[go: up one dir, main page]

Academia.eduAcademia.edu
Journal of Composite Materials http://jcm.sagepub.com/ Dynamics-based Damage Detection of Composite Laminated Beams using Contact and Noncontact Measurement Systems Pizhong Qiao, Wahyu Lestari, Mitali G. Shah and Jialai Wang Journal of Composite Materials 2007 41: 1217 DOI: 10.1177/0021998306067306 The online version of this article can be found at: http://jcm.sagepub.com/content/41/10/1217 Published by: http://www.sagepublications.com On behalf of: American Society for Composites Additional services and information for Journal of Composite Materials can be found at: Email Alerts: http://jcm.sagepub.com/cgi/alerts Subscriptions: http://jcm.sagepub.com/subscriptions Reprints: http://www.sagepub.com/journalsReprints.nav Permissions: http://www.sagepub.com/journalsPermissions.nav Citations: http://jcm.sagepub.com/content/41/10/1217.refs.html >> Version of Record - May 18, 2007 What is This? Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 Dynamics-based Damage Detection of Composite Laminated Beams using Contact and Noncontact Measurement Systems PIZHONG QIAO* Department of Civil and Environmental Engineering, Washington State University Pullman, WA 99164-2910, USA WAHYU LESTARI Department of Aerospace Engineering, Embry-Riddle Aeronautical University Prescott, AZ 86301, USA MITALI G. SHAH Department of Civil Engineering, The University of Akron Akron, OH 44325-3905, USA JIALAI WANG Department of Civil, Construction, and Environmental Engineering The University of Alabama, Tuscaloosa, Alabama 35487-0205, USA ABSTRACT: A reliable and effective damage detection technique is one of the significant tools to maintain the safety and integrity of structures. A dynamic response offers viable information for the identification of damage in the structures. However, the performance of such dynamics-based damage detection depends on the quality of measured data and the effectiveness of data processing algorithms. In this article, the experimentally measured data of two sensor systems, i.e., a surfacebonded piezoelectric sensor system and a noncontact scanning laser vibrometer (SLV) system, are studied, and their effectiveness in damage identification of composite laminated beams is compared. Three dynamics-based damage detection algorithms are evaluated using the data acquired from these two measurement systems. The curvature mode shape is selected as a parameter to locate damage due to its sensitivity. The piezoelectric sensors directly acquire the curvature mode shapes of the structures, while the SLV measures the displacement mode shapes. The difference in the measurement characteristics of these systems and their influence in the damage identification performance are addressed. The beam specimens *Author to whom correspondence should be addressed. E-mail: qiao@wsu.edu Figures 1–4, 7–9, 12 and 13 appear in color online: http://jcm.sagepub.com Journal of COMPOSITE MATERIALS, Vol. 41, No. 10/2007 0021-9983/07/10 1217–36 $10.00/0 DOI: 10.1177/0021998306067306 ß 2007 SAGE Publications Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1217 1218 P. QIAO ET AL. are made of E-glass/epoxy composites, and several different types of damages are introduced in the beams (i.e., delaminations, and impact and saw-cut damages). This study provides a thorough assessment of the two sensor systems in damage detection of composite laminated beams and verifies the validity of dynamics-based damage detection methodology in locating the local defects in composite structures. KEY WORDS: damage detection, dynamic response, curvature mode shapes, scanning laser vibrometer, piezoelectric sensors, delamination, laminated beams. INTRODUCTION DAMAGE DETECTION has gained increasing attention from the scientific community because unpredicted failure causes major economic loss and casualties. Higher operational loads, greater complexity of design (e.g., structures made of composite and hybrid materials), and longer lifetime periods imposed on civil, mechanical, and aerospace structures make it increasingly important to monitor the health of these structures. The availability of practical and robust non-destructive evaluation techniques for damage detection is critical to ensure acceptable performance of structures in terms of serviceability, reliability, durability, and prevention of catastrophic failure. Dynamic responses, which in many cases can be obtained easily, offer damage information such as the location and severity of the damage. The performance of the dynamics-based damage identification strongly depends on the quality of the measured dynamic responses. In general, the damage in the structure is identified by comparing the dynamic responses of original (pristine) structures with those of damaged structures or those after a certain period of time in service. When the original dynamic response is not available, the comparison can be performed by using the numerical simulation or mathematical approximation based on the current condition of the structures. A review on vibration-based delamination detection methods for composite structures by Zou et al. [1] discussed several approaches using different dynamic parameters, such as frequency, damping, and mode shape. The comparative study of various damage detection methods showed that the modal analysis method in general is global in nature, and it requires the data of original (or undamaged) structures as the benchmark. Although the frequency domain method seems cost-effective, it alone is not sufficient for locating structural damage. However, it can still be used as an indication of the presence of damage. The time domain method, which is usually in combination with frequency information, is capable of detecting damage events, both locally and globally, by changing the input frequencies. The impedance domain method offers a reliable approach, and it is particularly suitable for detecting planar defects such as delamination. Application of frequency in damage identification in the form of the response spectrum [2] showed that the use of vibration at higher frequencies allows the identification of delamination occurrence in the cantilever composite beams. Although the method can differentiate the size of delamination, it is unable to determine their quantity. The response spectrum of the sample was analyzed to infer the presence of delamination by comparing it with the baseline spectrum of the undamaged structure. S TRUCTURAL Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 Dynamics-based Damage Detection of Composite Laminated Beams 1219 Owing to the localized effect of stiffness change at the damage location, curvature mode shapes have been employed and studied for structural damage identification by many researchers since it was introduced by Pandey and Biswas [3]. A comparison of the effectiveness of the curvature mode shape method and frequency response function (FRF) curvature method as damage detection techniques was investigated by Sampaio et al. [4]. The damage indication parameters are determined by two methods called the ‘mode shape curvature method’ and the ‘damage index method’. The stiffness change of the damaged element was modeled as a percentage reduction of the undamaged one. Their results showed that the FRF curvature method performed well in detecting, locating, and quantifying damage, especially for a range of frequency before the occurrence of the first resonance or anti-resonance. Wahab and Roeck [5] investigated the application of modal curvature (MC) change to identify various levels of damage in a prestressed concrete bridge, while considering a currently intact bridge before it was artificially damaged as the reference measurement. The results confirmed the application of MC in detecting damage in civil engineering structures. However, MC was found to be more accurate for lower modes, and it was recommended that the fine measurement grid be used when the higher modes were desired for damage detection. The experimental study on carbon/epoxy composite laminated beams carried out by Hamey et al. [6] by using the piezoelectric sensor system provided information on the efficiency and drawback of several existing damage detection algorithms (e.g., absolute difference, damage factor, damage index, and FRF curvature methods). The effects of damage type (i.e., delamination, saw-cut, and impact damages) and size (i.e., delamination length) were studied as well. A comparative numerical study of damage identification algorithms applied to a bridge was performed by Farrar and Jauregui [7]. After benchmarking a finite element (FE) model against previously measured modal data from a bridge in its undamaged and damaged conditions, extensive numerical studies were performed to further evaluate various damage detection techniques, i.e., change in stiffness, damage index, change in mode curvature, change in uniform load surface (ULS) curvature, and change in flexibility. In general, all the methods identified the various damage locations correctly in the case of detecting severe damage like a cut completely through the bottom