edu.co ''peter.thomson@correounivalle.edu.co'']'' ''</span></div>
+
<span style="text-align: center; font-size: 75%;">Figure 2. Spatial resolution of mode shapes: mobile sensors</span></div>
-->
+
  
'''Abstract:''' '' ''Civil structures, once built and released for use, are expected to fulfil operational standards for extended periods. The fact, a civil structure´s life, is determined by temporary deterioration, harsh ambient, misuse, destruction, and, eventual rehabilitation for current use or conservation for historical reasons. Initially, vibrational analysis with stationary sensors, broad logistics and a large array of sensors to provide densely detailed data were required. The next evolutionary step; the simplification of the while-under-use data-acquisition stage; better, faster, more detailed parameters for diagnosis. With the structure under normal operation, a stationary reference sensor, a mobile sensor able to travel through specific trajectories over the structure, deep computational data analysis provides better, deeper, clearer frequencies, mode shapes and damping ratios. A structural health monitoring system should be able to reliably assess structural wellness under operational conditions. Such are the benefits of a mobile sensor monitoring and diagnosis system. This paper utilizes the time–scale analysis method and the Continuous Wavelet Transform (CWT) through a complex Morlet wavelet to extract modal parameters from the signal responses. The theoretical development is validated by numeric simulations on a simply supported beam and by double experimental confirmation on a pedestrian bridge at Universidad del Valle, in Cali (Colombia), by the technique herein, and by a stationary sensor scheme.
+
Recent implementation of mobile sensor networks has been diverse, although limited, and is still under development. Zhu et al. (2010, 2012) proposed a sensing device. Such device will record data at selected nodes, but will not do so while the sensor is moving.  Sibley et al. (2002) and Dantu et al. (2005) implemented Robomote, miniature low cost robots for mobile sensors networks. Partial modal identification include the use of a moving vehicle to identify frequencies of a single span bridge using frequency domain techniques (Cerda F. et al., 2012; Lin and Yang, 2005). Marulanda et al. (2016) and Marulanda (2014) developed a novel technique that uses two sensors, one mobile and the other stationary, in which the mobile sensor records continuously, and, in a subsequent procedure, performs the modal identification through spectrograms. Matarazzo et al. (2014, 2016) proposed a method of collecting data through continuous mobile detection in the presence of missing time and space observations. As the corresponding matrices have incompatible data records, they developed and successfully implemented an algorithm, called STRIDE, to cover for such “empty spaces”.
  
'''Keywords:''' ''Modal Identification, Mobile Sensors, Time-Scale Analysis, Continuous Wavelet Transform. ''
+
The use of mobile sensors, typically accelerometers, require an analysis of non-stationary response signals. Time-Frequency analysis has been proposed by several authors to analyse non-stationary responses in structural identification and in the assessment of mechanical damage. Staszewski et al. (1994) and Rioul et al. (1991) have  devoted several papers to modal identification using  the Short-Time Fourier Transform (STFT) and Wavelet Transform (WT).  The STFT transforms the signal in a two-dimensional time-frequency plane, assuming that the signal is stationary when viewed through a limited extension window. The Fourier Transform of the windowed signal leads to a time-frequency distribution (Gabor, 1946; Arango, 2009; Oppenheim, 2010). The STFT is linear and therefore provides poor energy information about the signal. Quadratic Time-Frequency Distributions (TFDs), on the other hand, allow the interpretation of representations from a much richer energy point of view (Hlawatsch and Boudreaux-Bartels, 1992; Loughlin et al., 1993).  Marginal properties express such interpretation. However, in view of the Uncertainty Principle, such properties are not sufficient to identify an energy density at each point on the time-frequency plane. (Ceravolo, 2009). A spectrogram, or quadratic representation of the amplitude of the STFT, is a three dimensional representation of the time-averaged Fourier Transform over adjacent time segments of a random process (Bendat and Piersol, 2011). It does not fulfil marginal properties and violates the Linearity Principle, generating crossed or interference terms. These spurious terms are restricted to those regions of the time-frequency domain where the auto-terms overlap. In the specific case of the spectrogram, for two sufficiently separated components in the time-frequency plane, its interference terms are practically nil (Ceravolo, 2009). Spectrograms can be used to obtain modal coordinates from the signals recorded from a mobile and a stationary sensor. These signals are divided into quasi-stationary segments and then the auto-spectral density and cross-spectral density functions for segments of both records are calculated. Finally  a modal identification procedure  is performed (Marulanda et al., 2016; Marulanda, J, 2014).
  
==1. Introduction==
+
In this paper, the non-stationary signals analysis recorded from sensors is performed by using Cohen’s class TFDs. The adoption of a quadratic representation helps in overcoming the limitations from  time-frequency resolution, as these transformations, the energetic and the correlative, are not based on the segmentation of signals. (Cohen, 1995). Invariance to time and frequency shifts characterize the Cohen’s Class Distributions (Hlawatsch and Auger, 2013). The system response is recorded in the time-frequency plane as the evolution of the spectral components corresponding to the energy of individual vibration modes whenever the Cohen’s Class Distributions is in use (Bonato et al., 1998, 1999; Bonato et al., 2000; Bonato P. et al., 1997; Cohen, 1995; Hammond and White, 1996; Hlawatsch and Boudreaux-Bartels, 1992). The proposed technique assumes that the frequency range of the system input spans the vibration modes. As the input spectrum increases, so does the identification accuracy. The instantaneous amplitude ratios directly determine the modal amplitude ratios for the time-frequency representations of signals recorded from a mobile sensor and a stationary sensor. Additionally, the phase of the cross-time- frequency representation between the two signals estimate the phase relationships. The estimators retain their dependence on time since they are derived directly from the two-dimensional functions of frequency and time variables. modal responses in linear time invariant systems show a constant relationship between the amplitude and the phase, and, therefore, the identified mode shapes are characterized by their stability over time (Bonato et al., 1998).
  
Structural Health Monitoring (SHM) is an essential tool to determine the safety, usability and maintainability of critical structures such as bridges, buildings or dams for example. Structural modal parameters, that is natural frequencies, mode shape vectors, and damping ratios, are the most commonly used condition indexes for vibration based SHM and damage detection. Currently, a modal model of a structure can be derived by an Operational Modal Analysis (OMA) showing significant advantages over procedures requiring known excitation inputs; low cost, broadband excitation, and the continuous use of the structures in real-life operational conditions [1 – 4].
+
==2. Theoretical Background==
  
In a traditional modal identification method (Fig. 1<span id='cite-Figure_1a'></span>[[#Figure_1a|(a)]]), a commonly large number of stationary sensors are distributed in pre-defined locations along the structure. The modal coordinates are identified over such discrete points. To improve the spatial resolution additional gear, additional sensor setups as well as experienced staff for proper configuration would be required, a process which is further challenged by the physical constraints of a structure.
+
''' Cohen’s Class Distributions'''
  
In mobile sensor setups (Fig. 1<span id='cite-Figure_1b'></span>[[#Figure_1b|(b)]]), sensors are mounted on mobile platforms to move along the structure as desired. A mobile sensing scheme offers clear improvements where a fixed sensor scheme falls short. With mobile sensing, few sensors, almost basic connectivity and lesser setup testing time can provide more spatial information than a dense fixed sensors network [5 – 8].
+
Cohen’s class distributions display a clear invariance of their representations for time and frequency shifts, a desirable property when correlating signal characteristics to phenomena occurring in the physical system generating the signal. The following equation  describes all the Cohen’s class distributions (Hlawatsch and Auger, 2013):
  
{| style="width: 100%;border-collapse: collapse;"  
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"
 +
|-
 +
|
 +
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
|  style="text-align: center;vertical-align: top;width: 100%;"|[[Image:Draft_Marmolejo_177547576-image2.png|444px]]
+
| <math>{D}_{ss}\left( t,f\right) =\int_{-\infty }^{+\infty }\int_{-\infty }^{+\infty }\int_{-\infty }^{+\infty }s\left( u+\frac{1}{2}\tau \right) {s}^{\ast }\left( u-\frac{1}{2}\tau \right) \Phi (\theta ,\tau ){e}^{-j2\pi \theta t-j2\pi \tau f+j2\pi \theta u}dud\tau d\theta</math>
 +
|}
 +
|  style="text-align: right;width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(1)</span>
 +
|}
  
<span id='Figure_1a'></span>(a)
 
  
 +
<span style="text-align: center; font-size: 75%;">where </span> <math display="inline">s\left( t\right)</math> <span style="text-align: center; font-size: 75%;"> is the signal, </span> <math display="inline">{s}^{\ast }\left( t\right)</math> <span style="text-align: center; font-size: 75%;">is its complex conjugate, </span> <math display="inline">t</math><span style="text-align: center; font-size: 75%;"> is time, </span> <math display="inline">f</math><span style="text-align: center; font-size: 75%;"> is frequency and </span> <math display="inline">\tau</math> <span style="text-align: center; font-size: 75%;"> is the time delay, </span> <math display="inline">\omega =</math><math>2\pi f</math><span style="text-align: center; font-size: 75%;"> is the circular frequency; </span> <math display="inline">\Phi (\theta ,\tau )</math><span style="text-align: center; font-size: 75%;"> is the kernel of the distribution and </span> <math display="inline">D\left( t,f\right)</math> <span style="text-align: center; font-size: 75%;">is the Cohen Class TFD. The cross-time-frequency distribution between two signals </span> <math display="inline">s\left( t\right)</math> <span style="text-align: center; font-size: 75%;">and </span> <math display="inline">y\left( t\right)</math> <span style="text-align: center; font-size: 75%;">can be defined as follows:</span>
  
 +
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"
 
|-
 
|-
| style="text-align: center;vertical-align: top;width: 100%;"|[[Image:Draft_Marmolejo_177547576-image3.png|438px]]
+
|  
 +
{| style="text-align: center; margin:auto;width: 100%;"
 +
|-
 +
| <math>{D}_{sy}\left( t,f\right) =\int_{-\infty }^{+\infty }\int_{-\infty }^{+\infty }\int_{-\infty }^{+\infty }s\left( u+\frac{1}{2}\tau \right) {y}^{\ast }\left( u-\frac{1}{2}\tau \right) \Phi (\theta ,\tau ){e}^{-j2\pi \theta t-j2\pi \tau f+j2\pi \theta u}dud\tau d\theta</math>
  
<span id='Figure_1b'></span>(b)
 
  
 
+
|}
|-
+
|  style="text-align: right;width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(2)</span>
|  style="text-align: center;vertical-align: top;"|Figure 1. Spatial resolution of mode shapes: (a) stationary sensors and (b) mobile sensors
+
 
|}
 
|}
  
  
Structural dynamic properties identification through Mobile Sensing has been a most concerned issue for the last decade. Recent studies have shown that mobile sensing schemes can record as much spatially dense information as stationary sensor schemes.
+
<span style="text-align: center; font-size: 75%;">The Wigner–Ville distribution (WVD) is the basic TFD and is a quadratic form that measures a local time-frequency energy. This distribution is obtained from equation (1) using </span> <math display="inline">{\Phi }_{WV}\left( \theta ,\tau \right) =</math><math>1</math><span style="text-align: center; font-size: 75%;">. Although the WVD satisfies a several important mathematical properties (Bonato et al., 2001; Flandrin, 1984; Wang and McFadden, 1993), it is not very suitable for applications with multiple components signals since the bi-linearity of the transform causes the appearance of interference terms . These are spurious terms when using the time-frequency characteristics of the signal for modal identification. The smoothed pseudo-Wigner-Ville transform (SPWVD) results from </span> <math display="inline">{\Phi }_{SPWV}\left( \theta ,\tau \right) =</math><math>\pi {e}^{-{T}^{2}{\pi }^{2}{\theta }^{2}}{e}^{-{B}^{2}{\pi }^{2}{t}^{2}}</math><span style="text-align: center; font-size: 75%;">, where  </span> <math display="inline">B</math><span style="text-align: center; font-size: 75%;"> and  </span> <math display="inline">T</math><span style="text-align: center; font-size: 75%;"> are the standard deviations in time and frequency, respectively. With the exception of real value, time-shift-invariance, and frequency invariance, the application of the smoothed function causes the loss of most of the mathematical properties of the WVD, (Hlawatsch and Boudreaux-Bartels, 1992). The foregoing independence between windows makes the SPWVD very versatile for reducing cross—terms in the WVD. This property is useful for modal identification.The Choi-Williams Distribution (CWD) (Choi and Williams, 1989; Papandreou and Boudreaux-Bertels, 1993) is defined by </span> <math display="inline">{\Phi }_{CW}\left( \theta ,\tau \right) =</math><math>{e}^{\frac{-{\theta }^{2}{\tau }^{2}}{\sigma }}</math><span style="text-align: center; font-size: 75%;">. This kernel is defined in the ambiguity plane </span> <math display="inline">\left( \theta ,\tau \right)</math> <span style="text-align: center; font-size: 75%;">.  If </span> <math display="inline">\sigma</math> <span style="text-align: center; font-size: 75%;">is very large, the kernel vanishes, therefore leading to the WVD, otherwise, the kernel application will have a filtering effect. For small values ​​of </span> <math display="inline">\sigma</math> <span style="text-align: center; font-size: 75%;">, the effect of the multiplication is to preserve the ambiguity function close the origin of the plane </span> <math display="inline">\left( \theta ,\tau \right)</math> <span style="text-align: center; font-size: 75%;">, therefore, this kernel satisfies the minimization property (Cohen, 1995). In multicomponent signals the authentic terms are usually close to the ambiguity plane axis, and the interference terms are dispersed away. The resulting characteristic function reduces the interference terms without significantly affecting the signal components (Bonato et al., 1997). Therefore, although the CWD violates the time and frequency support properties, it can be shown that the introduced error is generally negligible (Lippmann, 1989; Papandreou and Boudreaux-Bertels, 1993), so the </span> <math display="inline">\sigma</math> <span style="text-align: center; font-size: 75%;"> value should be adapted to each individual case, since the location of the interference terms and the signal components change with respect to the characteristics of the signal (Bonato et al., 1997). In this paper, SPWVD distribution is used. In this case, by parameterizing the windows, as they attenuate the interference terms without significantly modifying the signals, good modal identification results are obtained.</span>
  
Zhu et al. (2010) [9], with the introduction of flexural based mobile node sensors with magnetic capabilities of climbing over steel structures, facilitated gathering relevant isolated structural data with a minimum toil, automating spatial arrangements as with a stationary sensor network. Later, in 2012 [10], Zhu et al. redesigned the flexural sensors to attain better navigation over the structural elements. An experiment with five configurations to obtain information from four nodes, each mobile sensor stopping, gathering and recording ambient vibration. From the data acquired using those flexural mobile node sensors, three modal shapes were completely and precisely determined. Despite the sensing configuration restricted spatial information potential, there was a resolution enhancement of the record sets, when compared to a stationary sensor scheme.  In consequence, the data was treated as arising from stationary sensors, and the NExT-ERA technique used for its dynamic identification (James et al., 1993) [11].
+
''' Modal Identification Using Mobile Sensors'''
  
Yang et al. (2004) [12] developed the equation that describe the dynamic response of a vehicle as it passes over a bridge. The paper demonstrates that it is possible to determine the bridge´s fundamental frequencies from the vehicle´s registered acceleration data. Those structural dynamic properties are confirmed through a finite element analysis. Lin and Yang (2005) [13] towed an instrumented dolly truck over the Da-Wun-Lan Bridge in Taiwan, used three different speeds (13, 17 and 35 km/h) and experimentally double-checked the equations. They successfully extracted the bridge´s fundamental frequencies from the dolly power spectrums. Later simulations and parametric studies deeply explored two of the bridge´s fundamental frequencies (Yang and Chang (2009)) [14]. Siringoringo and Fujino (2012) [15] developed a bridge-vehicle analytical model. Parametric studies, Finite elements simulations and a field test at three at 10, 20 and 30 km/h were run over the vehicle speed, the vehicle frequency, and the bridge frequencies. An estimation of an instrumented dolly acceleration data Power Spectrum Density successfully identified the first frequency.
+
The proposed methodology requires a mobile sensor that acquires data continuously along the structure to be identified and simultaneous acquisition by a stationary sensor. Under ambient excitation using a zero mean Gaussian white noise process, the outputs of the system are also zero mean Gaussian processes, exciting the fundamental frequencies (Bendat and Piersol, 2011). Since the recorded signal from the mobile sensor is non-stationary, a procedure based on the Cohen Class Distributions (i.e. SPWVD) is proposed. These distributions are invariable to time and frequency shifts, which is required for proper system identification.
  
