University
stringclasses
19 values
Text
stringlengths
458
20.7k
ADE
Chapter 2 surface of the one-side water-immersed plate, which is excited by a circular piezoceramic transducer, is dominated by quasi-Scholte waves. The accuracy of the simulation results has been verified by comparing the phase velocities with the theoretical dispersion curves as well as the experimental measurements. The numerical simulations of wave scattering of the quasi-Scholte mode at a circular blind hole have been compared with the experimental measurements. A good agreement has been observed between the FE simulations and experimental measurements. It has been concluded the 3D FE model is able to accurately simulate quasi-Scholte wave propagation and its scattering characteristics for non-regular geometries. Further numerical studies have demonstrated that the scattering directivity patterns (SDPs) depend on both the diameter and the depth of the circular blind hole. At a given depth of the damage, the amplitudes of the backward scattered waves are comparable to the forward scattering amplitudes for small values of RDW. For larger RDW, the forward scattering amplitudes increase quickly with slight variation while the backward scattering magnitudes fluctuate following a sinusoidal pattern with the overall trend being a slow increase. In general, the forward scattered waves are more suitable to be used for identifying the size of the damage since they have larger amplitudes and follow a relatively simple scattering pattern. For the local damage of the same diameter, the forward and backward scattering amplitudes increase with the depth of the damage. The backward scattering amplitudes increase faster than the forward scattered waves. Also, for the directions perpendicular to the incident wave, the scattering amplitudes are weak for damage whose depth is less than half of the plate thickness but significantly increase for deeper damage. Finally, this study has provided a comprehensive investigation of the scattering phenomena due to low-frequency quasi-Scholte waves interacting with a circular blind hole. The findings of this study can be used to provide a guide on selecting appropriate excitation frequencies, guided wave modes, and transducer locations, and hence, it will help to improve the performance of in-situ damage detection techniques for structures exposed to the corrosive environment. 35
ADE
Chapter 3 Chapter 3. Numerical and experimental investigations on mode conversion of guided Waves in partially immersed plates Abstract This paper numerically and experimentally investigates the guided wave propagation in a steel plate with one side partly exposed to water. The fundamental anti-symmetric Lamb wave (A ) is excited on the dry plate section and travels to 0 the water-immersed plate section, where the generated A wave is mode converted 0 to the quasi-Scholte (QS) wave. The results demonstrate that the energy of QS wave converted by the A wave decreases when the excitation frequency increases. In 0 addition, it is revealed that the guided wave energy can shift in the frequency domain if the phase velocity of the incident A wave is larger than the sound speed 0 of water. The frequency shift phenomenon should be noticed in practical applications because the behaviors of guided waves vary with frequency. Finally, discussions are provided on the frequency selection for exciting guided waves to detect damage on partially immersed structures and assess liquid properties. Keywords: Quasi-Scholte waves; Lamb waves; leaky guided waves; mode conversion; submerged structures. 39
ADE
Chapter 3 3.2. Introduction Ultrasonic guided waves are elastic waves that travel along the boundary of a structure and have been widely used for identifying damage in structures [1-5], detecting debonding in adhesively bonded structures [6-9], sensing liquid levels and properties [10-12], and assessing coatings on the substrate surface [13, 14]. The advantages of ultrasonic guided waves are that they can propagate for a long distance, enabling an efficient large-area inspection. Lamb waves are guided waves in thin-walled structures, such as plates, shells, and pipes. When the plate is surrounded by air, Lamb waves are composed of multiple symmetric and anti- symmetric wave modes [15]. When one or both sides of the plate are exposed to liquid, there is a substantial increase in the energy leakage into the surrounding liquid medium [16, 17]. Therefore, guided waves in immersed plates are called leaky Lamb waves and they behave differently from their counterparts in the plates surrounded by air. In addition to the symmetric and antisymmetric leaky Lamb wave modes, there is an interface mode called quasi-Scholte (QS) wave in the plate immersed in liquid [18]. Recently, studies have been conducted on the QS wave for a wide range of applications because of its ability to propagate along the plate-fluid interface over a long distance and high sensitivity to changes in the properties of both the plate and fluid. Tietze et al. [19] experimentally demonstrated that QS wave propagating at the interface between electrode and electrolyte is able to remove the diffusion boundary layer, which can be employed to accelerate the electrochemical process. Aubert et al. [20] invented a low-cost fluid manipulation device that employed the generation of QS wave to sort living cells and separate plasma from a blood microdroplet. Through schlieren imaging, the acoustic fields of the QS wave were experimentally visualized to be evanescent in the direction normal to the plate surface, which is a promising characteristic for microfluidic applications. Hayashi and Fujishima [21] experimentally confirmed that QS wave could be excited by applying the normal vibration directly on the surface of a plate loaded with water. The generated QS wave was shown to be sensitive to the change of the physical conditions on the plate surface. Thus, the QS wave is feasible for non-destructive testing (NDT) of water-filled storage tanks and pipes. 41
ADE
Chapter 3 There are other studies focused on the application of QS wave, in which the QS wave is excited by mode conversion from the fundamental anti-symmetric mode (A ) of Lamb waves. Cegla et al. [18] developed a novel method for sensing fluid 0 property by exciting guided waves on a plate that was partially immersed in the fluid. A transducer was attached at the end of the dry plate section (outside the fluid) to excite A wave. When the generated A wave traveled from the dry section of 0 0 the plate to the section immersed in the fluid, part of the wave energy was reflected backward, and the rest of the wave energy was converted into the leaky A and QS 0 waves. Leaky A wave decayed rapidly and disappeared after a short propagation 0 distance. While QS wave propagated along the immersed plate with low attenuation, and then reached the end of the immersed section and reflected back to the measurement location. At the point where the plate was outside the fluid, the QS wave was converted back to A wave. The time-of-arrival and amplitude of the 0 measured signals changed with the viscosity and bulk longitudinal velocity of the fluid, and hence, they could be employed to measure the fluid properties. Yu et al. [22] proposed a Lamb wave-based method for assessing liquid levels in the nuclear cooling pipe system. The method used a pair of piezoelectric wafer transducers that were mounted on the wall of a test tank. One of the transducers was used as an actuator and the other was used as a receiver. The wave signals were measured on the test tank filled with different amounts of water. It was found that the fundamental symmetric mode (S ) of Lamb waves was not influenced by the change 0 of water level. In contrast, the presence of water significantly changed both the amplitude and phase of A wave. The phase change was shown to have a linear 0 relationship with the change of water level. It should be noted that this study did not take into account the QS and leaky A waves, which also exist in the water- 0 immersed plate [21, 23]. Guo et al. [24] developed two-dimensional (2D) finite element (FE) models to simulate guided wave propagation along an empty steel vessel and the steel vessel filled with water, respectively. At the selected excitation frequency, A and 0 QS waves were identified on the water-free vessel and the water-filled vessel, respectively. The latter was found to propagate more slowly than the former. Therefore, the traveling time of the guided waves between two transducers could be also utilized for measuring the liquid level in the steel vessel. This study only 42
ADE
Chapter 3 considered a single excitation frequency, at which the leaky A wave mentioned in 0 [18] was not detected in the water-immersed plate-like structure by both the 2D FE simulations and the experimental measurements [24]. The aforementioned studies employed the mode conversion between QS and A waves at different excitation frequencies to achieve different applications, 0 where the interactions among various guided wave modes were shown to be different. To date, there are very limited studies on the variation with frequency of the mode conversion phenomenon. However, studying the influence of excitation frequency on guided wave propagation is very important because the behaviors of guided waves are frequency-dependent. For example, it was reported that the displacements of the QS wave mainly occur in the liquid, and the majority of studies on QS wave had been focused on the fluid properties sensing [10, 18]. Only in recent years, its applications were extended to detect damage for plate structures submerged in liquid due to the observation that the QS wave at low frequencies has most of its wave motions conserved in the immersed plate [21, 25]. QS wave is dispersive at low frequencies and becomes nondispersive at high frequencies. The wave structure of QS wave at a low frequency significantly differs from that at a high frequency. Between the high and low frequencies, there is a frequency range, at which QS wave transitions from dispersive to nondispersive. In this frequency range, the wave structure of the QS wave changes rapidly with frequency, while that of the A wave does not change much. Therefore, the mode conversion between 0 QS and A waves should also vary significantly with frequency, which has not been 0 discussed in the literature. In the present study, the frequency dependence of the mode conversion from A wave to QS wave is studied numerically and experimentally. The findings of 0 this study complement the current knowledge about guided wave propagation in partially immersed plates and provide a guide on selecting appropriate excitation frequencies for NDT of partially immersed structures and assessing liquid properties and levels. The numerical method using a three-dimensional (3D) FE model is proposed to portray guided wave propagation in a steel plate, of which one side is partly exposed to water. A wave is excited at different frequencies on the 0 dry section of the plate and travels to the immersed section. The simulation results 43
ADE
Chapter 3 provide a visualization of the interaction of different guided wave modes in both the plate and water. Then, experiments are conducted on a steel tank that is partially filled with water. The time-space wave fields are captured by a scanning laser Doppler vibrometer (SLDV) before and after guided waves travel from the dry section of the plate into the immersed section. The mode conversion process is graphically shown with the use of 2D Fourier transform (FT). The experimental results show a good agreement with the numerical simulations. After that, the variation of the mode conversion from A wave to QS wave with the excitation 0 frequency is analyzed based on the theoretical dispersion curves and mode shapes of guided waves. Furthermore, it is observed that the energy of guided waves can shift in the frequency domain during the mode conversion process if the phase velocity of the incident A wave is larger than the sound speed of the surrounding 0 liquid medium. The energy shift in frequency can change the behaviors of guided waves, which should be paid attention to in practical applications. The paper is organized as below. Section 3.3 compares the theoretical dispersion curves and mode shapes of guided waves in a plate surrounded by air and the plate with one side exposed to water. Section 3.4 describes the 3D FE model and presents the simulated guided wave fields in the partially immersed plate. After that, Section 3.5 shows the experimental setup and the configuration of measurement points. Section 3.6 illustrates the signal processing techniques and the analysis of experimentally measured signals. Then, Section 3.7 summarizes the frequency dependence of the mode conversion from A wave to QS wave and 0 explains the mechanism of the energy shift in frequency phenomena according to the theoretical dispersion curves and mode shapes of the guided waves. Based on the findings, the selection of appropriate excitation frequency is discussed for reliable testing through the mode conversion from A wave to QS wave. Finally, 0 conclusions are drawn in Section 3.8. 44
ADE
Chapter 3 3.3. Guided waves in plates surrounded by air and plates with one side exposed to water Guided waves behave differently in plates surrounded by air and plates immersed in water. When guided waves propagate in a plate in gaseous environments, there is a very small energy leakage from the plate to the air. The energy leakage to the air is not modeled in this study because the resistance of air to the displacements of particles at the plate surface is very small. As shown in Figure 3.1, traction-free boundary conditions are applied to the plate surface open to the air. In comparison, when one side of the plate is exposed to water, the out-of-plane displacements and stresses at the plate-water interface become continuous. The shear stresses are disconnected because water cannot sustain shear forces [26]. The energy leakage to the water layer is substantially larger than that to the air. Figure 3.1. Guided wave propagation models and boundary conditions for (a) a dry plate surrounded by air, and (b) a plate with one side exposed to water The properties of guided waves vary with frequency, which can be theoretically predicted by dispersion curves. The present study employed the global matrix method to calculate the dispersion curves and the theoretical results were used to interpret the following numerical and experimental data. Two guided wave propagation models were constructed using the commercially available software DISPERSE [27]. They were a 2 mm thick steel plate and the plate with water loaded 45
ADE
Chapter 3 on one side, respectively. Table 3.1 gives the material properties of the steel plate and water. The water layer was defined as a semi-infinite half-space. The boundary conditions of the plate-air interface and plate-water interface were modeled by the solid-vacuum and solid-liquid interfaces, respectively. Based on the geometry and material properties, the stresses and displacements in the plate and water layers could be determined in terms of the partial waves. Then, a global matrix equation representing the whole model was assembled by matching the boundary conditions of each interface. The global matrix equation is a function of frequency, wavenumber, and attenuation. Solving this global matrix equation gives a series of combinations of frequency, wave number, and attenuation, at which the partial waves can combine to a guided wave mode that propagates on the plates in the directions as shown in Figure 3.1. Table 3.1. Material properties of the steel and water Young’s Bulk Bulk wave Density Poisson’s Material modulus modulus velocity (kg m-3) ratio (GPa) (GPa) (m s-1) Steel 7800 212.038 0.287 -- Water 1000 -- -- 2.2 1480 Figure 3.2 compares the dispersion curves of guided waves for the 2 mm thick steel plate surrounded by air and the plate with water loaded on its one side. At the frequency range up to 500 kHz, only A and S waves exist in the plate 0 0 without water. They are represented by the green and blue dash-dot lines in Figure 3.2, respectively. The black solid lines, red dashed lines, and magenta dotted lines denote QS, leaky A , and leaky S waves in the one-side water-immersed plate, 0 0 respectively. The phase velocity C and the group velocity C can be related to p g the angular frequency w and the real wavenumber k as C w k and p C w k. As shown in Figure 3.2(c), the wavenumber dispersion curves of A 0 g wave and S wave are almost overlapped with those of leaky A wave and leaky S 0 0 0 wave, respectively. It should be noted that leaky A wave appears only after 150 0 kHz where its phase velocity is greater than the sound speed of the surrounding 46
ADE
Chapter 3 water [28]. At a frequency lower than 150 kHz, the wavenumber of QS wave in the one-side water-immersed plate (black solid line) is just slightly larger than that of A wave in the dry plate (green dash-dot line) and the difference between them 0 increases with frequency. Figure 3.2. Comparison of dispersion curves of a 2 mm thick steel plate and the plate with one side exposed to water: (a) phase velocity curves; (b) group velocity curves; (c) wavenumber curves; (d) attenuation curves (the legends in Figure 3.2(d) are applied for Figures 3.2(a)-3.2(d)). Figure 3.2(d) shows the attenuation dispersion curves where significant deviations can be observed between the dry plate and the one-side water-immersed plate. Obviously, leaky A wave (red dashed line) has a much higher attenuation 0 than any other wave mode. This is because leaky A wave is dominated by the out- 0 of-plane displacements so that the wave energy can easily and massively radiate into the surrounding water [29]. The attenuation dispersion curve of leaky A wave 0 declines sharply in the selected frequency range. This means that the low-frequency leaky A wave has larger attenuation than the high-frequency leaky A wave. The 0 0 levels of attenuation of other wave modes are close for the frequency lower than 47
ADE
Chapter 3 200 kHz. Over 200 kHz, the attenuation of QS wave drops to almost zero, while the attenuations of A , S , and leaky S waves slowly increase with frequency. From 0 0 0 the above observations, it can be concluded that when guided waves propagate from the dry plate to the water-immersed plate, S wave is converted to leaky S wave 0 0 that has the same wavenumber but slightly higher attenuation. A wave is converted 0 to QS wave at a frequency lower than 150 kHz where both the wavenumber and attenuation of QS wave in the immersed plate are similar to those of A wave in the 0 dry plate. However, leaky A wave appears when the excitation frequency exceeds 0 150 kHz. A wave can be mode converted to both QS wave and leaky A wave. 0 0 Thus, the mode conversion process can be different with frequency. The similarity between A wave in the dry plate and QS wave in the one- 0 side water-immersed plate is also studied by their mode shapes. The mode shape of a guided wave mode shows the distributions of the displacements through the thickness of the structure and it can be calculated using DISPERSE [27]. The frequency band of interest is selected from 100 kHz to 200 kHz, in which the wavenumber dispersion curves of the A and QS waves gradually separate as the 0 frequency increases. Figure 3.3 shows the mode shapes of A wave at 100 kHz, 150 0 kHz, and 200 kHz for the 2 mm thick steel plate that is not in contact with water. The deformation of the dry plate is dominated by out-of-plane displacements denoted by the blue solid lines. As the frequency increases, the mode shape diagrams of A wave do not display much difference. 0 Figure 3.4 shows the mode shapes of QS wave at 100 kHz, 150 kHz, and 200 kHz for the 2 mm thick steel plate with one side exposed to water. The water layer was defined as a semi-infinite half-space. As shown in Figure 3.4, the normalized displacement fields of the mode shapes in the water layer monotonically decrease with the distance away from the plate-water interface. To better compare the wave structures in the plate with and without water, the mode shape diagrams only show the 2 mm water regions near the plate-water interface. As shown in Figures 3.3(a) and 3.4(a), the deformation of QS wave in the immersed plate is similar to that of A wave in the dry plate when the frequency is lower than 150 0 kHz. However, the deformation in the immersed plate of QS wave decreases rapidly with frequency. At frequencies above 150 kHz, most of the displacements of QS 48
ADE
Chapter 3 3.4. Numerical simulation of guided wave propagation To portray the guided wave propagation in partially water-immersed plates, a 3D FE model was developed using the commercial FE software ABAQUS. Table 3.1 gives the material properties used for the FE simulation. A 300 mm × 150 mm × 2 mm steel plate was modeled with symmetry boundary conditions applied to the top and right edges and absorbing regions attached to the left and bottom edges, as shown in Figure 3.5. The absorbing regions were 50 mm wide and were divided into 50 layers. The mass-proportional damping of the material in the absorbing regions gradually increased layer by layer from zero at the innermost layer to 4×106 at the outmost layer. The absorbing region by increasing damping can reduce unwanted waves reflected from the plate edges and has been widely used for ultrasonic guided wave simulation analysis [25, 30-33]. Figure 3.5. Schematic diagram of the 3D FE model for a steel plate with one side partly exposed to water: (a) front view and (b) side view. 50
ADE
Chapter 3 A wave was generated by applying the out-of-plane displacements to the 0 plate surface covered by a 5 mm diameter quarter-circle located at the top right corner of the plate [34]. Figure 3.9(b) shows the waveform of the excitation signal, which is a 10-cycle narrowband tone burst. To define the locations of the measurement points, a one-dimensional coordinate, scanning distance (SD), was defined along the right edge of the plate vertically downward as shown in Figure 3.5. The origin (SD = 0 mm) was set at the position of 50 mm below the excitation center. Then, the out-of-plane displacements were collected at five measurement points, which were evenly distributed at 50 mm apart from SD = 0 mm to SD = 200 mm. The other side of the plate was partially in contact with water as shown in Figure 3.5(b). The water level was located at the second measurement point (SD = 50 mm). The thickness of the water layer was chosen to separate the pressure wave reflections from the incident wave signals. The steel plate and the water layer were meshed using 3D eight-node solid elements with reduced integration (C3D8R) and 3D eight-node acoustic elements with reduced integration (AC3D8R), respectively. The fluid and solid interface was simulated by node-surface tie constraints [16, 25]. The element size was set as 0.5 mm, which ensured approximately fifteen elements exist per wavelength of QS wave at 200 kHz. The simulation results were calculated using the central-difference integration by ABAQUS/Explicit [35]. Figure 3.6 presents the snapshots of the simulation results with the excitation frequency of 120 kHz. The color in the water regions denotes the acoustic pressure. At this excitation frequency, a large proportion of the QS wave energy is conserved in the one-side water-immersed plate, of which the deformation is similar to that of A wave in the dry plate as shown in Figures 3.3 and 3.4. When A wave 0 0 travels from the dry section of the plate into the water-immersed section, part of the wave energy is converted to the pressure waves in the water, and the rest of the wave energy continues to propagate along the water-immersed plate at a speed slightly quicker than the pressure waves in the water. After a short propagation distance in the water-immersed plate, the first wave packet decays slowly and the acoustic field is tethered to the plate-water interface as shown in Figure 3.6(d). These are the typical characteristics of QS wave. 51
ADE
