Next Article in Journal
An Efficient Improved Harris Hawks Optimizer and Its Application to Form Deviation-Zone Evaluation
Next Article in Special Issue
Estimation Method of an Electrical Equivalent Circuit for Sonar Transducer Impedance Characteristic of Multiple Resonance
Previous Article in Journal
Study on Multi-Heterogeneous Sensor Data Fusion Method Based on Millimeter-Wave Radar and Camera
Previous Article in Special Issue
All-Optical, Air-Coupled Ultrasonic Detection of Low-Pressure Gas Leaks and Observation of Jet Tones in the MHz Range
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Embedded Ultrasonic Inspection on the Mechanical Properties of Cold Region Ice under Varying Temperatures

1
College of Mechanics and Materials, Hohai University, Nanjing 210024, China
2
College of Water Conservancy and Hydropower Engineering, Hohai University, Nanjing 210098, China
3
College of Civil and Transportation Engineering, Hohai University, Nanjing 210098, China
*
Author to whom correspondence should be addressed.
Sensors 2023, 23(13), 6045; https://doi.org/10.3390/s23136045
Submission received: 2 June 2023 / Revised: 16 June 2023 / Accepted: 19 June 2023 / Published: 29 June 2023
(This article belongs to the Special Issue Acoustic Sensors and Their Applications)

Abstract

:
The mechanical properties of ice in cold regions are significantly affected by the variation in temperature. The existing methods to determine ice properties commonly rely on one-off and destructive compression and strength experiments, which are unable to acquire the varying properties of ice due to temperature variations. To this end, an embedded ultrasonic system is proposed to inspect the mechanical properties of ice in an online and real-time mode. With this system, ultrasonic experiments are conducted to testify to the validity of the system in continuously inspecting the mechanical properties of ice and, in particular, to verify its capabilities to obtain ice properties for various temperature conditions. As an extension of the experiment, an associated refined numerical model is elaborated by mimicking the number, size, and agglomeration of bubbles using a stochastic distribution. This system can continuously record the wave propagation velocity in the ice, giving rise to ice properties through the intrinsic mechanics relationship. In addition, this model facilitates having insights into the effect of properties, e.g., porosity, on ice properties. The proposed embedded ultrasonic system largely outperforms the existing methods to obtain ice properties, holding promise for developing online and real-time monitoring techniques to assess the ice condition closely related to structures in cold regions.