González et al (2012) [16] proposed a model to estimate a bridge´s damping relation based on the acceleration data from a moving vehicle. Via a vehicle-bridge interaction model (VBI) technique that assumes Rayleigh´s structural damping through a double acceleration data integration, the displacement of the bridge under the vehicle´s wheel is estimated. This method was verified by means of simulated finite element models, considering several bridge lengths, 21 vehicle speeds, and 9 damping ratios. The sensitivity of the estimation with respect to surface roughness, measuring noise, and particular modelling inaccuracies, were also studied.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image4.png|600px]] </span></div>
  
Marulanda et al. (2016) [5] developed a mode shape identification procedure requiring two sensors: one mobile and one stationary. Using short time Fourier Transform, the mode shapes are densely extracted from a space-frequency representation. The fundamental frequencies required to obtain the mode shapes can be determined through some of the known techniques: Stochastic Subspace Identification (SSI); Peak-Picking (PP); Natural Excitation Technique - Eigensystem Realization Algorithm (NExT-ERA). A theoretical example, a simple supported beam geared with a stationary and a mobile sensor, assuming noise free data and known natural frequencies, precisely identified the first three modes shapes through 479 coordinates, showing how much richer the mobile sensor spatial information is. A laboratory experimental configuration: a 7.010 mt. simple supported steel beam; a mobile wireless iMote2 sensor car; a stationary sensor. The structure is then stimulated with white Gaussian noise from a dynamic shaker and a rubber hammer. Three mode shapes with 21 vertical coordinates each one was successfully identified, and a dense network of stationary sensors verified the result.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;">Figure 3. Schematic description of the proposed setup on a simply supported beam.</span></div>
  
Li et al (2018) [17] proposes a bridge dynamic parameter determination technique based on Stochastic Subspace Identification (SSI), on a two-vehicle dynamic response setup. At the bridge supports ambient excitation and regular traffic are simulated as white Gaussian Noise.  To estimate the modal shapes the technique requires two instrumented vehicles, one acting as a stationary reference, and the other as a moving sensor. The modal frequencies are obtained from the vehicle´s output. The mode shapes are obtained from the vehicles´ divided and processed recorded acceleration via Reference Based SSI method. Bridge deck roughness, measuring noise and vehicle properties were also mode shape identified and numerically analyzed and what their influence was. A vehicle-bridge interaction model was used in the experimental validation of the proposed technique. The first two mode shapes were identified in the numerical example and in the experiment. A very relevant conclusion from the paper and the technique is that the slower the velocity and the lighter the car, the more precise the results.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image5.png|600px]] </span></div>
  
Matarazzo and Pakzad (2016) [18] introduced a formulation for structural identification using expectation maximization (STRIDE). STRIDE is an algorithm that can process from mobile sensor networks which record time data simultaneously while moving in the space. An incomplete time series is created regarding a unique sampling location. The paper considers a hypothetical network of dense sensors (a spatial network) whose sampling locations coincide with the mobile sensor network. Two trials of the STRIDE at the Golden State Bridge exemplified the implementation using ambient vibration data: One application of missing packages (MP), with 20% of missing data; one 10-mobile-sensor simulated mobile sensors network, with 82% of missing data. Both applications successfully identified 19 mode shapes (vertical and torsional), a frequency accuracy below 1 Hz, a satisfactory damping consistency when STRIDE in a minimum order (p=2) was used. Additionally, the spatial resolution of the estimated mode shapes was comparable with one obtained through 49 stationary sensors, with MAC values of 0.97 and above, when calculated to these mode shapes.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;">Figure 4. Equivalent model of the proposed setup showed in Fig. 3.</span></div>
  
A mobile sensor´s signal is non-stationary. Therefore, Modal Identification sourced from such sensors should be performed by time-frequency domain techniques, i.e. Short-time Fourier Transform (STFT) [5, 6], Wigner–Ville Transform (WVT), Continuous Wavelet Transform (CWT), etc. This paper covers an investigation based on CWT.
+
<span style="text-align: center; font-size: 75%;">Assuming that the structure to be identified is the one-dimensional beam shown in Figure 3 has </span> <math display="inline">n</math><span style="text-align: center; font-size: 75%;">discrete degrees of freedom (''d.o.f.''), wich each ''d.o.f.'' corresponds to a segment of the signal from the mobile sensor (Figure 4). </span>
  
<span id='_Hlk484106739'></span>The CWT of non-stationary signals reveals time-localized frequency details in multiple scales [19]. Ruzzene ''et al.'' [20] used the CWT to analyze free vibration responses of MDOF systems, estimate natural frequencies and damping ratios. The Random Decrement Technique was introduced to extract the impulse response from the measurements. Kijewski and Kareem investigated the time and frequency resolutions of the Wavelet-based System Identification method showing how to ensure the modal separation and to minimize the edge‐effect errors [21]. Within this context, previous researchers used different basis functions (e.g., Morlet, Complex Morlet, Cauchy, Harmonic) to identify modal parameters [19, 21, 22]. However, the performance of different basis functions varies in different applications and restricts their generalization.
+
<span style="text-align: center; font-size: 75%;">Let </span> <math display="inline">{s}_{i}\left( t\right)</math> <span style="text-align: center; font-size: 75%;"> be the response (displacement) in the </span> <math display="inline">i</math><span style="text-align: center; font-size: 75%;">th position of the system, </span> <math display="inline">{d}^{\left( k\right) }(t)</math><span style="text-align: center; font-size: 75%;"> the response (displacement) associated with the </span> <math display="inline">k</math><span style="text-align: center; font-size: 75%;">th vibration mode, and </span> <math display="inline">{\varnothing }_{i}^{\left( k\right) }</math><span style="text-align: center; font-size: 75%;"> a term of the normalized eigenvectors matrix, then </span> <math display="inline">{s}_{i}\left( t\right)</math> <span style="text-align: center; font-size: 75%;">can be in decouple form as,</span>
  
<span id='_Hlk484090487'></span>Yan ''et al''. [23] proposed an OMA method for a linear system using continuous-wavelet transmissibility with a special feature; the operational modal frequencies and mode shapes were extracted by combining different continuous-wavelet  transmissibilities at several wavelet scales with different transferring outputs. Le [1] developed the Frequency-scale Domain Decomposition technique (FSDD) using the Morlet mother wavelet for modal identification through ambient vibration responses. Le demonstrated the link between modal parameters and the local maxima of the CWT modulus.
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"
 
+
|-
This paper uses the time-scale method, Continuous Wavelet Transform, for modal identification trough a mobile sensing scheme. The proposed methodology includes the extraction of the natural frequencies from the Wavelet Power Spectrum peaks and the determination of the mode shapes via the transmissibility function between the mobile and the stationary sensor´s CWT.
+
|
 
+
{| style="text-align: center; margin:auto;width: 100%;"
This paper not only addresses the determination modal shapes and natural frequencies. The fundamental issue is the determination of close modes through the use a time-scale formulation and a mobile sensing scheme.
+
|-
 
+
| <math>{s}_{i}=\sum _{k}^{}{\varnothing }_{i}^{(k)}{d}^{(k)}~(t)</math>
A demonstration of the effectiveness of the proposed method via a simulated numerical system example and an experimental validation on a pedestrian bridge, using the proposed technique and the classical stationary sensor scheme [2, 24 – 26].
+
|}
 
+
|  style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;"> (3)</span>
==2. Time-Scale Analysis Based Mode Shape Identification Using Mobile Sensors==
+
|}
 
+
Consider a simply supported beam. Also consider a pair of sensors: a mobile one travelling along the structure; a stationary one placed in any given point of the structure. The signal coming from the mobile sensor is non-stationary, covering the length of the path, and therefore is a function of the speed of the sensor and the length of the structural element. Such signals must be processed under a joint time-frequency domain function. For the purpose of this paper the Continuous Wavelet Transform (CWT) function is used.
+
 
+
<span id='_GoBack'></span>If the structural element is a linear, slightly damped system, under ambient excitation loads, it can be modelled by a zero mean Gaussian white noise process. The system responses <math display="inline">x\left( t\right)</math> are also Gaussian processes with zero mean [23 – 25].
+
  
A CWT decouples a multi-component random signal into complex-monocomponent signals: a random deterministic response (impulse or step response); a random component with a null mean value and no significance. This decoupling capacity of multicomponent signals is adopted for estimating modal properties of a randomly excited system [1, 27].
 
  
The Continuous Wavelet Transform <math display="inline">{W}_{\psi }^{x}\left( a,b\right)</math> '' ''of a signal <math display="inline">x\left( t\right)</math> with respect to a mother wavelet <math display="inline">\psi \left( t\right)</math> is defined as:
+
<span style="text-align: center; font-size: 75%;">The term of the sum in equation (3) extends to the vibration modes that are detected in the </span> <math display="inline">i</math><span style="text-align: center; font-size: 75%;">th position corresponding to a discretized ''d.o.f.''   The structure under ambient excitation is equipped with a mobile sensor with constant velocity that continuously records the acceleration and, simultaneously, with a stationary sensor also recording the acceleration vibrations. For each </span> <math display="inline">ith</math><span style="text-align: center; font-size: 75%;">''d.o.f., ''the individual modal components  appear in the form of an energy peak in the time-frequency plane that can be written in complex form as (Bonato et al., 2000):</span>
  
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{W}_{\psi }^{x}\left( a,b\right) =\frac{1}{\sqrt{a}}\int_{-\infty }^{+\infty }x\left( t\right) {\psi }^{\ast }\left( \frac{t-b}{a}\right) dt</math>
+
| <math>{s}_{i}^{(k)}={\varnothing }_{i}^{(k)}{d}^{(k)}~\left( t\right) ={\varnothing }_{i}^{\left( k\right) }{\tilde{d}}^{\left( k\right) }(t){e}^{j2\pi {f}^{(k)}t}</math>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (1)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(4)</span>
 
|}
 
|}
  
  
where <math display="inline">a</math>  is the dilation or scale parameter, <math display="inline">b</math> localizes the dilated wavelet decomposition in time. In this study, the Morlet complex wavelet function, used as a mother wavelet, is formulated as:
+
<span style="text-align: center; font-size: 75%;">where </span> <math display="inline">{\tilde{d}}^{\left( k\right) }(t)=</math><math>{A}^{\left( k\right) }(t){e}^{j{\varphi }^{\left( k\right) }(t)}</math><span style="text-align: center; font-size: 75%;"> is a baseband signal related to the </span> <math display="inline">kth</math><span style="text-align: center; font-size: 75%;"> mode; </span> <math display="inline">{A}^{\left( k\right) }\left( t\right)</math> <span style="text-align: center; font-size: 75%;"> and </span> <math display="inline">{\varphi }^{\left( k\right) }(t)</math><span style="text-align: center; font-size: 75%;"> are the amplitude and phase modulation waveforms respectively; </span> <math display="inline">{f}^{(k)}</math><span style="text-align: center; font-size: 75%;"> is the natural frequency; </span> <math display="inline">{\varnothing }_{i}^{\left( k\right) }</math><span style="text-align: center; font-size: 75%;"> is the amplitude of the </span> <math display="inline">kth</math><span style="text-align: center; font-size: 75%;">  mode shape in the </span> <math display="inline">ith</math><span style="text-align: center; font-size: 75%;"> position. The complex form of equation (4) represents an analytical signal and, therefore, the spectrum of </span> <math display="inline">{d}^{(k)}~\left( t\right)</math> <span style="text-align: center; font-size: 75%;"> is single-sided. The WVD will transform the signal into:</span>
  
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">\psi \left( t\right) =\frac{1}{\sqrt{2\pi }}{e}^{-i{\omega }_{0}t}{e}^{-(\frac{{t}^{2}}{2})}</math>
+
| <math display="inline">{WVD}_{{s}_{i}^{\left( k\right) }{s}_{i}^{\left( k\right) }}\left( t,f\right) =</math><math>{\varnothing }_{i}^{{\left( k\right) }^{2}}{WVD}_{{\tilde{d}}^{\left( k\right) }{\tilde{d}}^{\left( k\right) }}\left( t,f\right) {\ast }_{f}\delta (f-</math><math>{f}^{(k)}</math><span style="text-align: center; font-size: 75%;">'')''</span>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (2)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">  (5)</span>
 
|}
 
|}
  
  
where <math display="inline">i</math>'' ''is the imaginary unit and <math display="inline">{\omega }_{0}</math> the central frequency.
+
<span style="text-align: center; font-size: 75%;">where </span> <math display="inline">{WVD}_{{\tilde{d}}^{\left( k\right) }{\tilde{d}}^{\left( k\right) }}</math><span style="text-align: center; font-size: 75%;"> is the WVD of the signal </span> <math display="inline">{\tilde{d}}^{\left( k\right) }(t)</math><span style="text-align: center; font-size: 75%;">, </span> <math display="inline">{\ast }_{f}</math><span style="text-align: center; font-size: 75%;"> is the convolution in frequency, and </span> <math display="inline">\delta (f-</math><math>{f}^{(k)}</math><span style="text-align: center; font-size: 75%;">'')'' is the Dirac delta (impulse) located in </span> <math display="inline">{f}^{(k)}</math><span style="text-align: center; font-size: 75%;">. In multi-component signals, when auto-terms and interference overlap, using a kernel capable of attenuating interference terms leads to an increased reliability of the estimated amplitude ratio. The use of highly selective kernels, ambiguity domain functions (Bonato et al., 2000), improves the capacity of the cross-time-frequency-transform to separate signal components, and therefore, allow amplitude ratio and phase difference estimators become close to each modal shape frequency. </span>
  
The Morlet wavelet provides a remarkable flexibility in obtaining a good time–frequency resolution since a harmonic frequency can be closely controlled by the correct selection of <math display="inline">{\omega }_{0}</math>, as well as by the scale parameter <math display="inline">a</math>.
+
The following expression represents the Cohen’s Class Transforms:
  
For mode shapes identification, a <math display="inline">\Delta t</math> size segmentation of the recorded signals, equivalent to the structural span, generates a vector used in determining the <math display="inline">n</math> discrete modal coordinates. It is assumed recorded signals are modulated amplitude.
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
+
Let  <math>\left[x_{1j}\left(t\right),x_{2j}\left(t\right),\ldots ,x_{nj}\left(t\right)\right]</math> and  <math display="inline">\left[x_{1j}^{ref}\left(t\right),x_{2j}^{ref}\left(t\right),\ldots ,x_{nj}^{ref}\left(t\right)\right]</math> be the signal segments coming from the mobile and the stationary sensors respectively, identified by the CWT for mode <math display="inline">j</math>'' ''. The wavelet ridge of the wavelet coefficients of mode <math display="inline">j</math> for each given signal segment  <math>\left[x_1\left(t\right),x_2\left(t\right),\ldots ,x_n\left(t\right)\right]</math> and  <math display="inline">\left[x_1^{ref}\left(t\right),x_2^{ref}\left(t\right),\ldots ,x_n^{ref}\left(t\right)\right]</math> will be:
+
 