Chapter 3 Figure 3.7 shows the snapshots of the simulation results with the excitation frequency of 170 kHz. The mode shape of QS wave at 170 kHz is dominated by the in-plane displacements of water and the deformation of QS wave in the water- immersed plate is no longer similar to that of A wave in the dry plate (see Figures 0 3.3 and 3.4). The first wave packet in Figure 3.7 continuously radiates wave energy into the surrounding liquid medium, as shown by the skewed acoustic fields (skewed yellow lines) in the water layer in Figures 3.7(b) and 3.7(d). This wave packet with continuous wave energy leakage cannot be detected in Figure 3.6. Following the first wave packet, another wave packet propagates along the immersed plate at a speed slightly quicker than the pressure waves in water. Unlike the first wave packet, the second wave packet propagates with most of the energy confined to the plate-water interface. Based on the propagation speeds and the acoustic pressure in the water, the first and second wave packets in Figure 3.7 are identified as leaky A and QS waves, respectively. 0 To better observe how the signals change with the propagation distance, Figure 3.8 presents the simulated out-of-plane displacements at the five measurement points that are denoted by the red dots in Figure 3.5. Figures 3.8(a) and 3.8(b) show the time-domain data for the excitation frequencies of 120 kHz and 170 kHz, respectively. Their corresponding frequency spectrums are given in Figures 3.8(c) and 3.8(d), respectively. The amplitudes are normalized by the maximum absolute amplitudes of the signals measured at the first measurement point (SD = 0 mm). It should be noted that when the excitation frequency is 170 kHz, the time-domain signals measured at SD = 100 mm, 150 mm, and 200 mm are so small that they are magnified by a factor of four and shown in Figure 3.8(b). The normalized amplitudes of the signals collected in the immersed section of the plate (SD > 50 mm) significantly decrease with the excitation frequency. In addition, when the excitation frequency is 170 kHz, the simulated signals show apparent frequency shifts in the frequency spectrums. For example, the central frequency of the signal measured at SD = 100 mm, denoted by the orange dash-dot line in Figure 3.8(d), shifts to a frequency slightly higher than the central excitation frequency of 170 kHz. Subsequently, the wave energy shifts to a lower frequency as shown by the signals measured at SD = 150 mm (purple solid line) and SD = 200 mm (yellow solid line) in Figure 3.8(d). The simulation results demonstrate that the guided wave 53
ADE
Chapter 3 propagation in the partially water-immersed plate varies significantly with the excitation frequency. The following sections present experimental investigations to validate the simulation results, and the phenomenon of energy shift in the frequency domain due to the presence of water is discussed in detail. Figure 3.8. Simulated time-domain signals for a steel plate with one side partly exposed to water (the same conditions as Case B presented in the experimental section) with the excitation frequency of (a) 120 kHz and (b) 170 kHz, respectively; (c) and (d) are the frequency spectrums of (a) and (b), respectively (the time-domain signals measured at SD = 100 mm, 150 mm, and 200 mm in Figure 3.8(b) are magnified by a factor of four). 3.5. Experiment setup Experiments were conducted on a steel tank, of which the front wall was used as the test plate. The test plate was 2 mm thick and made of mild steel. A circular piezoceramic wafer (Ferroperm Pz27, Denmark) was used as the guided wave actuator and it was bonded on the external surface of the test plate as indicated by the PZT transducer in Figure 3.9(a). The diameter and the thickness of the 54
ADE
Chapter 3 defining measurement points. Then, the out-of-plane displacements were recorded by the SLDV at a sampling rate of 10.24 MHz. Each measurement was improved by averaging the signals with 800 acquisitions and applying a low-pass filter with the cut-off frequency being 1MHz. Figure 3.9(c) shows the schematic diagram of the experiment setup. Similar to the 3D FE model, a one-dimensional coordinate, SD, was defined on the external surface of the test plate. The origin (SD = 0 mm) was set at the position of 50 mm vertically below the center of the piezoceramic wafer. The experimental study included two parts, which were five-point scan tests and line scan tests, respectively. Firstly, the five-point scan tests were conducted on the steel tank to validate the simulation results. The signals were measured at SD = 0 mm, 50 mm, 100 mm, 150 mm, and 200 mm, which were at the same locations as the simulations as shown in Figure 5. Then, line scan tests were carried out to visualize guided wave fields on the test plate. The objective of the line scan test was to experimentally demonstrate the interaction of each guided wave mode during the mode conversion process. According to the simulation results shown in Figures 3.6, 3.7, and 3.8, the mode conversion process mainly occurs between SD = 50 mm and SD = 150 mm. After SD = 150 mm, only the QS wave can be detected as shown by the signals measured at SD = 150mm and SD = 200 mm in Figure 3.8. Therefore, the line scan tests focused on the region between SD = 0 mm and SD = 150 mm as shown in Figure 3.10(a). The signals were collected at 127 measurement points, which were evenly distributed along the scan line from SD = 0 mm to SD = 150 mm. With these measurements, the time-space wave fields could be constructed by plotting the amplitudes of the measured signals versus time and SD. 56
ADE
Chapter 3 Figure 3.10. Schematic diagram of the scan line on the test plate (a) front view; (b) side view of Case A (empty tank); (c) side view of Case B (partially water-filled tank). In order to demonstrate the influence of water on the guided wave propagation, both the five-point scan tests and line scan tests were carried out on the empty tank (Case A) and the partially water-filled tank (Case B), respectively. Figures 3.10(b) and 3.10(c) show the side view of the scan line on the test plate for Case A and Case B, respectively. In Case B, the steel tank was partially filled with water with the water level set at SD = 50 mm. Therefore, one-third of the scan line was located in the non-immersed section of the plate (from SD = 0 mm to SD = 50 mm), called “dry plate”, and the rest of the scan line was located in the section of the one-side water-immersed plate (from SD = 50 mm to SD = 150 mm), which was denoted as “immersed plate”. Guided waves were generated on the test plate by the piezoceramic wafer located at 100 mm above the water level. The time-space wave fields were captured before and after guided waves traveled from the dry plate into the immersed plate. 3.6. Experimental results and analysis 3.6.1. Validation of numerical simulations This section presents the experimental measurements to validate the accuracy of the 3D FE model. Figure 3.11 presents the experimental results with the central excitation frequency of 120 kHz. Figures 3.11(a) and 3.11(b) show the time-space 57
ADE
Chapter 3 wave fields for the empty tank and the partially water-filled tank, respectively. The amplitudes are normalized by the maximum absolute amplitudes of the signals measured at the first scan point (SD = 0 mm). The black dashed line in Figure 3.11(b) denotes the water level located at SD = 50 mm. The generated guided wave fields in the dry section of the plate (between SD = 0 mm and SD = 50 mm) are similar for both the empty tank in Figure 3.11(a) and the partially water-filled tank in Figure 3.11(b). However, the amplitudes of the signals measured in the immersed plate (SD > 50 mm) of the partially water-filled tank in Figure 3.11(b) are smaller than their counterparts in the empty tank in Figure 3.11(a). This indicates that part of the wave energy leaks from the immersed plate into the water as shown in the snapshots of the simulation results (see Figures 3.6(a) and 3.6(b)). Figures 3.11(c) and 3.11(d) show typical examples of the signals measured at SD = 0 mm, 50 mm, 100 mm, 150 mm, and 200 mm for the empty tank and the partially water-filled tank, respectively. The wave speed in the immersed plate of the partially water-filled tank, as denoted by the dark blue dash-dot line in Figure 3.11(b), is slightly slower than the wave speed in the dry plate that is marked by the red dashed line. From the wave speed evaluation, the wave packets measured from the empty tank in Figure 3.11(a) and the dry section of the partially water-filled tank in Figure 3.11(b) are identified as A wave [24]. S wave can be not observed 0 0 because it has negligible out-of-plane motions at the selected excitation frequency [25, 36]. The changes of the wave speed and amplitude in the immersed section of the test plate demonstrate that A wave is mode converted to QS wave. Figures 0 3.11(e) and 3.11(f) present the frequency spectrums of Figures 3.11 (c) and 3.11(d), respectively. The amplitudes are normalized by the signal measured at the first scan point (SD = 0 mm). The peaks of all measured signals are concentrated around the central frequency of the excitation signal of 120 kHz. There is a good agreement between the experimental measurements shown in Figures 3.11 (d) and 3.11(f) and the simulation results (see Figures 3.8(a) and 3.8(c)). 58
ADE
Chapter 3 Figure 3.11. Experimental results with the excitation frequency of 120 kHz: (a) and (b) are time-space wave fields for the empty tank and the partially water-filled tank, respectively; (c) and (d) are typical examples of the time-domain signals for the empty tank and the partially water-filled tank, respectively; (e) and (f) are the frequency spectrums of (c) and (d), respectively. For comparison, Figure 3.12 presents the experimental results with the central excitation frequency of 170 kHz. Figures 3.12(a) and 3.12(b) present the time-space wave fields for the empty tank and the partially water-filled tank, respectively. For the empty tank, the waves propagate at a consistent speed as represented by the red dashed line in Figure 3.12(a). The same wave fields are observed in the dry section of the plate (SD < 50 mm) of the partially water-filled tank as shown in Figure 3.12(b). However, when guided waves propagate into the water-immersed section (SD > 50 mm), the wave amplitudes decrease rapidly and 59
ADE
Chapter 3 Figures 3.12(c) and 3.12(d) show typical time-domain signals measured at SD = 0 mm, 50 mm, 100 mm, 150 mm, and 200 mm for the empty tank and the partially water-filled tank, respectively. To provide better observation, the signals measured from the partially water-filled tank at SD = 100 mm, 150 mm, and 200 mm are magnified by a factor of four and shown in Figure 3.12(d). Guided waves propagate as a single wave packet along the dry section of the plate (between SD = 0 mm and SD = 50 mm). However, after a short propagation distance in the immersed plate, the signal measured at SD = 100 mm shows two wave packets, each propagating at different speeds. It should be noted that the mode split phenomenon is not observed on the water-immersed plate with the central excitation frequency of 120 kHz (see Figure 3.11(d)). The first wave packet decays quickly and disappears as shown by the signal measured at SD = 150 mm in Figure 3.12(d), while the second wave packet propagates with low attenuation at a slower wave speed (dark blue dash-dot line). Figures 3.12(e) and 3.12(f) show the frequency spectrums of Figures 3.12(c) and 3.12 (d), respectively. The signals measured at SD = 100 mm, 150 mm, and 200 mm from the partially water-filled tank in Figures 3.12(f) have much smaller amplitudes than their counterparts from the empty tank in Figures 3.12(e). In addition, Figure 3.12(f) displays the energy shift in frequency, which has a good agreement with the simulation results (see Figure 3.8(d)). It is confirmed that the energy shift in the frequency domain is due to the presence of water because this phenomenon does not occur in the case of the empty tank as shown in Figure 3.12(e). To further investigate the accuracy of the 3D FE model, Figure 3.13 compares the peak amplitudes of the simulated and experimentally measured signals in the frequency domain for both the empty tank (Case A) and the partially water-filled tank (Case B), respectively. The magnitudes are normalized by their corresponding peak amplitudes of the signals measured at SD = 0 mm. Figures 3.13(a) and 3.13 (b) present the results with the excitation frequency of 120 kHz and 170 kHz, respectively. In the dry section of the plate (SD < 50 mm), the amplitudes of the simulated and experimentally measured signals are identical for both cases. When guided waves just pass the water level (from SD = 50 mm to SD = 100 mm), the measured signals of Case A decrease slowly and smoothly, but the signal amplitudes of Case B drop substantially. After SD = 100 mm, the guided 61
ADE
Chapter 3 wave amplitudes in both Case A and Case B decrease slowly with distance. When the excitation frequency increases from 120 kHz to 170 kHz, the amplitudes of the signals of Case A do not show obvious changes as shown by the red hexagons and circles in Figure 3.13. However, the amplitudes of the signals of Case B significantly decrease when the excitation frequency increases. In general, the proposed 3D FE model well predicts the frequency shift phenomena and wave attenuation characteristics and hence, the simulation results are validated to interpret the experimental data. Figure 3.13. Normalized amplitudes of the simulated and experimentally measured signals in the frequency domain for the empty tank (Case A) and the partially water- filled tank (Case B) with the excitation frequency of (a) 120 kHz and (b) 170 kHz, respectively. 3.6.2. Segmented frequency wavenumber analysis Although the time-space analysis displays the variation of wave amplitudes with the time and propagation distance, it cannot determine the wave mode conversion characteristics such as mode identities and their corresponding frequency components. To graphically demonstrate the mode conversion process, 2D FT is employed to identify the mode information of the experimental data collected along the scan line on the partially water-filled tank (Case B). The scan line is divided into three segments, each of which is 50 mm long and comprises 43 measurement points as shown in Figure 3.10(c). The generated guided waves first propagate through Segment 1 (Dry plate) and then to Segment 2 (Water-immersed plate) and 62
ADE
Chapter 3 finally to Segment 3 (Water-immersed plate). The water level is between Segment 1 and 2. The time-space data of each segment is converted to the frequency- wavenumber spectrum through 2D FT, which is defined as: uk, f = ux,te-i(2ft-kx)dtdx (3.1) where uk, f  and ux,t are the data in the frequency-wavenumber domain and time-space domain, respectively. k and f denote the wavenumber and frequency, respectively. x and t represent the space and time coordinate, respectively. Figures 3.14(a)-3.14(c) present the experimentally measured data in the time-space domain for the three segments and their corresponding frequency- wavenumber spectrums are given in Figures 3.14(d)-3.14(f), respectively. The excitation frequency is 120 kHz. The color in the frequency-wavenumber spectrums denotes the wave energy of the experimentally measured signals, which is calculated by 2D FT. The black solid lines are the theoretical wavenumber dispersion curves of A , S , and QS waves calculated by DISPERSE. As mentioned 0 0 in Section 3.3, the wavenumber dispersion curves of leaky A and leaky S waves 0 0 overlap with A and S waves. For convenience, leaky A and leaky S waves are 0 0 0 0 labeled as A and S in the figures, respectively. In Segment 1, where the plate is 0 0 not immersed in water, only the energy of A wave is identified in the frequency- 0 wavenumber spectrum as shown in Figure 3.14(d). The energy of S wave is absent 0 because it has negligible out-of-plane displacements [25, 36]. Next, the wave energy is converted from A wave to QS wave immediately in Segment 2 as shown 0 in Figure 3.14(e). The mode conversion occurs rapidly and the energy of QS wave decays slowly with distance and dominates the frequency-wavenumber spectrums of Segments 2 and 3. At this excitation frequency, leaky A wave is not detected 0 and the energy shift in frequency is not observed. 63
ADE
Chapter 3 QS wave. Considering that the deformation in the immersed plate of QS wave at low frequencies is greater than that at high frequencies (see Figure 3.4), the anti- symmetrical excitation of leaky A wave is more likely to produce QS wave at lower 0 frequencies. This can be also manifested by the frequency spectrums (see Figures 3.8(d) and 3.12(f)) where the wave energy is progressively transferred from a higher frequency at SD = 100 mm to a frequency lower than the central excitation frequency at SD = 150 mm. After that, the wave energy is conserved at the frequency (lower than the excitation frequency of 170 kHz) and propagates with small attenuation. 3.6.3. Further study by sweeping the excitation frequency To further investigate the phenomenon of energy shift in the frequency domain, the five-point scan tests were conducted on the partially water-filled tank using the excitation signals with the central frequencies of 110 kHz, 130 kHz, 140 kHz, 150 kHz, 160 kHz, and 180 kHz. The collected signals were transformed to the frequency domain and shown in Figure 3.16. The amplitudes were normalized by the signals measured at the first scan point (SD = 0 mm). As it is confirmed from the segmented frequency wavenumber plots (see Figures 3.14 and 3.15), the signals measured at SD = 150 mm to 200 mm, denoted by the purple and yellow solid lines in Figures 3.11(f), 3.12(f), and 3.16, are dominated by the low-attenuation QS wave. Therefore, it can be concluded that the normalized amplitude of QS wave converted by A wave decreases when the excitation frequency increases. 0 For excitation frequencies below 140 kHz, the signals measured from both the dry section and water-immersed section of the plate are concentrated around the central excitation frequency (see Figure 3.11(f) and Figures 3.16(a)-3.16(c)). The energy shift in frequency can be observed for excitation frequencies over 150 kHz, where leaky A wave appears as the phase velocity of the incident A wave becomes 0 0 larger than the sound speed of the surrounding water. Under the central excitation frequency of 150 kHz and 160 kHz, the signals measured at SD = 100 mm, 150 mm, and SD = 200 mm have most of their wave energy concentrated at a frequency lower than the central excitation frequency (see Figures 3.16(d) and 3.16(e)). The range of the frequency shift also increases when the central excitation frequency 67
ADE
Chapter 3 increases. However, the signals measured at SD = 100 mm under the central excitation frequency of 170 kHz and 180 kHz are shown to have more energy conserved at higher frequencies (see Figures 3.12(f) and 3.16(f)). This is because the amplitudes of QS wave converted by A wave are so small that leaky A wave 0 0 can be detected clearly at this measurement point. Leaky A wave at higher 0 frequencies decays more slowly than that at lower frequencies as discussed in Section 3.3, making the central frequency of the signal shift to a relatively higher frequency. However, leaky A wave completely disappears after a short 0 propagation distance, and only the QS wave can be detected at SD =150 mm and SD = 200 mm, as shown in Figure 3.15. Since the low-attenuation QS wave at lower frequencies has a larger deformation fraction in the immersed plate, the measured signals from the surface of the immersed plate eventually concentrate at a frequency lower than the central excitation frequency. The next section summarizes the frequency dependence of the mode conversion from A wave to QS wave and 0 further explains the mechanism of the frequency shift phenomena. 3.7. Discussion and application 3.7.1. The influence of excitation frequency on the mode conversion process The mode conversion from A wave to QS wave has been numerically and 0 experimentally shown to be dependent on the excitation frequency. The frequency dependence is summarized in this section and analyzed according to the dispersion behaviors of the guided waves. The theoretical dispersion curves and mode shapes of guided waves have been derived from the global matrix theory and are present in Section 3.3. In the following discussion, the term “high frequency” means the frequency range, in which the phase velocity of the incident A wave is larger than 0 the sound speed of the surrounding liquid medium and the leaky A wave appears. 0 The term “low frequency” indicates the frequency range, in which the phase velocity of the incident A wave is smaller than the sound speed of the surrounding 0 liquid medium and the quasi-Scholte wave is dispersive. The transition frequency between the high and low frequency ranges can be estimated from the phase velocity dispersion curves. For example, the phase velocity of A wave for a 0 metallic plate monotonically increases with frequency until it reaches the Rayleigh 68
ADE
Chapter 3 wave speed, which is around 3000 m/s for steel. The sound speed of water is around 1500 m/s, which is constant for all frequencies. Therefore, the transition frequency is around 150 kHz, at which the phase velocity of A wave traveling along the 2 0 mm thick steel plate is around 1500 m/s, as shown in Figure 3.2(a). The phase velocity dispersion curves can be calculated by the commercial software DISPERSE. The input data includes the material properties of the plate, the plate thickness, and the sound speed of the surrounding liquid medium. At low frequencies, the difference in wavenumber between A wave and QS 0 wave is small. In addition, the deformation of QS wave in the immersed plate is similar to that of A wave in the dry plate (see Figures 3.3 and 3.4). Thus, A wave 0 0 can be mode converted to QS wave rapidly with most of the wave energy conserved in the plate. When the excitation frequency increases, the energy distribution of QS wave in the immersed plate decreases sharply and the similarity reduces between A and QS waves (see the dispersion curves in Figure 3.2 and mode shapes 0 diagrams in Figures 3.3 and 3.4). Therefore, the amplitude of QS wave converted by A wave significantly decreases with frequency. 0 At high frequencies, leaky A wave appears when the phase velocity of the 0 incident A wave becomes larger than the sound speed of the water. A wave is 0 0 mode converted to both QS wave and leaky A wave with more energy transferred 0 to the latter. Leaky A wave that has the flexural mode shape in the plate 0 continuously radiates compressional waves in the liquid and also excites QS wave. After a short propagation distance, leaky A wave decays quickly and disappears so 0 that only QS wave can be detected. 3.7.2. The mechanism of the energy shift in frequency phenomenon The energy shift in frequency occurs during the mode conversion process when the incident A wave is generated at high frequencies. The mechanism of the energy 0 shift in frequency phenomenon can be explained by the dispersion curves and the mode shapes of guided wave modes as follows. Firstly, the central frequency of the signal can shift to a frequency higher than the center frequency of the excitation, when the wave fields are dominated by 69
ADE
Chapter 3 the leaky A wave (see signals at SD = 100 mm in Figures 3.12(f) and 3.16(f)). This 0 is due to the fact that the attenuation dispersion curve of leaky A wave declines 0 sharply in the selected frequency region (see Figure 3.2). The low-frequency components of leaky A wave decay much quicker than the high-frequency 0 components. Therefore, leaky A wave at high frequencies can travel longer 0 distances, making the central frequency of the signals measured in the immersed plate near the water level relatively higher than the central excitation frequency. Secondly, the central frequency of the mode converted QS wave is relatively lower than the central frequency of the excitation. One reason is that the deformation in the immersed plate of QS wave at low frequencies has a flexural mode shape, which is similar to A wave in the dry plate. But the similarity between 0 QS and A waves reduces with frequency. This makes the mode conversion from 0 A wave to QS wave (at the intersection of the dry plate and the water-immersed 0 plate) much easier at low frequencies than that at high frequencies, with more energy transferred and conserved in the plate. The other reason is that the deformation of QS wave in the plate is rapidly reduced with frequency (see Figure 3.4). Thus, QS wave at low frequencies can be excited on the plate more easily by the out-of-plane motions of leaky A wave after guided waves propagate into the 0 water-immersed plate. Since the generated QS wave has low attenuation, the measured signals in the immersed plate eventually shift to a frequency lower than the central excitation frequency (see signals at SD = 150 mm and SD = 200 mm in Figure 3.12(f) and Figures 3.16(d)-3.16(f)). 3.7.3. Implications for practical applications The findings of the present study suggest that the low-attenuation QS wave can be easily excited by mode conversion from A wave that is generated at low 0 frequencies on the dry plate section. This phenomenon can be employed to detect damage for the plate structures that are partially immersed in liquid, such as partially water-filled tanks and pipelines. These structures generally experience uniform corrosion and pitting corrosion. The latter is more critical because it damages the deep structures with little loss of metal [37]. 70