flange. But the methods were found to be inconsistent and failed to clearly identify the damage location when they were applied to less-severe damage cases. Results showed that the damage index method performed the best and was found to be the most convincing of all methods. An improved damage index formulation was presented by Kim et al. [8]. Both, the natural frequency- and mode-shape-based damage detection were developed and evaluated to locate and estimate the size of damage in structures. A damage index algorithm to localize and estimate the severity of damage was developed from monitoring the changes in the modal strain energy, and the required natural frequencies and mode shapes were generated from the FE models. The frequency-based method was observed to locate damage with some localization error, whereas the mode-shape-based method was able to detect damage accurately. A procedure for locating variability in structural stiffness using only the data obtained from the damaged structure was introduced by Yoon et al. [9]. Assuming the original healthy mode shapes are smooth without irregularity and using a curvefitting technique, the original mode shapes were approximated. A structural irregularity index was generated for each measurement data, and was later averaged over the obtained modes to increase the sensitivity of the indices. Numerical analysis and Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1220 P. QIAO ET AL. experimental vibration tests were performed to demonstrate the proposed method, which produced successful results in detecting the size and location of small, localized stiffness reduction. A damage identification approach using the changes in the strain energy based on measured modal parameters was discussed by Cornwell et al. [10]. This method requires that the mode shapes before and after damage be known, but the modes need not be mass-normalized and only a few modes are required. The 1-D strain energy method could be applied to only beam-type structures; however, a modification in the method can extend the application to the plate-type structures. This algorithm was found to be effective in locating areas with a stiffness reduction as low as 10% using relatively few modes. The strain energy-based damage identification using axial or torsional response was studied by Duffey et al. [11]. The changes of strain energy in discrete elements of the structure from the undamaged to damaged states were used as the damage parameter that was illustrated by numerical simulation and experimental spring–mass systems. The overall identification results depend strongly upon damage location, damage level, and the number of modes included. To minimize the sensitivity of the mode shapes to experimental error, truncation effects, and selection of mode shapes in damage identification, Zhang and Akhtan [12] proposed a new deformation parameter – ULS, which is based on the assumption that a structure is subjected to a uniform loading. Calculation of the ULS that uses the modal flexibility and measured mode shapes has an averaging effect. This process reduces the sensitivity of the damage identification method to the truncation effects, experimental error, and selection of mode shapes. Application of the ULS curvature to 2-D structures was demonstrated by Wu and Law [13]. The ULS curvature was calculated based on the Chebyshev polynomial, instead of the central difference method, to reduce the edge effects. The numerical examples considering different support conditions, measurement noises, mode truncation, and sensor sparsity were studied to evaluate the effectiveness of the proposed method. When the modal parameter of healthy structures was approximated by using the gapped-smoothing method, the results of the numerical examples revealed that the change in ULS curvature was found to be very sensitive to local damages and robust to truncation effects. Thus, the combination of the gapped smoothing with ULS was highly successful in detecting and locating damage in the structure. Direct use of the modal flexibility for identification of damage was proposed by Gao et al. [14], and it was called the damage localization vector (DLV) method. The technique was based on the determination of a special set of DLVs that have the property that, when they are applied to the structure as static forces at the sensor locations, no stress is induced in the damaged elements. The experimental demonstration using the truss structure indicated that the flexibility-based DLV method could be utilized to detect and locate damage using a limited number of sensors and truncated modes of the structure. Lu et al. [15] demonstrated numerically that the changes in flexibility and flexibility curvature have high sensitivity to closely distributed damages and can be effectively used for detecting multiple damage locations. Essentially, the success of damage identifications depends strongly on the quality and selection of the parameters involved, such as the number of mode shapes, mode selection, and frequency bandwidth. This article reports a comparative study of the performance of two measurement systems in providing the dynamic parameters for structural damage identification of composite laminated beams. One is a non-contact Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 Dynamics-based Damage Detection of Composite Laminated Beams 1221 scanning laser vibrometer (SLV) system, and the other utilizes a surface-bonded piezoelectric (polyvinylidenefluoride, PVDF) sensor array system. While the SLV is very convenient and efficient in acquiring dynamic responses, it does not have the feasibility that the piezoelectric sensors can offer to be developed for an onboard sensor system. On the other hand, the piezoelectric sensor is sensitive to interference and requires considerable time for employment. The results measured from both sensor systems are evaluated to identify damage in composite laminated beams with various damage configurations (i.e., delaminations, saw-cut damage, and impact damage). Three damage detection algorithms (i.e., gapped-smoothing method (GSM), generalized fractal dimension (GFD), and strain energy method (SEM)) are used in analyzing the measured data. The theoretical background of the damage detection algorithms is described briefly, and the numerical FE analysis is performed to verify the validity of the algorithms and guide the experimental procedure. The efficiency and effectiveness of the two sensor systems are examined, and the shortcomings are discussed. DAMAGE DETECTION ALGORITHMS The quality of the measurement systems in providing data for damage identification is evaluated by three newly modified damage detection algorithms. All three approaches use the curvature mode shapes of the structure as the main parameter to extract the damage information. The first approach is known as the GSM, which considers damage as a disturbance in the mode shape that is otherwise a smooth one. The second one is a GFM [16], developed based on the fractal theory with modifications in the variable. The last one, SEM, is based on the concept of strain energy, in which the beam bending stiffness and curvature are interrelated, and the change in mode shapes is associated with the change of beam bending stiffness due to damage. The modal parameters obtained from the PVDF sensor system are directly in the form of curvature mode shapes, and the mode shapes can thus be utilized as the input data to the damage detection algorithms, whereas the measured results from SLV system are in the form of displacement mode shapes. Therefore, the required curvature mode shapes are calculated by the second-order derivatives based on the fourth-order central difference approximation as 0 ¼   wiþ2 þ 16wiþ1  30wi þ 16wi1  wi2 12h2 ð1Þ where 0 is the curvature mode shape calculated at the ith grid point; wi is the displacement measured at node i for a mode j; and h is the distance between the measurement nodes i and i þ 1, where i denotes the node under observation along the length of the beam. The curvature mode shapes are then normalized before being further utilized as input for damage detection algorithms. Instead of employing the curvature mode shapes individually, a new surface deformation parameter is constructed based on the measured frequencies and curvature mode shapes as well as the assumption that the structure is subjected to uniform loading. Since this parameter is calculated from the collective mode shapes, it has an averaging effect that makes it less sensitive to mode shape selection and measurement errors. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1222 P. QIAO ET AL. The obtained parameter, which was introduced by Zhang and Akhtan [12], is known as uniform surface loading (ULS). This approach is especially useful to improve the performance of measured experimental data. Uniform Load Surface (ULS) Method Assuming that all degrees of freedom are subjected to a unit load together, the resulting displacement vector can be obtained by multiplying the flexibility matrix with the loading vector of the uniform load along the structure. Although it is difficult to achieve this condition in the actual structure, a linear system approximation of the mode shape under uniform loading based on the flexibility matrix is valid. For a structural system with m mode shapes data and n degrees of freedom, the deflection vector U of the structure under ULS can be defined as 2 f1, 1    6 6 f2, 1 U ¼ fi, j lj ¼ 6 . 