1. Introduction

In recent decades, research on ice has grown remarkably. Intensive research is being performed on whether to examine global climate change [1,2,3], to break ice sheets in the arctic with icebreakers [4], to investigate ice-covered electric antennas [5], to design ships and offshore structures [6,7], to study the icing conditions in aviation [8], etc. In those investigations, it is essential to obtain the mechanical properties of ice, which will change with variations in ambient temperature.
Currently, diverse measurement methods of ice mechanical properties have been developed in an experimental way, including uniaxial compression [9,10,11], triaxial compression, and flexural strength tests [12]. Aksenov et al. [13] attained the preliminary temperature-related stress-strain properties of the freshwater ice by performing short-time uniaxial compression tests for cylindrical ice specimens at various specific temperatures. Moslet [14] conducted the uniaxial compression strength tests on columnar sea ice in the field on Svalbard, with the indication of a strong relationship between Young’s modulus and ice porosity. Qiu et al. [15] obtained the compressive and tensile plastic properties of ice based on the triaxial compression test of columnar ice. However, due to the brittleness of the ice material itself, fracture is inevitable in the experiment. Moreover, the time between the testing starting and the fracture ending is very short. Therefore, in a single experiment, it is not possible to obtain changes in ice properties caused by continuous temperature changes. Additionally, those destructive tests are costly and inappropriate for structures in service. In order to obtain the mechanical properties of ice that continuously change with temperature without damaging the existing ice structure, a non-destructive testing method is needed that can achieve real-time monitoring and is suitable for ice materials. Accordingly, a non-destructive method for estimating the temperature-related mechanical properties of ice is demanded.
The ultrasonic technique is one of the most commonly used non-destructive methods in characterizations of material properties [16,17], derivations of dynamic responses [18], and condition monitoring [19,20,21,22,23]. Relevant sensing techniques for wave generation and measurement have been developed rapidly. Representative research works are as follows: Bayón et al. [24] employed the measured Rayleigh wave velocity and the aspect ratio of the elliptical trajectory amplitudes to obtain the Young’s constants of isotropic linear materials. Further, Medina and Bayón [25] proposed a method for calculating the dynamic Young’s constants of an anisotropic plate through the measured impact-echo resonance and Rayleigh wave velocity. Medina and Bayón [25] determined the mechanical properties and damage properties of a multilayer composite board by combining experimental and numerical data simultaneously. Ultrasonic technology has also been implemented in the estimation of ice properties. Bock and Polach [26] explored the applicability of the collected long-period surface wave dispersions in the inversion of ice shell thickness. Gagnon [27] presented an impulse-echo method to evaluate the longitudinal ultrasonic velocities of three different types of specimens to calculate Young’s modulus. Noteworthy, a state-of-the-art literature review specified that only a limited number of studies dealt with the utilisation of ultrasonic techniques in the determination of ice properties, and there are still some crucial issues to be solved. In addition, after obtaining the monitoring data, this study used the methods mentioned in [28,29] to conduct the corresponding data modelling and analysis, including regression analysis and spectral analysis.
In addition to experimental testing methods, numerical simulation is a valuable alternative method for estimating the ice properties by taking advantage of wave propagation simulation techniques [22,30,31,32,33]. Various numerical models were established to represent the ice materials, including Young’s non-linear behaviour model [34], the crushable foam model [35,36], and the user-defined elastic-plastic material model [37,38]. Liu et al. [6] proposed a quasi-static model on the basis of the strain rate-dependent plasticity theory to simulate ice structural behaviour by combining the Tsai Wu yield surface criteria and the associated flow rule. Bock and Polach [26] analysed the non-linear behaviour of the aqueous model ice, concluding that the non-linear behaviour of ice is independent of its crystal structure and chemical dopant. Gagnon et al. [27] adopted a “crushable foam” material type in LS-DYNA to model the ice behaviour, with the numerical results of the ‘calibrated stress-volumetric strain relationship’ showing a good agreement with the experimental ones. Noticeably, the existing numerical cases lacked studies on the influence of temperature variations and ice porosities.
From the surveyed literature, it is observed that the current prevalent methods to determine ice properties commonly rely on one-off and destructive compression and strength tests, which are incapable of acquiring the temperature variation-induced changes in ice properties. Moreover, ultrasonic technology, as one of the most commonly used non-destructive methods for characterising material properties, is rarely applied to ice mechanical properties. In addition, due to the unique nature of ice materials, it is necessary to study how ultrasonic sensors can be embedded in ice. Furthermore, there is a lack of numerical models for wave propagation in ice with different temperatures and porosities. To address these issues, an embedded ultrasound system is proposed to test the mechanical properties of ice, and its feasibility and effectiveness have been verified through ultrasound experiments and numerical models.
The rest of this paper is organised as follows: Section 2 formulates a novel ultrasonic method for determining the mechanical properties of ice. Section 3 builds the 2D and 3D ice models, with the numerical results at different temperatures and porosities being counted. Section 4 exhibits comparisons and discussions of experimental and numerical results. Section 5 presents the conclusions.

2. Test-Based Identification of the Mechanical Properties of Ice

The current prevalent methods to acquire the mechanical properties of ice are typically uniaxial compression, triaxial compression, and flexural strength experiments. However, due to the brittleness of the ice material itself, fracture is inevitable in the experiment. Moreover, the time between the testing starting and the fracture ending is very short. Therefore, in a single experiment, it is not possible to obtain changes in ice properties caused by continuous temperature changes. To address the deficiencies of existing methods, an embedded ultrasonic system is built as an alternative to compression tests to obtain the mechanical properties of ice at different temperatures.