ADE
Chapter 3 Previous studies have characterized corrosion damage using A wave for 0 plates surrounded by air [38-40] and QS wave for plates in contact with liquid [21, 41]. For partially immersed plates, it is also possible to evaluate the defects by sending A waves on the dry section of the plate and measuring the QS wave signals 0 on the immersed section. This method is very promising for long-range inspection because QS wave does not radiate energy in the liquid (see Figures 3.6 and 3.7) and is able to travel along the plate-fluid interface with low attenuation. The measurement range through using the dispersive QS wave by mode conversion from A wave can be of the order of several meters based on the low attenuation 0 characteristics as shown in Figure 3.13. However, the actual propagation distance is dependent on the material properties of the plate and the surrounding liquid medium as well as the excitation frequency. As discussed in Section 3.7.1, the lower the excitation frequency, the more wave energy conserved in the plate during the mode conversion process, and therefore the longer the propagation distance. Another advantage is that this method has the potential to characterize the structural defects entirely based on the QS wave, which is appealing for accurate detection and imaging of the defects [42]. After a short propagation distance, leaky A wave 0 decays quickly and disappears due to high attenuation. The low-attenuation QS wave can be well separated from the leaky S wave, of which the propagation speed 0 is three times that of QS wave (see Figure 3.2(b)). Although leaky S wave has low 0 attenuation, it is not sensitive enough to identify small and shallow corrosion damage in the early stage [43-45]. In contrast, the QS wave has the shortest wavelength at a given frequency and provides better sensitivity than the leaky S 0 wave to shallow hidden corrosion pits in immersed plates [41]. It should be noted that the damage detection algorithm for the partially immersed structures should consider the change of wave behaviors due to the mode conversion phenomenon and the presence of liquid. It is recommended to select an excitation frequency below the transition frequency, at which the proposed phenomenon of guided wave energy shift in frequency can be avoided. (see Figures 3.11 and 3.14). Otherwise, the effect of the potential wave energy shift in the frequency domain should also be carefully considered and compensated. For instance, it is observed in the present study that the wave energy of the mode converted QS wave in the immersed plate moves to a frequency below the central 71
ADE
Chapter 3 excitation frequency, making the actual wavenumber of the QS wave smaller than that at the central excitation frequency (see Figure 3.15(f)). This also indicates a smaller phase velocity, a higher group velocity, and a larger wavelength (see Figure 3.2). The change of the propagation characteristics of guided waves will affect the performance of conventional damage detection and imagining algorithms [16, 46]. The mode conversion between QS and A waves in partially immersed 0 plates was widely used for liquid-level assessing [22, 24, 47] and fluid-property sensing [18, 48, 49]. The behaviors of A wave depend on the geometry and 0 material properties of the plate, while QS wave reflects the properties of both the plate and the surrounding fluid medium. Generally, the difference between QS and A waves becomes larger with frequency. Therefore, increasing the excitation 0 frequency can result in larger deviations in the signals measured from the partially immersed plate in terms of the time of arrival, amplitude, and phase angle, and hence, it can potentially increase the sensitivity of the signals to the variation of liquid level and fluid properties. However, the results of the present study show that the amplitude of the QS wave converted by A wave significantly decreases with 0 frequency. Therefore, the optimal excitation frequency is a trade-off between the sensitivity and the amplitudes of the measured signals. For simplicity, it is also recommended to excite the guided waves at a low frequency to ensure that the phase velocity of the incident A wave is smaller than the sound speed of the surrounding 0 liquid medium. Without the interference of the leaky A wave, the mode conversion 0 process is simple and the frequency shift phenomena can be avoided. Lastly, QS wave at high excitation frequencies becomes nondispersive and is promising for fluid manipulation [20] and removing diffusion boundary layer [19, 50], where the focus is directed on the movement of the fluid particles and the deformation in the plate is not interested. The low-attenuation QS wave with most of the wave energy concentrated at the plate-fluid interface has the potential to cover a large area of the fluid near the plate surface. However, attention should be paid to the potential frequency shift of the measured signals. As shown in Figures 3.15 and 3.16, guided wave energy can shift in the frequency domain during the mode conversion process, which gives rise to the change of wave behavior such as group and phase velocities. 72
ADE
Chapter 3 3.8. Conclusion This paper has provided an insight into the measurement of ultrasonic guided waves in partially immersed plates. The main contributions are concluded as follows: (1). Global matrix method is employed to derive the theoretical dispersion curves and modes shapes of guided waves for a 2 mm thick steel plate and the plate with one side loaded with water. It is found that the low- frequencies QS wave in the one-side water-immersed plate and the A wave 0 in the dry plate have similar wavenumbers and deformations. But the similarity reduces with frequency. (2). A 3D FE model is developed to simulate the guided wave field in the steel plate with one side partially immersed in water. The simulation results are validated by the experimental data. The frequency shift phenomenon and the guided wave amplitudes with propagation distance can be well predicted. (3). The experimental studies are conducted on the empty tank and the partially water-filled tank, respectively. It is confirmed that the frequency shift phenomenon is due to the presence of water. The further investigation presents a segmented frequency wavenumber analysis to graphically demonstrate the mode conversion process, which is divided into three segments: (i) before guided waves propagate into the water-immersed plate, (ii) guided waves just propagate into the water-immersed plate, and (iii) after a short propagation distance in the water-immersed plate. The experimental data are compared with the theoretical dispersion curves, through which the mode identities and the corresponding experimentally measured wave energy can be determined. (4). The guided wave energy shift in the frequency occurs during the mode conversion process, which is not caused by the material nonlinearity (micro cracks) of the plate. The amplitudes of the guided wave signals measured from the water-immersed plate section are much smaller than those obtained from the water-free plate section, on which the frequency shift phenomenon is not observed. Then, the mechanism of the energy shift in frequency 73
ADE
Chapter 3 phenomenon is explained by the attenuation dispersion curves of leaky A 0 wave and the mode shapes of QS wave. (5). Based on the findings, the selection of appropriate excitation frequency is discussed for damage detection of partially submerged structures, assessing liquid properties and levels, and fluid manipulations. In summary, comprehensive investigations have been carried out for the frequency dependence of the mode conversion from A wave to QS wave in a steel 0 plate with one side partially immersed in water. The findings can provide support for the further development of guided wave-based techniques for damage detection on partially immersed structures, liquid-level assessing, and fluid-property sensing. This paper has only focused on the plate partially immersed in water. Future work can study the partially immersed structures with different geometries and investigate the effect when the structure is immersed in other types of liquid. In addition, the influence of damage such as corrosion pits or stress cracking on the guided wave propagation and mode conversion can be investigated. 3.9. Acknowledgment This work was funded by the Australia Research Council (ARC) under grant numbers DP200102300 and DP210103307. The authors are grateful for this support. 3.10. Reference [1] J. Moll, C.P. Fritzen, Guided waves for autonomous online identification of structural defects under ambient temperature variations, Journal of Sound and Vibration, 331 (2012) 4587-4597. [2] J. He, C.A.C. Leckey, P.E. Leser, W.P. Leser, Multi-mode reverse time migration damage imaging using ultrasonic guided waves, Ultrasonics, 94 (2019) 319-331. [3] A. Aseem, C.T. Ng, Debonding detection in rebar-reinforced concrete structures using second harmonic generation of longitudinal guided eave, NDT & E International, (2021) 102496. [4] S. He, C.-T. Ng, C. Yeung, Time-domain spectral finite element method for modeling second harmonic generation of guided waves induced by material, 74
ADE
Chapter 4 Chapter 4. Early damage detection of metallic plates with one side exposed to water using the second harmonic generation of ultrasonic guided waves Abstract Metallic plates are the main structural components in a wide range of thin-walled structures, such as nuclear cooling pipes, pressure vessels, rocket fuel tanks, and submarine hulls. These structures operate in extreme environments and are subjected to transient and repetitive loads. Real-time health monitoring of these structures is indispensable because they are vulnerable to fatigue and corrosion damage. Second harmonic generation is one of the reliable damage detection approaches to evaluate microstructural evolution and has been successfully applied to characterize initial damage on different structures in gaseous environments. However, there have been very limited studies on the second harmonics generation on the structures submerged in liquid. This paper experimentally and numerically investigates the feasibility of using second harmonic generation to evaluate the material degradation of metallic plates with one side exposed to water. The fundamental leaky symmetric Lamb wave mode (leaky S ) at low frequencies is 0 selected because it has low-attenuation and weakly dispersive features, enabling approximate internal resonance. The experimental results show that the second harmonics of leaky S waves grow linearly with the propagation distance. The 0 growth rate of the relative nonlinearity parameter (') can be related to the material nonlinearity of the one-side water-immersed plate. In addition, this study proposed a three-dimensional (3D) finite element (FE) model to simulate the generation of second harmonics in the one-side water-immersed plate. The material properties of the plate are modeled by the Murnaghan strain energy function. The Murnaghan constants of aluminum that describe the material nonlinearity at different fatigue levels are obtained from a previous experimental study. The simulation results demonstrate that the values of ' change significantly with the material properties 79
ADE
Statement of Authorship Title of Paper Early damage detection of metallic plates with one side exposed to water using the second harmonic generation of ultrasonic guided waves Publication Status Published Accepted for Publication Unpublished and Unsubmitted work written in Submitted for Publication manuscript style Publication Details X. Hu, C.T. Ng, A. Kotousov, (2022). Early damage detection of metallic plates with one side exposed to water using the second harmonic generation of ultrasonic guided waves. Thin- Walled Structures, 176, 109284. Principal Author Name of Principal Author (Candidate) Xianwen Hu Contribution to the Paper Conceptualization, Developing and validating numerical models, Conducting experimental measurements, Signal processing and data analysis, Writing the original draft and editing. Overall percentage (%) 80% Certification: This paper reports on original research I conducted during the period of my Higher Degree by Research candidature and is not subject to any obligations or contractual agreements with a third party that would constrain its inclusion in this thesis. I am the primary author of this paper. Signature Date 07/03/2022 Co-Author Contributions By signing the Statement of Authorship, each author certifies that: i. the candidate’s stated contribution to the publication is accurate (as detailed above); ii. permission is granted for the candidate in include the publication in the thesis; and iii. the sum of all co-author contributions is equal to 100% less the candidate’s stated contribution. Name of Co-Author Ching-Tai Ng Contribution to the Paper Supervision, writing – review & editing. Signature Date 9/3/2022 Name of Co-Author Andrei Kotousov Contribution to the Paper Supervision, writing – review & editing Signature Date 07/03/2022 Please cut and paste additional co-author panels here as required.
ADE
Chapter 4 4.2. Introduction Thin-walled structures are widely used in energy, petrochemical, aerospace, civil, and ocean engineering, such as nuclear cooling pipes, pressure vessel, rocket fuel tanks, storage tanks, and submarine hulls. These structures usually have one side exposed to liquid and serve in extreme environments. They are subjected to cyclic loads due to the draining and refilling process, liquid sloshing impacts, and temperature fluctuations [1, 2]. Over time, fatigue damage takes place even though the peak values of the cyclic loads are much smaller than the loading capacity of the structures [3, 4]. In the early damage stage, fatigue is distributed in the material and appears as microscopic imperfections, such as dislocations, persistent slip bands, precipitates, and short cracks at the micro-scale. Optical images of the microstructural features can be found in [5, 6]. As the number of loading cycles increases, these microscopic defects accumulate, coalesce, and form macroscopic cracks, which continuously grow to their critical sizes and cause catastrophic failures [7]. In addition to the cyclic loading, the corrosive operational conditions of the one-side submerged plate can accelerate the damage-accumulation process by corrosion, which results in the metal wear through electrochemical reaction [8, 9]. To mitigate the risk of in-service failure of the structures, real-time health monitoring of the submerged plate structures is indispensable. Existing non- destructive testing (NDT) for submerged structures has visual inspection by divers and robots [10], acoustic emission [11, 12], hydro test [13], magnetic flux leakage test [14] and eddy current test [15]. Most of these approaches require periodic , shutdown of the devices and can only inspect a localized area. They are costly, inconvenient, and inefficient for scanning a large structure. Guided waves have been extensively studied as a potential alternative to overcome the aforementioned limitations of existing NDT techniques. They have the capability of fast propagation over long distances, volumetric inspection of a relatively large area with a small number of sensors, the ability to scan structures with coatings and insulations, and the ability to monitor defects in inaccessible regions [16, 17]. The majority of the studies on guided wave applications were carried out on the structures in gaseous environments. However, guided waves behave differently when the structures are exposed to liquid. The liquid coupling 82
ADE
Chapter 4 can change the dynamic properties of the plate structures [18] and provides a way for the guided wave energy to leak into the surrounding liquid medium [19, 20]. Due to the energy leakage, guided waves that propagate in submerged plates are called leaky Lamb waves, which have multi-modal and dispersive features. At any excitation frequency, multiple leaky symmetric and antisymmetric Lamb wave modes can exist simultaneously. Each of these wave modes behaves differently and varies with the excitation frequency. Compared with structures surrounded by air, guided wave applications on the submerged structures are much more challenging because most of the leaky Lamb wave modes decay quickly and disappear after a short propagation distance. Only a limited number of leaky Lamb wave modes at their corresponding low-attenuation frequency bands can travel a long distance, which has the potential to enable large-area inspection for the submerged structure [19]. Therefore, a good understanding of the wave propagation characteristics is desired for the practical application of leaky Lamb wave-based techniques. Several researchers studied different leaky Lamb wave modes to evaluate various defects in the submerged plate structures. Santos and Perdigao [21] investigated the fundamental leaky symmetric Lamb wave mode (leaky S ) in a 0 pitch and catch configuration to detect and estimate the size of circular hole defects in bonded aluminum lap joints fully immersed in water. An empirical parameter that was defined based on the amplitudes of the received signals was shown to have a linear correlation with the dimensions of the defects. Chen, Su, and Cheng [22] investigated the propagation characteristics of the fundamental leaky anti- symmetric Lamb wave mode (leaky A ) in a submerged plate. Circular holes 0 created on the submerged plate mechanically and chemically were accurately detected by leaky A waves with the appropriate rectification for the medium 0 coupling effect. Rizzo et al. [23] used a pulsed laser focusing on the upper surface of the plate to excite leaky Lamb waves in an immersed aluminum plate. The signals received by immersed transducers included leaky S waves, quasi-Scholte waves, 0 and pressure waves. The leaky S mode that has the fastest propagation speed was 0 well separated from the other wave modes. Then, the leaky S waves were extracted 0 from the rest of the signals for further processing with continuous wavelet transform. Artificial defects, such as notches and circular holes, which were as small as a few millimeters, were successfully captured. Sharma and Mukherjee [24] studied leaky 83
ADE
Chapter 4 Lamb waves on an underwater steel plate using two immersed transducers inclined at specific angles. Three leaky Lamb wave modes were generated at their corresponding low-attenuation frequencies, which were the leaky S wave mode, 0 the first-order leaky symmetric Lamb wave mode (leaky S ), and the first-order 1 leaky anti-symmetric Lamb wave mode (leaky A ). Each of these wave modes 1 showed different sensitivities in monitoring the progressive notch damage machined on the underwater steel plate. They concluded that leaky S and leaky A 1 1 waves were more sensitive to surface defects, of which the depth was less than 37.5% of the plate thickness. In contrast, leaky S waves were more suitable for evaluating 0 the deeper defects since the amplitudes of the transmitted leaky S waves 0 consistently decreased with the notch depth. Sharma and Mukherjee [25] used similar techniques to monitor corrosion damage in an underwater plate. The initial surface degradation could be successfully identified by leaky S waves. Further 1 progression of the corrosion was evaluated better by leaky S waves. Takiy et al. 0 [19] conducted experimental measurements on a submerged aluminum plate to confirm the existence of leaky S , leaky A , leaky S , and the second-order leaky 0 1 1 symmetric Lamb wave mode (leaky S ) at their corresponding low-attenuation 2 frequency bands. Then, leaky S waves at 3.4 MHz-mm were selected for 1 characterizing damage. An image of the submerged plate was obtained to precisely identify the locations of five drilled holes. Xie, Ni, and Shen [26] proposed an experimental method to generate pure leaky S waves by applying the pulsed laser 0 radiation laterally at the whole side of an aluminum plate submerged in water. Through interacting with the damage, the generated leaky S waves can mode 0 convert into leaky A waves, which have mostly out-of-plane wave motions. Hence, 0 the damage can be easily recognized. However, clear edges of the real structures are not always accessible, making this pure wave mode excitation at the structural edges very challenging in practice. These studies demonstrated that it is feasible to use leaky Lamb waves for identifying damage in plate structures submerged in liquid. The presence of defects with sizes in the order of millimeters can be detected and quantified based on the linear features of leaky Lamb waves, such as the change of wave speed and amplitude. It was reported that the sensitivity of linear ultrasonic guided waves is limited to damage of a size comparable to the wavelength of the selected guided 84
ADE
Chapter 4 wave mode [27, 28]. However, they are insensitive to smaller defects in the initial damage stage. Before macroscopic cracks nucleate, the evolution of microstructural defects with load accounts for a major part of the total service life of a structure. In many cases, when the microstructural damage grows into macro scale, the remaining life of the structures is very short [29]. Therefore, it is better to identify defects sooner than later. Earlier detection of damage allows more time to characterize the evolution of the damage and schedule the maintenance actions for improving safety. Recent studies have proposed several nonlinear guided wave techniques to capture the microstructure evolution and early-stage material degradation [30]. Second harmonic generation is one of the most popular nonlinear techniques and has been successfully applied to different structures in gaseous environments to evaluate plasticity-induced damage [31, 32], thermal degradation [33, 34], precipitation [35], small fatigue cracks [36-38], debonding [39] and bolt loosening [40]. In the early damage stage (before the appearance of macro cracks), the generation of second harmonics takes advantage of the fact that the microstructural features in real materials distort the passing sinusoidal ultrasonic waves. The distortion can generate new wave components at frequencies other than the excitation frequencies, which provides a way for the evaluation of the microstructural defects. However, the measurement of second harmonics is challenging due to the fact that the material nonlinearity is weak [41]. To ensure measurable generation of second harmonics, the incident waves should have finite amplitudes so that there is sufficient wave energy to interact with the microstructural features. In addition, wave mode selection is required to ensure that the primary waves and second harmonics conform to non-zero power transfer and phase velocity matching conditions [42-47]. The aforementioned studies on the second harmonic generation were carried out on the structures in gaseous environments. Although the nonlinear characteristics of guided waves have been demonstrated by a number of studies to be more sensitive to the microstructural defects in the early stage of damage and are less influenced by environmental changes, there have been very limited studies on the use of nonlinear guided waves for damage detection on the submerged structures. Undoubtedly, the vibration of the plate submerged in liquid behaves differently from that in the air [48-50]. Compared with structures in gaseous 85