+
<math display="inline">\left\{ \begin{matrix}{W}_{\psi }^{{x}_{1j}}\left( {a}_{j},b\right) =\frac{\sqrt{{a}_{j}}}{2}{A}_{1j}{e}^{-{\hat{\sigma }}_{j}b}{e}^{i\left( {\hat{\omega }}_{j}b+{\theta }_{01j}\right) }\\\vdots \\{W}_{\psi }^{{x}_{nj}}\left( {a}_{j},b\right) =\frac{\sqrt{{a}_{j}}}{2}{A}_{nj}{e}^{-{\hat{\sigma }}_{j}b}{e}^{i\left( {\hat{\omega }}_{j}b+{\theta }_{0nj}\right) }\end{matrix}\right.</math>      for mobile sensor (3)
+
 
+
<math display="inline">\left\{ \begin{matrix}{W}_{\psi }^{{x}_{1j}^{ref}}\left( {a}_{j},b\right) =\frac{\sqrt{{a}_{j}}}{2}{A}_{1j}^{ref}{e}^{-{\hat{\sigma }}_{j}b}{e}^{i\left( {\hat{\omega }}_{j}b+{\theta }_{01j}^{ref}\right) }\\\vdots \\{W}_{\psi }^{{x}_{nj}^{ref}}\left( {a}_{j},b\right) =\frac{\sqrt{{a}_{j}}}{2}{A}_{nj}^{ref}{e}^{-{\hat{\sigma }}_{j}b}{e}^{i\left( {\hat{\omega }}_{j}b+{\theta }_{0nj}^{ref}\right) }\end{matrix}\right.</math>      for stationary sensor (4)
+
 
+
The instantaneous ratios between <math display="inline">{W}_{\psi }^{{x}_{kj}}\left( {a}_{j},b\right)</math> '' '', and <math display="inline">{W}_{\psi }^{{x}_{kj}^{ref}}\left( {a}_{j},b\right)</math>  , are  <math display="inline">\frac{{W}_{\psi }^{{x}_{kj}}\left( {a}_{j},b\right) }{{W}_{\psi }^{{x}_{kj}^{ref}}\left( {a}_{j},b\right) }=</math><math>\frac{{A}_{kj}}{{A}_{kj}^{ref}}{e}^{i\left( {\theta }_{0kj}-{\theta }_{0kj}^{ref}\right) }</math>. Then the complex vector
+
 
+
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{U}_{j}\left( b\right) ={\left[ \frac{{W}_{\psi }^{{x}_{1j}}\left( {a}_{j},b\right) }{{W}_{\psi }^{{x}_{1j}^{ref}}\left( {a}_{j},b\right) },\frac{{W}_{\psi }^{{x}_{2j}}\left( {a}_{j},b\right) }{{W}_{\psi }^{{x}_{2j}^{ref}}\left( {a}_{j},b\right) },\ldots ,\frac{{W}_{\psi }^{{x}_{nj}}\left( {a}_{j},b\right) }{{W}_{\psi }^{{x}_{nj}^{ref}}\left( {a}_{j},b\right) }\right] }^{T}=</math><math>{\left[ \frac{{A}_{1j}}{{A}_{1j}^{ref}}{e}^{i\left( {\theta }_{01j}-{\theta }_{01j}^{ref}\right) },\frac{{A}_{2j}}{{A}_{2j}^{ref}}{e}^{i\left( {\theta }_{02j}-{\theta }_{02j}^{ref}\right) },\ldots ,\frac{{A}_{nj}}{{A}_{nj}^{ref}}{e}^{i\left( {\theta }_{0nj}-{\theta }_{0nj}^{ref}\right) }\right] }^{T}</math>
+
| <math>{D}_{ss}\left( t,f\right) ={\varnothing }_{i}^{{\left( k\right) }^{2}}\Gamma \left( t,f\right) {{\ast }_{t,f}WVD}_{ss}\left( t,f\right)</math>  
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (5)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">  (6)</span>
 
|}
 
|}
  
  
is defined by putting all these ratios together at each time <math display="inline">b</math>. Eq. 28 is an estimation of the time-evolution of the complex mode shape <math display="inline">{U}_{j}</math> [23]. The Real Part of <math display="inline">{U}_{j}(b)</math> corresponds to the <math display="inline">j</math>th mode shape <math display="inline">{\varnothing }_{j}(b)</math> [1, 22, 26, 28], which is:
+
<span style="text-align: center; font-size: 75%;">where </span> <math display="inline">{D}_{ss}\left( t,f\right)</math> <span style="text-align: center; font-size: 75%;"> is a Cohen class TFD, </span> <math display="inline">\Gamma \left( t,f\right)</math> <span style="text-align: center; font-size: 75%;"> is the kernel of the TFD in time-frequency plane.  </span> <math display="inline">{\ast }_{t,f}</math><span style="text-align: center; font-size: 75%;"> is the double convolution in time and frequency. From equation (6) it is evident that the energy of the transform concentrates close to the modal frequency </span> <math display="inline">{f}^{(k)}</math><span style="text-align: center; font-size: 75%;">. The distribution shape is determined by the time-frequency transform of the modulating waveform </span> <math display="inline">{\tilde{d}}^{\left( k\right) }(t)</math><span style="text-align: center; font-size: 75%;">. Because the shape of the modulated waveform is kept in the time–frequency domain. The amplitude ratio between two modal components can be determined from auto-time-frequency representations, as follows (Bonato et al., 1998; Bonato et al., 2000):</span>
  
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{\varnothing }_{j}\left( b\right) =\mathrm{Re}\,({U}_{j}\left( b\right) )</math>
+
| <math>{\left. AR\left( t\right) \right| }_{f={f}^{(k)}}={\left. \sqrt{\frac{{D}_{{s}_{i}^{\left( k\right) }{s}_{i}^{\left( k\right) }}\left( t,f\right) }{{D}_{{s}_{j}^{\left( k\right) }{s}_{j}^{\left( k\right) }}\left( t,f\right) }}\right| }_{f={f}^{(k)}}=</math><math>\frac{{\varnothing }_{i}^{(k)}}{{\varnothing }_{j}^{(k)}}</math>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (6)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(7)</span>
 
|}
 
|}
  
  
A minimization procedure is required to define the optimized mode shape ( <math display="inline">{\hat{\varnothing }}_{j}</math>):
+
<span style="text-align: center; font-size: 75%;">And the phase can be  deteremined from cross-time-frequency representtions, as follows:</span>
  
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{\hat{\varnothing }}_{j}=\mathrm{min}\,\left( \int_{b\epsilon {\Omega }_{j}}^{\, }{\left[ {\varnothing }_{j}\left( b\right) -{\overline{\varnothing }}_{j}\right] }^{2}db\right)</math>
+
| <math>{\left. PH\left( t\right) \right| }_{f={f}^{\left( k\right) }}=phase{\left. \left\{ {D}_{{s}_{i}^{\left( k\right) }{s}_{j}^{\left( k\right) }}\left( t,f\right) \right\} \right| }_{f={f}^{\left( k\right) }}=</math><math>\Delta {\varphi }_{ij}^{(k)}</math>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (7)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(8)</span>
 
|}
 
|}
  
  
where <math display="inline">{\hat{\varnothing }}_{j}</math> is the optimized mode shape, <math display="inline">\, {\varnothing }_{j}\left( b\right)</math> '' ''is the mode shape identified at time <math display="inline">b</math>, <math display="inline">{\overline{\varnothing }}_{j}</math> indicate the mean value of the <math display="inline">{\varnothing }_{j}\left( b\right)</math> , and <math display="inline">\, {\Omega }_{j}</math> is the segment time interval.
+
<span style="text-align: center; font-size: 75%;">The TFDs of the signals from both sensors, mobile and stationary, were used in separating modal and residual components and in estimating the modal Amplitudes Ratios (AR); while cross-time-frequency representations were used in estimating the Phase Ratios (PR) (Bonato et al., 1998). The system is displayed in the time-frequency plane as the unfolding of spectral components corresponding to the energy of the predominant vibration modes. The ARs, gathered from their time-frequency representations, are determined from the ratio between the instantaneous amplitudes of the signal segments. The PRs are the ratio between the imaginary and the real part of the cross- time-frequency representation of two signal segments (Bonato et al., 2000). Once the cross terms have been filtered, it is possible to identify a potential error source in the vibration modes due to a close coupling between the components. In contrast to classical frequency analysis, the cross time-frequency transform between channels can filter the products among various components. Although being closely coupled in frequency, they are uncorrelated in time (Bonato P. et al., 1997). </span>
  
The fundamental frequencies, required in determining the mode shapes, are obtained via The Wavelet Power Spectrum (WPS), as shown below.
+
<span style="text-align: center; font-size: 75%;">In the frequency ranges in which a single modal component is predominant, the estimators tend to be a constant value over time.  Because the variance (</span> <math display="inline">{\sigma }_{TFD}</math><span style="text-align: center; font-size: 75%;">) is a measure of the density of the time-frequency representation, the peaks of the variance of the cross-time-frequency transform are concentrated close to fundamental frequencies and, therefore, such frequencies are located in the maximum of the following function:</span>
  
The square modulus of the Continuous Wavelet Transform of <math display="inline">x</math> is the scalogram, defined as:
+
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
+
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{SC}_{x}\left( a,b;\psi \right) =\frac{1}{\left| a\right| }{\left| \int_{-\infty }^{+\infty }x\left( t\right) {\psi }^{\ast }\left( \frac{t-b}{a}\right) dt\right| }^{2}</math>
+
| <math>{\sigma }_{TFD}=\int_{0}^{{T}_{S}}{\left[ {D}_{{s}_{i}^{\left( k\right) }{s}_{j}^{\left( k\right) }}\left( t,f\right) -\overline{{D}_{{s}_{i}^{\left( k\right) }{s}_{j}^{\left( k\right) }}\left( t,f\right) }\right] }^{2}dt</math>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (8)
+
| style="text-align: right;width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(9)</span>
 
|}
 
|}
  
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
 
 +
<span style="text-align: center; font-size: 75%;">where  </span> <math display="inline">{T}_{S}</math><span style="text-align: center; font-size: 75%;"> is the length of the signal, and  </span> <math display="inline">\overline{{D}_{{s}_{i}^{\left( k\right) }{s}_{j}^{\left( k\right) }}\left( t,f\right) }</math><span style="text-align: center; font-size: 75%;">  indicates the mean value of the cross-time-frequency distribution.</span>
 +
 
 +
==3. Numerical Simulations==
 +
 
 +
<span style="text-align: center; font-size: 75%;">A proposed example to evaluate the described procedure using a numerical model of a simply supported beam (Figure 3). The beam’s </span> <math display="inline">n</math><span style="text-align: center; font-size: 75%;">th natural vibration mode shape and corresponding natural frequency are given by the following equation:</span>
 +
 
 +
{| class="formulaSCP" style="width: 100%;border-collapse: collapse;width: 100%;text-align: center;"  
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{E}_{x}=\int_{-\infty }^{+\infty }\int_{-\infty }^{+\infty }{\left| {SC}_{x}\left( a,b;\psi \right) \right| }^{2}db\frac{da}{{a}^{2}}</math>
+
| <math display="inline">{\phi }_{n}\left( x\right) =\mathrm{sin}\,(\frac{n\pi x}{L})</math><span style="text-align: center; font-size: 75%;"> ; </span> <math display="inline">\, \, {\omega }_{n}=</math><math>\frac{{n}^{2}{\pi }^{2}}{{L}^{2}}\sqrt{\frac{EI}{{m}_{l}}}</math>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (9)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(10)</span>
 
|}
 
|}
  
  
Where <math display="inline">{E}_{x}</math> is an energy distribution of the signal in the time-scale domain, associated with the measure <math display="inline">\frac{da}{{a}^{2}}</math> [29]'''. ''' Additionally, the wavelet power spectrum is a measure of the energy density of the time-scale representation where the fundamental frequencies are concentrated close to the peaks of the frequency marginals [25, 30]:
+
<span style="text-align: center; font-size: 75%;">where </span> <math display="inline">L</math><span style="text-align: center; font-size: 75%;"> is the total length of the beam, </span> <math display="inline">E</math><span style="text-align: center; font-size: 75%;"> is the elasticity modulus of the material, </span> <math display="inline">I</math><span style="text-align: center; font-size: 75%;"> is the moment of inertia of the cross section, and </span> <math display="inline">{m}_{l}</math><span style="text-align: center; font-size: 75%;"> is the constant mass per unit length (Chopra, 2012). In the following numerical simulations. In this simulation is assumed that </span> <math display="inline">L=</math><math>60\, m</math><span style="text-align: center; font-size: 75%;">, </span> <math display="inline">E=</math><math>25\, GPa</math><span style="text-align: center; font-size: 75%;">, </span> <math display="inline">I=</math><math>6.75\, {m}^{4}</math><span style="text-align: center; font-size: 75%;">, </span> <math display="inline">{m}_{l}=</math><math>150\, \frac{kN}{\frac{g}{m}}</math><span style="text-align: center; font-size: 75%;">, and </span> <math display="inline">{\zeta }_{n}=</math><math>0.05</math><span style="text-align: center; font-size: 75%;">; and, therefore, the first three natural frequencies of the beam are </span> <math display="inline">1.45</math><span style="text-align: center; font-size: 75%;">, </span> <math display="inline">5.80</math><span style="text-align: center; font-size: 75%;">, and </span> <math display="inline">13.05\, Hz</math><span style="text-align: center; font-size: 75%;">. The stationary sensor was located at </span> <math display="inline">25\, m</math><span style="text-align: center; font-size: 75%;">from the left support and the mobile sensor had a velocity of </span> <math display="inline">\, 0.0167\frac{m}{s}</math><span style="text-align: center; font-size: 75%;">. The excitation of the beam corresponds to a Gaussian white noise acceleration at the supports. The simulated response signals were sampled at </span> <math display="inline">600\, Hz</math><span style="text-align: center; font-size: 75%;">, and then conditioned using a Hilbert transform and sub-sampled at </span> <math display="inline">30\, Hz</math><span style="text-align: center; font-size: 75%;">. SPWVD was used for the modal identification. The computed SPWVD was performed with a Hanning time smoothing window of </span> <math display="inline">39</math><span style="text-align: center; font-size: 75%;"> samples, Hanning frequency smoothing window of </span> <math display="inline">125</math><span style="text-align: center; font-size: 75%;"> samples and 256 frequency bins. The identified mode shapes </span> <math display="inline">{\Phi }_{j}</math><span style="text-align: center; font-size: 75%;">are compared to the exact mode shapes </span> <math display="inline">{\Phi }_{k}</math><span style="text-align: center; font-size: 75%;">using the Modal Assurance Criterion (MAC) via  equation (11) (Pastor et al., 2012). The damping is identified using the Half-Power Bandwidth method (Bonato P. et al., 1997; Chopra, 2012). The influence of noise in the identification is evaluated. The software used for signals processing using TFDS was The Time-Frequency Toolbox (TFTB) (Auger F. et al., 1996). </span>
  