2.1. Test Specimen

2.1.1. Actuators and Sensors

Piezoelectric material can be used as both actuators and transducers, which is beneficial for its direct and inverse piezoelectric effects [39]. Figure 1c shows a piezoelectric ceramic transducer (PZT), with its dimensions and material properties shown in Table 1. In the PZT, the electric wires are welded to the positive and negative poles on their different sides. Both sides are waterproof with the transparent insulating glues. The temperature sensor is made of a 4 mm-diameter, waterproof platinum probe (shown in Figure 1d). The measurement range is −50–+200 °C.

2.1.2. Ice Specimen and Sensing Strategy

(1)
Ice specimen
The ice specimen is fabricated using distilled water with the purpose of avoiding excessive impurities. The dimensions of the ice specimen are 150 × 150 × 200 mm3. Before freezing, all the required sensors during the ultrasonic tests are fixed in the middle of the mould with steel wires, as shown in Figure 1b.
The process for making ice samples is as follows: first, place the mould with distilled water and sensors into a refrigerator with an initial temperature of −5 °C; second, when the surface water is frozen, progressively decrease the temperature of the refrigerator by 5 °C per two hours; next, when the refrigerator temperature reaches −35 °C, the ice specimen should be fully made; and finally, the newly-made ice specimen needs to be stored in a refrigerator at −35 °C for at least 48 h to ensure a uniform temperature distribution in the ice specimen. In this way, the ice cracks caused by big temperature variations could be avoided.
(2)
Sensing strategy
The PZTs and temperature sensor are set in the middle of the ice specimen. The distance between the two PZTs is considered to be 70 mm. Additionally, the temperature sensor is set at the uniform level of the PZTs due to the effect of temperature stratification in ice.
(3)
Considerations
Preliminary tests with a distance between the two PZTs are conducted to determine the appropriate properties. Comparing the results at 25 mm, 50 mm, and 70 mm, it is found that when the distance between the two PZTs is small, the wave packets of the longitudinal wave and the shear wave cannot be distinguished. The reason is that the propagation speed of the wave is very fast. When the shear wave propagates, the longitudinal wave has not been fully received, which causes the longitudinal wave and shear wave to be superimposed on each other, so it is difficult to distinguish a separate shear wave packet.

2.2. Velocity of the Transmitted Waves

Wave velocity, one of the key properties of wave propagation research, is defined as the velocity at which a disturbance propagates in specified materials. It mainly depends on the material properties, structural geometries, and external excitations. The longitudinal and shear wave velocities are the most widely used variables in ultrasonic-structural analysis. Their relationships with structural properties are formulated as follows:
V l = E 1 μ ρ 1 + μ 1 2 μ ,
V s = E 2 ρ 1 + μ
According to Equations (1) and (2), the material properties (such as E, G, and μ) can be achieved with the measured longitudinal and shear wave velocities (Vl and Vs) (shown in Equations (3)–(6)). These relationships, which are the fundamentals of the characterizations of the ice properties, can be written as follows:
μ = V l 2 2 V s 2 2 V l 2 V s 2 ,
E = V s 2 ρ 3 V l 2 4 V s 2 V l 2 V s 2 ,
G = ρ V s 2 ,
K = ρ V l 2 4 3 V s 2 ,
where ρ is the density, E is Young’s modulus, μ is Poisson’s ratio, and Vl and Vs represent the velocity of the longitudinal wave and the shear wave, respectively.