ADE
Chapter 4 environments, the generation and measurement of second harmonics on the submerged structure are more challenging because the wave energy can be absorbed by the surrounding liquid medium. So, the application of nonlinear guided waves for submerged structures deserves separate and careful studies considering the high reward for the earlier detection of material degradation. This paper presents experimental and numerical investigations on the feasibility of using second harmonic generation to evaluate material degradation in metallic plates with one side in contact with water. Leaky S mode is selected to 0 generate second harmonic leaky S waves due to the observations from previous 0 studies that leaky S mode at low frequencies has very low attenuation and its phase 0 velocity decreases slowly with frequency [51-53]. These features make the primary and second harmonic leaky S waves satisfy non-zero power flux and approximate 0 phase velocity matching conditions. Then, experiments are carried out on a metal tank filled with water. Leaky Lamb waves are generated on the wall of the water- filled tank by a piezoceramic transducer and measured by a scanning laser vibrometer. It is demonstrated that second harmonics can be generated by leaky S 0 waves at low excitation frequencies and the corresponding relative nonlinearity parameters are growing linearly with the propagation distance. The growth rate of the relative nonlinearity parameters can be used to characterize the material stratus of the one-side water-submerged plate. After that, a three-dimensional (3D) finite element (FE) model is developed with the material nonlinearity of the submerged plate simulated by the Murgnahan strain energy function. The material properties of aluminum at different levels of fatigue damage are obtained from previous experimental results [54]. The numerical simulations are validated through the experimental data. Next, the experimentally validated 3D FE model is employed in the parametric study to analyze the second harmonic generation in the submerged plate at different levels of fatigue damage. The results show that leaky S mode at 0 low frequencies can generate measurable second harmonics, which are sensitive to the change of material properties of the one-side water-immersed plate at the initial stages of the fatigue damage. This paper is organized as follows. Section 4.3 introduces the second harmonic generation techniques and discusses the selection of leaky Lamb modes for the generation of second harmonics in the plate with one side exposed to water. Section 4.4 describes the experimental study. Section 4.5 86
ADE
Chapter 4 the nonlinear material can be written in terms of the displacement gradient as a Taylor series expansion truncated at order 2 1 Eu' Eu'2 (4.1) 2 where , u , E , and  are the stress, displacement, Young’s modulus of the medium, and nonlinearity parameters, respectively. u'u x is the displacement gradient. The particle motions can be described as 'u (4.2) where  represents the mass density; ' x and u2u t2 . Perturbation theory is employed to solve Eqs. (4.1) and (4.2), leading to the final solution [7] u A cos(kxwt)A cos(2kx2wt) (4.3) 1 2 where k and w are the wavenumber and angular frequency of the excited primary 1 waves. A is the amplitude of the primary waves at w. A  k2A2x represents 1 2 8 1 the amplitude of the second harmonics at 2w. x is the propagation distance. Two observations can be obtained from Eq.(4.3). Firstly, the primary waves and the second harmonics should have the same phase velocity, i.e., c w k 2w 2k. Secondly, when the structure is excited by waves with a fixed p wavenumber value k , the nonlinearity parameter  that correlates to the material nonlinearity can be determined by measuring the magnitudes of the primary waves and the second harmonics as 8 A  2 (4.4) k2x A2 1 For practical applications, the change in  with its initial value is more important than its absolute value. Therefore, a relative nonlinearity parameter ' is defined as A ' 2 x (4.5) A2 1 88
ADE
Chapter 4 It can be seen from Eq.(4.5) that ' is proportional to  and grows linearly with the propagation distance x.  can be evaluated by the gradient of accumulation of '. Therefore, any abnormal increase in ' indicates an increase in the material nonlinearity and progress in material degradation [36]. Although Eq. (4.1)-Eq. (4.4) are derived for longitudinal waves, Eq.(4.5) has been widely considered to be applicable for characterizing the second harmonic generation of Lamb waves [37, 44], Rayleigh waves [55], and Edge waves [56]. Therefore, ' is employed in this study to quantify the change of the second harmonics of the leaky Lamb waves. 4.3.2. Selection of primary leaky Lamb wave modes Although the generation of second harmonics by guided waves has been studied for evaluating the incipient damage in various structures that are open to the air, the feasibility of using leaky Lamb wave modes to generate second harmonics in the plate with one side exposed to water has not been explored. Considering the multi- modal and dispersive features, the selection of primary leaky Lamb waves is important. This section introduces the theoretical derivation of the dispersion curves, which describe the number of leaky Lamb wave modes and their corresponding properties with the frequency. Based on the dispersion curves, the leaky Lamb wave modes that have low attenuation and low dispersion characteristics are identified. Then, the primary wave mode and excitation frequency are selected to meet the following three conditions. Firstly, the primary wave mode should have sufficient wave energy propagating in the submerged structure to interact with the microstructural features so that the generation of second harmonics best reflects the material nonlinearity. Secondly, the primary waves should be of the same type of wave mode as the second harmonics to ensure that the wave energy can be transferred between the primary waves and the second harmonics [46]. Thirdly, the primary waves and the second harmonics should have similar phase velocities. Consider a plate loaded with water on its bottom surface, as shown in Figure 4.2. Traction-free boundary conditions apply to the top surface of the plate as Plate_0 Plate_0 0 (4.6) 33 31 89
ADE
Chapter 4 where Plate_0 and Plate_0 are the normal and shear stress on the top surface of the 33 31 plate, respectively. The bottom surface of the plate is coupled to the water layer. Under the non-viscosity assumption that water cannot sustain shear forces, the boundary conditions at the plate-water interface can be described as uPlate_d uWater_d 33 33 Plate_d Water_d (4.7) 33 33 Plate_d 0 31 where uPlate_d , Plate_d , and Plate_d represent the normal displacement, normal 33 33 31 stress, and shear stress of the plate at the plate-water interface, respectively. uWater_d 33 and Water_d are the normal displacement and normal stress of the water at the plate- 33 water interface, respectively. Figure 4.2. Schematic diagram of a plate loaded with water on its bottom surface Previous studies have derived the characteristic equation of the leaky Lamb waves for the one-side water-immersed plate [57] q2 k2 q2 k2 2qk 2qk 0 2pk 2pk q2 k2 q2 k2 0 w2 (q2 k2)eipd (q2 k2)eipd 2qkeiqd 2qkeiqd w (4.8)  0 2pkeipd 2pkeipd (q2 k2)eiqd (q2 k2)eiqd 0 w2 peipd peipd keiqd keiqd k2 c2 w 90
ADE
Chapter 4 where p w2 c2 k2 , q w2 c2 k2 , c  2  , c   . L S L S E  112  and  E  21  are the first and second Lame constants of the plate, respectively. E and  are Young’s modulus and Poisson’s ratio of the plate, respectively.  and  represent the density of the plate and the w surrounding water, respectively. c is the speed of the bulk wave in the water. w Eq.(4.8) can be solved numerically and the solutions can be presented by a series of dispersion curves. Table 4.1. Material properties of the aluminum plate and water Aluminum density  (kg/m3) 2700 Aluminum 1st Lame parameter  (GPa) 51.64 Aluminum 2nd Lame parameter  (GPa) 26.60 Water density  (kg/m3) 1000 w Water bulk wave velocity c (m/s) 1500 w Figure 4.3 shows the dispersion curves of a 1.6 mm thick aluminum plate with one side in contact with water. The material properties are given in Table 4.1. Within the frequency range up to 1MHz, there are only three wave modes, which are leaky S , leaky A , and quasi-Scholte waves. Other higher-order wave modes 0 0 appear when the frequency increases. The quasi-Scholte wave that propagates along the plate-water interface has very low attenuation (close to zero) for the entire frequency bandwidth as shown in Figure 4.3(c). This wave mode is highly dispersive in the low-frequency range (below 250 kHz) as denoted by the blue dotted lines in Figures 4.3(a) and 4.3(b). The phase velocity of the quasi-Scholte wave increases monotonically with frequency until its value reaches the speed of the bulk wave in the surrounding water (around 1500 m/s). Further increasing the frequency, the quasi-Scholte wave becomes nondispersive. Although the quasi- Scholte wave at the frequency range over 250 kHz has low attenuation and satisfies phase velocity matching conditions, it is not considered in this study. The reason is that the quasi-Scholte wave in the nondispersive frequency range has most of its 91
ADE
Chapter 4 In addition to quasi-Scholte waves, leaky S waves also have low 0 attenuation and low dispersion characteristics within the frequency range up to 600 kHz. The phase velocity of leaky S waves decreases slowly with frequency. For 0 example, the phase velocity of the leaky S mode at 100 kHz is 5435 m/s, and that 0 at 500 kHz is 5383 m/s. The deviation is around 0.96%. A previous study defined the approximate phase velocity matching condition as the relative phase velocity deviation less than 1% [59]. Therefore, if the leaky S waves below 250 kHz are 0 selected as the primary waves, the second harmonics at twice the frequency should satisfy the approximate phase velocity matching condition with the primary waves. Another key factor is that the low-attenuation frequency band of the leaky S mode 0 is limited to the frequency range up to 600 kHz as shown by the black solid line in Figure 4.3(c). As the frequency increases beyond 600 kHz, the attenuation of the leaky S wave increases exponentially. Also, the phase velocity and group velocity 0 decrease quickly with the frequency. Figure 4.4 presents the mode shapes of leaky S waves for the aluminum 0 plate with one side in contact with water. The mode shape diagrams show the distributions of the displacements of leaky S waves through the thickness of the 0 plate. The x-axis denotes the magnitudes that are normalized by the maximum displacement amplitudes. The y-axis denotes the thickness location x that is 3 defined in Figure 4.2. The red dashed lines represent the particle’s displacements in the direction parallel to the plate surface (in-plane displacements). The blue solid lines show the displacement in the direction normal to the plate surface (out-of- plane displacements). It can be seen that the out-of-plane displacements between the plate and the water areas are continuous. In comparison, the in-plane displacements are disconnected. When the frequency is below 600 kHz, the mode shape of leaky S waves is dominated by the in-plane displacements in the plate as 0 indicated by the red dashed lines in Figure 4.4(a). The vast majority of the wave energy is conserved in the one-side water-immersed plate with minimal loss, making it ideal for using the leaky S wave to scan the one-side water-immersed 0 plate. However, the out-of-plane displacements in both the plate and water regions increase with the frequency. This indicates more energy leakage from the structure into the surrounding liquid medium. From these observations, the excitation 93
ADE
Chapter 4 frequencies are chosen to be below 250 kHz in the following experimental and numerical studies. Figure 4.2. Mode shapes of leaky S wave at (a) 300 kHz, (b) 600 kHz, and (c) 900 0 kHz for a 1.6 mm thick aluminum plate loaded with water on its bottom surface (the red dashed lines represent the in-plane displacements and the blue solid lines denote the out-of-plane displacements) 4.4. Experimental study 4.4.1. Experimental setup This section presents an experimental study on the leaky Lamb wave propagation in a metal plate with one side in contact with liquid, which aims to simulate a variety of thin-walled structures operating in extreme conditions, such as nuclear cooling pipes, pressure vessels, rocket fuel tanks, and submarine hulls. The experiments were carried out using a metallic tank fully filled with water. The front wall of the tank was used as the test plate, which was a 1.6 mm thick aluminum plate. The internal surface of the test plate was in contact with water, while the outer surface was exposed to air. Figure 4.5 illustrates the overall experiment setup and the top view of the water-filled tank. A computer-controlled signal generator (NI PIX-5412) was employed to generate a six-cycle Hanning window modulated sinusoidal tone burst pulse. Then, the signal was sent to a power amplifier (Ciprian HVA-800-A) and the voltage was increased to 160 Vp-p. After that, the amplified signal was sent to a 94
ADE
Chapter 4 piezoceramic transducer (Ferroperm Pz27) which was bonded to the outer surface of the test plate. The circular piezoceramic transducer has a diameter of 10 mm and a thickness of 0.5 mm and can convert the electric signals to mechanic motions, exciting leaky lamb waves on the test plate. The thin and circular piezoceramic wafer deforms mainly in the radial direction which is parallel to the plate surface. As a result, the excitation should be dominated by the in-plane motions of the plate and leaky S waves could be generated effectively. 0 The response signals were collected on the water-free surface by a non- contact scanning laser Doppler vibrometer (Polytec PSV-400-M2-20). Taking the center of the piezoceramic transducer as the origin, a Cartesian coordinate system was defined as shown in Figure 4.5. The x -axis denotes the in-plane direction 1 parallel to the plate surface, and the x -axis is the out-of-plane direction that is 3 normal to the plate surface. The measurement points are defined along a line parallel to x -axis. The signals were collected at a sampling rate of 25.6 MHz. To improve 1 the quality of measurements, each signal was averaged by 1000 recordings and filtered by a low-pass filter with a cut-off frequency of 1MHz. Figure 4.5. Schematic diagram of the experiment setup 95
ADE
Chapter 4 4.4.2. Experimental results Figure 4.6(a) presents an example of the experimentally measured signals. The excitation frequency was 170kHz, at which there were only leaky S , leaky A , and 0 0 quasi-Scholte waves, as shown by the dispersion curves in Figure 4.3. The measurement point was 200 mm away from the excitation center. The first wave packet was identified as the leaky S wave mode, which has the fastest group 0 velocity, as shown in Figure 4.3(b). The phase velocity of the leaky S wave mode 0 at 170 kHz is 5431 m/s, and that at 340 kHz is 5413 m/s. So, the primary leaky S 0 wave is almost phase matched with its second harmonic with a deviation of 0.33%. It should be noted that the experimental data mainly captured the out-of-plane motions on the plate surface because the laser beam was perpendicular to the test plate during the test. Since the out-of-plane displacement of the leaky S wave was 0 small as shown by the blue solid lines in Figure 4.4, the actual magnitudes of the leaky S wave should be strong enough so that it could be measured by the scanning 0 laser Doppler vibrometer. The following wave packet should be dominated by the quasi-Scholte wave mode because the leaky A wave mode decays quickly due to high attenuation [51]. 0 At this excitation frequency, the quasi-Scholte wave is highly dispersive. The phase velocity of the quasi-Scholte mode at 170 kHz is 1317 m/s, and that at 340 kHz is 1488 m/s. The deviation of the phase velocity between the primary waves and the second harmonics is around 13%. Therefore, the quasi-Scholte waves do not satisfy the approximate phase velocity matching condition [59]. To further investigate, the first wave packet was cut from the rest of the signal to exclude the quasi-Schole waves and the unwanted reflections in the data. Figure 4.6(b) presents the window-cut signal. Then, the chopped signal was transferred to the frequency domain by FFT and is shown in Figure 4.6(c). There are two peaks for the primary waves (at 170 kHz) and the second harmonics (at 340 kHz), respectively, as highlighted by the black dotted line in Figure 4.6(c). There is also a small peak at three times the excitation frequency, which is the third harmonic [29, 60]. However, the signal-to-noise ratio of the third harmonics is much lower than that of the second harmonics, and the third harmonics are not the focus of the paper, so the third harmonics are not discussed in this study. After that, ' were 96
ADE
Chapter 4 4.5. Finite element simulation 4.5.1. Model description A 3D FE model was developed to simulate the leaky Lamb wave propagation in the one-side water-immersed plate. By applying the symmetry boundary conditions to the left and bottom edges of the plate, only the top right part of the plate was modeled. The plate modeled in the FE was 430 mm long, 250 mm wide, and 1.6 mm thick. The bottom surface was in contact with a water layer of the same planar area. The thickness of the water layer was 90 mm, which was chosen to avoid unwanted reflections from the bottom of the water layer. Figure 4.7 presents the schematic diagram of the FE model, which was modeled using the commercial software, ABAQUS. The bottom surface of the plate and the top surface of the water layer were tied together using the surface-based tie constraint, which connected the acoustic pressure of the water and the out-of-plane translations of the plate. Previous studies of ultrasonic guided waves in the solid-liquid coupled medium experimentally validated that the tie constraint could accurately model the solid-liquid interactions [22, 51, 58, 61]. The plate and the water layers were modeled by 3D eight-node reduced integration solid elements and 3D eight-node reduced integration acoustic elements, respectively. The largest dimension of the element size was less than 0.5 mm, ensuring that there were at least 20 FE nodes within the wavelength of the leaky S wave [59, 62]. 0 Figure 4.7. Schematic diagram of the 3D FE model. 98
ADE
Chapter 4 To simulate the material nonlinearity of the one-side water-immersed plate, the Murnaghan strain energy function was introduced to define the material properties of the aluminum plate using VUMAT subroutine in ABAQUS. Murnaghan strain energy function includes the third-order Taylor series expansion of the strain potential and has been widely used in the analysis of second harmonic generation for modeling the nonlinear material behaviors [62-64]. This function can be written as 1 1 WE  trE2 tr E2  l2m trE3 2 3 (4.9) mtrE  trE2 tr E2 ndetE where l, m, and n are the Murnaghan constants which are related to the third- 1 order elastic constants; E  FTFI is the Lagrangian strain; F and I denote 2 the deformation gradient and the identity tensor, respectively. Table 4.2 presents the values of Murnaghm constants of aluminum at different levels of fatigue damage, which are obtained from previous experimental results [54]. Table 4.2. Murnaghan constants of aluminum at different levels of fatigue damage [54] Murnaghan constants Fatigue life l (GPa) m (GPa) n (GPa) 0% -252.2 -325.0 -351.2 40% -266.8 -332.8 -358.3 80% -271.2 -335.8 -359.8 The leaky S wave was excited by applying nodal displacements at the 0 circumference of a circular transducer represented by the quarter-circle of 10 mm diameter located at the left-bottom corner of the plate. The excitation signal was a six-cycle Hanning window modulated sinusoidal tone burst pulse. The explicit module of ABAQUS was employed to solve the dynamic simulations. Figure 4.8 99
ADE
Chapter 4 shows a snapshot of the simulation results with the excitation frequency of 170 kHz. The rainbow color represents the acoustic pressure in the water layer. It can be seen that the leaky S wave propagates fastest with minimum wave energy leaking into 0 the liquid. Following the leaky S wave, the quasi-Scholte wave propagates at a 0 speed slightly faster than the pressure wave in water. The acoustic pressure of quasi- Scholte waves is concentrated around the plate-water interface [58]. The leaky A 0 wave dominated by the out-of-plane displacements was not observed from the surface of the submerged plate. In general, the simulation results have a good agreement with the experimental data as shown in Figure 4.6(a). The numerically calculated acoustic wavefields provide additional information to interpret the experimental data. The simulation results of the 3D FE model are further validated in the following sections to gain physical insights into the second harmonic generation on the one-side water-immersed plate. Figure 4.8. Snapshot of the simulation results at 72 s 4.5.2. Experimental validation of the FE model In this section, the 3D FE model is validated by comparing the simulation results with the experimental data. The simulations carried out using the 3D FE model with the VUMAT subroutine are labeled as nonlinear FE because the material nonlinearity was modeled by the Murnaghan strain energy function. Firstly, the linear features of the leaky S waves that were simulated by nonlinear FE and 0 100
ADE
Chapter 4 measured from experiments were investigated and compared. From both the experiments and the nonlinear simulations, the out-of-plane displacements were obtained at 21 measurement points along a line from 200 mm to 300 mm away from the excitation center. Figure 4.9(a) shows the signal simulated by the nonlinear FE. The measurement point was 200 mm away from the excitation center. Figure 4.9(b) compares the waveforms of the window-cut signals obtained from the experiments and the simulations from the nonlinear FE at the same measurement point. The amplitudes are normalized by the maximum peak magnitudes of the signals. In general, the nonlinear FE well predicts the waveform and the time of arrival of leaky S waves. 0 Figure 4.9. (a) Signal simulated by the nonlinear FE (b) Comparison of the window- cut signals measured from the experiment and simulated by the nonlinear FE. To further validate the accuracy of the nonlinear FE, the group and phase velocities were calculated using the simulated signals, and the simulation results were compared with theoretical values and experiment data. The excitation frequency was swept from 150 kHz to 390 kHz in steps of 20 kHz. The signals were collected at the first 10 points for each excitation frequency to calculate the averaged phase and group velocities. The distance between the two consecutive measurement points was 5 mm and it was less than half of the wavelength of the C selected leaky S wave. The phase velocity was calculated by 0 p C  f 2f  x, where f represents the central frequency of the excitation. p c c c 101
ADE
Chapter 4  and x are the phase difference and the distance between the two measurement points, respectively. The group velocity C was calculated by g C  f x t, where t is the time lag between the two measurement points. g c Figures 4.10(a) and 4.10(b) present the phase velocity and group velocity dispersion curves for the aluminum plate loaded with water on the single side, respectively. In both figures, there are three wave modes represented by three lines, which are calculated based on the global matrix theory by the commercial software DISPERSE [65]. The black solid lines on the top represent the leaky S mode. The 0 values calculated by the experimental data and the nonlinear FE simulations are denoted by cycles and stars, respectively. It can be seen that the simulation results have a good agreement with the theoretical derivations and experimental measurements. The maximum deviation is less than 2%. Figure 4.10. (a) Phase velocity dispersion curves and (b) group velocity dispersion curves calculated by the theoretical derivations (black solid line, black dashed line, and blue dotted line), nonlinear finite element simulations (stars), and experimental measurements (circles) Then, the nonlinear features of the simulated and experimentally measured leaky S waves were analyzed. The simulations were carried out using the 3D FE 0 model with and without the VUMAT subroutine. The simulations, solved using only the linear elastic material properties as shown in Table 4.1, are labeled as linear FE because they do not consider the inherent material nonlinearity. The nonlinear 102