{| class="formulaSCP" style="width: 100%; text-align: center;"  
+
{| class="formulaSCP" style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;width: 100%;text-align: center;"  
 
|-
 
|-
 
|  
 
|  
{| style="text-align: center; margin:auto;"  
+
{| style="text-align: center; margin:auto;width: 100%;"
 
|-
 
|-
| <math display="inline">{\left| X\left( \frac{{\omega }_{0}}{a}\right) \right| }^{2}=</math><math>\int_{-\infty }^{+\infty }{SC}_{x}\left( a,b;\psi \right) db</math>
+
| <math>{MAC}_{jk}=\frac{{({\Phi }_{j}^{T}{\Phi }_{k})}^{2}}{({\Phi }_{j}^{T}{\Phi }_{j})({\Phi }_{k}^{T}{\Phi }_{k})}</math>
 
|}
 
|}
| style="width: 5px;text-align: right;white-space: nowrap;" | (10)
+
| style="width: 5px;text-align: right;white-space: nowrap;"|<span style="text-align: center; font-size: 75%;">(11)</span>
 
|}
 
|}
  
  
where <math display="inline">a=\frac{{\omega }_{0}}{\omega }</math>, <math display="inline">\omega =</math><math>2\pi f</math>, and <math display="inline">{\left| X\left( \frac{{\omega }_{0}}{a}\right) \right| }^{2}</math> are the frequency marginals of the CWT.
+
<span style="text-align: center; font-size: 75%;">'''TFD-based mode shapes identification'''</span>
  
==3. Numerical Simulation (Closed Modal Frequencies)==
+
<span id='_Hlk488654953'></span><span id='_Hlk488655038'>Figure 5 shows the time history of the signal recorded with the mobile sensor traveling at 0.0167 m/s, along with the spectral density and auto-SPWVD. Figure 6 shows the time history of the signal recorded with the stationary sensor, along with its spectral density and auto-SPWVD. Figure 7 shows the time history of the signals recorded with the mobile and the stationary sensors along with their spectral density and cross-SPWVD. Figure 8 shows the standard deviation of the cross-SPWVD. The modal frequencies and damping ratios were obtained via Equation (9) by selecting peaks and using the Half-Power Bandwidth method (Figure 8). The mode shapes were obtained from TFDs (i.e. SPWVD), as given by Equations (7) and (8). Table 1 summarizes results of the modal parameter identification, where it is shown that errors in natural frequencies are generally less than 1.5 %. The identified mode shapes are also compared to the exact mode shapes. In this example, 59 modal coordinates were identified, and MAC values were all equal to one for the first three modes (Figure 9). MAC values for different mobile sensor velocity values, ranging from </span> <math display="inline">0.0167\frac{m}{s}</math><span style="text-align: center; font-size: 75%;"> to </span> <math display="inline">0.25\frac{m}{s}</math><span style="text-align: center; font-size: 75%;">, were calculated. Figure 10 shows the variations in the identified modal coordinates (IMC) per unity of beam’s length  versus  the mobile sensor’s velocity, it is assumed that two modes are considered correlated when the MAC value is  equal to 0.95, which corresponds to an angle of  18 degrees (Pastor et al., 2012). From Figure 11, it can be observed that good identification results have been obtained from the simulated ambient response data contaminated with different intensity levels of white gaussian noise.</span>
 
+
The numerical example is a simple supported beam which is added with <math display="inline">8000\, kg/m</math> symmetrically distributed masses (Fig. <span id='cite-Figure_2'></span>[[#Figure_2|2]]). For this beam <math display="inline">L=</math><math>60\, m</math>, <math display="inline">E=60\, GPa</math>, <math display="inline">I=</math><math>6.75\, {m}^{4}</math>, <math display="inline">{m}_{l}=150\, \frac{kN}{\frac{g}{m}}</math>, and <math display="inline">\zeta =</math><math>0.05</math>. The first five natural frequencies are: <math display="inline">0.194</math>, <math display="inline">0.435</math>, <math display="inline">2.511</math>, <math display="inline">2.961</math>, and <math display="inline">6.491\, Hz</math>.
+
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
[[Image:Draft_Marmolejo_177547576-image8.png|600px]] </div>
+
<span style="text-align: center; font-size: 75%;">Table 1. Summary of results obtained using equations (7), (8), (9) and (11)</span></div>
 
+
<div id="Figure_2" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Figure 2. Simple supported beam with symmetrically added masses </div>
+
 
+
Stationary sensor <math display="inline">19.5\, m</math> away from the left support. Mobile sensor velocity  <math display="inline">\, 0.0167\frac{m}{s}</math>. Gaussian white noise acceleration excitation input at the beam supports. Response signals conditioned using Hilbert Transform and sub-sampling. Morlet wavelet used for the modal identification. 256 frequency bins for this wavelet; Wavelength approximately equal to the square root of length of the signals (half-length the Morlet analyzing wavelet scale); lower and the upper normalized frequency bounds of the analyzed signal, <math display="inline">{f}_{min}=</math><math>0.01</math> and <math display="inline">{f}_{max}=0.50</math>.
+
 
+
Fig. <span id='cite-Figure_3'></span>[[#Figure_3|3]] and Fig. <span id='cite-Figure_4'></span>[[#Figure_4|4]] show the stationary sensor recorded signal time history along with its scalogram, and the mobile sensor recorded signal time history traveling at 0.0167 m/s along with its scalogram, respectively.
+
 
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
[[Image:Draft_Marmolejo_177547576-image9.png|438px]] </div>
+
 
+
<div id="Figure_3" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Figure 3. Stationary sensor recorded signal: time history, energy spectral density and Scalogram</div>
+
 
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
[[Image:Draft_Marmolejo_177547576-image10.png|450px]] </div>
+
 
+
<div id="Figure_4" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Figure 4. Mobile sensor recorded signal: time history, energy spectral density, and Scalogram</div>
+
 
+
Modal frequencies obtained by selecting peaks (Fig. <span id='cite-Figure_5'></span>[[#Figure_5|5]]); damping ratios obtained using the stationary sensor signal scalogram [1]; mode shapes obtained using the response signals wavelet transforms via Eqs. 6 and 7; comparison made between the identified and the exact mode shapes; in this example 59 modal coordinates were identified.
+
 
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
[[Image:Draft_Marmolejo_177547576-image11.png|450px]] </div>
+
 
+
<div id="Figure_5" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Figure 5. Mobile sensor Frequency Marginals (i.e. Natural Frequencies)</div>
+
 
+
Comparative results performed from the numerical simulation and the theoretical example; identified mode shapes <math display="inline">{\phi }_{j}</math> compared to the exact mode shapes <math display="inline">{\phi }_{k}</math> using the Modal Assurance Criterion (MAC) [31] (Fig. <span id='cite-Figure_6'></span>[[#Figure_6|6]]); damping ratio identified using technique identification by Le [1]; The Time-Frequency Toolbox software (TFTB) used for the wavelet signals processing [29, 32]. Table <span id='cite-Table_1'></span>[[#Table_1|1]] summarizes the results of the modal parameter identification. As shown, the errors in natural frequencies are less than 1.6%, the maximum modal damping error is 10.0 %, and MAC values are all very close to 1.0.
+
 
+
Fig. <span id='cite-Figure_7'></span>[[#Figure_7|7]]: The X axis is the mobile sensor’s velocity. The Y axis is the variations in the Identified Modal Coordinates (IMC) per beam’s length unit. Two modes are considered correlated when the MAC value is at least  equal to 0.95, corresponding to 18 degrees angle [31]. For instance, if the speed of the mobile sensor is 0.024 m/s, its correspondent IMC is 2 (Identified Modal Coordinates per meter).
+
 
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
[[Image:Draft_Marmolejo_177547576-image12.png|450px]] </div>
+
 
+
<div id="Figure_6" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Figure 6. Identified mode shapes</div>
+
 
+
<div id="Table_1" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Table 1. Numerical simulation results summary  </div>
+
  
 
{| style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;"  
 
{| style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;"  
 
|-
 
|-
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|MODE
+
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">MODE</span>
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|Theoretical frequencies
+
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">Theoretical frequencies</span>
  
''(Hz)''
+
<span style="text-align: center; font-size: 75%;">''(Hz)''</span>
|  colspan='3'  style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|Identification by wavelets
+
|  colspan='3'  style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">Identification by TFD (SPWVD)</span>
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|Frequency
+
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">Frequency Error</span>
  
error
+
<span style="text-align: center; font-size: 75%;">(%)</span>
 
+
(%)
+
|  rowspan='2' style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|Damping
+
 
+
error
+
 
+
(%)
+
 
|-
 
|-
|  style="text-align: center;vertical-align: top;"|'''Natural '''
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">'''Natural '''</span>
  
'''frequencies (Hz)'''
+
<span style="text-align: center; font-size: 75%;">'''frequencies (Hz)'''</span>
|  style="text-align: center;vertical-align: top;"|'''Damping'''
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">'''Damping'''</span>
  
'''ratios'''
+
<span style="text-align: center; font-size: 75%;">'''ratios'''</span>
|  style="text-align: center;vertical-align: top;"|'''MAC'''
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">'''MAC'''</span>
  
'''values'''
+
<span style="text-align: center; font-size: 75%;">'''values'''</span>
|-
+
style="text-align: center;vertical-align: top;"|1
+
|  style="text-align: center;vertical-align: top;"|0.194
+
|  style="text-align: center;vertical-align: top;"|0.191
+
|  style="text-align: center;vertical-align: top;"|0.047
+
|  style="text-align: center;vertical-align: top;"|1.000
+
|  style="text-align: center;vertical-align: top;"|1.546
+
|  style="text-align: center;vertical-align: top;"|6.000
+
|-
+
|  style="text-align: center;vertical-align: top;"|2
+
|  style="text-align: center;vertical-align: top;"|0.435
+
|  style="text-align: center;vertical-align: top;"|0.439
+
|  style="text-align: center;vertical-align: top;"|0.046
+
|  style="text-align: center;vertical-align: top;"|0.999
+
|  style="text-align: center;vertical-align: top;"|0.920
+
|  style="text-align: center;vertical-align: top;"|8.000
+
 
|-
 
|-
|  style="text-align: center;vertical-align: top;"|3
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1</span>
style="text-align: center;vertical-align: top;"|2.511
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1.450</span>
|  style="text-align: center;vertical-align: top;"|2.536
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1.465</span>
|  style="text-align: center;vertical-align: top;"|0.055
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">0.060</span>
|  style="text-align: center;vertical-align: top;"|0.999
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1.00</span>
|  style="text-align: center;vertical-align: top;"|0.996
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1.03</span>
|  style="text-align: center;vertical-align: top;"|10.000
+
 
|-
 
|-
|  style="text-align: center;vertical-align: top;"|4
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">2</span>
style="text-align: center;vertical-align: top;"|2.961
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">5.798</span>
|  style="text-align: center;vertical-align: top;"|2.969
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">5.801</span>
|  style="text-align: center;vertical-align: top;"|0.053
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">0.029</span>
|  style="text-align: center;vertical-align: top;"|0.998
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1.00</span>
|  style="text-align: center;vertical-align: top;"|0.270
+
|  style="text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">0.05</span>
|  style="text-align: center;vertical-align: top;"|6.000
+
 
|-
 
|-
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|5
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">3</span>
style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|6.491
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">13.046</span>
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|6.413
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">13.008</span>
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|0.052
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">0.025</span>
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|0.999
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">1.00</span>
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|1.202
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|<span style="text-align: center; font-size: 75%;">0.29</span>
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|4.000
+
 
|}
 
|}
  
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
[[Image:Draft_Marmolejo_177547576-image13.png|390px]] </div>
+
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image6.jpeg|378px]] </span></div>
  
<div id="Figure_7" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
Figure 7. IMC/m values versus mobile sensor velocity.</div>
+
<span style="text-align: center; font-size: 75%;">Figure 5. Signal recorded with the mobile sensor: time history, spectral density and SPWVD</span></div>
  
The noise influence was evaluated too. As per observable in Fig. <span id='cite-Figure_8'></span>[[#Figure_8|8]], signal-to-noise ratio, results for values greater than <math display="inline">15\, dB</math> lie very close to 1 (MAC value). Such good identification results were obtained from the simulated ambient contaminated response data.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image7.jpeg|378px]] </span></div>
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
[[Image:Draft_Marmolejo_177547576-image14.png|378px]] </div>
+
<span style="text-align: center; font-size: 75%;">Figure 6. Signal recorded with the stationary sensor: time history, spectral density and SPWVD</span></div>
  
<div id="Figure_8" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
Figure 8. MAC values and signal-to-noise ratio variation. (first five identified modes)</div>
+
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image8.jpeg|378px]] </span></div>
  
<span id='_Toc520379411'></span>
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;">Figure 7. Signals recorded with the stationary and the mobile sensors: time history, spectral density and SPWV cross-distribution.</span></div>
  
==4. Experimental Validation==
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image9.jpeg|378px]] </span></div>
  
The above presented procedure is applied to acceleration responses on a pedestrian bridge, at Universidad del Valle, Cali, Colombia. The bridge structure is a simply supported steel truss on concrete frames as shown in Figure  <span id='cite-Figure_9'></span>[[#Figure_9|9]]. The bridge superstructure has 12 m span and 1.2 m width. Its mass and effective flexural rigidity are approximately uniform over the entire span. The mass is about 310 kg/m.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;">Figure 8. Standard deviation from the SPWVD cross-distribution.</span></div>
  
The setup (Fig. <span id='cite-Figure_10'></span>[[#Figure_10|10]]), a mobile sensing platform (Fig. <span id='cite-Figure_11'></span>[[#Figure_11|11]]) and a stationary sensor (Fig. <span id='cite-Figure_12'></span>[[#Figure_12|12]]) located 5 meters from the support. Forced vibration tests carried out using white noise input. Vertical acceleration response along the axis of the bridge measured and recorded via a mobile sensing accelerometer and a 16-bits data acquisition system.
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image10.jpeg|378px]] </span></div>
  
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
[[Image:Draft_Marmolejo_177547576-image15.jpeg|390px]] </div>
+
<span style="text-align: center; font-size: 75%;">Figure 9. First three mode shapes identified using equations (7) and (8): 59 modal coordinates were identified.  </span></div>
  
<span id='Figure_18'></span><span id='_Toc520379453'></span><div id="Figure_9" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
Figure 9. Pedestrian Bridge</div>
+
<span style="text-align: center; font-size: 75%;"> [[Image:Marmolejo_et_al_2019b-image11.jpeg|378px]] </span></div>
  
{| style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;"  
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
|-
+
<span style="text-align: center; font-size: 75%;">Figure 10. IMC/m values versus mobile sensor velocity.</span></div>
style="text-align: center;vertical-align: top;width: 100%;"|[[Image:Draft_Marmolejo_177547576-image16.png|600px]]
+
|-
+
|  style="text-align: center;vertical-align: top;"|<span id='Figure_22'></span><span id='_Toc520379458'></span><span id='Figure_10'></span>Figure 10.  Mobile sensing setup
+
  
 +
'''Signal-to-Noise Ratio'''
  
|}
+
In most experimental applications, the acceleration record from the mobile sensor would be more contaminated with noise than the stationary sensor due to the motion system. To evaluate the effect of such noise on the identified modes through the MAC values, artificial normally distributed white noise was added to the acceleration response of the mobile to generate signal-to-noise ratios from -5 to 20 dB. Figure 11 shows the MAC values versus SNR for the first three identified modes. It can be seen from Figure 11 that the identification accuracy of all mode shapes can be maintained at a high level using the proposed method for SNR values higher than 15 dB.
  