4 .. fn, 1 f1, 2 f2, 2 .. .   .. . fn, 2  3 9 f1, n 8 1> > > = < > f2, n 7 7 1 .. 7 .. > >.> . 5> : ; 1 fn, n ð2Þ with uðiÞ ¼ n X fi, j ¼ m n X r ðiÞ X r¼1 j¼1 !2r r ð jÞ ð3Þ j¼1 or U¼FL ð4Þ where F is the modal flexibility matrix and L ¼ f1, . . . ,1gT1n is the unit vector representing the uniform load acting on the structure. Each term in the summation for the deflection vector u(i) has the corresponding frequency in the denominator, and this results in a rapid decrease of the higher-mode term-contribution. Hence, the ULS can be well-approximated by only a few lower-mode terms. Because the technique considering the ULS reduces the sensitivity of the damage identification to the mode shape choices, such a technique is adapted in this study to evaluate the location and relative magnitude of damage. In this study, the mode shapes are in the form of a curvature, and correspondingly the ULS approximation results are denoted as . The modal flexibility matrix for a structural system with n degrees of freedom can be determined by the following expression [12]: F¼ n X r T r r¼1 !2r ð5Þ where r is the rth normalized mode shape and !r is the corresponding rth natural frequency. In practice, the modal testing can only yield several low modes. With m modes Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 Dynamics-based Damage Detection of Composite Laminated Beams 1223 available from experimental data, the modal flexibility matrix of the structure can be approximated as m   X r Tr ð jÞ F ¼ fi, j ¼ !2r r¼1 ð6Þ in which fi,j describes the modal flexibility at the ith point under the unit load at the point j and is the summation of the product of the related term from available modes. Hence, the displacement vector can be obtained by multiplying the flexibility matrix with the loading vector. Gapped-Smoothing Method (GSM) In some cases such as for structures that are already in service for a long period of time, data of healthy or undamaged structures are rarely available. These data can be approximated by using a gapped-smoothing technique, where the basic assumption is that the mode shape of a healthy structure has a smooth surface [17]. Using the mode shape data of a damaged structure and an interpolation technique with polynomial approximation, the smooth mode shape surfaces of healthy structures are estimated. In this study, the ULS of a healthy beam is approximated as a fourth-order polynomial as ðxÞ ¼ c0 þ c1 x þ c2 x2 þ c3 x3 þ c4 x4 ð7Þ where the coefficients of the polynomial c0, c1, c2, c3, and c4 can be determined using the regression analysis technique based on either the numerical or experimental data. The damage parameter (GSM) based on this approach is calculated as the square of the difference between the measured data of the damaged structure (measured) and the fitted value represented the healthy data (GSM) from Equation (7): GSM ¼ ½measured  GSM ðxÞ2 ð8Þ Generalized Fractal Dimension (GFD) Method The concept of a fractal curve is based on the principle that, if the ruler length is reduced by 1/r, then the length of the curve would correspondingly increase to L ¼ rD multiplied by the original value of length. The power D in the expression is known as the fractal dimension (FD) of the curve. A regular smooth curve has an FD of one, while a curve with irregularity will have a FD greater than one. Thus, as the value of D increases, irregularity in the curve increases, which enables the prediction of the location of damage. The FD expression given by Katz [18] is stipulated as FDM ðxÞ ¼ logðnÞ logðnÞ þ logðdðxi , MÞ=lðxi , MÞÞ Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 ð9Þ 1224 P. QIAO lðxi , MÞ ¼ M qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi X ððxiþj Þ  ðxiþj1 ÞÞ2 þ ðxiþj  xiþj1 Þ2 ET AL. ð10Þ j¼1 M qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi X ððxiþj Þ  ðxi ÞÞ2 þ ðxiþj  xi Þ2 dðxi , MÞ ¼ ð11Þ 1jM is the average distance between successive where x ¼ 1/2 (x1 þ xiþM), n ¼ 1/ , and points. The term M represents the sliding-window dimension length. However, the fractal dimension technique has the drawback of representing an abnormality in the mode shape as being more severe near the maximum curvature point as compared to other points on the mode shape [16]. As a result, a modification to the above-mentioned algorithm is carried out to produce more accurate results. The modified algorithm defined as GFD was recently presented by Wang and Qiao [16] as logðnÞ logðnÞ þ logðds ðxi , MÞ=ls ðxi , MÞÞ M qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi X ls ðxi , MÞ ¼ ððxiþj Þ  ðxiþj1 ÞÞ2 þ S2 ðxiþj  xiþj1 Þ2 GFDM ðxÞ ¼ ð12Þ ð13Þ j¼1 ds ðxi , MÞ ¼ M qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi X ððxiþj Þ  ðxi ÞÞ2 þ S2 ðxiþj  xi Þ2 : ð14Þ 1jM The new S parameter is a scale parameter that enables the GFD to have a multi-scale feature. As the abnormality in the deformation mode shape is more localized in nature, the extra peaks aroused at the other parts of the structure are eliminated by choosing a proper scale S for the deformation abnormality peak. The GFD method receives prime importance in online data processing because it requires only a small segment of measured signal to detect the damage and gives a sharp peak at the location of damage. Strain Energy Method (SEM) The strain-energy-based damage detection algorithm aims to use the strain energy changes as a damage indicator for structural health monitoring. As the first step, the strain energy pertaining to each mode shape for the entire beam is calculated using the following equation: 1 U¼ 2 Z l o  2 2 @ w EI dx: @x2 ð15Þ As @2w/@x2 refers to the curvature  of the beam, the above equation can be rewritten as U¼ 1 2 Z l EIðÞ2 dx ) U / 2 : o Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 ð16Þ 1225 Dynamics-based Damage Detection of Composite Laminated Beams Based on the above analogy, we can observe that the strain energy is proportional to the square of the curvature; thus, the square of the curvature can be used in damage detection, and it can reflect the change of beam bending stiffness due to damage under the condition of strain energy continuity along the beam length. Considering the measured curvature and the resulting GSM approximation, the damage parameter based on the strain energy approach is defined as   SEM ¼ 2measured  ðxÞ2  ð17Þ where measured is the curvature measured experimentally or obtained numerically; (x) is the smooth-curve fit of measured using a fourth-order polynomial (Equation (7)). Thus, SEM enables the use of beam strain energy as an alternative to the curvature for damage detection. NUMERICAL ANALYSIS The numerical FE analysis is conducted to provide insight and understanding of the structural dynamic behavior of damaged structures and to verify the validity of the abovementioned damage detection algorithms. Dynamic characteristics resulting from numerical analysis can also be used to guide the experimental study. The modeling of composite laminated beams and the analysis of dynamic characteristics are performed using commercial software ANSYS. The structure was a rectangular beam section of length 0.5588 m (22.0 in.), width 50.8 mm (2 in.), and a thickness of 4.8 mm (0.189 in.). The lamination of the composite had a [CSM/90(0/90)3]S lay-up configuration with a total of 16 layers (where, CSM refers to continuous-strand mat). The thickness of the outer layer is about 0.5 mm, while the thickness of the remaining individual layer is about 0.268 mm (0.0117 in.). The beam was discretized into 350 elements along the length, 32 elements along the width, and two elements along the thickness. The beam specimen is modeled using the SOLID46 layered element, and the material properties of the orthotropic composite are listed in Table 1. The FE model for a healthy cantilevered beam is shown in Figure 1. Table 1. Material properties for the E-glass–epoxy composite beam model. Elastic modulus (E1) Elastic modulus (E2) Shear modulus (G12) Poisson’s ratio (12) Mass density () 24.0 GPa 14.2 GPa 2.17 GPa 0.4 2500 kg/m3 Solid 46 0.0508m 0.5588 m Figure 1. Finite element model of a healthy composite beam. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1226 P. QIAO ET AL. 0.5588m (a) 0.3302m 0.0508m 44 56 0.5588m (b) 0.2540m 0.1016m 38 58 0.5588 m (c) 0.0508 m 0.3810 m 62 (d) 72 150 0.001 m 0.2794m 0.5588m (e) 0.0508m 0.1778m 65 75 Figure 2. FE model of damage beams: (a) Delamination A model; (b) Delamination B model; (c) Delamination C model; (d) Saw-cut damage model; and (e) Impact damage model. Case Study A numerical study of dynamic response in composite laminated beams with five different damage cases is performed based on the numerical modal analysis. The five damage cases are three delaminations with different sizes or locations, saw-cut, and impact damages, which correspond to the physical specimens introduced in the experimental program. The FE models of the beams for all the damage configurations are shown in Figure 2. The delamination conditions prevalent in the beam specimens are simulated in the FE model by using the bilinear LINK10 elements. LINK10 is a 3-D spar element having the unique feature of a bilinear stiffness matrix resulting in a uniaxial tension-only (or compression-only) element. The tension-only option is utilized in the FE model, where the stiffness goes to zero if the element goes into tension, whereas the stiffness is set to 2 GPa when it goes into compression. The opening of the delamination for all the cases Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1227 Dynamics-based Damage Detection of Composite Laminated Beams SOLID46 Bilinear LINK10 Figure 3. Detail view of the delaminated beam model. 0.5588m 0.1778m 20 40 60 1 2 3 4 65 60 40 20 5 6 Stiffness reduction 7 8 75 Figure 4. Detailed view of stiffness reduction on an impact-damaged beam. Table 2. Details of the stiffness reduction in an impact-damaged beam. Beam section (percentage reduction) 1 2 3 4 5 6 7 8 (0) (20) (40) (60) (60) (40) (20) (0) Stiffness, E1 (GPa) 24.00 19.20 14.40 9.60 9.60 14.40 19.20 24.00 was taken as 0.1 mm. An enlarged view of the delaminated area and the arrangement of the LINK10 element are detailed in Figure 3. The saw-cut damage is induced in the beam specimen by creating a transverse notch in the beam. It runs along the beam length from 0.2794 to 0.2804 m. The notch has a width of around 1.0 mm (0.0394 in.) and is cut to about 50% of the beam thickness through the width of the beam. The formulation of the saw-cut beam model remains the same as the delaminated beams with respect to the dimensions, elements used, and the meshing size adopted. However, the total number of FE elements along the beam length is 285. The schematic representation of the saw-cut-damaged beam is shown in Figure 2(d). The impact damage is modeled by reducing the stiffness of the beam elements gradually along the beam length in the damaged area. The beam is divided into eight sub-elements in the area of damage, as highlighted in Figure 4. The longitudinal stiffness (E1) in the element is reduced in steps of 20% and the stiffness properties for the impact-damaged beam elements are organized in Table 2. The impact damage ranged from 0.1778 Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1228 P. QIAO ET AL. Table 3. Natural frequencies obtained from the numerical modal analysis. Specimen Second mode (Hz) Third mode (Hz) Fourth mode (Hz) 62.55 60.63 57.45 59.20 58.15 55.75 160.50 156.22 152.30 155.00 154.10 153.30 315.25 311.10 304.80 310.00 309.90 308.10 Healthy Delamination A Delamination B Delamination C Saw-cut Impact Mode 2 Mode 3 Mode 4 4 (b) Curvature Displacement (a) 2 0 0.2 Mode 2 Mode 3 Mode 4 0.1 0.0 −0.1 −2 0 20 40 60 −0.2 80 0 20 Mode 2 Mode 3 Mode 4 4 2 (d) 60 80 0 0.3 Mode 2 Mode 3 Mode 4 0.2 Curvature Displacement (c) 40 Sensor location Sensor location 0.1 0.0 −0.1 −2 0 20 40 60 −0.2 80 0 20 Sensor location Mode 2 Mode 3 Mode 4 4 2 (f) 60 0 0.3 80 Mode 2 Mode 3 Mode 4 0.2 Curvature Displacement (e) 40 Sensor location 0.1 0.0 −0.1 −2 −0.2 0 20 40 60 Sensor location 80 0 20 40 60 80 Sensor location Figure 5. Displacement and curvature mode shapes obtained from the numerical analysis: (a) Healthy – displacement – FE; (b) Healthy – curvature – FE; (c) Delamination A – displacement – FE (44–56); (d) Delamination A – curvature – FE (44–56); (e) Delamination B – displacement – FE (38–58); (f) Delamination B – curvature – FE (38–58); (g) Delamination C – displacement – FE (62–72); (h) Delamination C – curvature – FE (62–72); (i) Saw-cut damage – displacement – FE (150); (j) Saw-cut damage – curvature – FE (150); (k) Impact damage – displacement – FE (65–75); and (l) Impact damage – curvature – FE (65–75). Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1229 Dynamics-based Damage Detection of Composite Laminated Beams (g) Mode 2 Mode 3 Mode 4 (h) 2 0 Mode 2 Mode 3 Mode 4 0.1 0.0 −0.1 −0.2 −2 0 20 40 60 Sensor location (i) −0.3 80 Mode 2 Mode 3 Mode 4 (j) Curvature 4 Displacement 0.3 0.2 Curvature Displacement 4 2 0 0 20 40 60 Sensor location 0.4 Mode 2 Mode 3 Mode 4 0.2 0.0 −0.2 −2 0 50 100 150 200 250 0 50 100 Sensor location (k) 150 200 250 Sensor location Mode 2 Mode 3 Mode 4 4 2 (l) 0 −2 0.3 Mode 2 Mode 3 Mode 4 0.2 Curvature Displacement 80 0.1 0.0 −0.1 0 50 100 150 0 Sensor location 50 100 150 Sensor location Figure 5. Continued. to 0.2286 m (elements 65–75). The FE mesh used comprises a total of 185 elements along the beam length as shown in Figure 2(e). Numerical Results The natural frequencies obtained from the numerical modal analysis are presented in Table 3. The curvature mode shapes are obtained by using the central difference derivation of displacement mode shapes from the numerical modal analysis (Equation (1)). Figure 5 shows the displacement and curvature mode shapes obtained from the numerical analysis, side by side for each case. The actual locations of damage in terms of node points are given in parentheses. The damages are hardly visible in the displacement mode shapes, and only the difference or shift of nodal points are noticeable. While in the curvature mode shapes, the damage effects on the mode shapes are very discernable by the discontinuity of the curvature at the boundaries of the damages. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1230 P. QIAO ET AL. UM 1208 (CSM+unidirectional) C1800 [0/90] cross-ply Figure 6. Lay-up of composite samples. 4.8mm 0.6096 m 50.8 mm Figure 7. Dimension of beam specimens. EXPERIMENTAL PROGRAM Beam Specimens The experiment is conducted on six composite laminated beam samples, i.e., one healthy and five damaged beams. Each test sample was made of E-glass fiber and epoxy resins and has a [CSM/0(90/0)3]S lay-up for a total of 16 layers as shown in Figure 6. UM 1208 (CSM þ unidirectional) stitched combo layer has a thickness of 0.49 mm (0.0189 in.) for unidirectional ply and 0.23 mm (0.0087 in.) for CSM ply. The thickness of C1800 is 0.28 mm/ply (0.0117 in.) and 0.56 mm/mat (0.0220 in.). The 16 layers lead to a total composite thickness of 4.8 mm (0.189 in.). A composite plate is fabricated using a vacuum bagging process, and it is then cut in the beam samples with dimensions of 0.0508 m (2 in.) wide and 0.6096 m (24 in.) long as shown in Figure 7. The length of the beam for a cantilevered condition is 0.5588 m (22 in.). Figure 8 shows the picture of the beam specimens with different damage types. As aforementioned, six E-glass–epoxy composite laminated beam specimens are tested in this study. Experimental modal analysis is first conducted on an intact (healthy) beam specimen. The intact curvature modes were recorded for reference. To study the effect of the location and size of different types of damage on the modal parameters, three delamination damage cases of different length and location (i.e., delamination A, delamination B and delamination C) as well as saw-cut and impact damages are studied. The delamination is introduced in the beam by inserting a Teflon film between the second and third layer of the composite laminate during the manufacturing process of the composite plate. After curing, the plate is cut into beam samples, and the Teflon film is Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1231 Dynamics-based Damage Detection of Composite Laminated Beams (a) Delamination C (b) Delamination B (c) Delamination A (d) Impact damage (e) Saw-cut damage Figure 8. Specimens with various types of damage. Table 4. Damage location and size of the specimens. No. 1 2 3 4 5 Damage position according to sensor location Damage type Damage location from the fixed end (m) PVDF SLV FE Delamination A Delamination B Delamination C Saw-cut Impact 0.3302–0.3810 0.2540–0.3556 0.3810–0.4318 0.2794–0.2804 0.1778–0.2286 11–13 8–12 13–15 9 5–7 37–43 27–39 53–59 48 22–28 44–56 38–58 62–72 150 65–75 pulled out, leaving a debonded area (delamination) in the beam sample between the second and third layers. For example, the beam with delamination A has a 50.8 mm (2 in.) delamination beginning at 0.3302 m from the cantilevered end. The saw-cut damage is introduced at 0.2794 m from the cantilevered end. The depth of the through width saw-cut is about half the specimen thickness. To simulate impact damage, the beam sample is impacted in the region of 0.1778–0.2286 m from the supporting end, using the material testing systems (MTS) machine. Details of all the beam specimens and damage configurations are summarized in Table 4, and the corresponding specimen geometry and PVDF sensor locations are also shown in Figure 9. Sensor Systems The complexity and sensitivity of damage detection methods depend on the quality of the measurement of the parameter and also on the sensitivity of that parameter to possible Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1232 P. QIAO 1 (a) 2 3 4 5 6 7 8 9 ET AL. 10 11 12 13 14 15 16 17 18 19 0.3302 m 0.0508 m 0.5588 m 1 (b) 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 