ADE
Chapter 4 FE simulates the material nonlinearity of the aluminum plate by introducing the Murnaghan constants of zero fatigue damage in Table 4.2. Figure 4.11(a) compares the window-cut signals obtained from the experimental measurements and the simulated signals from both the linear and nonlinear FE models. The amplitudes are normalized by their corresponding peak magnitudes. The simulated linear and nonlinear signals do not show much difference in the time domain, and both have a good agreement with the experimental data. Figure 4.11(b) shows the corresponding data in the frequency domain. It can be seen that nonlinear FE captures the second harmonics at twice the excitation frequency, as highlighted by the black dotted lines in the figure. In general, the second harmonics simulated by the nonlinear FE have a good agreement with the experimental data (see the red dash-dotted line and black solid line in Figure 4.11(b)). However, the linear FE that uses only the second-order elastic constants could only predict the primary waves as shown by the blue dashed lines. The second harmonics are not observable in the linear FE. The comparison between the linear and nonlinear FE further confirms that the generated second harmonics are due to the material nonlinearity by considering the Murnaghan strain energy function and Murnaghan constants. Figure 4.11. (a) Comparison of the time domain signals obtained from the experimental measurement, linear finite element simulation, and nonlinear finite element simulation; and (b) their corresponding frequency spectra. It should be noted that the Murnaghan strain energy function, incorporating the third-order approximation of the constitutive relation, could be used to predict 103
ADE
Chapter 4 the generation of only up to the second harmonics. To study the third-order harmonics, the fourth-order expansion of the constitutive relation and fourth-order elastic material constants should be included in the constitutive model [60]. This explains why the nonlinear FE is unable to capture the peak located at three times the excitation frequency. Since this study focuses on the generation of the second harmonics, the Murnaghan strain energy function is sufficient for analysis. Therefore, the simulation results are validated. 4.5.3. Second harmonic generation in the submerged plate at different levels of fatigue damage The experimentally validated 3D FE model was employed to explore the influence of evenly distributed fatigue damage on the generation of the second harmonics. Stobbe [54] experimentally measured the values of Murnaghan constants for aluminum at different fatigue levels. In his experimental studies, a series of dog- bone samples made of aluminum were fatigued by repeated uniaxial tensile loads. One sample was loaded to 52800 load cycles and failed, which was defined as 100% fatigue damage. The rest samples were then fatigued to different cycles and were referenced to different percentages of fatigue damage. The numerical simulations in the present study employed the experimentally measured Murnaghan constants for the undamaged aluminum and the aluminum at 40% and 80% fatigue damage [54], as shown in Table 4.2. They have been used in previous studies to simulate the generation of nonlinear guided waves in aluminum plates [59] and pipes [64] in gaseous environments. For the first time, this paper presents the numerical simulations of an aluminum plate with one side loaded with water and the sensitivity of the second harmonics generated by leaky S waves to fatigue damage 0 is investigated. Three simulations were carried out using the experimentally validated 3D FE model with three different sets of Murnaghan constants (see Table 4.2), which represent the aluminum plates at 0%, 40%, and 80% fatigue damage. The other settings remained unchanged. The simulated out-of-plane displacements were obtained at 21 points from 200 mm to 300 mm away from the excitation center to calculate '. For direct comparison, the values are normalized by the initial value at 200 mm for the 104
ADE
Chapter 4 aluminum plate at zero fatigue damage. Figure 4.12(a) presents the normalized nonlinear parameters versus the propagation distance. It can be seen that the normalized nonlinearity parameters increase linearly with the propagation distance for the three cases. In addition, the slopes of the best-fit lines increase with the fatigue damage levels. Figure 4.12 (a) The normalized nonlinearity parameters versus propagation distance and (b) normalized slopes of the best-fit lines at different levels of fatigue damage Figure 4.12(b) compares the slopes of the best fit lines at different levels of fatigue damage. The values are normalized by the initial value at zero fatigue damage. As mentioned in Section 4.3.1, the material nonlinearity can be evaluated by the gradient of accumulation of '. Thus, it shows that material nonlinearity increases quickly during the initial stage of fatigue damage and the increasing rate becomes much slower after 40% fatigue damage. This behavior is in agreement with the preceding studies on fatigue damage evaluation by longitudinal waves [66, 67], Rayleigh waves [68, 69], and Lamb waves [33, 44, 70]. Therefore, the numerical simulations reveal that second harmonics generated by the low- attenuation leaky S waves have the potential to characterize the material 0 nonlinearity of the plate when one side of the plate is exposed to water. 105
ADE
Chapter 4 4.5.4. Comparison between the free plate and water-immersed plate The experimentally validated 3D FE model was also employed to explore the influence of the surrounding liquid on the generation of the second harmonics. Two simulations were carried out for the undamaged aluminum plate with and without the water layer, respectively. The simulation for the plate without the water layer is labeled as the free plate, while that for the plate with one side exposed to water is labeled as the water-immersed plate. The other settings remained the same. Guided waves that propagate in the free plate are called Lamb waves, which consist of multiple symmetric and antisymmetric Lamb wave modes. Figure 4.13(a) shows the simulated signals for the free plate. The first wave packet is identified as the fundamental symmetric Lamb (S ) mode that propagates fastest in the selected 0 excitation frequency. The second wave packet that arrives around 80 s is the fundamental antisymmetric Lamb (A ) mode. It propagates slightly faster than the 0 quasi-Scholte wave as shown in Figure 4.9(a). Figure 4.13(b) compares the window-cut signals for the free plate and water-immersed plate, respectively. Their corresponding frequency spectra are shown in Figure 4.13(c). The amplitudes are normalized by the corresponding peak magnitudes for comparison. It can be seen that the leaky S wave has a similar waveform as the S wave in the time domain 0 0 due to the fact that they have similar group and phase velocities [71, 72]. However, the ratios of the second harmonics to the primary waves are smaller for the leaky S 0 wave compared to that of the S wave. This is because the attenuation of the leaky 0 S wave gradually increases with the frequency as discussed in Section 4.3.2. This 0 phenomenon also has a significant influence on the growing trends of the nonlinearity parameters as shown in Figure 4.13(d). The slope of the best-fit line of the free plate is 7.0876, which is nearly double the value of the one-side water- immersed plate. Thus, the simulation results reveal that the surrounding liquid can reduce the accumulation rate of the nonlinearity parameters. In addition, the influence of water is much stronger than that caused by fatigue damage as discussed in Section 4.5.3. So, the liquid coupling effects should be considered when the nonlinear guided wave techniques are used for damage detection for the immersed plate structures. 106
ADE
Chapter 4 Figure 4.13 (a) Signal simulated by the free plate FE; (b) Comparison of the window-cut signals simulated by the free plate FE and water-immersed plate FE; (c) frequency spectra of the signals in (b), and (d) comparison of the normalized nonlinearity parameters versus propagation distance. 4.6. Conclusion This paper has investigated experimentally and numerically the second harmonic generation by guided waves in plates immersed in liquid on one side, which has the potential to characterize the microstructural evolution before the appearance of macroscale damage and fraction. The findings can provide support for the further development of NDT techniques for partially submerged structures, such as nuclear cooling pipes, pressure vessels, rocket fuel tanks, storage tanks, and submarine hulls. Earlier damage detection of these partially immersed structures allows more time to schedule the maintenance actions and reduces the risks of in-service failure. Firstly, the dispersion behavior of guided waves has been analyzed for metallic plates with one side immersed in water. It has been found that the leaky S 0 mode at low frequencies has low attenuation and low dispersion features. This 107
ADE
Chapter 4 analysis leads to the selection of leaky S to generate second harmonics of the same 0 type of wave mode in the one-side water-immersed plate because the primary and second harmonic leaky S waves satisfy approximate phase velocity matching and 0 non-zero power flux conditions. Next, experimental studies have been conducted on the metal tank filled with water. A case study using experimentally measured signals at the excitation frequency of 170 kHz has been presented. Both the primary and second harmonic leaky S waves can be identified in the frequency spectrums. 0 In addition, the relative nonlinearity parameter ' has been calculated and shown to grow linearly with the propagation distance. The experimental results confirm that leaky S waves can generate measurable second harmonics due to the material 0 nonlinearity of the one-side water-immersed plate. After that, numerical simulations have been carried out using a 3D FE model and validated through experimental measurements. The experimentally validated 3D FE model has been employed in parametric studies to explore the second harmonic generation in the one-side water-immersed plate at different levels of fatigue damage. The results have shown that the second harmonic generation techniques are promising for non- destructively evaluating microstructural defects in plate structures with one side immersed in liquid. The present study only demonstrates that the interaction between guided waves and microstructural features of partially immersed plates can generate measurable and low-attenuation second harmonics. When the microscopic defects grow into macro scale, there can be a substantial increase in amplitudes for the nonlinear guided waves due to the clapping behavior between the surfaces as the primary guided waves pass through. For the structures in gaseous environments, the clapping effect of macro cracks is classified as contact-type nonlinearity, which has been demonstrated to generate second harmonics with much larger amplitudes than the material nonlinearity [28, 37]. Future studies can investigate the effect of the size, shape, and location of macro cracks (e.g. stress corrosion cracking) on the nonlinear guided wave features for the structures immersed in liquid. 108
ADE
Chapter 5 Chapter 5. Structural health monitoring of partially immersed metallic plates using nonlinear guided wave mixing Abstract Metallic plates are important structural components of many liquid containment structures, such as liquid storage tanks and sewer pipes. Time-dependent loads can result in fatigue and degradation of the metallic material. Nonlinear guided wave mixing has been demonstrated to be sensitive to microstructural change at the early stage of material degradation. Previous studies have been carried out using the nonlinear guided wave mixing technique on various structures in gaseous environments. However, its application to structures immersed in liquid has not been explored. This paper numerically and experimentally investigates the nonlinear guided wave mixing in an aluminum plate loaded with water on one side. Experiments are carried out with an empty metal tank and the tank filled with water, respectively. The results show that cumulative generation of harmonics at the sum frequency due to the material nonlinearity of the partially immersed plate can be achieved by mixing the fundamental leaky symmetrical Lamb (leaky S ) waves at 0 two different frequencies. Under the same experimental conditions, the amplitudes of the guided wave signals and the values of the relative nonlinearity parameters on the partially immersed plate are different from their counterparts on the plate without water. Finally, numerical simulations are performed with the material nonlinearity of the test plate simulated by the Murnaghan constitutive model. The numerical results reveal that both the second harmonics and the combination harmonics are sensitive to the material nonlinearity of the plate loaded with water on one side. Keywords: Structural health monitoring; Water containment structures; Metallic plates; Leaky Lamb waves; Nonlinear guided waves; Guided wave mixing; Second harmonics 115
ADE
Statement of Authorship Title of Paper Structural health monitoring of partially immersed metallic plates using nonlinear guided wave mixing Publication Status Published Accepted for Publication Unpublished and Unsubmitted work written in Submitted for Publication manuscript style Publication Details X. Hu, T. Yin, H. Zhu, C.T. Ng, A. Kotousov, (2022). Structural health monitoring of partially immersed metallic plates using nonlinear guided wave mixing. Construction and Building Materials (In-print). Principal Author Name of Principal Author (Candidate) Xianwen Hu Contribution to the Paper Conceptualization, Developing and validating numerical models, Conducting experimental measurements, Signal processing and data analysis, Writing the original draft and editing. Overall percentage (%) 80% Certification: This paper reports on original research I conducted during the period of my Higher Degree by Research candidature and is not subject to any obligations or contractual agreements with a third party that would constrain its inclusion in this thesis. I am the primary author of this paper. Signature Date 07/03/2022 Co-Author Contributions By signing the Statement of Authorship, each author certifies that: i. the candidate’s stated contribution to the publication is accurate (as detailed above); ii. permission is granted for the candidate in include the publication in the thesis; and iii. the sum of all co-author contributions is equal to 100% less the candidate’s stated contribution. Name of Co-Author Tingyuan Yin Contribution to the Paper Analytical derivations, Writing – review & editing. Signature Date 07/03/2022 Name of Co-Author Hankai Zhu Contribution to the Paper Experimental measurements, Writing – review & editing. Signature Date 07/03/2022
ADE
Chapter 5 5.2. Introduction Metallic plates are commonly used for constructing undersea tunnels [1], storage tanks [2], sewer pipes [3], and containment buildings [4]. These structures have one side immersed in liquid and are subjected to cyclic loads with varying amplitudes. Material degradation and fatigue are the primary culprits for the failure of these partially immersed metallic structures [5]. In the early damage stage, dislocations and slip bands occur in the materials and then micro cracks are formed. With the increase in loading cycles, the micro cracks continue to accumulate and grow to a critical point, which can cause catastrophic failures [6, 7]. Continuous evaluation of material properties of partially immersed metallic plates is crucial to maintain the structural integrity of high-valued infrastructures. Guided wave testing is a non-destructive inspection technique that has attracted extensive research interest. It outperforms other non-destructive testing methods, such as eddy current testing, acoustic emission, and conventional ultrasonic testing, because guided waves can travel relatively long distances on various structures and have a high sensitivity to different kinds of damage [8]. The structural health can be monitored by both linear and nonlinear features of guided waves. Conventional guided wave testing is based on linear features. Specifically, the presence of defects changes the transmitted guided wave signals, typified as scattering, mode conversion, attenuation, and change in wave velocity. Linear guided wave testing was successfully applied to immersed structures to characterize cracks [9], pits [10], notches [11], and corrosion [12, 13]. In these studies, guided waves in immersed structures were shown to behave differently from their counterparts in structures without exposure to liquid. In addition, the sizes of the defects were around a few millimeters, which were comparable to the wavelength of the selected guided wave modes. However, linear guided waves are insensitive to smaller defects such as micro cracks and dislocations in the early damage stage. When the micro cracks evolve into macro cracks and become identifiable through the linear guided wave features, the metallic structures, in many cases, reach more than 80% of its total service life [14]. Recent studies have focused on the nonlinear features of guided waves, which provide much better sensitivity than the linear features in detecting 118
ADE
Chapter 5 microstructural defects that precede the macro-scale damage [15]. The second harmonic approach is one of the most popular nonlinear guided wave methods. When the structural material is excited by guided waves with finite amplitudes, the ultrasonic guided wave energy can be transferred from the excitation frequency to twice the excitation frequency due to the interaction of the primary guided waves with the microstructural features of the material. This phenomenon provides a way to identify and characterize material degradation at its early stage. For example, the second harmonics generated by guided waves were used to evaluate the evolution of thermal aging [16, 17] and fatigue [18, 19] for metallic plates in gaseous environments. A comprehensive review of the second harmonic guided wave approach can be found in [15]. The major difficulty hindering the applications of second harmonics is that the instrumentation of the measurement system can also produce nonlinear signals at the integer multiples of the excitation frequency [20, 21]. It is difficult to distinguish the nonlinearity due to the material from the nonlinearity caused by the instruments. To tackle this limitation of the second harmonic approach, a number of researchers proposed the nonlinear guided wave mixing technique, in which the structural material is excited by guided waves with two different frequencies. Hasanian and Lissenden [22] conducted a wave vector analysis for the mutual interaction of two guided waves with different frequencies. The mutual interaction can generate combination harmonics at the sum and difference frequencies that are far from the nonlinear waves produced by the instrumentation. Hasanian and Lissenden [23] further studied the internal resonance criteria for the non-collinear guided wave interaction, where the guided waves propagate in different directions and meet in a localized mixing zone. They concluded that the amplitudes of the generated combination harmonics are dependent on the size of the wave mixing areas. Jiao et al [24] demonstrated both experimentally and numerically that the combination harmonics of guided waves at the sum frequency are sensitive to micro cracks in metallic plates. Metya et al [25] revealed that the nonlinear guided wave mixing technique can also evaluate localized deformation of a steel plate during creep. Shan et al [26] mixed two shear horizontal waves that propagated in the same direction to generate cumulative combination harmonics at the sum frequency. The combination harmonics demonstrated a high sensitivity to degradation of the 119
ADE
Chapter 5 aluminum plate during early fatigue stages. Cho et al [27] proposed a novel technique to detect localized fatigue damage in aluminum plates by the interaction of two counter-propagating shear horizontal waves. The wave mixing area can be controlled and moved to different locations on the sample by adjusting the time delays of the input signals. Thus, the whole area of the plate can be scanned. Li et al [28] employed the guided wave mixing and mixing frequency peak counting techniques to assess low-velocity impact damage in CFRP composite laminates. The value of the mixing frequency peak count could be correlated with the impact energy in the test. Guan et al [29] developed a three-dimension (3D) finite element (FE) model to demonstrate that the directions and locations of the micro cracks in plates can affect the amplitudes of the nonlinear waves generated by nonlinear guided wave mixing. All the aforementioned studies demonstrated that the guided wave mixing technique has many advantages over the second harmonic approach. One of the most important merits is that the combination harmonics generated by mixing two guided waves are less affected by the higher harmonics produced by the instrumentations, such as amplifiers and transducers. In addition, mutual interaction between guided waves propagating in different directions provides more flexibility for the selection of guided wave modes and their corresponding excitation frequencies. The majority of the work on nonlinear guided waves has focused on the structures in gaseous environments. Before the initiation of macro-scale damage, the material nonlinearity due to the microstructural features is very weak, making the generation and measurement of nonlinear guided waves very challenging. Only a few guided wave modes within limited frequency bandwidths, satisfying the phase velocity matching and non-zero power flux conditions, have the potential to generate cumulative and measurable nonlinear guided waves and can be used to characterize the material nonlinearity [15]. Phase velocity matching refers to that the primary and the corresponding nonlinear guided waves should have the same phase velocity. Non-zero power flux means that there must be nonzero power flow from the primary waves to the nonlinear guided waves. They have been widely recognized as the criteria for selecting guided wave modes and excitation frequencies for the nonlinear guided wave methods for the structures in gaseous environments. When the structures are exposed to liquid, the fluid-solid coupling 120
ADE
Chapter 5 allows the guided wave energy to leak from the structures into the surrounding liquid medium. On the immersed structures, the generation and measurement of nonlinear guided waves are more challenging because most of the guided wave modes have higher attenuation [10]. In addition, the fluid-solid coupling makes the guided waves in the immersed structures behave differently from their counterparts in the structures surrounded by air [30-33]. Therefore, a comprehensive investigation is desired for the nonlinear guided waves in the immersed structures. This paper presents a series of experimental and numerical investigations on the nonlinear guided wave mixing in partially immersed metallic plates. In the experiments, the fundamental leaky symmetric Lamb (leaky S ) waves are excited 0 at two different frequencies on an aluminum plate, of which one side is exposed to water. The response signals display the combination harmonics at the sum frequency. Next, the amplitudes of the combination harmonics are investigated with varying excitation magnitudes and propagation distances. The effect of the liquid- structure coupling is also explored by comparing the guided wave signals measured from the test plate with and without water. Then, numerical simulations using a 3D FE model are implemented to further investigate the sensitivity of the combination harmonics to the material nonlinearity of the metallic plate partially immersed in water. The findings of this study can provide support for the development of structural health monitoring techniques for metallic liquid containment structures. The remainder of the paper is organized as follows. Section 5.3 introduces the theoretical background of the mutual interaction of ultrasonic waves in materials with weak nonlinearity. Section 5.4 presents the experimental study, including the overall experimental setup, preliminary tests to select the guided wave modes and excitation frequencies, and the results of the experimental investigations. The numerical study is illustrated in Section 5.5. The 3D FE model is described and validated through experimental measurements. This section also includes a parametric study, in which the experimentally validated 3D FE model is employed to investigate the characteristics of the nonlinear guided wave mixing in materials with different levels of fatigue. Finally, conclusions are drawn in Section 5.6. 121