 +
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
 +
[[Image:Marmolejo_et_al_2019b-image12.jpeg|378px]] </div>
  
{| style="width: 100%;border-collapse: collapse;"  
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
|-
+
<span style="text-align: center; font-size: 75%;">Figure 11. MAC values of the first three identified modes for variation in the signal-to-noise ratio.</span></div>
style="text-align: center;vertical-align: top;width: 50%;"|[[Image:Draft_Marmolejo_177547576-image17-c.jpeg|186px]]
+
style="text-align: center;vertical-align: top;width: 50%;"|[[Image:Draft_Marmolejo_177547576-image18.png|222px]]
+
|-
+
|  style="text-align: center;vertical-align: top;"|<span id='Figure_11'></span>Figure 11. Mobile sensing platform
+
  
 +
==4. CONCLUSIONS==
  
|  style="text-align: center;vertical-align: top;"|<span id='Figure_21'></span><span id='Figure_12'></span><span id='_Hlk11270928'></span>Figure 12.<big>''' '''</big>1-Axis Accelerometer Module
+
This paper presents the use of mobile sensors for modal identification through Cohen’s Class TFDs, providing identified modal shapes with high spatial resolution. Mobile sensors can be used in structures and mechanical systems under normal operating conditions and unknown excitation. It is envisaged that mobile sensors not only offer flexible but adaptable spatial resolutions that may enhance the theory and practice of future SHM and damage detection research and technological development.
  
 +
The mobile sensor continuously records vibration data as it travels along trajectories on the structure, therefore, modal coordinates are identified with a higher spatial density than using stationary sensors. The quantity of identified modal coordinates depends on the velocity of the sensor, the selected kernels, and the identified frequencies.
  
|}
+
The TFD-based modal identification utilized to identify the structural system in terms of natural frequencies, damping ratios and mode shapes, along with the use of mobile sensors complied successfully. The mode shapes are estimated through the auto and cross time-frequency transforms of the signals from mobile and stationary sensors (i.e. modulating waveforms). The proposed methodology is highly accurate when calculating mode shapes by determining the amplitude and phase ratios between TFDs of signals recorded with the mobile and the stationary sensors, respectively. The standard deviations of cross time-frequency transforms of the recorded signals are related to modal frequencies and damping ratios.
  
 +
<span style="text-align: center; font-size: 75%;">The effectiveness of the technique is clear, based on auto and cross-time-frequency estimators for structural identification when non-stationary signals are used. For velocities between </span> <math display="inline">0.25\frac{m}{s}</math><span style="text-align: center; font-size: 75%;"> to </span> <math display="inline">0.25\frac{m}{s}</math><span style="text-align: center; font-size: 75%;">,  the variations in the identified modal coordinates per unity of beam’s length ranges from </span> <math display="inline">20\, IMC/m</math><span style="text-align: center; font-size: 75%;"> to </span> <math display="inline">1.5\, IMC/m</math><span style="text-align: center; font-size: 75%;">. For a velocity of </span> <math display="inline">0.0167\frac{m}{s}</math><span style="text-align: center; font-size: 75%;">, 59 modal coordinates were identified, and MAC values were all equal to one for the first three modes using only two acceleration records. For signal-to-noise ratios higher than 15 dB, all identified modal shapes are noise stable. </span>
  
One ±4g 130 dB (at 1 Hz) triaxial accelerometer and three ±4g 2µg<sup>2</sup>/Hz uniaxial accelerometers were used. Four configurations were used for a total of 11 equidistant points along the deck axis. Vertical acceleration: sets of 100 samples/s, 5 minutes recording time. Response signals processed via Stochastic Subspace Identification (SSI) technique [33].
+
==REFERENCES==
  
The process with the mobile sensing setup (Figure <span id='cite-Figure_9'></span>[[#Figure_9|9]]). Firstly, the mobile platform frequency was individually identified (Figure <span id='cite-Figure_13'></span>[[#Figure_13|13]]). Then the mobile sensor set was put to travel along the bridge axis at <math display="inline">4.1328\, cm/s</math>. Both sensors, the mobile and the stationary, achieved a sampling rate of <math display="inline">100\, \frac{samples}{s}</math>. The white noise input was supplied from a shaker, and the recording time was <math display="inline">t=</math><math>290.36\, s</math>. With all Response  Signals available the fundamental frequencies were extracted from the peaks of each PSD graph ( Figure <span id='cite-Figure_14'></span>[[#Figure_14|14]]), and, finally, a Crossed PSD (CPSD) used to clear off process alien peaks ( Figure <span id='cite-Figure_15'></span>[[#Figure_15|15]]).
+
Auger F., Flandrin P., Goncalves P., and Lemoine O (1996), “Time-Frequency Toolbox for Use with MATLAB Reference Guide”, ''Centre National de la Recherche Scientifique''.
  
{| style="width: 100%;border-collapse: collapse;"
+
Auger F., Flandrin P., Goncalves P., and Lemoine O (1996). “Time-Frequency Toolbox for Use with MATLAB Tutorial” ''Centre National de la Recherche Scientifique''.
|-
+
|  style="text-align: center;vertical-align: top;width: 52%;"|[[Image:Draft_Marmolejo_177547576-image19.png|312px]]
+
  
(1) Time history - Mobile platform signal
+
Bendat, J.S., and Piersol, A.G. (2011), “Random Data: Analysis and Measurement Procedures”, ''John Wiley & Sons.''
|  style="text-align: center;vertical-align: top;width: 47%;"|[[Image:Draft_Marmolejo_177547576-image20.png|288px]]
+
  
(2) PSD – mobile platform signal
+
Bonato P., Ceravolo R., De Stefano A., and Knaflitz M. (1997), “Bilinear Time-Frequency Transformations in The Analysis of Damaged Structures”, ''Mech. Syst. Signal Process.'', vol. 11, no. 4, pp. 509–527.
  
 +
Bonato, P., Ceravolo, R., De Stefano, A., and Molinari, F. (1998). “Time-frequency and cross-time-frequency based techniques for the structural identification of systems”, ''Proceedings of the IEEE-SP International Symposium on Time-Frequency and Time-Scale Analysis'', 1998, pp. 445–448.
  
|-
+
Bonato, P., Ceravolo, R., De Stefano, A., and Molinari, F. (1999). “Adaptive kernel cross-time-frequency transformations for the identification of structural systems”, ''SPIE ''Vol. 3807 pp. 582–590.
|  colspan='2' style="text-align: center;vertical-align: top;"|<span id='Figure_13'></span>Figure 13. Mobile platform record
+
|}
+
  
 +
P. Bonato, R. Ceravolo, A. De Stefano, and F. Molinari (2000), “Use of Cross-Time–Frequency Estimators for Structural Identification in Non-Stationary Conditions and Under Unknown Excitation”, ''J. Sound Vib.'', vol. 237, no. 5, pp. 775–791.
  
Table <span id='cite-Table_2'></span>[[#Table_2|2]] summarizes the  identified frequencies from Figures <span id='cite-Figure_13'></span>[[#Figure_13|13]], <span id='cite-Figure_14'></span>[[#Figure_14|14]] and <span id='cite-Figure_15'></span>[[#Figure_15|15]]. Figure <span id='cite-Figure_16'></span>[[#Figure_16|16]], the mobile sensor signal marginals, its peaks show the fundamental frequencies, extracted via Eq. (10). Modal Identification performed via Eq. (6) and Eq. (7) and Damping Ratios obtained using [1]. Table <span id='cite-Table_3'></span>[[#Table_3|3]] is a comparative table between ''SSI'' results (stationary sensor dynamic identification) and time-scale based identification results (mobile sensor dynamic identification).
+
Bonato, P., Roy, S.H., Knaflitz, M., and Luca, C.J. de (2001), “Time-frequency parameters of the surface myoelectric signal for assessing muscle fatigue during cyclic dynamic contractions”. IEEE ''Trans. Biomed. Eng.'' 48, 745–753.
  
<span id='_Ref520370875'></span><span id='_Toc520379471'></span><div id="Table_2" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Bonato P., Ceravolo R., and De Stefano A. (1997). “Time-Frequency and Ambiguity Function Approaches in Structural Identification”.'' J. Eng. Mech.'' 123, 1260–1267.
Table 2. Summary – PSD-based frequency identification</div>
+
  
{| style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;"
+
Bonato P., Ceravolo R., De Stefano A., and Molinari F. (1998), “A New Cross-Time-Frequency Method for the Structural Identification of Mechanical Systems in Non-Stationary Conditions”, ''NATO Advanced Study Institute on Modal Analysis and Testing'' (Sesimbra, Portugal), pp. 725–741.
|-
+
|  style="border-top: 1pt solid black;text-align: center;"|Identified Element
+
|  style="border-top: 1pt solid black;text-align: center;"|Mobile Platform Hz.
+
  
(Figure <span id='cite-Figure_13'></span>[[#Figure_13|13]])
+
Brownjohn, J.M.W. (2007), “Structural health monitoring of civil infrastructure”. ''Philos. Trans. R. Soc. Math. Phys. Eng. Sci.'' 365, 589–622.
|  style="border-top: 1pt solid black;text-align: center;"|Mobile Sensor Hz.
+
  
(Figure <span id='cite-Figure_14'></span>[[#Figure_14|14]]b)
+
Ceravolo, R. (2009), “Time–Frequency Analysis”Encyclopedia of Structural Health Monitoring, ''John Wiley & Sons''.
|  style="border-top: 1pt solid black;text-align: center;"|Stationary Sensor Hz.
+
 
+
(Figure <span id='cite-Figure_14'></span>[[#Figure_14|14]]d)
+
|  style="border-top: 1pt solid black;text-align: center;"|Stationary- Mobile Sensors Hz.
+
 
+
(Figure <span id='cite-Figure_15'></span>[[#Figure_15|15]])
+
|-
+
|  style="vertical-align: top;"|Bridge
+
|  style="text-align: center;vertical-align: top;"|---
+
|  style="text-align: center;vertical-align: top;"|3.37
+
|  style="text-align: center;vertical-align: top;"|3.37
+
|  style="text-align: center;vertical-align: top;"|3.37
+
|-
+
|  style="vertical-align: top;"|Platform
+
|  style="text-align: center;vertical-align: top;"|9.77
+
|  style="text-align: center;vertical-align: top;"|9.77
+
|  style="text-align: center;vertical-align: top;"|---
+
|  style="text-align: center;vertical-align: top;"|---
+
|-
+
|  style="vertical-align: top;"|Bridge
+
|  style="text-align: center;vertical-align: top;"|---
+
|  style="text-align: center;vertical-align: top;"|12.01
+
|  style="text-align: center;vertical-align: top;"|12.01
+
|  style="text-align: center;vertical-align: top;"|11.96
+
|-
+
|  style="vertical-align: top;"|Platform
+
|  style="text-align: center;vertical-align: top;"|19.53
+
|  style="text-align: center;vertical-align: top;"|19.48
+
|  style="text-align: center;vertical-align: top;"|---
+
|  style="text-align: center;vertical-align: top;"|---
+
|-
+
|  style="vertical-align: top;"|Bridge
+
|  style="text-align: center;vertical-align: top;"|---
+
|  style="text-align: center;vertical-align: top;"|21.92
+
|  style="text-align: center;vertical-align: top;"|21.92
+
|  style="text-align: center;vertical-align: top;"|21.92
+
|-
+
|  style="border-bottom: 1pt solid black;vertical-align: top;"|Platform
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|23.44
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|23.34
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|---
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|---
+
|}
+
 
+
 
+
<span id='_Hlk519691862'></span>
+
 
+
{| style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;"
+
|-
+
|  style="text-align: center;vertical-align: top;width: 51%;"|[[Image:Draft_Marmolejo_177547576-image21.png|294px]]
+
 
+
(1) Time-history: mobile sensor signal
+
|  colspan='2'  style="text-align: center;vertical-align: top;width: 48%;"|[[Image:Draft_Marmolejo_177547576-image22.png|264px]]
+
 
+
(2) PSD – mobile sensor signal
+
|-
+
|  style="text-align: center;vertical-align: top;width: 51%;"|[[Image:Draft_Marmolejo_177547576-image23.png|294px]]
+
 
+
(3) Time-history: stationary sensor
+
|  colspan='2'  style="text-align: center;vertical-align: top;width: 48%;"|[[Image:Draft_Marmolejo_177547576-image24.png|270px]]
+
 
+
(4) PSD - stationary sensor signal
+
|-
+
|  colspan='3'  style="text-align: center;vertical-align: top;"|<span id='Figure_14'></span>Figure 14.  Time-history signals and spectral representations
+
 
+
 
+
|-
+
|  colspan='2'  style="text-align: center;vertical-align: top;width: 100%;"|[[Image:Draft_Marmolejo_177547576-image25.png|348px]]
+
|-
+
|  colspan='2'  style="text-align: center;vertical-align: top;"|<span id='Figure_15'></span>Figure 15.  PSD - stationary-mobile signals (CPSD)
+
|}
+
 
+
 
+
{| style="width: 71%;margin: 1em auto 0.1em auto;border-collapse: collapse;"
+
|-
+
|  style="text-align: center;vertical-align: top;width: 100%;"|[[Image:Draft_Marmolejo_177547576-image26.png|420px]]
+
 
+
<span id='Figure_16'></span>Figure 16.  Marginals mobile sensor signal
+
 
+
 
+
|}
+
 
+
 
+
Estimations of modal frequencies obtained from the marginal peaks (Fig. <span id='cite-Figure_16'></span>[[#Figure_16|16]]). Estimated modal frequencies via Wavelets were notoriously close to those found through SSI, showing a maximum error of 0.811%. Mode shapes were identified through equations (6) and (7). MAC values found notoriously close to 1.00 (Fig. <span id='cite-Figure_17'></span>[[#Figure_17|17]])
+
 
+
<span id='Table_4'></span><span id='_Toc520379473'></span><span id='Table_3'></span>Table 3: Summary of the Dynamic identification with Mobile Sensors and stationary sensors. f: frequency, ζ: damping ratio.
+
 
+
{| style="width: 100%;margin: 1em auto 0.1em auto;border-collapse: collapse;"
+
|-
+
| colspan='8'  style="border-top: 1pt solid black;text-align: center;vertical-align: top;"|DYNAMIC IDENTIFICATION (MOBILE SENSORS – STATIONARY SENSORS)
+
|-
+
|  rowspan='2' style="text-align: center;vertical-align: top;"|MODE
+
|  colspan='2'  style="text-align: center;vertical-align: top;"|'''SSI'''
+
|  colspan='2'  style="text-align: center;vertical-align: top;"|'''WAVELETS'''
+
|  rowspan='2' style="text-align: center;"|'''MAC'''
+
|  rowspan='2' style="text-align: center;vertical-align: top;"|'''Frequency'''
+
 
+
'''diff.'''
+
 
+
'''(%)'''
+
|  rowspan='2' style="text-align: center;vertical-align: top;"|'''Damping'''
+
 
+
'''diff.'''
+
 
+
'''(%)'''
+
|-
+
|  style="text-align: center;vertical-align: top;"|'''f (Hz)'''
+
|  style="text-align: center;vertical-align: top;"|'''ζ (%)'''
+
|  style="text-align: center;vertical-align: top;"|'''f (Hz)'''
+
|  style="text-align: center;vertical-align: top;"|'''ζ (%)'''
+
|-
+
|  style="text-align: center;vertical-align: top;"|1
+
|  style="text-align: center;vertical-align: top;"|3.330
+
|  style="text-align: center;vertical-align: top;"|0.720
+
|  style="text-align: center;vertical-align: top;"|3.357
+
|  style="text-align: center;vertical-align: top;"|0.840
+
|  style="text-align: center;vertical-align: top;"|0.999
+
|  style="text-align: center;vertical-align: top;"|0.811
+
|  style="text-align: center;vertical-align: top;"|16.667
+
|-
+
|  style="text-align: center;vertical-align: top;"|2
+
|  style="text-align: center;vertical-align: top;"|11.920
+
|  style="text-align: center;vertical-align: top;"|0.350
+
|  style="text-align: center;vertical-align: top;"|12.000
+
|  style="text-align: center;vertical-align: top;"|0.240
+
|  style="text-align: center;vertical-align: top;"|0.998
+
|  style="text-align: center;vertical-align: top;"|0.671
+
|  style="text-align: center;vertical-align: top;"|31.428
+
|-
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|3
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|21.740
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|0.900
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|21.870
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|0.910
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|0.988
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|0.598
+
|  style="border-bottom: 1pt solid black;text-align: center;vertical-align: top;"|1.111
+
|}
+
 