0.1016 m 0.2540 m 0.5588 m 1 (c) 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 0.3810 m 0.0508 m 0.5588 m (d) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 0.001 m 0.2794 m 0.5588 m (e) 1 2 0.1778 m 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 0.0508 m 0.5588 m Figure 9. Damage configuration and PVDF sensor location in the specimens: (a) Delamination A; (b) Delamination B; (c) Delamination C; (d) Saw-cut damage; and (e) impact damage. structural damage. When damage occurs, a structure suffers a change (a decrease in most cases) in stiffness, and as a consequence, a change is reflected in the dynamic parameters, such as the frequencies, mode shapes, FRF, etc. Structural vibration measurements can be conveniently divided into two major types: (1) discrete point measurements, using an accelerometer or equivalent transducer; and (2) spatial field measurements, of which speckle pattern interferometry is a modern example. In this study, two discrete-point measurement-based sensor systems (i.e., PVDF sensor and SLV) are employed to acquire the dynamic response of the composite beam samples. The dynamic response obtained from the SLV system is in the form of displacement mode shapes; whereas the PVDF sensor system directly yields the desired curvature mode shapes. Piezoelectricity is a phenomenon observed in certain crystal (e.g., quartz) PZT (lead–zirconate–titanate) ceramic materials, and PVDF (polyvinylidenefluoride) polymer. In both cases, the vibration excitation is generated through a piezoelectric actuator. The PZT ceramic actuators have the dimensions Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1233 Dynamics-based Damage Detection of Composite Laminated Beams Sensor dSPACE connector Actuator Data acquisition system Beam specimen Power amplifier Figure 10. Experimental setup for the PVDF sensor system. of 15  10  0.25 mm, and are attached near the cantilever-supporting location of the beam. PVDF SENSOR ARRAY SYSTEM The PVDF sensor films, from MSI (Measurement Specialties Inc.), are 30 mm (1.19 in.) in length, 12 mm (0.484 in.) in width, and have a thickness of 28 mm. The beam sample is divided into 19 sensor locations (Figure 9) to best accommodate the films. Each point is aligned with the center of the PVDF film during testing. The PVDF sensor locations are illustrated in Figure 9. The experiment is first conducted for the intact (undamaged) specimen with the cantilevered boundary conditions, followed by performing similar tests on damaged beam specimens. Continuous-sweep sine excitation generated by a waveform generator with a magnitude of 140 V is run through the PZT actuators to excite the beams, and the responses are recorded by a dSPACE data acquisition system as time domain responses. The experimental setup of the dynamic testing is presented in Figure 10. The dynamic responses are measured by roving the position of the sensor at 19 locations along the beam length as shown in Figure 9. The measured time domain data is then post-processed to construct the input data for modal analysis calculation, which is performed in commercial software ME-Scope. The curvature mode shapes and the corresponding natural frequencies are acquired after processing the data in the ME-Scope. This data is then further post-processed using MATLAB and Microsoft Excel to evaluate the modal parameters and the aforementioned damage detection algorithms are employed to locate the damage. The above-mentioned experimental procedure is summarized in the form of a flowchart, as presented in Figure 11. SCANNING LASER VIBROMETER (SLV) SYSTEM On the other hand, SLV is a noncontact measurement technique that has the capability of handling a large number of measurement points and at the same time eliminates the problem with wiring management of conventional or surface-bonded sensor systems. The measurement capabilities of SLV, such as sensitivity, accuracy, and reduced Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1234 P. QIAO ET AL. Experimental setup dSPACE data acquisition system Experimentation Processing data using MATLAB Data collection ME-Scope Calculation of modal parameters (natural frequencies and mode shapes) Employment of damage detection algorithms Damage evaluation Figure 11. Schematic of experimental procedure. intrusivity, allow such systems to establish themselves as an important diagnostic instrument in structural health monitoring and damage detection. A PSV 400 SLV from Polytec is used to obtain the dynamic response, which is later used for damage detection in the composite laminated beams. The same excitation source and signal as the vibration testing with the PVDF sensor system are applied in this testing. Each testing specimen is scanned using almost 530 scanning points. The experimental setup for laser vibrometry testing is shown in Figure 12. The PSV 400 measures the 2-D distribution of vibration velocities on the basis of laser interferometry. The main system components of PSV 400 are controller, junction box, scanning head, and the data management system. The junction box is the central connection between the system components and provides the interfaces for peripheral devices. The scanning head consists of the interferometer, the scanner mirrors to deflect the laser beam, and a video camera to visualize the measurement object. The measurement data is digitally recorded in the data management system. The PSV software controls the data acquisition and offers user-friendly functions to evaluate the measurement data. The obtained responses are the natural frequencies and the displacement mode shapes. The curvature mode shapes are calculated by using the central difference method of Equation (1), similar to the approach applied for the numerical FE analysis. The SLV sensor system provides the output in the form of displacement mode shapes, which is sorted to acquire the displacement mode shape pertaining to the centerline of the beam (Figure 13) and is used in further evaluation. The displacement mode shapes are further processed using the central difference method (Equation (1)) to arrive at the curvature mode shapes. The curvature mode shapes derived from the measured displacement mode shapes are very sensitive to slight measurement discrepancies and noise. The data shows lots of undulations owing to the local and global errors accumulated Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1235 Dynamics-based Damage Detection of Composite Laminated Beams (a) Laser beam Reference sensor Actuator Beam specimen PSV400 scanning head Controller/computer system (b) Figure 12. Experimental setup for the SLV sensor system: (a) experimental measurement using the SLV system and (b) laboratory testing using the SLV system. in the calculation of the second derivative of displacements. To minimize these calculation errors, the filtering process of the calculated curvature mode shapes is performed [19]. The Savitzky–Golay smoothing filters, also called digital smoothing polynomial filters or least-squares smoothing filters, are used to ‘smooth out’ a noisy signal whose frequency span (without noise) is large. The Savitzky–Golay filters are optimal, i.e., they minimize the least-squares error in fitting a polynomial to frames of noisy data. Experimental Results The extracted modal frequencies of the healthy and damaged composite beams for the first three bending modes are listed in Table 5. However, as the quality of the first mode Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1236 P. QIAO (a) 37 0.3302 m ET AL. 43 0.0508 m 0.5588 m (b) 27 39 0.2540 m 0.1016 m 0.5588 m (c) 53 0.3810 m 59 0.0508 m (d) 48 0.001 m 0.2794 m 0.5588 m (e) 22 0.1778 m 28 0.0508 m 0.5588 m Figure 13. Scanning array and damage configurations for the SLV system: (a) Delamination A; (b) Delamination B; (c) Delamination C; (d) Saw-cut damage; and (e) Impact damage. Table 5. Natural frequencies obtained from experimental analysis. PVDF – f (Hz) SLV – f (Hz) Specimen type Second mode Third mode Fourth mode Second mode Third mode Fourth mode Healthy Delamination A Delamination B Delamination C Saw-cut Impact 61.75 58.13 55.00 59.10 57.75 59.30 158.20 150.60 148.95 151.63 155.00 152.50 316.90 309.35 302.10 305.15 311.60 308.80 59.00 56.17 53.20 55.75 54.00 56.50 155.20 149.15 146.90 150.30 151.00 149.55 318.25 302.25 297.10 308.60 311.60 305.20 shape obtained from experimental analysis results is not good, the three consecutive mode shapes starting from the second mode are presented and considered for further analysis of the results. The presence of damage or deterioration in the structure causes changes in the natural frequencies of the structure. The existence of damage in a section of the beam is equivalent Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1237 Dynamics-based Damage Detection of Composite Laminated Beams (a) (b) Mode 2 Mode 3 Mode 4 0.6 0.4 Curvature Curvature 0.4 Mode 2 Mode 3 Mode 4 0.6 0.2 0.0 −0.2 0.2 0.0 −0.2 0 5 10 15 20 0 5 Sensor location (c) 20 15 20 15 20 Mode 2 Mode 3 Mode 4 0.6 0.4 0.4 Curvature Curvature 15 (d) Mode 2 Mode 3 Mode 4 0.6 10 Sensor location 0.2 0.0 0.2 0.0 −0.2 −0.2 −0.4 0 5 10 15 20 0 5 Sensor location (e) (f) Mode 2 Mode 3 Mode 4 0.6 Mode 2 Mode 3 Mode 4 0.6 0.4 0.4 Curvature Curvature 10 Sensor location 0.2 0.0 0.2 0.0 −0.2 −0.2 −0.4 0 5 10 Sensor location 15 20 0 5 10 Sensor location Figure 14. Curvature mode shapes of the specimens obtained from the PVDF sensor system: (a) Healthy – curvature–PVDF; (b) Delamination A – curvature – PVDF (11–13); (c) Delamination B – curvature – PVDF (8–12); (d) Delamination C – curvature – PVDF (13–15); (e) Saw-cut – curvature – PVDF (9); and (f) Impact damage – curvature – PVDF (5–7). to a reduction in the second moment of area of the beam. This leads to a reduction in the local bending stiffness at that cross section. The consequence of reduced local bending stiffness in lowering the values of the natural frequencies in bending is justified by the results in Table 5. The observation of the results shown in Table 5 indicates that the natural frequency of the beam decreases with the increase in the damage (e.g., delamination). Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1238 P. QIAO (a) (b) Mode 2 Mode 3 Mode 4 0.3 Mode 2 Mode 3 Mode 4 0.3 0.2 Curvature Displacement 0.2 0.1 0.0 ET AL. 0.1 0.0 −0.1 −0.1 −0.2 −0.2 0 (c) 10 20 30 40 50 Sensor location 60 70 (d) 0.2 40 60 0.3 Mode 2 Mode 3 Mode 4 0.2 Curvature Displacement 20 Sensor location Mode 2 Mode 3 Mode 4 0.3 0 0.1 0.0 0.1 0.0 −0.1 −0.1 −0.2 −0.2 0 (e) 10 20 30 40 50 Sensor location 60 70 40 60 (f) Mode 2 Mode 3 Mode 4 0.3 0.2 0.2 Curvature Displacement 20 Sensor location Mode 2 Mode 3 Mode 4 0.3 0 0.1 0.0 0.1 0.0 −0.1 −0.1 −0.2 0 10 20 30 40 50 Sensor location 60 70 −0.2 0 20 40 60 Sensor location Figure 15. Displacement and curvature mode shapes of the specimens obtained from the SLV measurements and numerical derivation, respectively: (a) Healthy – displacement – SLV; (b) Healthy – curvature – SLV; (c) Delamination A – displacement – SLV (37–43); (d) Delamination A – curvature – SLV (37–43); (e) Delamination B – displacement – SLV (27–39); (f) Delamination B – curvature – SLV (27–39); (g) Delamination C – displacement – SLV (53–59); (h) Delamination C – curvature – SLV (53–59); (i) Saw-cut damage – displacement – SLV (48); (j) Saw-cut damage – curvature – SLV (48); (k) Impact damage – displacement – SLV (22–28); and (l) Impact damage – curvature – SLV (22–28). RESULTS FROM THE PVDF SENSOR SYSTEM The PVDF sensor system, which has the potential to be implemented as an efficient, automatic, real-time, and onboard measurement technique, acquires the curvature mode shapes directly. For comparison, the mode shapes are first normalized by evaluating their root mean square. Figure 14 shows the curvature mode shapes for all the testing specimens, i.e., the healthy and the damaged ones. The actual locations of Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1239 Dynamics-based Damage Detection of Composite Laminated Beams (g) (h) Mode 2 Mode 3 Mode 4 0.3 0.2 Curvature Displacement 0.2 Mode 2 Mode 3 Mode 4 0.3 0.1 0.0 0.1 0.0 −0.1 −0.1 −0.2 −0.2 0 10 20 30 40 50 60 70 0 20 Sensor location (i) Mode 2 Mode 3 Mode 4 0.3 0.2 Curvature Displacement 0.2 60 (j) Mode 2 Mode 3 Mode 4 0.3 40 Sensor location 0.1 0.0 0.1 0.0 −0.1 −0.1 −0.2 0 10 20 30 40 50 60 −0.2 70 0 20 Sensor location (k) Mode 2 Mode 3 Mode 4 0.3 0.2 0.2 Curvature Displacement 60 (l) Mode 2 Mode 3 Mode 4 0.3 40 Sensor location 0.1 0.0 0.1 0.0 −0.1 −0.1 −0.2 0 10 20 30 40 50 60 70 −0.2 0 Sensor location 20 40 60 Sensor location Figure 15. Continued. the damage, represented by the number of sensor locations, are given in parentheses. As the curvature mode shapes are the second derivative of the displacement, it is observed that the curvature mode shapes provide a relatively clear indication of the presence of damage. RESULTS FROM THE SCANNING LASER VIBROMETER (SLV) SYSTEM Both the measured displacement mode shapes and the derived curvature mode shapes are presented in Figure 15. Again, the actual locations of the damage, represented by the number of sensor locations (scanning points), are given in parentheses. As shown in Figure 15, the locations of various damages cannot be proved by simply examining the displacement mode shapes; whereas they are discernable in the derived curvature mode shapes. These derived curvature mode shapes from the displacement mode shapes Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1240 (a) P. QIAO (b) 1.0 ULS GSM 1.0 ULS GSM 0.8 0.6 ULS/GSM ULS/GSM 0.8 0.4 0.2 0.0 0.6 0.4 0.2 0.0 −0.2 0 20 40 60 −0.2 80 0 20 Sensor location (c) ET AL. 0.8 40 60 80 Sensor location (d) ULS GSM 0.6 ULS GSM 0.5 0.6 ULS/GSM ULS/GSM 0.4 0.4 0.2 0.3 0.2 0.1 0.0 0.0 −0.2 0 20 40 60 0 80 50 (e) 150 200 250 1.0 ULS GSM 0.8 ULS/GSM 100 Sensor location Sensor location 0.6 0.4 0.2 0.0 0 50 100 150 Sensor location Figure 16. Results of ULS and GSM calculation based on numerical FE analysis: (a) Delamination A – FE (44–56); (b) Delamination B – FE (38–58); (c) Delamination C – FE (62–72); (d) Saw-cut damage – FE (150); and (e) Impact damage – FE (65–75). are later used as inputs in all the aforementioned damage detection algorithms for further damage evaluation. APPLICATION OF DAMAGE DETECTION ALGORITHMS Based on the curvature mode shapes obtained from either the PVDF or SLV sensor systems, ULS is calculated. In this study, the first three mode shapes beginning from the second mode (i.e., the second, third, and fourth modes) are considered. Even the damage Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1241 Dynamics-based Damage Detection of Composite Laminated Beams (a) (b) 1.0 ULS GSM ULS GSM 0.8 0.6 ULS/GSM ULS/GSM 0.8 1.0 0.4 0.6 0.4 0.2 0.2 0.0 0.0 0 5 10 15 20 5 Sensor location (c) (d) 1.2 1.0 ULS GSM 10 Sensor location 15 10 15 1.0 ULS GSM 0.8 ULS/GSM ULS/GSM 0.8 0.6 0.4 0.2 0.6 0.4 0.2 0.0 0.0 5 10 Sensor location (e) 5 20 Sensor location 1.0 ULS GSM 0.8 ULS/GSM 0 15 0.6 0.4 0.2 0.0 0 5 10 15 20 Sensor location Figure 17. Results of ULS and GSM calculation based on piezoelectric (PVDF) sensor data: (a) Delamination A – PVDF (11–13); (b) Delamination B – PVDF (8–12); (c) Delamination C – PVDF (13–15); (d) Saw-cut damage – PVDF (9); and (e) Impact damage – PVDF (5–7). location can be detected in the experimentally based curvature mode shapes directly (Figures 14 and 15 for the PVDF and SLV systems, respectively), the damage indications are not very obvious on the curves and sometimes the exact locations are difficult to discern. To further facilitate the damage detection effort, damage location is determined based on the aforementioned ULS technique and the three damage detection algorithms (i.e., GSM, GFD, and SEM). The detection results based on the numerical data are first presented, followed by the results using the experimental data. To indicate the validity of the detection results, the actual locations of damage in terms of sensor points (PVDF), scanning points (SLV), or node points (FE) are given in parentheses. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1242 (a) P. QIAO (b) 1.0 ULS GSM 0.6 0.4 0.2 0.0 10 20 30 40 50 Sensor location 60 0.4 0.2 70 0 (d) 1.0 ULS GSM 0.8 0.6 0.4 10 20 30 40 50 Sensor location 60 70 20 30 40 50 Sensor location 60 70 1.0 ULS GSM 0.8 ULS/GSM ULS/GSM 0.6 −0.2 0 0.6 0.4 0.2 0.2 0.0 0.0 −0.2 ULS GSM 0.0 −0.2 (c) 1.0 0.8 ULS/GSM ULS/GSM 0.8 ET AL. −0.2 0 10 20 30 40 50 Sensor location (e) 70 0 10 1.0 ULS GSM 0.8 ULS/GSM 60 0.6 0.4 0.2 0.0 −0.2 0 10 20 30 40 50 Sensor location 60 70 Figure 18. Results of ULS and GSM calculation based on SLV system data: (a) Delamination A – SLV (37–43); (b) Delamination B – SLV (27–39); (c) Delamination C – SLV (53–59); (d) Saw-cut damage – SLV (48); and (e) Impact damage – SLV (22–28). Uniform Load Surface (ULS)/Gapped Smoothing Method (GSM) The results from the GSM based on the numerical FE analysis data, as well as the approximated ULS, are presented in Figure 16. The results based on the experimental data from the PVDF sensor and SLV measured data are presented in Figures 17 and 18, respectively. The GSM results acquired using the ULS curvatures show the location of damage in the form of distinct peaks or discontinuities in the curve. However, the results from experimentally measured data also show multiple peaks at locations other than the actual Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1243 Dynamics-based Damage Detection of Composite Laminated Beams (a) (b) 3.0 2.5 2.5 GFD GFD 2.0 2.0 1.5 1.5 1.0 1.0 0 (c) 10 20 30 40 50 60 Sensor location 70 80 0 90 (d) 2.4 10 20 30 40 50 Sensor location 60 70 4 2.2 3 1.8 GFD GFD 2.0 1.6 2 1.4 1 1.2 1.0 0.8 0 0 20 40 60 Sensor location (e) 80 0 50 100 150 200 Sensor location 250 8 GFD 6 4 2 0 0 50 100 150 Sensor location Figure 19. Results of GFD calculation based on numerical FE analysis data: (a) Delamination A – FE (44–56); (b) Delamination B – FE (38–58); (c) Delamination C – FE (62–72); (d) Saw-cut damage – FE (150); and (e) Impact damage – FE (65–75). damage, which challenges the accuracy and efficiency of the method in determining the exact location of damage. The results from the piezoelectric (PVDF) sensor have broad peaks due to the coarse sensor grid. However, it