ADE
Chapter 5 5.3. The theoretical background of the wave mixing technique This section describes the theoretical derivations for the mutual interaction of ultrasonic waves with two different frequencies in nonlinearly elastic materials. The material nonlinearity is small and attributed to the microstructural features such as dislocations, microvoids, and micro cracks. For one-dimensional problems, the stress-strain relationship of the nonlinear elastic material can be expressed as [15]     Eu u2  (5.1)  2  where  represent the stress. uu x with u and x representing the displacement and the position, respectively. E and denote the linear elastic modulus and the second order nonlinear parameters, respectively. The quadratic term accounts for the weak nonlinearity of the material, which is ignored in the linear theory. Considering that two waves with different frequencies ( f and f , f  f ) 1 2 2 1 travel in the material, the equation of motions can be described as u (5.2) where  is the mass density;  x and u2u t2 ; t denotes the time. Substituting Eq (5.1) into Eq (5.2) gives E uuuu (5.3)  where u2u x2. Eq (5.3) can be solved using the perturbation approach with the assumption that the solution form of the total displacements is the sum of the primary waves and the nonlinear waves uu u (5.4) P N where u and u represent the primary and nonlinear components of the total P N displacements, respectively. The primary waves are also called linear waves because they have the same frequencies as the input signals. The nonlinear waves are generated by the interaction of the primary waves with the material nonlinearity. 122
ADE
Chapter 5 The amplitude of the nonlinear waves is much smaller than that of the primary waves ( u u ). Then, the governing equation can be obtained by substituting P N Eq (5.4) into Eq (5.3) as follows E  u u u u u u  u u (5.5)  P N P N P N P N Since the amplitudes of the nonlinear waves are very small, the derivatives with respect to u can be neglected [15]. Then the governing equation can be further N divided into two differential equations as follows. E u  u 0 P  P (5.6) E E u  u  uu N  N  P P Finally, the solution of the primary waves is u  A cos(wtk x)A cos(wtk x) (5.7) P f 1 1 1 f 2 2 2 1 2 The primary waves combine the two excitation frequencies f and f . A and 1 2 f 1 A are the amplitudes of the primary waves at f and f , respectively. w, k , and, f 1 2 2  represent the angular frequency, wavenumber, and phase shift, respectively, with the subscripts denoting the first and second frequency components. The solution of the nonlinear waves at the frequencies other than the excitation frequencies can be written as u  A cos(2wt2k x) A cos(2w t2k x) N 2f 1 1 1 2f 2 2 2 1 2  A cos(w w tk k x) (5.8) f f 1 2 1 2 1 2 1 2  A cos(w w tk k x) f f 1 2 1 2 1 2 1 2 where A A2k2x 8 and A A2k2x 8 are the amplitudes of the second 2f f 1 2f f 2 1 1 2 2 harmonics at 2f and 2f , respectively. A A A kk x 4 is the amplitude 1 2 f f f f 1 2 1 2 1 2 of the combination harmonics (sum harmonics) at the sum frequency ( f  f ) and 1 2 A A A kk x 4 is the amplitude of the combination harmonics (difference f f f f 1 2 1 2 1 2 harmonics) at the difference frequency ( f  f ). Therefore, when the structural 1 2 123
ADE
Chapter 5 material is excited by waves with two different frequencies, the nonlinear parameters can be estimated as follows 8  A 1    2f 1  for the second harmonics at2f 2f 1 k2  A2  x 1 1  f  1 8  A 1    2f 2  for the second harmonics at2f 2f 2 k2  A2  x 2 2  f  2 (5.9) 4  A 1    f 1f 2  for the sum harmonics at f  f f 1f 2 k k  A A  x 1 2 1 2  f f  1 2 4  A 1    f 2f 1  for the difference harmonics at f - f f 2f 1 k k  A A  x 2 1 1 2  f f  1 2 It can be seen that when the propagation characteristics of the primary waves do not change (e.g. fixed wavenumber values k and k ), the material 1 2 nonlinearity at any location can be correlated to the amplitudes of the primary waves and the nonlinear waves. For simplicity, the material properties can be characterized by relative nonlinearity parameters and they are defined as follows A   2f i for the second harmonics at2f 2f i A2 i f i A   f 1f 2 for the sum harmonics at f  f (5.10) f 1f 2 A A 1 2 f f 1 2 A   f 2f 1 for the difference harmonics at f - f f 2f 1 A A 2 1 f f 1 2 where the relative nonlinearity parameter  is proportional to the second order nonlinear parameter  and the propagation distance x. The above derivations consider only the simplest case that ultrasonic waves travel in the isotropic material in one direction. However, the relative nonlinearity parameters defined by Eq. (5.10) have been widely recognized to be applicable for characterizing the material nonlinearity of various structures using different guided wave modes. In the literature, there are two popular approaches for evaluating nonlinear elastic properties. The first approach is to excite structures with the input signals of various magnitudes and measure the response signals at a single location 124
ADE
Chapter 5 [34-37]. If the measured nonlinear guided waves are generated due to the material nonlinearity, the amplitudes of the second harmonics (A ) should increase linearly 2f i with the square of the amplitudes of the corresponding primary waves ( A2 ). In f i contrast, the amplitudes of the combination harmonics, including the sum harmonics (A ) and difference harmonics (A ), should have a positive linear ff f f 1 2 2 1 relationship with the product of the primary waves at the two excitation frequencies (A A ). The increasing rate of the nonlinear guided wave magnitudes to the f f 1 2 corresponding primary waves can be correlated to the material nonlinearity of structures. The second approach estimates the material nonlinearity by measuring signals at several locations with different propagation distances [26, 38-41]. Based on Eqs (5.9) and (5.10), the relative nonlinearity parameters () will grow linearly with the propagation distance, provided that the primary waves have sufficiently large motion magnitudes to interact with the microstructural features of the material [40, 42, 43]. The material's nonlinear elastic properties can be characterized by the accumulation gradient of  with the propagation distances. Within the same sample and identical experimental setup, any abnormal increase of the nonlinear guided waves and the relative nonlinearity parameters indicate the growth of the material nonlinearity and degradation. In the present study, the relative nonlinearity parameters are calculated by Eq. (5.10) to characterize the nonlinear guided wave features on the partially immersed plate. The growing trends of the nonlinear guided wave features with varying excitation magnitudes and propagation distance are investigated, respectively. 5.4. Experimental study 5.4.1. Experimental setup for actuating and sensing guided waves Experiments were conducted on a 1.6 mm thick aluminum plate, which was fixed to the front of a metal tank with bolts. To investigate the effect of the liquid coupling on the nonlinear guided wave mixing, experimental measurements were collected on the external surface of the aluminum plate when the tank was empty and when 125
ADE
Chapter 5 it was filled with water, respectively. Due to the isotropic features of the metal materials, the findings of this study could be also applicable to the isotropic plates made of other metal materials such as steel and alloy. Figure 5.1 shows a photo of the experimental setup. A computer-controlled signals generator (NI PIX-5412) was employed to generate the excitation signals. The waveforms of the excitation signals and the frequency selection are discussed in detail in Section 5.4.2. Then, the voltage of the input signals was magnified by a voltage amplifier (Ciprian HVA- 800-A). After that, the amplified excitation signals were sent to the piezoceramic transducer (Ferroperm Pz27, 10 mm diameter and 0.5 mm thick) that was bonded to the outer surface of the test plate. The piezoceramic transducer could convert the electric signals to mechanical motions and generate guided waves on the test plate. The response signals on the plate surface were measured by a scanning laser vibrometer (Polytec PSV-400-M2-20) and then further processed using the software MATLAB. Each measurement was collected at 25.6 MHz and averaged by 1000 acquisitions. The signal-to-noise ratios were improved by applying a low- pass filter, of which the cut-off frequency was set to 1MHz. Figure 5.1. Experimental setup 126
ADE
Chapter 5 5.4.2. Mode tuning and frequency selection This section describes the selection of guided wave modes and excitation frequencies for the nonlinear guided wave mixing in partially immersed plates. To begin with, the dispersion features of guided waves were studied using DISPERSE. Based on the global matrix method, the dispersion curves were derived for the 1.6 mm thick aluminum plate surrounded by air (when the metallic tank was empty) and the same plate with one side exposed to water (when the metallic tank was filled with water), respectively [44]. The material properties of the plate and the water are shown in Table 5.1. The water layer was defined as a non-viscous semi-infinite acoustic medium. The air properties were not considered in modeling because the influence of the air on the guided wave propagation was very small. The air-coupled plate surfaces were assumed to be traction-free. The water-coupled plate surface was defined by the structural-liquid boundary conditions, which connected the normal stresses and displacements at the plate-water interface [44]. Through the out-of-plane motions, the guided waves in the liquid-coupled structures could continuously radiate wave energy into the surrounding liquid medium. Table 5.1: Material properties for the aluminum plate and the water layer Density Young’s Poisson’s Longitudinal (kg/m3) modulus ratio velocity (m/s) (GPa) Aluminum 2704 70.76 0.33 -- Water 1000 -- -- 1500 Figure 5.2 presents the dispersion curves for the 1.6 mm thick aluminum plate without water and loaded with water on one side, respectively. The frequency range was selected to be below the cut-off frequency of higher-order guided wave modes. Therefore, only the fundamental guided wave modes could be excited by the piezoceramic transducer. When the tank is empty, the test plate is surrounded by air on both surfaces. There are only the fundamental symmetric (S ) and 0 antisymmetric (A ) wave modes on the air-coupled metallic plate as shown in 0 Figures 5.2(a)-5.2(c). Previous studies demonstrated that S wave on air-coupled 0 127
ADE
Chapter 5 plates satisfies approximate phase velocity matching and nonzero power flow conditions [19, 45]. Specifically, the phase velocity of S mode decreases very 0 slowly with frequency. When the phase velocity of S mode at the excitation 0 frequency matches that of S mode at twice the excitation frequency with a relative 0 deviation of less than 1%, the interaction of the primary S wave at the excitation 0 frequency with the material nonlinearity of a metallic plate can generate measurable second harmonic S waves at twice the excitation frequency. In addition, the 0 calculated relative nonlinearity parameters can grow linearly with the propagation distance and can be used to characterize the microstructural change of the material before the initiation of macro-scale damage [19, 45, 46]. In contrast, the phase velocity of the A wave increases rapidly with 0 frequency, making the phase velocity of the A wave at the excitation frequency 0 significantly different from that of A wave at other frequencies. To date, there are 0 very limited studies using A wave to evaluate the microstructural changes of 0 material in the early damage stage. Chillara and Lissenden [47] numerically demonstrated that the interaction of A wave with the material nonlinearity of a 0 metallic plate can only generate second harmonic S waves, which propagate 0 independently and separate from the primary A wave. The generated second 0 harmonic S waves are so small that it is difficult to measure in practical 0 applications. However, A wave has been extensively employed to identify contact- 0 type damage, such as open fatigue cracks [18], delamination [48], and the bonding effects of bolts [49]. The clapping behaviors between the surfaces as guided waves pass through can generate measurable nonlinear guided waves. Since the contact- type nonlinearity is much larger than the material nonlinearity, the evaluation of the contact-type defects does not require phase velocity matching and nonzero power flow conditions. When the tank is filled with water, one side of the test plate is in contact with water, and the other side is exposed to air. Guided waves in the partially immersed plate include the leaky S wave and the fundamental leaky antisymmetric 0 (leaky A ) wave as well as the quasi-Scholte wave as shown in Figures 5.2(d)-5.2(f). 0 The leaky A wave has high attenuation and is unable to propagate a long distance 0 in the immersed plates, as shown in Figure 5.2(f). The quasi-Scholte wave has low 128
ADE
Chapter 5 black solid lines in Figures 5.2(d)-5.2(f), the phase and group velocities of the leaky S wave decrease very slowly with frequency, which is similar to those of the S 0 0 wave in the dry plate. In addition, the attenuation of the leaky S wave is close to 0 zero at the frequency range below 600 kHz. These features enable the leaky S 0 waves to have similar propagation characteristics across a relatively wide frequency range, which provides good flexibility for the selection of excitation frequencies for guided wave mixing. For further investigation, mode shapes were extracted by DISPERSE for the S mode and the leaky S mode, respectively, as shown in Figure 5.3. The mode 0 0 shape diagrams display the relative displacements of guided wave modes across the thickness of the structure. The red dashed lines and the blue solid lines represent the particle displacements in the in-plane direction (parallel to the wave propagation) and the out-of-plane direction (normal to the plate surface), respectively. The amplitudes are normalized by the maximum absolute magnitudes. Figures 5.3(a) and 5.3(b) show the mode shapes for the S mode at 100 kHz and 500 kHz, 0 respectively. Within the selected frequency range, the S mode is dominated by the 0 in-plane displacement that is uniformly distributed across the plate thickness. The out-of-plane displacement component is small at low frequency and gradually increases with frequency. For comparison, Figures 5.3(c) and 5.3(d) show the mode shapes for the leaky S mode at 100 kHz and 500 kHz, respectively. Generally, the 0 wave structure of leaky S mode in the immersed plate is similar to that of S mode 0 0 in the plate without water. The leaky S mode has predominately wave motions 0 conserved in the plate structure, which ensures sufficient wave motions to interact with the material microstructures. The in-plane displacement between the plate and water is disconnected, while the out-of-plane displacement is continuous. Therefore, as the frequency increases, more wave energy can leak into the liquid medium through the out-of-plane wave motions. For these observations, the frequency range of interest was chosen to be below 600 kHz to ensure that the generated nonlinear guided waves have low attenuation and are measurable. 130
ADE
Chapter 5 Figure 5.3. Mode shapes for S mode in a 1.6 mm thick aluminum plate at (a) 100 0 kHz and (b) 500kHz and mode shapes for leaky S mode for the plate loaded water 0 on one side at (c) 100 kHz and (d) 500kHz. (the red dashed lines represent the in- plane displacements and the blue solid lines denote the out-of-plane displacements) Preliminary tests were implemented with single-frequency excitation signals on the empty tank and the water-filled tank, respectively, to evaluate the excitability of the piezoceramic transducers. The preliminary tests aimed to select two excitation frequencies, at which the selected guided wave modes have comparable wave motions. The single-frequency excitation signals were 6-cycle narrow-band tone burst pulses modulated by Hanning window [15]. The excitation frequency swept from 90 kHz to 410 kHz in steps of 20 kHz. Figure 5.4(a) shows typical examples of the guided wave signals measured at 250 mm away from the excitation center. The excitation frequency was 230 kHz. The red dash-dot and black solid lines denote the signals obtained from the empty tank and water-filled tank, respectively. From the wave speed evaluation, the first wave packets between 131
ADE
Chapter 5 45 s and 80 s were identified as the S wave for the empty tank and the leaky S 0 0 wave for the water-filled tank, respectively. It should be noted that the scanning laser vibrometer measures the normal displacements on the plate surface [32]. Although S and leaky S waves are relatively small in the received signals, their 0 0 actual wave motions in the plate should be strong because both wave modes have mostly in-plane motions as shown in Figure 5.3. Following the first wave packets, the second wave in the empty tank should be the A wave, while the second wave 0 in the water-filled tank is the quasi-Scholte wave. The latter travels much slower than the former, which is in good agreement with the theoretical predictions by the group velocity dispersion curves as shown in Figures 5.2(b) and 5.2(e). Figure 5.4. (a) Comparison of the time-domain signals experimentally collected from the empty tank and water-filled tank and (b) the peak amplitudes of the extracted signals across various frequencies. For further signal processing, the first wave packets were extracted from the rest of the signals. Figure 5.4(b) shows the peak amplitudes of the first wave packet for different excitation frequencies. In general, the magnitudes of the signals measured from the empty tank and water-filled tank change with the excitation frequency, following a similar pattern. When the frequency increases from 90 kHz to 250 kHz, the amplitudes of the signals increase with frequency. For the frequency range over 250 kHz, the amplitudes of the signals decrease with frequency. The decreasing rates of the signals measured from the water-filled tank are much quicker than those obtained from the empty tank. From these observations, f and 1 132
ADE
Chapter 5 f were selected to be 170 kHz and 270 kHz, respectively, for the following three 2 reasons. Firstly, the difference between f and f was chosen to be 100 kHz, 1 2 which enables good separation between the second harmonics and the combination harmonics. Secondly, the signals collected at both excitation frequencies had relatively high signal-to-noise ratios. Thirdly, the received signals show the normal displacement components on the plate surface. For the S and leaky S waves, the 0 0 out-of-plane displacements increase with frequency as shown in Figure 5.3. Therefore, the signal amplitude at f should be slightly lower than that at f to 1 2 ensure the actual wave motions at the two selected excitation frequencies have comparable magnitudes on the plate. Figure 5.5. Merging two single-frequency tone burst signals to generate a mixed frequency signal. Next, the mixed frequency excitation signals were generated by merging two single-frequency signals, which were a 6-cycle Hanning window-modulated tone burst at a central frequency of 170 kHz and a 9-cycle Hanning window- modulated tone burst at a central frequency of 270 kHz. The higher frequency signal had more cycles, which was to ensure that the two frequency components had similar duration in the time domain and comparable energy contents in the 133
ADE
Chapter 5 frequency spectrums. A mix ratio of 1:1 (where 1:1 mixing means 1 part 170 kHz and 1 part 270 kHz) is promising to generate larger combination harmonics than other mix ratios [52]. Figure 5.5 shows the waveforms of the single-frequency and mixed frequency signals and their corresponding frequency spectrums. 5.4.3. Experimental results 5.4.3.1. Guided wave mixing in partially immersed metal plates A demonstration of the guided wave mixing phenomenon was presented for the partially immersed plate. Firstly, experiments were carried out with the water-filled tank using the two single-frequency excitation signals as shown in Figure 5.5. The measurements were collected at a fixed location that was 250 mm away from the actuator center. The voltage of the input signals was increased to 160 V. Figures 5.6(a) and 5.6(b) show the time-domain signals for 170 kHz and 270 kHz, respectively. The first wave packets in the two figures are identified as the leaky S 0 waves that propagate faster than any other wave mode as shown by the group velocity dispersion curves in Figure 5.2(e). Secondly, experiments with the same settings were performed using the mixed frequency signal. The response signals measured at the same location are shown in Figure 5.6(c). It can be seen that the waveform of the first wave packet is similar to the input of the mixed frequency signal as shown in Figure 5.5. This is because leaky S waves have similar phase 0 and group velocities and low attenuation within the selected frequency range. The leaky S waves at 170 kHz and 270 kHz can propagate together and the wave- 0 mixing zone is maximized. Following the first wave packet, there are other wave components in Figures 5.6(a)-5.6(c). Since leaky A waves have high attenuation at the selected 0 frequencies, the remaining wave components should be dominated by the low- attenuated quasi-Scholte waves. It can be seen that the waveform of the quasi- Scholte wave in Figure 5.6(c) is similar to that at 170 kHz as shown in Figure 5.6(a). This is because the deformation fraction in the immersed plate of the quasi-Scholte wave decreases rapidly with frequency [53]. As a result, when the structure is excited by the mixed frequency signal, the quasi-Scholte waves measured on the plate surface are dominated by the low-frequency components. The high-frequency 134
ADE
Chapter 5 components of the quasi-Scholte wave are too weak to interact with the low- frequency components. Therefore, the quasi-Scholte wave can be filtered out for the study of nonlinear guided wave mixing. Figure 5.6. Experimental signals from the water-filled tank with the excitation (a) at f = 170 kHz, (b) at f = 270 kHz, (c) at mixed frequencies; and (d) the frequency 1 2 spectrum of their corresponding window-cut data. Then, the first wave packets were extracted from the rest of the signals and transferred into the frequency domain by fast Fourier transfer (FFT) as shown in Figure 5.6(d). When the one-side water-immersed plate is excited separately by the single-frequency signals, the 170 kHz and 270 kHz leaky S waves can generate 0 second harmonics, as manifested by the peaks at double the excitation frequencies (2f 340kHzand 2f 540kHz). For comparison, when the partially immersed 1 2 plate is excited by the mixed frequency signal, the received signal has an additional peak at the sum frequency ( f  f 440kHz ), as shown by the black solid line in 1 2 Figure 5.6(d). The combination harmonics at the difference frequency ( f  f 100kHz) cannot be observed clearly for the mixed frequency excitation. 1 2 135
ADE