+
 
+
<div class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
[[Image:Draft_Marmolejo_177547576-image27.png|432px]] </div>
+
 
+
<span id='Figure_28'></span><div id="Figure_17" class="center" style="width: auto; margin-left: auto; margin-right: auto;">
+
Figure 17. Identified mode shapes, wavelets versus SSI</div>
+
 
+
<span id='_Ref520266438'></span><span id='_Toc520379412'></span>
+
 
+
==5. Conclusions ==
+
 
+
This paper shows the simplicity, exactness and effectiveness of using mobile sensors setups in determining structural dynamic properties, carried out under normal operation.
+
 
+
The technique proved its usefulness for high spatial resolution mode shape identification. Given the time and frequency variability nature of the signals, they were processed via a time-scale analysis, i.e. the Continuous Wavelet Transform (CWT). For the simulation case, five vertical closed mode shapes were identified from the mobile and stationary sensors data using the technique outlined in this document. When SNR is greater 15 dB, acceptable MAC values are obtained. The mobile sensor velocity influence was evaluated too. Acceptable results are obtained for velocities less than <math display="inline">5\, cm/s</math>.
+
 
+
An experimental test was carried out to validate the methodology with a pedestrian bridge. The results of the proposed methodology were compared with the mode shapes obtained from a modal identification procedure using stationary sensors. A mobile sensor was built to collect data in motion while a reference sensor remained stationary. The mobile sensor consists of a platform with a 1-Axis Accelerometer Module (MEMS type). The mobile sensor velocity was <math display="inline">4.13\, cm/s</math>. Eleven vertical modal coordinates were identified for the first three modes shapes using signals recorded by the mobile sensor and the stationary sensor. MAC values of 0.999, 0.998 and 0.987 were obtained between the modes identified using the SSI technique and the based Time - Scale technique. Four configurations with stationary sensors were required to identify eleven modal coordinates.
+
 
+
==ACKNOWLEDGEMENTS==
+
 
+
The authors wish to thank Universidad del Valle and Universidad del Quindío in Colombia, for their support on the research.
+
 
+
==REFERENCES==
+
  
[1] T.-P. Le, ‘Use of the Morlet mother wavelet in the frequency-scale domain decomposition technique for the modal identification of ambient vibration responses’, ''Mech. Syst. Signal Process.'', vol. 95, pp. 488–505, Oct. 2017.
+
Cerda F., Garret J. H., Bielak J., Rizzo P., and Barrera J. (2012). Indirect structural health monitoring in bridges: scale experiments. ''Proceedings of the Seventh International Conference on Bridge Maintenance, Safety and Management''. Lago di Como.
  
[2] Z. Sun, W. Hou, L.M. Sun, ‘Close-Mode Identification Based on Wavelet Scalogram Reassignment – Society for Experimental Mechanics’, presented at the 2006 IMAC-XXIV: Conference & Exposition on Structural Dynamics, St. Louis, Missouri USA, 2006.
+
Choi, H.-I., and Williams, W.J. (1989), “Improved time-frequency representation of multicomponent signals using exponential kernels”, ''IEEE Trans. Acoust. Speech Signal Process.'' 37, 862–871.
  
[3] A. Abdelgawad and K. Yelamarthi, ‘Structural health monitoring: Internet of things application’, in ''2016 IEEE 59th International Midwest Symposium on Circuits and Systems (MWSCAS)'', 2016, pp. 1–4.
+
Chopra, A.K. (2012). “Dynamics of structures: theory and applications to earthquake engineering”, ''Prentice Hall'', Boston.
  
[4] L. Zhang, T. Wang, and Y. Tamura, ‘A frequency–spatial domain decomposition (FSDD) method for operational modal analysis’, ''Mech. Syst. Signal Process.'', vol. 24, no. 5, pp. 1227–1239, Jul. 2010.
+
Cohen L. (1995), Time-frequency Analysis, Englewood Cliffs, N.J:'' Prentice Hall'' PTR.
  
[5] J. Marulanda, J. M. Caicedo, and P. Thomson, ‘Modal Identification Using Mobile Sensors under Ambient Excitation’, ''J. Comput. Civ. Eng.'', p. 04016051, Aug. 2016.
+
Dantu, K., Rahimi, M., Shah, H., Babel, S., Dhariwal, A., and Sukhatme, G.S. (2005). Robomote: enabling mobility in sensor networks. In IPSN 2005.'' Fourth International Symposium on Information Processing in Sensor Networks'', pp. 404–409.
  
[6] J. Marulanda,  Modal Identification Using Smart Mobile Sensing Units. Cali, Colombia: Programa Editorial Universidad del Valle, 2014.
+
Flandrin, P. (1984). “Some features of time-frequency representations of multicomponent signals”. ''In Acoustics, Speech, and Signal Processing, IEEE International Conference on ICASSP ’84''., pp. 266–269.
  
[7] T. J. Matarazzo, M. Horner, and S. N. Pakzad, ‘High-Resolution Mode Shape Identification Using Mobile Sensors’, in ''Topics in Modal Analysis & Testing, Volume 10'', Springer, Cham, 2016, pp. 255–260.
+
Gabor, D. (1946). “Theory of communication”  ''Institution of Electrical Engineering'', London.
  
[8] B. Valeti, T. J. Matarazzo, and S. N. Pakzad, ‘Experimental Study on Wireless Mobile Sensor Configurations for Output-Only Modal Identification of a Beam Testbed’, in ''Sensors and Instrumentation, Volume 5'', Springer, Cham, 2017, pp. 71–77.
+
Hammond J. K.  and White P. R. , “The Analysis of Non-Stationary Signals Using Time-Frequency Methods”, ''J. Sound Vib''., vol. 190, no. 3, pp. 419–447, Feb. 1996.
  
[9] D. Zhu, X. Yi, Y. Wang, K.-M. Lee, and J. Guo, ‘A mobile sensing system for structural health monitoring: design and validation’, ''Smart Mater. Struct.'', vol. 19, no. 5, p. 055011, Mar. 2010.
+
Hlawatsch, F., and Auger, F. (2013). “Time-Frequency Analysis”,'' John Wiley & Sons''.
  
[10] D. Zhu, J. Guo, C. Cho, Y. Wang, and K.-M. Lee, ‘Wireless Mobile Sensor Network for the System Identification of a Space Frame Bridge’, ''IEEEASME Trans. Mechatron.'', vol. 17, no. 3, pp. 499–507, Jun. 2012.
+
Hlawatsch, F., and Boudreaux-Bartels, G.F. (1992). “Linear and quadratic time-frequency signal representations”. ''IEEE Signal Process. Mag''. 9, 21–67.
  
[11] G. H. I. James, T. G. Carne, and J. P. Lauffer, ‘The Natural Excitation Technique (next) for Modal Parameter Extraction from Operating Wind Turbines’, Sandia National Labs., Albuquerque, NM (United States), SAND--92-1666, Feb. 1993.
+
Arango H. (2009), “Análisis de señales con las transformadas de Fourier, Gabor y Onditas”. ''Instituto Tecnológico Metropolitano'', Medellín, Colombia, 2009.
  
[12] Y.-B. Yang, C. W. Lin, and J. D. Yau, ‘Extracting bridge frequencies from the dynamic response of a passing vehicle’, ''J. Sound Vib.'', vol. 272, no. 3, pp. 471–493, May 2004.
+
Karbhari, V.M., Guan, H., and Sikorsky, C. (2009).  “Operational modal analysis for vibration-based structural health monitoring of civil structures”. ''Structural Health Monitoring of Civil Infrastructure Systems,'' (Woodhead Publishing), pp. 213–259.
  
[13] C. W. Lin and Y. B. Yang, ‘Use of a passing vehicle to scan the fundamental bridge frequencies: An experimental verification’, ''Eng. Struct.'', vol. 27, no. 13, pp. 1865–1878, Nov. 2005.
+
Lin, C.W., and Yang, Y.B. (2005). “Use of a passing vehicle to scan the fundamental bridge frequencies: An experimental verification”. ''Eng. Struct''. 27, 1865–1878.
  
[14] Y. B. Yang and K. C. Chang, ‘Extracting the bridge frequencies indirectly from a passing vehicle: Parametric study’, ''Eng. Struct.'', vol. 31, no. 10, pp. 2448–2459, Oct. 2009.
+
Lippmann, R.P. (1989). “Pattern classification using neural networks”. ''IEEE Commun. Mag''. 27, 47–50.
  
[15] D. M. Siringoringo and Y. Fujino, ‘Estimating Bridge Fundamental Frequency from Vibration Response of Instrumented Passing Vehicle: Analytical and Experimental Study’, ''Adv. Struct. Eng.'', vol. 15, no. 3, pp. 417–433, Mar. 2012.
+
Loughlin, P.J., Pitton, J.W., and Atlas, L.E. (1993). “Bilinear time-frequency representations: new insights and properties”. ''IEEE Trans. Signal Process''. 41, 750–767.
  
[16] A. González, E. J. OBrien, and P. J. McGetrick, ‘Identification of damping in a bridge using a moving instrumented vehicle’, ''J. Sound Vib.'', vol. 331, no. 18, pp. 4115–4131, Aug. 2012.
+
Lynch, J.P. (2006). “A Summary Review of Wireless Sensors and Sensor Networks for Structural Health Monitoring”. ''Shock Vib. Dig''. 38, 91–128.
  
[17] J. Li, X. Zhu, S. Law, and B. Samali, ‘Indirect bridge modal parameters identification with one stationary and one moving sensors and stochastic subspace identification’, ''J. Sound Vib.'', vol. 446, pp. 1–21, Apr. 2019.
+
Magalhães, F., Cunha, Á., and Caetano, E. (2008). “Dynamic monitoring of a long span arch bridge”. ''Eng. Struct.'' 30, 3034–3044.
  
[18] T. J. Matarazzo and S. N. Pakzad, ‘Structural Identification for Mobile Sensing with Missing Observations’, ''J. Eng. Mech.'', vol. 142, no. 5, May 2016.
+
Magalhães, F., Cunha, Á., and Caetano, E. (2009). “Online automatic identification of the modal parameters of a long span arch bridge”. ''Mech. Syst. Signal Process.'' 23, 316–329.
  
[19] W. J. Staszewski, ‘Identification of Damping in Mdof Systems Using Time-Scale Decomposition’, ''J. Sound Vib.'', vol. 203, no. 2, pp. 283–305, Jun. 1997.
+
Marulanda, J., Caicedo, J.M., and Thomson, P. (2016). “Modal Identification Using Mobile Sensors under Ambient Excitation”. J. Comput. Civ. Eng. 04016051.
  
[20] M. Ruzzene, A. Fasana, L. Garibaldi, and B. Piombo, ‘Natural Frequencies and Dampings Identification Using Wavelet Transform: Application to Real Data’, ''Mech. Syst. Signal Process.'', vol. 11, no. 2, pp. 207–218, Mar. 1997.
+
Marulanda J. (2014), “Modal Identification Using Smart Mobile Sensing Units”. ''Programa Editorial Universidad del Valle. ''Cali, Colombia.
  
[21] T. Kijewski and A. Kareem, ‘Wavelet Transforms for System Identification in Civil Engineering’, ''Comput.-Aided Civ. Infrastruct. Eng.'', vol. 18, no. 5, pp. 339–355, Sep. 2003.
+
Matarazzo, T.J., and Pakzad, S.N. (2014). “Modal Identification of Golden Gate Bridge Using Pseudo Mobile Sensing Data with STRIDE”. ''Dynamics of Civil Structures'', Volume 4, F.N. Catbas, ed. (Springer International Publishing), pp. 293–298.
  
[22] A. Chakraborty and Biswajit Basu, ‘Nonstationary Response Analysis of Long Span Bridges under Spatially Varying Differential Support Motions Using Continuous Wavelet Transform’, ''Journal of Engineering Mechanics'', vol. 134, no. 2, Feb-2008.
+
Matarazzo, T.J., and Pakzad, S.N. (2016). “STRIDE for Structural Identification Using Expectation Maximization: Iterative Output-Only Method for Modal Identification”. ''J. Eng. Mech''. 142, 04015109.
  
[23] W.-J. Yan and W.-X. Ren, ‘Use of Continuous-Wavelet Transmissibility for Structural Operational Modal Analysis’, ''J. Struct. Eng.'', vol. 139, no. 9, pp. 1444–1456, Sep. 2013.
+
Matarazzo T. J., and  Pakzad S. N. (2013). “Mobile Sensors in Bridge Health Monitoring”. ''International Workshop on Structural Health Monitoring'', Stanford, CA.
  
[24] T.-P. Le and P. Paultre, ‘Modal identification based on the time–frequency domain decomposition of unknown-input dynamic tests’, ''Int. J. Mech. Sci.'', vol. 71, pp. 41–50, Jun. 2013.
+
Oppenheim, A.V., and Schafer, R. W (2010). “Discrete-time signal processing”, Upper Saddle River, ''Pearson''.
  
[25] X. Pan, F. Shang, and S. Yuan, ‘Wavelet-based electromechanical mode shape online identification from ambient data’, in ''2011 4th International Conference on Electric Utility Deregulation and Restructuring and Power Technologies (DRPT)'', 2011, pp. 1841–1847.
+
Papandreou, A., and Boudreaux-Bertels, G.F. (1993). “Generalization of the Choi-Williams Distribution and the Butterworth Distribution for Time-Frequency Analysis”. I''EEE Trans. Signal Process. ''41, 463-.
  
[26] X. Pan and V. Venkatasubramanian, ‘Multi-dimensional wavelet analysis for power system oscillation monitoring using synchrophasors’, in ''2012 IEEE PES Innovative Smart Grid Technologies (ISGT)'', 2012, pp. 1–10.
+
Pastor, M., Binda, M., and Harčarik, T. (2012). “Modal Assurance Criterion”. ''Procedia Eng.'' 48, 543–548.
  
[27] M. Bronzini, S. Bruno, M. D. Benedictis, and M. L. Scala, ‘Power system modal identification via wavelet analysis’, in ''2007 IEEE Lausanne Power Tech'', 2007, pp. 2041–2046.
+
Rioul, O., and Vetterli, M. (1991). “Wavelets and signal processing”. ''IEEE Signal Process. Mag.'' 8, 14–38.
  
[28] W. J. Staszewski and D. M. Wallace, ‘Wavelet-based Frequency Response Function for time-variant systems—An exploratory study’, ''Mech. Syst. Signal Process.'', vol. 47, no. 1–2, pp. 35–49, Aug. 2014.
+
Sibley, G.T., Rahimi, M.H., and Sukhatme, G.S. (2002). “Robomote: a tiny mobile robot platform for large-scale ad-hoc sensor networks”.  ''IEEE International Conference on Robotics and Automation, 2002. Proceedings. ICRA ’02'', pp. 1143–1148.
  
[29] Auger F., Flandrin P., Goncalves P., and Lemoine O., ''Time-Frequency Toolbox For Use with MATLAB Tutorial''. Centre National de la Recherche Scientifique, 1996.
+
Spencer, B.F., Ruiz-Sandoval, M.E., and Kurata, N. (2004). “Smart sensing technology: opportunities and challenges”. ''Struct. Control Health Monit.'' 11, 349–368.
  