has fewer small undulations in other places than in the damage location. On the contrary, the results from SLV measurements have sharp discontinuities, typically at the boundaries of the damages. The undulation at other places rather than at the damage location is quite significant (about one third of the peak at the damage location), and this occurs in a few places. This small discontinuity may Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1244 (b) 6 5 4 4 3 3 2 2 1 1 0 5 10 15 20 0 5 Sensor location (d) 6 5 4 4 3 2 1 1 10 15 20 0 5 Sensor location (e) 20 15 20 3 2 5 15 6 5 0 10 Sensor location GFD GFD (c) ET AL. 6 5 GFD GFD (a) P. QIAO 10 Sensor location 6 GFD 5 4 3 2 1 0 5 10 15 20 Sensor location Figure 20. Results of GFD calculation based on piezoelectric (PVDF) sensor data: (a) Delamination A – PVDF (11–13); (b) Delamination B – PVDF (8–12); (c) Delamination C – PVDF (13–15); (d) Saw-cut damage – PVDF (9); and (e) Impact damage – PVDF (5–7). be magnified during the derivative calculation process, because the SLV-based curvature mode shapes are obtained from the measured displacement mode shapes. Generalized Fractal Dimension (GFD) The results from the GFD algorithm using the ULS curves based on the numerical analysis data are presented in Figure 19. The results based on the experimental data Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1245 Dynamics-based Damage Detection of Composite Laminated Beams (a) (b) 3.5 3.5 3.0 3.0 GFD 2.5 GFD 4.0 2.0 2.5 2.0 1.5 1.5 1.0 1.0 0 10 20 30 40 50 60 70 0 10 20 Sensor location (d) 3.0 2.5 2.0 2.0 1.5 1.5 1.0 1.0 0 10 20 30 40 50 60 70 0 10 20 Sensor location (e) 40 50 60 70 60 70 3.0 2.5 GFD GFD (c) 30 Sensor location 30 40 50 Sensor location 3.0 GFD 2.5 2.0 1.5 1.0 0 10 20 30 40 50 60 70 Sensor location Figure 21. Results of GFD calculation based on SLV system data: (a) Delamination A – SLV (37–43); (b) Delamination B – SLV (27–39); (c) Delamination C – SLV (53–59); (d) Saw-cut damage – SLV (48); and (e) Impact damage – SLV (22–28). from the piezoelectric (PVDF) sensor and SLV measured data are presented in Figures 20 and 21, respectively. Although the FD curves indicate the location of damage and do not require the smooth curve fitting as in the GSM, the method has the abnormality of highlighting the regions of higher curvature with peaks other than those peaks at the location of damage when experimentally measured data (either PVDF or SLV) are used. Furthermore, the extra peaks in the results based on the measured SLV data have the same level as the peaks at the damage location and can easily be mistaken as other damage. The effect of numerical derivation to obtain curvature modes is amplified in the GFD algorithm. Although the Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1246 (b) 0.10 0.08 0.06 0.06 0.04 (c) 0.04 0.02 0.02 0.00 0.00 0 10 20 30 40 50 60 Sensor location 70 80 0 90 (d) 0.4 ET AL. 0.10 0.08 SEM SEM (a) P. QIAO 10 20 30 40 50 60 Sensor location 70 80 90 0.10 0.08 0.3 SEM SEM 0.06 0.2 0.04 0.1 0.02 0.0 0.00 0 10 20 30 40 50 60 Sensor location (e) 70 80 90 0 50 100 150 200 250 Sensor location 0.010 0.008 SEM 0.006 0.004 0.002 0.000 0 50 100 Sensor location 150 Figure 22. Results of SEM calculation based on numerical FE analysis data: (a) Delamination A – FE (44–56); (b) Delamination B (38–58); (c) Delamination C – FE (62–72); (d) Saw-cut damage – FE (150); and (e) Impact damage – FE (65–75). derivation procedure already includes the filtering process, calculating the curvature from the second derivative of experimentally measured displacement data is still a challenge. In addition, the scale parameter ‘S’ in the GFD algorithm [Equations (12)–(14)] involves some trial and error in order to achieve the results in the desired scale ratios. Strain Energy Method (SEM) The results from the SEM algorithm using the ULS from the numerical analysis data are presented in Figure 22, while the results based on experimental data, from the piezoelectric (PVDF) sensor and SLV measured data are presented in Figures 23 and 24, respectively. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1247 Dynamics-based Damage Detection of Composite Laminated Beams (a) (b) 0.020 0.16 0.14 0.12 0.015 SEM SEM 0.10 0.010 0.08 0.06 0.005 0.04 0.02 0.000 0.00 0 5 10 15 20 0 5 Sensor location 0.025 (d) 15 20 15 20 0.05 0.020 0.04 0.015 0.03 SEM SEM (c) 10 Sensor location 0.010 0.02 0.01 0.005 0.00 0.000 0 5 10 15 20 0 5 (e) 10 Sensor location Sensor location 0.030 0.025 SEM 0.020 0.015 0.010 0.005 0.000 0 5 10 15 20 Sensor location Figure 23. Results of SEM calculation based on piezoelectric (PVDF) sensor data: (a) Delamination A – PVDF (11–13); (b) Delamination B – PVDF (8–12); (c) Delamination C – PVDF (13–15); (d) Saw-cut damage – PVDF (9); and (e) Impact damage – PVDF (5–7). The identification results based on the modal strain energy shows more irregularity in the SEM curve. Although, the SEM algorithm indicates the presence and location of damage (with large peaks at the beginning and end of the damage), multiple small peaks at other locations near the actual damage invokes doubt regarding the actual damage location and small imperfections present in the corresponding locations. SUMMARY AND DISCUSSION Comparisons of the estimated damage location using the three proposed algorithms (i.e., GSM, GFD, and SEM) from each system methods, for the numerical FE analysis data, the piezoelectric PVDF sensor data, and the SLV system data are Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1248 P. QIAO (b) 0.008 0.008 0.006 0.006 SEM SEM (a) ET AL. 0.004 0.004 0.002 0.002 0.000 0.000 0 10 20 30 40 50 60 0 70 10 20 30 40 50 60 70 60 70 Sensor location Sensor location (d) (c) 0.008 0.006 0.006 SEM SEM 0.004 0.004 0.002 0.002 0.000 0.000 0 10 20 30 40 50 60 0 70 10 20 (e) 30 40 50 Sensor location Sensor location 0.008 SEM 0.006 0.004 0.002 0.000 0 10 20 30 40 50 60 70 Sensor location Figure 24. Results of SEM calculation based on SLV system data: (a) Delamination A – SLV (37–43); (b) Delamination B – SLV (27–39); (c) Delamination C – SLV (53–59); (d) Saw-cut damage – SLV (48); and (e) Impact damage – SLV (22–28). Table 6. Comparison of estimated and actual damage locations – FE. Estimated damage locations Damage Delamination A Delamination B Delamination C Saw-cut Impact Actual damage location GSM GFD SEM 44–56 38–58 62–72 150 65–75 44–56 38–59 62–72 150 65–75 40–58 36–60 58–74 150 55–75 44–56 40–58 62–72 150 65–75 Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1249 Dynamics-based Damage Detection of Composite Laminated Beams Table 7. Comparison of the actual and estimated damage locations –PVDF. Estimated damage locations Damage Actual damage location Delamination A Delamination B Delamination C Saw-cut Impact 11–13 8–12 13–15 9 5–6 GSM 12 8–12 14 9 6 GFD 11 8–12 11–15 9 5 SEM 12 8–12 14 8–9 6 Table 8. Comparison of the actual and estimated damage locations – SLV. Estimated damage locations Damage Delamination A Delamination B Delamination C Saw-cut Impact Actual damage location GSM GFD SEM 37–43 27–39 53–59 48 22–28 38–44 25–38 55–58 48 21–28 32–45 26–46 50–60 45 28–32 38–44 25–39 48–56 48 21–38 summarized in Tables 6–8, respectively. In all three damage detection algorithms, the ULS data are used to obtain the respective damage parameters. The data of healthy structures are not required when all three algorithms are implemented. The detection based on numerical FE analysis data shows good indication of the damage location and size of the damage for all three damage detection algorithms (Table 6), thus verifying the validity of the algorithms in damage detection implementation. It validates the promise that all three proposed algorithms are capable of performing the damage detection analysis if the validated input data are provided. As shown in Table 6 and also in Figures 16 (GSM), 19 (GFD), and 22 (SEM), damage detection from the FE data based on GSM and SEM seems more accurate and effective than the one from GFD. However, GFD does not involve the processing of smooth curve fitting; while both GSM and SEM have similar algorithms by comparing them with the smooth curve fits, which are assumed to be the healthy ones. While in the damage detection using the data measured from the piezoelectric (PVDF) sensor system, relatively smooth curves with few peaks (Figures 17 (GSM), 20 (GFD), and 23 (SEM)) are produced for all three algorithms, because of the limited number of PVDF sensors used (a total of 19 PVDF sensors along the beam length are used in the experiment, Figure 9). Since the PVDF sensor acquires the curvature over the sensor length (about 38.1 mm (1.5 in.) in length), the measured data are averaged in nature, thus smoothing out some local effect over the sensor length. But the PVDF sensor system is still viable in damage detection with one or two large peaks over the damage location, and it has the potential of performing real-time and onboard structural monitoring by embedding or surface-bonding the piezoelectric sensors in or on the structure. For the SLV sensor systems, denser scanning grids (sensor points) are employed (Figure 13), compared to those in the