Chapter 5 This is because the difference frequency is too close to the lower excitation frequency ( f 170kHz) with only 70 kHz spacing. So, the difference harmonics 1 can be overwhelmed by the side lobes [46, 54]. The same reason also applies to the second harmonics at (2f 340kHz) that are very close to the higher excitation 1 frequency ( f 270kHz ). Nevertheless, the experimental results indicate that 2 mixing leaky S waves with two different frequencies can generate low-attenuated 0 sum harmonics on the partially immersed plate. Figure 5.7. The actual excitation signals from the piezoceramic transducer (a) in the time domain, and (b) in the frequency domain. Next, the nonlinearity due to the instrumentations was investigated. Figure 5.7(a) shows the actual signal from the piezoceramic transducer measured by the scanning laser vibrometer. Figure 5.7(b) presents the corresponding frequency spectrum. The amplitudes are normalized by the maximum absolute magnitudes. As shown in Figure 5.7(b), there are no apparent peaks at the sum frequency ( f  f 440kHz ) and difference frequency ( f  f 100kHz) in the frequency 1 2 1 2 spectrum of the actual signal from the piezoceramic transducer. Therefore, the sum harmonics observed in Figure 5.6(d) should be generated mainly by the interaction of the guided waves with the material nonlinearity. However, the actuation system can produce second harmonics as manifested by the peak at (2f 540kHz) in 2 Figure 5.7(b). 136
ADE
Chapter 5 Figure 5.8. Experimental signals from the water-filled tank excited by mixed frequency signal (a) in the time domain, and (b) in the frequency domain; (c) the amplitudes of the sum harmonics versus the product of the primary wave amplitudes; and (d) the amplitudes of the second harmonics at 2f versus the 2 square of the corresponding primary wave amplitudes at f . 2 After that, the nonlinear response of the guided wave mixing was further investigated by varying the excitation voltages. The voltage of the input signal was increased to 40V, 80V, 120V, and 160V, respectively. For each voltage, the response signals were measured five times at 250 mm away from the excitation center. Figures 5.8(a) and 5.8(b) show the experimentally measured signals in the time domain and the frequency domain, respectively. Figure 5.8(c) shows the amplitudes of the sum harmonics versus the product of the primary waves at the two excitation frequencies. Figure 5.8(d) shows the amplitudes of the second harmonics at 2f versus the square of the amplitudes of the corresponding primary 2 waves at f . The black solid lines in Figures 5.8(c) and 5.8(d) represent the best-fit 2 lines by linear regressions with the error bars denoting the standard deviations of five measurements. As discussed in Section 2, the amplitudes of the second 137
ADE
Chapter 5 harmonics due to the material nonlinearity should increase linearly with the square of the amplitudes of the corresponding primary waves, while the amplitudes of the combination harmonics should grow linearly with the product of the primary waves at the two excitation frequencies. Overall, this analysis demonstrates that both the combination harmonics and the second harmonics could be generated due to the material nonlinearity of the specimen. However, the growing trends of the second harmonics in Figure 5.8(d) show relatively larger deviations from the best-fit line. This phenomenon indicates that the second harmonics are more susceptible to the nonlinearity generated by the instrumentations, which agrees well with Figure 5.7(b). 5.4.3.2. The effect of the surrounding liquid on the guided wave propagation The section compares the phenomenon of guided wave mixing in the test plate with and without water. Guided wave signals were collected from the empty tank and the water-filled tank, respectively, under the same experimental conditions. Figure 5.9(a) presents the time-domain signals measured at 250 mm away from the excitation center. The red dash-dot and black solid lines denote the signals obtained from the empty tank and water-filled tank, respectively. The first wave packets are identified as the mixed frequency S wave for the empty tank and the mixed 0 frequency leaky S wave for the water-filled tank, respectively. Both wave modes 0 have similar amplitudes and waveforms in the time domain. However, an obvious difference can be observed in the frequency domain. Figure 5.9(b) shows the frequency spectrums of the window-cut signals extracted from Figure 5.9(a). The combination harmonics at the sum frequency can be observed in both the empty tank and the water-filled tank. The amplitudes of the primary waves at fundamental excitation frequencies (170 kHz and 270 kHz) are similar for the empty tank and water-filled tank. Over 270 kHz, the signal obtained from the empty tank has larger amplitudes than that obtained from the water-filled tank. Further research investigated the growing trends of combination harmonics at the sum frequency with increasing excitation voltage and propagation distance, respectively. Figure 5.9(c) shows the amplitudes of the sum harmonics versus the product of the primary wave magnitudes at the fundamental frequencies. The 138
ADE
Chapter 5 response signals were collected at 250 mm away from the excitation center and the voltage of the input signal was increased to 40V, 80V, 120V, and 160V. Five measurements were collected for each voltage on both the empty tank and the water-filled tank, respectively. It can be seen that the amplitudes of both the primary waves (A A ) and the sum harmonics (A ) on the empty tank are relatively f f ff 1 2 1 2 larger than those on the water-filled tank. Also, when the primary waves (A A ) f f 1 2 increase, the sum harmonics (A ) on the empty tank increase much faster than ff 1 2 those on the water-filled tank. Figure 5.9. Comparison of experimental signals from the empty tank and water- filled tank excited by mixed frequency signal (a) in the time domain, and (b) in the frequency domain; (c) the amplitudes of the sum harmonics versus the product of the primary waves; and (d) the amplitudes of the nonlinearity parameters for the sum harmonics versus propagation distance. To investigate the relationship between the combination harmonics at the sum frequency and the propagation distance, 21 measurement points were defined on the external surface of the test plate and equally spaced between 200 mm and 139
ADE
Chapter 5 300 mm away from the excitation center. The voltage of the input signals was increased to 160 V. Five measurements using the same settings were performed on the empty tank and water-filled tank, respectively. Then, the relative nonlinearity parameters were calculated by Eq. (5.10) for the sum harmonics and plotted against the propagation distance as shown in Figure 5.9(d). For both the empty tank and the water-filled tank, the relative nonlinearity parameters grow linearly with the propagation distance from 200 mm to 270 mm. The cumulative propagation distances are limited to 270 mm, which may be caused by the small phase velocity difference between the primary waves and the sum harmonics. As shown in Figure 5.2, the phase velocities of S wave in the dry plate and leaky S wave in the partially 0 0 immersed plate have similar values and decrease slowly with frequency. The phase velocities of the primary S waves at 170 kHz and 270 kHz are 5430 m/s and 5422 0 m/s, respectively. They are very close to the phase velocity of the sum harmonic S 0 waves at 440 kHz, which is around 5397 m/s. The deviation is less than 1%. Previous studies have analytically and experimentally demonstrated that the cumulative propagation distances of second harmonic generation due to material nonlinearity decrease with the phase velocity difference between the primary S 0 waves and the second harmonic S waves [19, 45]. In the present study, the 0 cumulative propagation distances of sum harmonics determined by experimental measurements have a similar order of magnitudes to the theoretical predictions for the second harmonic generation by S waves due to material nonlinearity [45]. Thus, 0 it can be validated that the sum harmonics of the signals measured from 200 mm to 270 mm are generated due to the nonlinearity of the material. In the linearly cumulative range, the growth rate of the relative nonlinearity parameters with the propagation distance does not show an apparent difference between the plate surrounded by air and the plate partially immersed in water. However, the absolute values of the relative nonlinearity parameters on the empty tank are much larger than those on the water-filled tank. Therefore, these results indicate that the cumulative generation of combination harmonics due to the material nonlinearity of partially immersed plates can be achieved with a mixed frequency leaky S wave. Under the same experimental conditions, the amplitudes 0 of the guided wave signals and the values of the relative nonlinearity parameters on 140
ADE
Chapter 5 the partially immersed plate are different from their counterparts on the plate in gaseous environments. 5.5. Numerical study 5.5.1. Modeling material nonlinearity Numerical simulations were carried out with ABAQUS to further investigate the sensitivity of the combination harmonics to the material nonlinearity of the partially immersed metallic plate. The numerical methods have the advantage of eliminating unwanted effects of the noises from the measurement system. The nonlinearity of the material was simulated by incorporating a VUMAT subroutine that introduced a constitutive model proposed by Murnaghan [55]. This section presents the constitutive equations. To begin with, X and xare defined as the coordinates in the reference and current configurations, respectively. The deformation gradient can be expressed as [56] x F IH (5.11) X where I is the identity tensor; Hu X is the displacement gradient and u  xX is the displacement vector. The Lagrangian strain tensor can be written as 1 1 E FTF HHT HTH (5.12)   2 2 For a hyperelastic and homogeneous isotropic solid material, the strain energy function is [55] 1 1 W(E)  trE2 tr E2  l2m trE3 2 3 (5.13) mtrE  trE2 tr E2 ndetE where  and  are the lame constants; l , m, and n are Murnaghan constants that describe the second order material nonlinearity. Previous studies experimentally measured the Murnaghan constants from dog bone samples made 141
ADE
Chapter 5 of aluminum [57]. The samples were subjected to repeated uniaxial tensile loads with different cycles. One sample was loaded to a total of 52800 cycles and failed. The rest samples were loaded to various cycles and referenced to the percent of fatigue level. Table 5.2 summarizes the material properties of aluminum with 0, 40%, and 80% fatigue levels [57]. Table 5.2. Lame constants and Murnaghan constants for aluminum [57] Fatigue 𝜌 𝜆 𝜇 L M N level (kg m-3) (GPa) (GPa) (GPa) (GPa) (GPa) (%) 0 2704 51.6 26.6 -252.2 -325.0 -351.2 40 2704 51.6 26.6 -266.8 -332.8 -358.3 80 2704 51.6 26.6 -271.2 -335.8 -359.8 The second Piola-Kirchhoff stress tensor can be obtained by WE T  (5.14) PK2 E The Piola-Kirchhoff stresses are used to describe the reference configuration and correlated with the Cauchy stress tensor σ as T detFF1σ F1T (5.15) PK2 5.5.2. 3D FE model ABAQUS/CAE was employed to build and mesh the 3D FE model as shown in Figure 5.10. The model consisted of a test plate that was 250 mm wide and 430 mm long and had the same thickness as the experimental specimen. The bottom surface of the test plate was exposed to water. The red quarter circle at the bottom left corner of the plate represented a quarter of the piezoceramic transducer, which was perfectly bonded to the top surface of the plate. Symmetric boundary conditions were defined for the left and bottom edges. Firstly, the numerical simulations were implemented by considering only the linear elastic material properties of the test plate and water as shown in Table 5.1. The linear finite element (FE) simulations consider the metallic plate as a linear elastic material. Then, nonlinear FE 142
ADE
Chapter 5 Leaky S waves were generated by applying nodal displacements to the 0 circumference of the simulated piezoceramic transducer [18]. The displacements were assigned in the radial direction as shown in Figure 5.10(a). The excitation signals were the mixed frequency signals as shown in Figure 5.5. The magnitude of the displacement was 3 m. The simulated guided wave signals were calculated by the central-difference integration through ABAQUS/Explicit. In all simulations, the increment time step was automatically controlled by ABAQUS. The maximum time increment step is less than the ratio of the minimum element size to the dilatational wave speed [43]. The accuracy of the nonlinear constitutive model was validated by comparing the results with the outcomes of the linear FE and experimental measurements. 5.5.3. Experimental validation Figure 5.11(a) compares the experimental measurements and simulation results in the time domain. The time-domain signals, simulated by the nonlinear FE model incorporating the Murnaghan constants for the intact aluminum (zero fatigue in Table 2), are consistent with those simulated by the linear FE model. Both have a good agreement with the experimental signals. Next, the signals were windowed between 45 s and 80 s and transferred into the frequency domain as shown in Figure 5.11(b). The linear and nonlinear FE models well predict the primary guided wave components of the experimental signals at the fundamental excitation frequencies, as highlighted by the dotted lines at 170 kHz and 270 kHz in Figure 5.11(b). The linear FE signals have no nonlinear guided wave components at frequencies other than the excitation frequencies. But the nonlinear FE well predicts the location of the peaks for the combination harmonics at the sum frequency ( f  f ) and the second harmonics at 2f . The amplitudes of the simulated 1 2 2 nonlinear guided waves are comparable to the experimentally measured data. Although there is a discrepancy between the experimental signals and simulated signals at other frequency components, the nonlinear FE simulation well predicts the generation of the combination harmonics due to the material nonlinearity. 144
ADE
Chapter 5 Figure 5.11: Experimental validations by comparing the experimental and simulated signals (a) in the time domain and (b) in the frequency domain 5.5.4. Parametric study The 3D FE model validated by experiments was employed to further investigate the sensitivity of the nonlinear guided waves mixing to the material nonlinearity of the partially immersed metallic plate. Nonlinear FE simulations were implemented using the same 3D FE models, in which the material properties of the test plate were defined as aluminum at the three different fatigue levels as shown in Table 5.2, respectively. The out-of-plane displacements were collected at 11 measurement points that were equally distributed on the top surface of the test plate from 200 mm to 250 mm away from the excitation center. Then, the nonlinearity parameters were calculated using Eq. (5.10) for the combination harmonics at the sum frequency ( f  f ) and the second harmonics at 2f , respectively. To better observe the 1 2 2 influence of material nonlinearity evolution, the relative nonlinearity parameters were normalized by their corresponding minimum value at zero fatigue level. Figures 5.12(a) and 5.12(b) show the normalized nonlinearity parameters for the combination harmonics and the second harmonics, respectively. Within the selected measurement range, the normalized nonlinearity parameters grow linearly with propagation distance. The growth rate (slope values of the best-fitted line) increases as the material suffers more fatigue damage. 145
ADE
Chapter 5 5.6. Conclusion This paper presents experimental and numerical investigations on the nonlinear guided waves mixing in the partially immersed plates. The main contributions are summarized as follows: (1). According to the dispersion curves, leaky S waves at low excitation 0 frequencies have low attenuation and low dispersion effects, which provide good flexibility for the selection of excitation frequencies for guided wave mixing. (2). Experiments have been conducted on an aluminum plate loaded with water on one side using the single-frequency excitation and mixed frequency excitation, respectively. Leaky S waves with two different frequencies can 0 generate combination harmonics at the sum of the excitation frequencies that cannot be achieved by single-frequency excitations (see Figure 5.6). (3). The combination harmonics at sum frequency ( A ) grow linearly with f +f 1 2 the product of the primary guided waves at the fundamental excitation frequencies ( A A ) and are less affected by the nonlinearity due to f f 1 2 instrumentations (see Figures 5.7 and 5.8). (4). Under the same experimental conditions, the liquid-structure coupling of the partially immersed plate makes the amplitudes of the guided wave signals and the relative nonlinearity parameters different from those of the test plate without liquid coupling (see Figure 5.9). (5). Numerical studies have been carried out with the material nonlinearity of the test plate simulated by the Murnaghan constitutive model. The numerical results reveal that both the second harmonics and the combination harmonics are sensitive to the material nonlinearity of the partially immersed plate (see Figures 5.12 and 5.13). In conclusion, the current study has demonstrated that nonlinear guided wave mixing has the potential to evaluate the material nonlinearity of metallic plates with one side exposed to water. This new possibility can be significant considering the high rewards for earlier detection of the damage to maintain the structural integrity of high-valued infrastructures. To maximize the wave mixing zone, the selected 147
ADE
Chapter 5 guided waves propagate together in the same direction. Future studies are required to explore the feasibility of non-collinear guided wave mixing, where the selected guided waves propagate in different directions and meet in a localized mixing zone. The non-collinear guided wave mixing has the potential to identify the area of localized material degradation but the generation and measurement of the nonlinear waves are more challenging because the nonlinear wave amplitudes can be affected by the reduced wave mixing zone. In addition, more experimental studies need to be carried out to investigate the correlation between the nonlinear guided wave signals and the degree of damage. 5.7. Acknowledgment This work was funded by the Australia Research Council (ARC) under grant numbers DP200102300 and DP210103307. The supports are greatly appreciated. 5.8. Reference [1] R. Liu, S. Li, G. Zhang, W. Jin, Depth detection of void defect in sandwich- structured immersed tunnel using elastic wave and decision tree, Construction and Building Materials, 305 (2021) 124756. [2] R. Ignatowicz, E. Hotala, Failure of cylindrical steel storage tank due to foundation settlements, Engineering Failure Analysis, 115 (2020) 104628. [3] N. Balekelayi, S. Tesfamariam, Statistical inference of sewer pipe deterioration using Bayesian geoadditive regression model, Journal of Infrastructure Systems, 25 (2019) 04019021. [4] A. Heifetz, D. Shribak, X. Huang, B. Wang, J. Saniie, J. Young, S. Bakhtiari, R.B. Vilim, Transmission of images with ultrasonic elastic shear waves on a metallic pipe using amplitude shift keying protocol, IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control, 67 (2020) 1192-1200. [5] M. Abbas, M. Shafiee, An overview of maintenance management strategies for corroded steel structures in extreme marine environments, Marine Structures, 71 (2020) 102718. [6] M. Hong, Z. Su, Q. Wang, L. Cheng, X. Qing, Modeling nonlinearities of ultrasonic waves for fatigue damage characterization: Theory, simulation, and experimental validation, Ultrasonics, 54 (2014) 770-778. 148
ADE
Chapter 6 Chapter 6. Conclusions and remarks 6.1. Summary of contributions This thesis has presented a comprehensive study on the ultrasonic guided waves in thin-walled structures immersed in liquid on one side. The multimodal and dispersive features of guided waves have been analyzed and the influence of the surrounding liquid medium on the guided wave propagation characteristics has been demonstrated both numerically and experimentally. The quasi-Scholte wave and the fundamental leaky symmetric Lamb (leaky S ) wave have been intensively 0 studied for characterizing corrosion pits and microstructural evolution, respectively. Both linear and nonlinear guided wave features have been discussed. The findings complement the current knowledge about guided waves in submerged structures and provide support for safety inspections of high-valued and critical infrastructures, such as liquid storage tanks, vessels and pipelines, and submarine hulls. Chapter 2 (Paper 1) has investigated the interaction of guided waves with corrosion pits in a steel plate loaded with water on one side. Among many other guided wave modes, the quasi-Scholte mode has been selected to characterize the dimensions of circular blind holes that are the simplest representations of progressive corrosion pits. The results have indicated that the quasi-Scholte mode has a high sensitivity to the physical conditions of the plate-water interface and is promising for evaluating corrosion damage in submerged structures. In the literature, the quasi-Scholte wave was rarely used for damage detection. This chapter has highlighted that applying the quasi-Scholte wave for detecting damage in submerged structures is limited to the low frequency range, at which the quasi-Scholte wave is dispersive. A case study has been presented for using the quasi-Scholte wave at 100 kHz to characterize the blind holes in a 2 mm thick steel plate with one side immersed in water. The advantages of the quasi- Scholte wave at low frequencies have been demonstrated, which include easy excitation, low attenuation, strong signal-to-noise ratios, and shorter wavelength compared to other guided wave modes (higher sensitivity to smaller defects). However, the quasi-Scholte mode becomes nondispersive at a high frequency and 153
ADE
Chapter 6 has most of its wave motions conserved in the liquid medium. The scattered waves caused by the damage can be hardly measured on the plate surface. Chapter 3 (Paper 2) has further investigated and compared the behaviors of guided waves in a dry plate surrounded by air and the plate with one side immersed in water. The foregoing studies demonstrated that mode conversion occurs when guided waves propagate from the dry plate to the immersed plate. However, the variation with the excitation frequency of the mode conversion phenomenon has not been discussed before. In this chapter, the fundamental anti-symmetric Lamb wave (A ) is excited on the dry plate section and travels to the water-immersed plate 0 section, where the generated A wave is mode converted to the quasi-Scholte wave. 0 The frequency dependence of the mode conversion from A wave to QS wave has 0 been studied numerically and experimentally. It has been discovered that the guided wave energy can shift in the frequency domain when the phase velocity of the incident A wave is larger than the sound speed of the surrounding liquid medium. 0 Due to the dispersive features of guided waves, the change in frequency components can make the guided wave propagation properties different. Therefore, the findings of this study are important for practical applications (e.g. using guided waves for damage detection of partially immersed structures and assessing liquid properties and levels). Chapters 2 and 3 have also indicated that the defects (e.g. corrosion pits) on partially immersed plates can be evaluated by sending A wave on the dry section 0 of the plate and measuring the quasi-Scholte wave on the immersed section. This method is very promising for long-range inspection because the quasi-Scholte wave has very low attenuation and does not radiate energy in the liquid. However, future study is required to develop an effective damage detection algorithm for the partially immersed structures to consider the change of wave behaviors due to the mode conversion phenomenon and the presence of liquid. The targeted defects in the first two papers (Chapters 2 and 3) are thickness thinning in local areas and have a size of around a few millimeters. These macroscopic defects can change the propagation speeds and amplitudes of the transmitted guided wave signals, of which the wavelength is comparable to the dimension of the defects. Recent studies on guided wave applications have focused 154
ADE