[30] O. Rioul and P. Flandrin, ‘Time-scale energy distributions: a general class extending wavelet transforms’, ''IEEE Trans. Signal Process.'', vol. 40, no. 7, pp. 1746–1757, Jul. 1992.
+
Staszewski, W.J., and Tomlinson, G.R. (1994). “Application of the wavelet transform to fault detection in a spur gear”. ''Mech. Syst. Signal Process''. 8, 289–307.
  
[31] M. Pastor, M. Binda, and T. Harčarik, ‘Modal Assurance Criterion’, ''Procedia Eng.'', vol. 48, pp. 543–548, Jan. 2012.
+
Wang W. J. and  McFadden P. D. (1993), “Early detection of gear failure by vibration analysis I. Calculation of the time-frequency distribution”, ''Mech. Syst. Signal Process.,'' vol. 7, no. 3, pp. 193–203.
  
[32] Auger F., Flandrin P., Goncalves P., and Lemoine O., ''Time-Frequency Toolbox For Use with MATLAB Reference Guide''. Centre National de la Recherche Scientifique, 1996.
+
Zhu, D., Yi, X., Wang, Y., Guo, J., and Lee, K.-M. (2010). “Mobile Sensor Networks: A New Approach for Structural Health Monitoring”. ''American Society of Civil Engineers'', pp. 159–168.
  
[33] R. Brincker and P. Andersen, ‘Understanding stochastic subspace identification’, ''Proc. 24th IMAC St Louis Mo.'', 2006.
+
Zhu, D., Guo, J., Cho, C., Wang, Y., and Lee, K.-M. (2012). “Wireless Mobile Sensor Network for the System Identification of a Space Frame Bridge”.'' IEEE ASME Trans. Mechatron.'' 17, 499–507.

Latest revision as of 21:29, 20 August 2019


Abstract: Dynamic characterization of structures from field measurements is useful for different purposes (e.g. retrofit validation, model updating, structural health monitoring, etc.). The identification of high spatial density mode shapes has been recently a challenge tackled using mobile sensors. These sensors travel over the structure and continuously acquire vibration data that is used to identify modal coordinates with a higher spatial density than can generally be obtained using a limited number of stationary sensors. The recorded signal from a mobile sensor is non-stationary, thus, it has significant variations in its spectral content over time, requiring a suitable processing to extract the properties not only for the time but also for the frequency domain. In this paper, Cohen’s class Time-Frequency Distributions (TFD) are proposed for the output-only dynamics identification of structures based on non-stationary signals recorded with mobile sensors. Identification is achieved through cross-time-frequency estimators using Smoothed Pseudo-Wigner-Ville (SPWVD) distribution. Results from numerical simulations using a simply supported beam subject to ambient vibration are shown and the sensibility of the proposed identification to the presence of measurement noise is evaluated. Numerical results show that use of the cross-time-frequency estimators is effective in extracting modal properties of the structures and filtering noise.

Keywords: Modal Identification, Mobile Sensors, Non-Stationary Signals, Time-Frequency Distributions

1. Introduction

Structural Health Monitoring (SHM) Systems usually play a crucial duty in early failure or damage detection for construction, maintenance or catastrophe management. Civil infrastructure and many other sectors can greatly benefit from systems that monitor behaviour under normal operating conditions, without the need of interrupting regular day-to-day activity (Brownjohn, 2007; Karbhari et al., 2009).

The future trend in instrumentation for SHM is the use of smart mobile sensing along with wireless intelligent networks, as they not only provide similar information to conventional cabled systems, but can also be installed at a much lower cost and process data autonomously using their embedded microprocessors and software (Lynch, 2006; Spencer et al., 2004). Modal shapes are traditionally identified with stationary sensors fixed at locations that provide representative structural responses, however, these schemes contain restricted spatial information (Figure 1). If more spatial and temporal data is obtained from a vibrating structure, the structure’s response can be evaluated more accurately, and hence the advantage of implementing the structure with as many stationary sensors as possible during the acquisition of data. However, despite the improved accuracy in estimations, these or dense sensor array approaches can be impractical for many reasons including the cost of sensors and setup time, memory constrains, network reliability, power requirements, and physical limitations due to structure geometries (Magalhães et al., 2008, 2009; Matarazzo T. J. and Pakzad, S. N. 2013). Mobile sensing is a novel alternative in Structural Health Monitoring (SHM). The objective is to solve failings from fixed sensor array setups. Few sensors are usually used to collect dense spatial information in mobile sensing setups. Mobile sensing provides several advantages over schemes using stationary sensors, but the main advantage is that a single mobile sensor can be used to record signals continuously along the structure. Modal coordinates are identified with a higher spatial density than using stationary sensors because the mobile sensor continuously records vibration data as it travels along trajectories on the structure (Figure 2). Mobile sensing is not only easier to implement, but also more cost effective when compared to dense stationary sensor arrays (Marulanda et al., 2016; Marulanda, J., 2014; Matarazzo T. J. and Pakzad S. N., 2013).

Marmolejo et al 2019b-image2.png
Figure 1. Spatial resolution of mode shapes: stationary sensors
Marmolejo et al 2019b-image3.png
Figure 2. Spatial resolution of mode shapes: mobile sensors

Recent implementation of mobile sensor networks has been diverse, although limited, and is still under development. Zhu et al. (2010, 2012) proposed a sensing device. Such device will record data at selected nodes, but will not do so while the sensor is moving. Sibley et al. (2002) and Dantu et al. (2005) implemented Robomote, miniature low cost robots for mobile sensors networks. Partial modal identification include the use of a moving vehicle to identify frequencies of a single span bridge using frequency domain techniques (Cerda F. et al., 2012; Lin and Yang, 2005). Marulanda et al. (2016) and Marulanda (2014) developed a novel technique that uses two sensors, one mobile and the other stationary, in which the mobile sensor records continuously, and, in a subsequent procedure, performs the modal identification through spectrograms. Matarazzo et al. (2014, 2016) proposed a method of collecting data through continuous mobile detection in the presence of missing time and space observations. As the corresponding matrices have incompatible data records, they developed and successfully implemented an algorithm, called STRIDE, to cover for such “empty spaces”.

The use of mobile sensors, typically accelerometers, require an analysis of non-stationary response signals. Time-Frequency analysis has been proposed by several authors to analyse non-stationary responses in structural identification and in the assessment of mechanical damage. Staszewski et al. (1994) and Rioul et al. (1991) have devoted several papers to modal identification using the Short-Time Fourier Transform (STFT) and Wavelet Transform (WT). The STFT transforms the signal in a two-dimensional time-frequency plane, assuming that the signal is stationary when viewed through a limited extension window. The Fourier Transform of the windowed signal leads to a time-frequency distribution (Gabor, 1946; Arango, 2009; Oppenheim, 2010). The STFT is linear and therefore provides poor energy information about the signal. Quadratic Time-Frequency Distributions (TFDs), on the other hand, allow the interpretation of representations from a much richer energy point of view (Hlawatsch and Boudreaux-Bartels, 1992; Loughlin et al., 1993). Marginal properties express such interpretation. However, in view of the Uncertainty Principle, such properties are not sufficient to identify an energy density at each point on the time-frequency plane. (Ceravolo, 2009). A spectrogram, or quadratic representation of the amplitude of the STFT, is a three dimensional representation of the time-averaged Fourier Transform over adjacent time segments of a random process (Bendat and Piersol, 2011). It does not fulfil marginal properties and violates the Linearity Principle, generating crossed or interference terms. These spurious terms are restricted to those regions of the time-frequency domain where the auto-terms overlap. In the specific case of the spectrogram, for two sufficiently separated components in the time-frequency plane, its interference terms are practically nil (Ceravolo, 2009). Spectrograms can be used to obtain modal coordinates from the signals recorded from a mobile and a stationary sensor. These signals are divided into quasi-stationary segments and then the auto-spectral density and cross-spectral density functions for segments of both records are calculated. Finally a modal identification procedure is performed (Marulanda et al., 2016; Marulanda, J, 2014).

In this paper, the non-stationary signals analysis recorded from sensors is performed by using Cohen’s class TFDs. The adoption of a quadratic representation helps in overcoming the limitations from time-frequency resolution, as these transformations, the energetic and the correlative, are not based on the segmentation of signals. (Cohen, 1995). Invariance to time and frequency shifts characterize the Cohen’s Class Distributions (Hlawatsch and Auger, 2013). The system response is recorded in the time-frequency plane as the evolution of the spectral components corresponding to the energy of individual vibration modes whenever the Cohen’s Class Distributions is in use (Bonato et al., 1998, 1999; Bonato et al., 2000; Bonato P. et al., 1997; Cohen, 1995; Hammond and White, 1996; Hlawatsch and Boudreaux-Bartels, 1992). The proposed technique assumes that the frequency range of the system input spans the vibration modes. As the input spectrum increases, so does the identification accuracy. The instantaneous amplitude ratios directly determine the modal amplitude ratios for the time-frequency representations of signals recorded from a mobile sensor and a stationary sensor. Additionally, the phase of the cross-time- frequency representation between the two signals estimate the phase relationships. The estimators retain their dependence on time since they are derived directly from the two-dimensional functions of frequency and time variables. modal responses in linear time invariant systems show a constant relationship between the amplitude and the phase, and, therefore, the identified mode shapes are characterized by their stability over time (Bonato et al., 1998).

2. Theoretical Background

Cohen’s Class Distributions

Cohen’s class distributions display a clear invariance of their representations for time and frequency shifts, a desirable property when correlating signal characteristics to phenomena occurring in the physical system generating the signal. The following equation describes all the Cohen’s class distributions (Hlawatsch and Auger, 2013):

(1)


where is the signal, is its complex conjugate, is time, is frequency and is the time delay, is the circular frequency; is the kernel of the distribution and is the Cohen Class TFD. The cross-time-frequency distribution between two signals and can be defined as follows:


(2)


The Wigner–Ville distribution (WVD) is the basic TFD and is a quadratic form that measures a local time-frequency energy. This distribution is obtained from equation (1) using . Although the WVD satisfies a several important mathematical properties (Bonato et al., 2001; Flandrin, 1984; Wang and McFadden, 1993), it is not very suitable for applications with multiple components signals since the bi-linearity of the transform causes the appearance of interference terms . These are spurious terms when using the time-frequency characteristics of the signal for modal identification. The smoothed pseudo-Wigner-Ville transform (SPWVD) results from , where and are the standard deviations in time and frequency, respectively. With the exception of real value, time-shift-invariance, and frequency invariance, the application of the smoothed function causes the loss of most of the mathematical properties of the WVD, (Hlawatsch and Boudreaux-Bartels, 1992). The foregoing independence between windows makes the SPWVD very versatile for reducing cross—terms in the WVD. This property is useful for modal identification.The Choi-Williams Distribution (CWD) (Choi and Williams, 1989; Papandreou and Boudreaux-Bertels, 1993) is defined by . This kernel is defined in the ambiguity plane . If is very large, the kernel vanishes, therefore leading to the WVD, otherwise, the kernel application will have a filtering effect. For small values ​​of , the effect of the multiplication is to preserve the ambiguity function close the origin of the plane , therefore, this kernel satisfies the minimization property (Cohen, 1995). In multicomponent signals the authentic terms are usually close to the ambiguity plane axis, and the interference terms are dispersed away. The resulting characteristic function reduces the interference terms without significantly affecting the signal components (Bonato et al., 1997). Therefore, although the CWD violates the time and frequency support properties, it can be shown that the introduced error is generally negligible (Lippmann, 1989; Papandreou and Boudreaux-Bertels, 1993), so the value should be adapted to each individual case, since the location of the interference terms and the signal components change with respect to the characteristics of the signal (Bonato et al., 1997). In this paper, SPWVD distribution is used. In this case, by parameterizing the windows, as they attenuate the interference terms without significantly modifying the signals, good modal identification results are obtained.

Modal Identification Using Mobile Sensors

The proposed methodology requires a mobile sensor that acquires data continuously along the structure to be identified and simultaneous acquisition by a stationary sensor. Under ambient excitation using a zero mean Gaussian white noise process, the outputs of the system are also zero mean Gaussian processes, exciting the fundamental frequencies (Bendat and Piersol, 2011). Since the recorded signal from the mobile sensor is non-stationary, a procedure based on the Cohen Class Distributions (i.e. SPWVD) is proposed. These distributions are invariable to time and frequency shifts, which is required for proper system identification.

Marmolejo et al 2019b-image4.png
Figure 3. Schematic description of the proposed setup on a simply supported beam.
Marmolejo et al 2019b-image5.png
Figure 4. Equivalent model of the proposed setup showed in Fig. 3.

Assuming that the structure to be identified is the one-dimensional beam shown in Figure 3 has discrete degrees of freedom (d.o.f.), wich each d.o.f. corresponds to a segment of the signal from the mobile sensor (Figure 4).

Let be the response (displacement) in the th position of the system, the response (displacement) associated with the th vibration mode, and a term of the normalized eigenvectors matrix, then can be in decouple form as,

(3)


The term of the sum in equation (3) extends to the vibration modes that are detected in the th position corresponding to a discretized d.o.f. The structure under ambient excitation is equipped with a mobile sensor with constant velocity that continuously records the acceleration and, simultaneously, with a stationary sensor also recording the acceleration vibrations. For each d.o.f., the individual modal components appear in the form of an energy peak in the time-frequency plane that can be written in complex form as (Bonato et al., 2000):

(4)


where is a baseband signal related to the mode; and are the amplitude and phase modulation waveforms respectively; is the natural frequency; is the amplitude of the mode shape in the position. The complex form of equation (4) represents an analytical signal and, therefore, the spectrum of is single-sided. The WVD will transform the signal into:

)
(5)


where is the WVD of the signal , is the convolution in frequency, and ) is the Dirac delta (impulse) located in . In multi-component signals, when auto-terms and interference overlap, using a kernel capable of attenuating interference terms leads to an increased reliability of the estimated amplitude ratio. The use of highly selective kernels, ambiguity domain functions (Bonato et al., 2000), improves the capacity of the cross-time-frequency-transform to separate signal components, and therefore, allow amplitude ratio and phase difference estimators become close to each modal shape frequency.

The following expression represents the Cohen’s Class Transforms:

(6)


where is a Cohen class TFD, is the kernel of the TFD in time-frequency plane. is the double convolution in time and frequency. From equation (6) it is evident that the energy of the transform concentrates close to the modal frequency . The distribution shape is determined by the time-frequency transform of the modulating waveform . Because the shape of the modulated waveform is kept in the time–frequency domain. The amplitude ratio between two modal components can be determined from auto-time-frequency representations, as follows (Bonato et al., 1998; Bonato et al., 2000):

(7)


And the phase can be deteremined from cross-time-frequency representtions, as follows:

(8)


The TFDs of the signals from both sensors, mobile and stationary, were used in separating modal and residual components and in estimating the modal Amplitudes Ratios (AR); while cross-time-frequency representations were used in estimating the Phase Ratios (PR) (Bonato et al., 1998). The system is displayed in the time-frequency plane as the unfolding of spectral components corresponding to the energy of the predominant vibration modes. The ARs, gathered from their time-frequency representations, are determined from the ratio between the instantaneous amplitudes of the signal segments. The PRs are the ratio between the imaginary and the real part of the cross- time-frequency representation of two signal segments (Bonato et al., 2000). Once the cross terms have been filtered, it is possible to identify a potential error source in the vibration modes due to a close coupling between the components. In contrast to classical frequency analysis, the cross time-frequency transform between channels can filter the products among various components. Although being closely coupled in frequency, they are uncorrelated in time (Bonato P. et al., 1997).