PVDF sensor system. This grid measures the local displacement at the particular point, thus acquiring the displacement mode shapes. In order to implement the damage detection algorithms with the SLV measured data, the central difference method (see Equation (1)) is used to obtain the curvature mode Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1250 P. QIAO ET AL. shape from the measured displacement mode shapes. As shown in Shah [19], the implementation of the damage detection algorithms directly with the measured displacement mode shapes from SLV does not warrant the damage detection. As shown in Figures 18 (GSM), 21 (GFD), and 24 (SEM), more peaks and irregularities are observed for all three damage detection methods, and they may be caused by the refined scanning grids used in the experiment (Figure 13). Even though all three algorithms locate the damage relatively well, additional peaks in the damage parameter curves may promote false detection of the damage. In particular, more sharp peaks are observed in the GFD curve (Figure 21), indicating that the GFD algorithm based on the SLV measured data may not be capable of uniquely picking up the actual damage location. While for the GSM and SEM, multiple peaks appear, the peaks at the actual damage location are significantly larger than the ones from the non-damage locations, and are capable of locating the damage. In summary, with the validation of the proposed three damage detection algorithms based on numerical FE analysis, it is indicated that all the algorithms using the ULS curvature curve are capable of detecting the presence and size (e.g., the delamination length) of the damage if the validated mode shape data is available. Two sensor systems (PVDF vs. SLV) are both capable of acquiring such validated mode shape data useful for damage detection analysis. The results (Tables 6–8 and Figures 16–24) indicate that the detection based on GSM and SEM provides better outcome than the one based on GFD. However, GFD eliminates the necessity of smooth curve fitting as is required in GSM and SEM. CONCLUSIONS The present study focuses on the comparison of two sensor systems (i.e., PVDF sensors and SLV system) in providing reliable measured dynamic response data for damage detection purposes and verifies the validity of three damage detection algorithms. Two sensor systems of contact (by surface-bonding of PVDF sensors) and noncontact (by laser scanning of the specimens) are used to acquire the dynamic response (e.g., frequency and mode shapes). The numerical FE modal analysis is conducted to show the validity of the proposed damage detection algorithms. The experimental program comprises testing of six E-glass–epoxy composite beams with different damage configurations (i.e., delamination, saw-cut, and impact damage). The ULS curvature is used in this study to minimize the effect of data truncation effects and measurement error, as well as the selection of curvature mode shapes, because the use of individual curvature mode shapes may yield misleading results in detecting the location of damage. Once the ULS curvatures are obtained by the numerical FE analysis, PVDF, or SLV systems, they are implemented in the three damage detection algorithms (i.e., GSM, GFD, and SEM). In general, the damage identification results using the SLV measured data have lots of peaks or significant undulations that may cause a misunderstanding in locating damage and can be mistaken for a multiple-damage case. The probable sources of this irregularity may be from the numerical process in estimating the curvature mode shapes from measured displacement mode shapes or the inherent imperfection in the composite beams. Improvement in this process by applying suitable approximation algorithms that includes a proper filtering process will improve the detection results. However, the noncontact SLV Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 Dynamics-based Damage Detection of Composite Laminated Beams 1251 system enjoys the advantages such as simplicity in operation, ability to account for a large number of measurement points, and absence of wiring management problems. The damage detection using the SLV system is valid for the composite laminated beams used in this study because the peaks of damage parameters at the actual damage location are significantly larger than those shown at the nondamage locations. On the other hand, the identification process based on the PVDF sensor measurement shows less irregularity and has only one or two major peaks indicating the location of the damage. Due to the relatively coarse sensor grid, the damage identification parameters cover a large area (broad peak over the sensor length) that may miss the actual area of the damage locations. However, the PVDF sensor system has a great potential of being implemented as an automatic, real-time and onboard structural health monitoring technique. Successful implementation of damage detection algorithms and two sensor systems for damage detection of composite laminated beams demonstrates that the dynamics-based damage detection approach using the two sensor systems is a viable technology for structural health monitoring of composite structures. ACKNOWLEDGMENTS This study is partially supported by the Air Force Office of Scientific Research (AFOSR) (FA9650-04-C-0078), and their financial support is gratefully acknowledged. The assistance provided by Kan Lu and Luyang Shan on experimental testing and specimen fabrication is highly appreciated. REFERENCES 1. Zou, Y., Tong, L. and Steven, G.P. (2000). Vibration Based Model Dependent Damage (delamination) Identification and Health Monitoring for Composite Structures – a Review. Journal of Sound and Vibration, 230(2): 357–378. 2. Valdes, S.H.D. and Soutis, C. (1999). Delamination Detection in Composite Laminates from Variations of their Modal Characteristics, Journal of Sound and Vibration, 228(1): 1–9. 3. Pandey, A.K. and Biswas, M. (1991). Damage Detection in Structures using Changes in Flexibility, Journal of Sound and Vibration, 169(1): 3–17. 4. Sampaio, R.P.C., Maia, N.M.M. and Silva, J.M.M. (1999). Damage Detection using the Frequency Response Function Curvature Method, Journal of Sound and Vibration, 226(5): 1029–1042. 5. Wahab, M.M.A. and Roeck, G.D. (1999). Damage Detection in Bridges Using Modal Curvatures: Application to a Real Damage Scenario, Journal of Sound and Vibration, 226(2): 217–235. 6. Hamey, C.S., Lestari, W., Qiao, P. and Song, G. (2004). Experimental Damage Identification of Carbon/epoxy Composite Beams using Curvature Mode Shapes, Structural Health Monitoring, 3(4): 333–353. 7. Farrar, C.R. and Jauregui, D.A. (1997). Comparative Study of Damage Identification Algorithms Applied to a Bridge, Engineering Analysis Group, Los Alamos National Laboratory, Los Alamos, NM, pp. 720–731. 8. Kim, J.T., Ryu, Y.S., Hyun, M.C. and Stubbs, N. (2003). Damage Identification in Beam-type Structures: Frequency Based Method v/s Mode-shape Based Method, Engineering Structures, 25(1): 57–67. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013 1252 P. QIAO ET AL. 9. Yoon, M.K., Heider, D., Gillespie, J.W., Ratcliffe, C.P., Jr and Crane, R.M. (2005). Local Damage Detection using the Two-dimensional Gapped Smoothing Method, Journal of Sound and Vibration, 279(1–2): 119–139. 10. Cornwell, P., Doebling, S.W., and Farrar, C.R. (1999). Application of the Strain Energy Damage Detection Method to Plate like Structures, Journal of Sound and Vibration, 224(2): 359–374. 11. Duffey, T.A., Doebling, S.W., Farrar, C.R., Baker, W.E. and Rhee, W.H. (2001). Vibrationbased Damage Identification in Structures Exhibiting Axial and Torsional Response, Journal of Vibration and Acoustics, 123: 84–91. 12. Zhang, Z. and Akhtan, A.E. (1998). Application of Modal Flexibility and its Derivatives in Structural Damage Identification, Research in Non-destructive Evaluation, 10: 43–61. 13. Wu, D. and Law, S.S. (2003). Damage Localization in Plate Structures from Uniform Load Surface Curvature, Journal of Sound and Vibration, 276(1): 227–244. 14. Gao, Y., Spencer, B.F., Jr. and Bernal, D. (2004). Experimental Verification of the Damage Locating Vector Method, Journal of Engineering Mechanics, 128(1): 7–14. 15. Lu, Q., Ren, G. and Zhao, Y. (2001). Multiple Damage Location with Flexibility Curvature and Relative Frequency Change for Beam Structures, Journal of Sound and Vibration, 253(5): 1101–1114. 16. Wang, J. and Qiao, P. (2006). Improved Damage Detection of Beam Type Structures using Uniform Load Surface, Structural Health Monitoring (in press). 17. Ratcliffe, C.P. and Bagaria, W.J. (1998). A Vibration Technique for Locating Delamination in a Composite Beam, AIAA Journal, 36: 1074–1077. 18. Katz, M.J. (1988). Fractals and the Analysis of Waveforms, Computers in Biology and Medicine, 18(3): 145–156. 19. Shah, M.G. (2005). Dynamics-based Damage Detection of Beam-type Structures using Piezoelectric Sensors and Scanning Laser Vibrometer, MS Thesis, Department of Civil Engineering, The University of Akron, Akron, OH. Downloaded from jcm.sagepub.com at UNIV ALABAMA LIBRARY/SERIALS on June 18, 2013