Chapter 6 on nonlinear guided wave features which provide the potential to identify and track the evolution of damage in the microscale that precedes macroscopic defects. Although nonlinear guided wave features have better sensitivity to microstructural defects, their generation and measurements are very challenging because the microscopic defects in the material are quite small. In the literature, several conditions were proposed for the structures surrounded by air to ensure that the nonlinear guided waves generated due to the material nonlinearity are cumulative and measurable. However, these conditions have not been validated for the structures submerged in liquid. Chapter 4 (Paper 3) has explored the feasibility of the second harmonics generation by guided waves in metallic plates with one side exposed to water. The dispersive behaviors of multiple guided wave modes have been analyzed and three criteria have been proposed in regard to selecting appropriate guided wave modes and excitation frequencies. Firstly, the selected guided wave modes at the excitation frequency (primary waves) should have low attenuation and have most wave motions conserved in the submerged structures. Secondly, there should be nonzero power flux between the primary waves and the corresponding second harmonics. Thirdly, the phase velocity deviation between the primary waves and second harmonics should be less than 1%. These criteria can be also applied to different structures exposed to various liquids. Then, a case study has been presented using experimental signals with an excitation frequency of 170 kHz. The results have indicated that cumulative generation of second harmonics can be achieved by leaky S wave. Next, a three-dimensional (3D) finite element (FE) model has been 0 developed to simulate the nonlinear guided wave generation. The material nonlinearity of the immersed plate has been simulated by a VUMAT subroutine that incorporates Murnaghan’s strain energy function. The numerical simulations have been validated through experimental measurements. After that, a series of parametric studies have been carried out and the results have shown that the second harmonics are sensitive to the early-stage damage in plates with one side immersed in water. However, the shortcoming of the second harmonic approach is that the instrumentations for sensing and actuating signals can also produce higher harmonics at the integer multiples of the excitation signals. It is very difficult to 155
ADE
Chapter 6 extract the second harmonics due to the material nonlinearity from the instrument nonlinearity. To cope with this limitation, Chapter 5 (Paper 4) has numerically and experimentally investigated the nonlinear guided wave mixing in an aluminum plate loaded with water on one side. Leaky S waves are excited at two different 0 frequencies on the wall of a metal tank filled with water. The results have shown that the nonlinear interaction between leaky S waves at two different frequencies 0 can produce cumulative combination harmonics at the sum frequency. Compared to the second harmonics studied in Chapter 4, the combination harmonics are less affected by the higher harmonics produced by the instrumentations, such as amplifiers and transducers. In addition, mixing guided waves with different frequencies provides more flexibility for the selection of guided wave modes and excitation frequencies. Finally, Chapter 5 has also presented parametric studies using the experimentally validated nonlinear FE model. The combination harmonics have shown a better sensitivity to the early stage of fatigue damage than the second harmonics. 6.2. Future work and recommendations The current research only considers the fundamental guided wave modes at low excitation frequencies for the metallic plates partially immersed in water (non- viscous liquid). However, the current research has built a foundation for the development of guided wave-based techniques for safety inspection of submerged plates or thin-walled structures, based on the excellent ability of guided waves to travel at fast speed over long distances and the high sensitivity to the evolution of microscopic defects. Below are the possible research directions for future studies. 1. The propagation characteristics of guided waves in plates immersed in viscous liquids (e.g. honey, oil, and gas). The guided wave energy can leak into viscous liquids through both shear and longitudinal wave motions. This is different from the structures immersed in non-viscous liquids. Shear waves do not exist in non-viscous liquids because shear forces cannot be 156
ADE
Chapter 6 sustained. Therefore guided waves in the plates immersed in viscous liquids can be significantly different from that in the non-viscous liquids. 2. The propagation characteristics of guided waves in submerged structures with different geometries (e.g. curved plates and T-section joints) and various materials (e.g. plastics and composites). Guided waves behave differently on different structures and materials. This future work is very important because real-world structures are more complex. 3. The feasibility of using guided waves to characterize multiple corrosion pits with irregular shapes. The damage in real-world structures can vary significantly in size, shape, and locations. Therefore, the effect of damage with increasing numbers and dimensions should be considered. 4. The acoustoelastic effect of submerged plates due to the hydrostatic pressure. Submerged structures containing liquids are usually subjected to hydrostatic pressure with varying aplitudes that can potentially induce and incrase microstructural defects. The acoustoelastic effect can change the propagation properties of guided waves and the generation of nonlinear guided wave features. 5. The interaction between guided waves and stress corrosion cracking. Stress corrosion cracking is one of the major concerns for the immersed metallic plate structures that are subjected to time-dependent loads with varying amplitudes. The present study only demonstrated that the micro cracks in the early stage of damage can generate measurable and low-attenuation second harmonics. The effect of the size, shape, and location of stress corrosion cracking in macroscale on the guided wave propagation can be future work. 6. The development of damage detection tools using guided waves and machine learning approaches. Current algorithms cannot be applied to structures with one side partly immersed in water as mentioned in Chapter 3 (Paper 2). A data-derived approach such as machine learning can help consider the change of wave behaviors due to the mode conversion phenomenon and the variation of liquid levels. 157
ADE
Abstract This thesis aims to investigate novel approaches in the field of Machine learning and advanced data analytics that can handle large data volumes and open new doors in the field of reservoir characterization. To begin, a new approach for rock typing is introduced using fractal theory where conventional resistivity logs are the only required data. Fractal analysis of resistivity logs showed that the fractal dimension of these logs which is a measure of the variability of the signal, is related to the complexity of the rock fabric. the fractal dimension of multiple deep resistivity logs in the Cooper Basin, Australia was measured and compared with the fabric structure of cores from same intervals. The results showed that the fractal dimension of resistivity logs increases from 1.14 to 1.29 Ohm-meter for clean to shaly sands respectively, indicating that the fractal dimension increases with complexity of rock texture. The thesis continues with a machine learning application to augment/automate facies classification using resistivity image logs. Given the complexity of the application, a supervised learning strategy in combination with transfer learning was used to train a deep convolutional neural network on available data. The results show that in the absence of other information/logs, the trained network can detect image facies with a testing accuracy of 82% form electric image logs and a proposed post-processing method increases the final categorization accuracy even further. An important step in reservoir characterization is understanding and quantification of uncertainty in reservoir models. In the next section a novel Generative Adversarial Network (GAN) architecture is introduced which can generate realistic geological models while iv
ADE
maintaining the variability of the generated dataset. The concept of mode collapse and its adverse effect on variability is addressed in detail. The new architecture is applied to a binary channelized permeability distribution and the results compared with those generated by Deep Convolutional GAN (DCGAN) and Wasserstein GAN with gradient penalty (WGAN-GP). The results show that the proposed architecture significantly enhances variability and reduces the spatial bias induced by mode collapse, outperforming both DCGAN and WGAN-GP in the application of generating subsurface property distributions. Finally, an advanced analytics technique for efficient history matching is proposed in the appendix. In this part of the thesis, an ensemble of surrogates (proxies) with generation-based model-management embedded in CMA-ES is proposed to reduce the number of simulation calls efficiently, while maintaining the history marching accuracy. History matching for a real field problem with 59 variables and PUNQ-S3 with eight variables was conducted via a standard CMA-ES and the proposed surrogate-assisted CMA-ES. The results showed that up to 65% and 50% less simulation calls for case#1 and case#2 were required. v
ADE
Declaration I certify that this work contains no material which has been accepted for the award of any other degree or diploma in my name in any university or other tertiary institution and, to the best of my knowledge and belief, contains no material previously published or written by another person, except where due reference has been made in the text. In addition, I certify that no part of this work will, in the future, be used in a submission in my name for any other degree or diploma in any university or other tertiary institution without the prior approval of the University of Adelaide and where applicable, any partner institution responsible for the joint award of this degree. The author acknowledges that copyright of published works contained within this thesis resides with the copyright holder(s) of those works. I give permission for the digital version of my thesis to be made available on the web, via the University's digital research repository, the Library Search and also through web search engines, unless permission has been granted by the University to restrict access for a period of time. I acknowledge the support I have received for my research through the provision of an Australian Government Research Training Program Scholarship. Roozbeh Koochak 02/04/2023 vi
ADE
Acknowledgement I am grateful to all those whose belief in me, encouragement, and support have been the driving force behind my achievements. Sahba, my loving wife, your endless patience, understanding, and belief in me have been the bedrock of my success. Your constant support and sacrifices made it possible for me to focus on my research and overcome challenges. Your presence by my side has been a source of strength, motivation, and joy, and I am forever grateful for your unwavering love. To my dear mother, Farideh, and siblings, Atousa, Reza, and Parisa, your consistent support, encouragement, and understanding have been invaluable. Your belief in my abilities and unwavering support, both emotionally and practically, have been instrumental in my accomplishments. Your presence in my life has filled me with gratitude and motivation to excel. I would like to extend my deepest appreciation to my esteemed supervisors, Dr. Manoucheher Haghighi, Dr. Mohammad Sayyafzadeh, and Dr. Mark Buch. Your expert guidance, knowledge, and dedication have shaped my research, broadened my horizons, and enriched my understanding of the subject matter. Your tireless efforts to provide me with valuable feedback, constructive criticism, and mentorship have played a crucial role in my academic growth and the successful completion of my Ph.D. I am truly grateful for the opportunities you have given me to explore and contribute to the field. Lastly, I would like to thank my dear friends, Dr. Ali Nadian, Colin Jordan, and Dr. Martin Roberts for their unwavering support, interest in my work, and willingness to lend a helping hand whenever needed. Your intellectual discussions, feedback, and collaborative spirit have been instrumental in shaping my ideas and sharpening my research focus. vii
ADE
Thesis by publication This is a thesis by publication and is composed of multiple pieces of work. This includes 4 publications in total. 2 published in peer-reviewed journals, 1 submitted for publication to a peer-reviewed journal and 1 conference paper. These are the details below: Published Peer-reviewed Journal Papers: Koochak, R., Haghighi, M., Sayyafzadeh, M. and Bunch, M., 2018. Rock typing and facies identification using fractal theory and conventional petrophysical logs. The APPEA Journal, 58(1), pp.102-111. Koochak, R., Sayyafzadeh, M., Nadian, A., Bunch, M. and Haghighi, M., 2022. A variability aware GAN for improving spatial representativeness of discrete geobodies. Computers & Geosciences, 166, p.105188. Submitted to Peer-reviewed journal for publication: Roozbeh Koochak, Ali Nadian Ghomsheh, Manouchehr Haghighi, Mark Bunch, Mohammad Sayyafzadeh, A transfer learning approach for facies prediction using resistivity image well logs. Submitted to Geoscience Frontiers for publication. Published Conference paper: Sayyafzadeh, M., Koochak, R. and Barley, M., 2018, September. Accelerating cma-es in history matching problems using an ensemble of surrogates with generation-based management. In ECMOR XVI-16th European Conference on the Mathematics of Oil Recovery (Vol. 2018, No. 1, pp. 1-15). EAGE Publications BV. viii
ADE
1. Introduction 1.1 Problem statement Reservoir characterization includes all techniques and methods that enhance our understanding of the geologic, Geo-chemical and petrophysical controls of fluid flow. it is a constant process that Continuously evolves from field discovery to development and production down to abandonment phase. Throughout these stages, knowledge of the subsurface continually enhances as new data is received. This makes characterization of the reservoir a dynamic process. The challenge is that the dynamic properties of the reservoir such as Pressure, Saturation or Porosity continually change as the reservoir is produced. This together with the fact that most of the data measurements are indirect along with error of measurement, causes uncertainty in the reservoir characterization. Adequate characterization of a reservoir is increasingly important for optimising field development, reservoir evaluation and production. Also, in recent years with novel subsurface applications such as carbon capture and sequestration or hydrogen storage, detailed characterization is imperative to successful operations throughout the life a field. An integrated approach for characterization and/or modelling of the reservoir can tear down traditional disciplinary divides and lead to better understanding and handling of uncertainties in the reservoir. With this statement in mind, the methodologies presented in this thesis have multidisciplinary applications where the disciplines of Petrophysics, Reservoir Engineering and Geology have been brought together. Reservoir characterization generally involves estimating reservoir parameters at different locations by correlating collected data from a wide variety of sources. Sources of data for subsurface reservoirs include cores, well logs, Seismic surveys, production data and outcrop analogues. These sources of data all have different scales and vary in dimensionality. Logs have 1
ADE
high resolution but on a small scale, generally, 1D or 2D in dimension. Well logs provide a vertically high-resolution model at the well locations. However, the distribution of well locations is sparse and biased towards proven section of the field. Seismic data on the other hand is large scale and covers extensive areas but with low resolution. Dimensionality of this data type is 2D, 3D and sometime 4D. Combining well logs with geophysical and geological data will provide the necessary constrains required to extrapolate the high resolution well data beyond where they are measured. Core data, as 3-dimension datatype, is classed as hard data and presents the most accurate information with highest resolution however their availability is limited as the acquisition of cores, is expensive and requires a lot of effort. The most effective utilisation of these data sources is, therefore, combining these different data sources. This approach results in the best and most complete description of reservoir which is generally referred to as integration modelling. Integration of different data sources enhances understanding of the reservoir, reduces uncertainties and mitigates risks. The goal of characterization is to develop different models that can be used in analytical or numerical evaluation methods. The oil and gas industry are experiencing a surge in the amount of data they are receiving from their fields. Field data, in recent years, has expanded in volume, velocity and complexity. This includes significant increase in the number of sensors in the field and in the pipelines, connected through Internet of Things (IoT). Resolution and sampling rate of wireline data has increased and complex data such as 4D seismic are becoming more common and more frequently recorded. Advent of technologies such as fibre optics has made access to these data virtually instant. The industry has always been overwhelmed with large quantities of data but was never able to make efficient and productive use of this data. Traditional methods in subsurface energy generally need to compromise between data size and complexity on one 2
ADE
hand and fidelity of the model (they are used to construct) on the other. In recent years Data analytics and Artificial Intelligence (AI) - Machine Learning (ML) more specific in this thesis - have introduced new methodologies that not only handles large and complex data types but is able to process the data much faster than traditional methods. This in part is also driven by advances in commodity hardware. Furthermore, ability of these methods to identify and learn features and patterns in the data makes them a great tool to unlock new insights, extract more information, develop new usage and leverage and optimize untapped data. In this work, the aim is to develop and implement advanced analytics and machine learning techniques with a focus on reservoir characterization using multi-scale, multi-dimensional data. Firstly, a novel application of fractal dimension of resistivity logs is presented. In this one- dimensional, small-scale application porosity and permeability derived from cores was rock typed into categories and the fractal dimension of the corresponding deep resistivity log was calculated. The correlation between the rock fabric and fractal dimension of the resistivity logs was investigated and a new method to make direct use of these logs in rock typing is proposed. This research investigates the effect of pore structure on the variability of resistivity logs. It takes a step further than just interpreting resistivity logs based on change in average value and signature and reveals information at pore scale. In this study only the effect of pore structure on the variability and fractal dimension of logs was investigated. Further, research is required to determine the effect of other factors like fluid in the rock or different shale types. Bearing in mind that fractal dimension is rather a quality parameter and a flag for change in pore structure. Then in a 2-dimensional small-scale application, a machine learning algorithm complete with practical data pipeline is introduced to augment/automate the tedious process of resistivity image log interpretation. In this technique, a convolutional neural network (CNN) is trained to learn interpreted facies categories in one well, then the trained network is used 3
ADE
to detect the learned facies categories in a newly drilled well. The ability of CNNs to learn and identify features in geological facies images is thoroughly investigated by studying the confusion matrix. Multiple CNN architectures are compared and challenges in the application are identified, and solutions presented. Next, we apply machine learning to a 2-dimensional large-scale application. In this study, Generative Adversarial Networks (GANs) are used to quantify uncertainty in a field or basin wide scale. The concept of Training Image (TI) is reviewed. Methodology to generate network training data from a single TI is presented and compared with other traditional geo-statistics methods. The variability of generated realizations (which is detrimental in geological uncertainty quantification) using GANs is thoroughly investigated including the concept of mode collapse and the effect of input training data. A novel architecture specialized for maintaining the variability of geological realizations at the same level as the input training data is presented. Working with large data sets and complex algorithm, computation efficiency and cost becomes an important factor in popularity of a characterization method. We enhance the computational efficiency of a Covariance Matric Adaptation Evolutionary Strategy (CMA-ES) by proposing an online learning scheme to update an ensemble of proxies. The effective ness of the technique was evaluated on two different history matching cases and other techniques such as generation-based model management and evolution control were examined. 1.2 Thesis Structure This is a thesis by publication. Chapter 1 begins with an introduction to reservoir characterization and its challenges. It describes how novel advanced analytics techniques and machine learning algorithms can enhance the process of reservoir characterization and describes the aims of this thesis. The contribution of each publication to this thesis, is also 4
ADE
analytics and state-of-the-art techniques such as machine learning are utilized to better understand uncertainty, develop new methods, and potentially extract more information and propose new usage of the traditional data, while considering, different data sources with different scales and dimensionality. Furthermore, computational efficiency and automation is addressed where these techniques are adaptable. This work begins by introducing a new application for conventional resistivity logs which are one dimensional and on a centimetre scale. In the paper titled "Rock typing and facies identification using fractal theory and conventional petrophysical logs" a technique to use fractal dimension of resistivity logs for rock typing and flow unit classification is proposed. Rock typing is an integral part of reservoir characterization. In this process the reservoir is subdivided into layers based on similar properties and flow points. In other words, rock fabric of each layer, that is, pore throat dimensions, geometry, size, distribution, and capillary pressures must be similar. This enhances flow behaviour modelling and, significantly reduces uncertainty and risk of predicting production and/or injection in the field. In this study, porosity and permeability measured from cores were correlated with fractal dimension of corresponding deep resistivity logs. A methodology to determine fractal dimension from 1D data is proposed and the propagation of ions and electric current in the rock fabric along with its relationship and effect to fractal dimension of the resistivity logs is investigated. Traditionally, only the change in average of resistivity log over an interval is utilized when interpreting resistivity logs. The results of this investigation show that further information can be derived from these logs. For example, presence of layered beds with thicknesses less than the resolution of the tool can be flagged using is effect of these beds on fractal dimension of the resistivity log. While this is not possible using conventional interpretation methods. This study is presented in detail in chapter 3. 6