In the frequency ranges in which a single modal component is predominant, the estimators tend to be a constant value over time. Because the variance ( ) is a measure of the density of the time-frequency representation, the peaks of the variance of the cross-time-frequency transform are concentrated close to fundamental frequencies and, therefore, such frequencies are located in the maximum of the following function:

(9)


where is the length of the signal, and indicates the mean value of the cross-time-frequency distribution.

3. Numerical Simulations

A proposed example to evaluate the described procedure using a numerical model of a simply supported beam (Figure 3). The beam’s th natural vibration mode shape and corresponding natural frequency are given by the following equation:

 ;
(10)


where is the total length of the beam, is the elasticity modulus of the material, is the moment of inertia of the cross section, and is the constant mass per unit length (Chopra, 2012). In the following numerical simulations. In this simulation is assumed that , , , , and ; and, therefore, the first three natural frequencies of the beam are , , and . The stationary sensor was located at from the left support and the mobile sensor had a velocity of . The excitation of the beam corresponds to a Gaussian white noise acceleration at the supports. The simulated response signals were sampled at , and then conditioned using a Hilbert transform and sub-sampled at . SPWVD was used for the modal identification. The computed SPWVD was performed with a Hanning time smoothing window of samples, Hanning frequency smoothing window of samples and 256 frequency bins. The identified mode shapes are compared to the exact mode shapes using the Modal Assurance Criterion (MAC) via equation (11) (Pastor et al., 2012). The damping is identified using the Half-Power Bandwidth method (Bonato P. et al., 1997; Chopra, 2012). The influence of noise in the identification is evaluated. The software used for signals processing using TFDS was The Time-Frequency Toolbox (TFTB) (Auger F. et al., 1996).

(11)


TFD-based mode shapes identification

Figure 5 shows the time history of the signal recorded with the mobile sensor traveling at 0.0167 m/s, along with the spectral density and auto-SPWVD. Figure 6 shows the time history of the signal recorded with the stationary sensor, along with its spectral density and auto-SPWVD. Figure 7 shows the time history of the signals recorded with the mobile and the stationary sensors along with their spectral density and cross-SPWVD. Figure 8 shows the standard deviation of the cross-SPWVD. The modal frequencies and damping ratios were obtained via Equation (9) by selecting peaks and using the Half-Power Bandwidth method (Figure 8). The mode shapes were obtained from TFDs (i.e. SPWVD), as given by Equations (7) and (8). Table 1 summarizes results of the modal parameter identification, where it is shown that errors in natural frequencies are generally less than 1.5 %. The identified mode shapes are also compared to the exact mode shapes. In this example, 59 modal coordinates were identified, and MAC values were all equal to one for the first three modes (Figure 9). MAC values for different mobile sensor velocity values, ranging from to , were calculated. Figure 10 shows the variations in the identified modal coordinates (IMC) per unity of beam’s length versus the mobile sensor’s velocity, it is assumed that two modes are considered correlated when the MAC value is equal to 0.95, which corresponds to an angle of 18 degrees (Pastor et al., 2012). From Figure 11, it can be observed that good identification results have been obtained from the simulated ambient response data contaminated with different intensity levels of white gaussian noise.

Table 1. Summary of results obtained using equations (7), (8), (9) and (11)
MODE Theoretical frequencies

(Hz)

Identification by TFD (SPWVD) Frequency Error

(%)

Natural

frequencies (Hz)

Damping

ratios

MAC

values

1 1.450 1.465 0.060 1.00 1.03
2 5.798 5.801 0.029 1.00 0.05
3 13.046 13.008 0.025 1.00 0.29


Marmolejo et al 2019b-image6.jpeg
Figure 5. Signal recorded with the mobile sensor: time history, spectral density and SPWVD
Marmolejo et al 2019b-image7.jpeg
Figure 6. Signal recorded with the stationary sensor: time history, spectral density and SPWVD
Marmolejo et al 2019b-image8.jpeg
Figure 7. Signals recorded with the stationary and the mobile sensors: time history, spectral density and SPWV cross-distribution.
Marmolejo et al 2019b-image9.jpeg
Figure 8. Standard deviation from the SPWVD cross-distribution.
Marmolejo et al 2019b-image10.jpeg
Figure 9. First three mode shapes identified using equations (7) and (8): 59 modal coordinates were identified.
Marmolejo et al 2019b-image11.jpeg
Figure 10. IMC/m values versus mobile sensor velocity.

Signal-to-Noise Ratio

In most experimental applications, the acceleration record from the mobile sensor would be more contaminated with noise than the stationary sensor due to the motion system. To evaluate the effect of such noise on the identified modes through the MAC values, artificial normally distributed white noise was added to the acceleration response of the mobile to generate signal-to-noise ratios from -5 to 20 dB. Figure 11 shows the MAC values versus SNR for the first three identified modes. It can be seen from Figure 11 that the identification accuracy of all mode shapes can be maintained at a high level using the proposed method for SNR values higher than 15 dB.

Marmolejo et al 2019b-image12.jpeg
Figure 11. MAC values of the first three identified modes for variation in the signal-to-noise ratio.

4. CONCLUSIONS

This paper presents the use of mobile sensors for modal identification through Cohen’s Class TFDs, providing identified modal shapes with high spatial resolution. Mobile sensors can be used in structures and mechanical systems under normal operating conditions and unknown excitation. It is envisaged that mobile sensors not only offer flexible but adaptable spatial resolutions that may enhance the theory and practice of future SHM and damage detection research and technological development.

The mobile sensor continuously records vibration data as it travels along trajectories on the structure, therefore, modal coordinates are identified with a higher spatial density than using stationary sensors. The quantity of identified modal coordinates depends on the velocity of the sensor, the selected kernels, and the identified frequencies.

The TFD-based modal identification utilized to identify the structural system in terms of natural frequencies, damping ratios and mode shapes, along with the use of mobile sensors complied successfully. The mode shapes are estimated through the auto and cross time-frequency transforms of the signals from mobile and stationary sensors (i.e. modulating waveforms). The proposed methodology is highly accurate when calculating mode shapes by determining the amplitude and phase ratios between TFDs of signals recorded with the mobile and the stationary sensors, respectively. The standard deviations of cross time-frequency transforms of the recorded signals are related to modal frequencies and damping ratios.

The effectiveness of the technique is clear, based on auto and cross-time-frequency estimators for structural identification when non-stationary signals are used. For velocities between to , the variations in the identified modal coordinates per unity of beam’s length ranges from to . For a velocity of , 59 modal coordinates were identified, and MAC values were all equal to one for the first three modes using only two acceleration records. For signal-to-noise ratios higher than 15 dB, all identified modal shapes are noise stable.

REFERENCES

Auger F., Flandrin P., Goncalves P., and Lemoine O (1996), “Time-Frequency Toolbox for Use with MATLAB Reference Guide”, Centre National de la Recherche Scientifique.

Auger F., Flandrin P., Goncalves P., and Lemoine O (1996). “Time-Frequency Toolbox for Use with MATLAB Tutorial” Centre National de la Recherche Scientifique.

Bendat, J.S., and Piersol, A.G. (2011), “Random Data: Analysis and Measurement Procedures”, John Wiley & Sons.

Bonato P., Ceravolo R., De Stefano A., and Knaflitz M. (1997), “Bilinear Time-Frequency Transformations in The Analysis of Damaged Structures”, Mech. Syst. Signal Process., vol. 11, no. 4, pp. 509–527.

Bonato, P., Ceravolo, R., De Stefano, A., and Molinari, F. (1998). “Time-frequency and cross-time-frequency based techniques for the structural identification of systems”, Proceedings of the IEEE-SP International Symposium on Time-Frequency and Time-Scale Analysis, 1998, pp. 445–448.

Bonato, P., Ceravolo, R., De Stefano, A., and Molinari, F. (1999). “Adaptive kernel cross-time-frequency transformations for the identification of structural systems”, SPIE Vol. 3807 pp. 582–590.

P. Bonato, R. Ceravolo, A. De Stefano, and F. Molinari (2000), “Use of Cross-Time–Frequency Estimators for Structural Identification in Non-Stationary Conditions and Under Unknown Excitation”, J. Sound Vib., vol. 237, no. 5, pp. 775–791.

Bonato, P., Roy, S.H., Knaflitz, M., and Luca, C.J. de (2001), “Time-frequency parameters of the surface myoelectric signal for assessing muscle fatigue during cyclic dynamic contractions”. IEEE Trans. Biomed. Eng. 48, 745–753.

Bonato P., Ceravolo R., and De Stefano A. (1997). “Time-Frequency and Ambiguity Function Approaches in Structural Identification”. J. Eng. Mech. 123, 1260–1267.

Bonato P., Ceravolo R., De Stefano A., and Molinari F. (1998), “A New Cross-Time-Frequency Method for the Structural Identification of Mechanical Systems in Non-Stationary Conditions”, NATO Advanced Study Institute on Modal Analysis and Testing (Sesimbra, Portugal), pp. 725–741.

Brownjohn, J.M.W. (2007), “Structural health monitoring of civil infrastructure”. Philos. Trans. R. Soc. Math. Phys. Eng. Sci. 365, 589–622.

Ceravolo, R. (2009), “Time–Frequency Analysis”. Encyclopedia of Structural Health Monitoring, John Wiley & Sons.

Cerda F., Garret J. H., Bielak J., Rizzo P., and Barrera J. (2012). Indirect structural health monitoring in bridges: scale experiments. Proceedings of the Seventh International Conference on Bridge Maintenance, Safety and Management. Lago di Como.

Choi, H.-I., and Williams, W.J. (1989), “Improved time-frequency representation of multicomponent signals using exponential kernels”, IEEE Trans. Acoust. Speech Signal Process. 37, 862–871.

Chopra, A.K. (2012). “Dynamics of structures: theory and applications to earthquake engineering”, Prentice Hall, Boston.

Cohen L. (1995), Time-frequency Analysis, Englewood Cliffs, N.J: Prentice Hall PTR.

Dantu, K., Rahimi, M., Shah, H., Babel, S., Dhariwal, A., and Sukhatme, G.S. (2005). Robomote: enabling mobility in sensor networks. In IPSN 2005. Fourth International Symposium on Information Processing in Sensor Networks, pp. 404–409.

Flandrin, P. (1984). “Some features of time-frequency representations of multicomponent signals”. In Acoustics, Speech, and Signal Processing, IEEE International Conference on ICASSP ’84., pp. 266–269.

Gabor, D. (1946). “Theory of communication”  Institution of Electrical Engineering, London.

Hammond J. K.  and White P. R. , “The Analysis of Non-Stationary Signals Using Time-Frequency Methods”, J. Sound Vib., vol. 190, no. 3, pp. 419–447, Feb. 1996.

Hlawatsch, F., and Auger, F. (2013). “Time-Frequency Analysis”, John Wiley & Sons.

Hlawatsch, F., and Boudreaux-Bartels, G.F. (1992). “Linear and quadratic time-frequency signal representations”. IEEE Signal Process. Mag. 9, 21–67.

Arango H. (2009), “Análisis de señales con las transformadas de Fourier, Gabor y Onditas”. Instituto Tecnológico Metropolitano, Medellín, Colombia, 2009.

Karbhari, V.M., Guan, H., and Sikorsky, C. (2009).  “Operational modal analysis for vibration-based structural health monitoring of civil structures”. Structural Health Monitoring of Civil Infrastructure Systems, (Woodhead Publishing), pp. 213–259.

Lin, C.W., and Yang, Y.B. (2005). “Use of a passing vehicle to scan the fundamental bridge frequencies: An experimental verification”. Eng. Struct. 27, 1865–1878.

Lippmann, R.P. (1989). “Pattern classification using neural networks”. IEEE Commun. Mag. 27, 47–50.

Loughlin, P.J., Pitton, J.W., and Atlas, L.E. (1993). “Bilinear time-frequency representations: new insights and properties”. IEEE Trans. Signal Process. 41, 750–767.

Lynch, J.P. (2006). “A Summary Review of Wireless Sensors and Sensor Networks for Structural Health Monitoring”. Shock Vib. Dig. 38, 91–128.

Magalhães, F., Cunha, Á., and Caetano, E. (2008). “Dynamic monitoring of a long span arch bridge”. Eng. Struct. 30, 3034–3044.

Magalhães, F., Cunha, Á., and Caetano, E. (2009). “Online automatic identification of the modal parameters of a long span arch bridge”. Mech. Syst. Signal Process. 23, 316–329.

Marulanda, J., Caicedo, J.M., and Thomson, P. (2016). “Modal Identification Using Mobile Sensors under Ambient Excitation”. J. Comput. Civ. Eng. 04016051.

Marulanda J. (2014), “Modal Identification Using Smart Mobile Sensing Units”. Programa Editorial Universidad del Valle. Cali, Colombia.

Matarazzo, T.J., and Pakzad, S.N. (2014). “Modal Identification of Golden Gate Bridge Using Pseudo Mobile Sensing Data with STRIDE”. Dynamics of Civil Structures, Volume 4, F.N. Catbas, ed. (Springer International Publishing), pp. 293–298.

Matarazzo, T.J., and Pakzad, S.N. (2016). “STRIDE for Structural Identification Using Expectation Maximization: Iterative Output-Only Method for Modal Identification”. J. Eng. Mech. 142, 04015109.

Matarazzo T. J., and  Pakzad S. N. (2013). “Mobile Sensors in Bridge Health Monitoring”. International Workshop on Structural Health Monitoring, Stanford, CA.

Oppenheim, A.V., and Schafer, R. W (2010). “Discrete-time signal processing”, Upper Saddle River, Pearson.

Papandreou, A., and Boudreaux-Bertels, G.F. (1993). “Generalization of the Choi-Williams Distribution and the Butterworth Distribution for Time-Frequency Analysis”. IEEE Trans. Signal Process. 41, 463-.

Pastor, M., Binda, M., and Harčarik, T. (2012). “Modal Assurance Criterion”. Procedia Eng. 48, 543–548.

Rioul, O., and Vetterli, M. (1991). “Wavelets and signal processing”. IEEE Signal Process. Mag. 8, 14–38.

Sibley, G.T., Rahimi, M.H., and Sukhatme, G.S. (2002). “Robomote: a tiny mobile robot platform for large-scale ad-hoc sensor networks”. IEEE International Conference on Robotics and Automation, 2002. Proceedings. ICRA ’02, pp. 1143–1148.

Spencer, B.F., Ruiz-Sandoval, M.E., and Kurata, N. (2004). “Smart sensing technology: opportunities and challenges”. Struct. Control Health Monit. 11, 349–368.

Staszewski, W.J., and Tomlinson, G.R. (1994). “Application of the wavelet transform to fault detection in a spur gear”. Mech. Syst. Signal Process. 8, 289–307.

Wang W. J. and  McFadden P. D. (1993), “Early detection of gear failure by vibration analysis I. Calculation of the time-frequency distribution”, Mech. Syst. Signal Process., vol. 7, no. 3, pp. 193–203.

Zhu, D., Yi, X., Wang, Y., Guo, J., and Lee, K.-M. (2010). “Mobile Sensor Networks: A New Approach for Structural Health Monitoring”. American Society of Civil Engineers, pp. 159–168.

Zhu, D., Guo, J., Cho, C., Wang, Y., and Lee, K.-M. (2012). “Wireless Mobile Sensor Network for the System Identification of a Space Frame Bridge”. IEEE ASME Trans. Mechatron. 17, 499–507.

Back to Top

Document information

Published on 28/08/19
Submitted on 20/08/19

Volume 19 (1), 2019
Licence: CC BY-NC-SA license

Document Score

0

Views 58
Recommendations 0

Share this document