2.3. Experimental Setup and Methods

(1)
Test process
The experimental apparatus for the ultrasonic test is presented in Figure 2.
The excitation part includes the wave generator and desktop. Firstly, the waveform of excitation is generated on the desktop. Then, a tone burst signal, consisting of sinusoids modulated by the Hanning window [40], is utilised as the external excitation. This signal can be expressed as follows [41]:
x t = A 2 sin 2 π f c t 1 cos 2 π f c t n ,
where A is the amplitude, n denotes the number of signal cycles, and fc represents the central frequency.
The designated excitation is then transferred into the Agilent 33250A arbitrary wave generator, which converts the excitation from a digital signal to an analogue one. The output terminal of the waveform generator is connected to the PZT through a Bayonet Nut Connector cable. Because of the positive piezoelectric effect, the PZT deforms in the longitudinal direction of the ice, leading to the generation of longitudinal waves. The output voltage of the waveform generator, which regulates the excitation amplitude, is set to be 10 V in order to ensure a fully deformed piezoelectric ceramic sheet.
In the receiving part, the PZT transducer generates current due to the inverse piezoelectric effect. Based on the characteristics of PZT, this generated current has a linear relationship with its deformation. Thus, the captured current from the receiver can be regarded as the propagated longitudinal wave in the ice. This received current is then amplified by a fixed-gain universal preamplifier, PXPA3, to increase its amplitude. The transfer gain of the charge amplifier is 10 mv/pc. Then the current is introduced into the Agilent DSO7034B oscilloscope to be converted into a digital signal. The sampling frequency of the oscilloscope is set to 50 MHz. In particular, the application of the waterproof glue on the surface of PZT not only makes the sensor insulated but also prevents the generated current from leaking into the ice. In addition, there is no extra circuit other than the experimental equipment. Therefore, the measured current from the receiver thoroughly originates from the inverse piezoelectric effect caused by the longitudinal wave.
During the ultrasonic test, the ice specimen is covered with a foam box, which is a common thermal insulation material in daily life and is used to slow down the impact of external temperature on water, thus slowing down the speed of water icing. If the freezing speed is too fast, it can cause the ice surface to expand and crack. Cracking ice will badly affect the experimental results. The internal temperature is monitored in real-time through the temperature sensor. By continuously recording these internal temperatures and the corresponding received signals, the non-destructive monitoring of wave propagation in ice under the condition of temperature variation is realised.
(2)
Determination of signal properties
Two properties (n and fc) on the right-hand side of Equation (7) need to be determined in advance. Preliminary tests with different n and fc are conducted to determine the appropriate properties. The measured ultrasonic signals with different n and fc are shown in Figure 3 and Figure 4, respectively.
In the preliminary test, the response signal results of 2 periods, 3 periods, 4 periods, and 5 periods are compared and analysed. The experimental results in Figure 3 indicate that the second peak of the measured signals can be more easily distinguished in cases of having a smaller number of cycles (n = 2, 3) than in cases of having a larger number of cycles (n = 4, 5).
On the other hand, the results in Figure 4 imply that the second peak of the measured signals is easier to identify in cases of having larger central frequencies (fc = 250 kHz, 300 kHz, 350 kHz, 400 kHz, 450 kHz, and 500 kHz) than in cases of having smaller central frequencies (fc = 150 kHz and 200 kHz).
Accordingly, a normalised 2-cycle 250 kHz tone burst signal (shown in Figure 5) is employed as the external excitation in the experimental tests below.

2.4. Experimental Results

2.4.1. The Threshold Denoising Based on the Wavelet Transform

The mode reflection/conversion is inevitable at the external and internal boundaries of the ice specimen, contributing to a complex multi-mode wave signal. Moreover, the measured signals could be adversely affected by environmental noise. Therefore, it is necessary to rectify the measured signals to extract the inherent characteristics of the interested signal rather than those induced by mode reflection/conversion and noise. In this study, a wavelet transform-based threshold denoising process is adopted, with the results presented in Figure 6.

2.4.2. Experimental Results

Considering the faster propagation of the longitudinal wave compared with the shear wave, the first and second peaks of the ultrasonic waveform can be regarded as the longitudinal wave and the shear wave, respectively. Figure 7 shows four denoised ultrasonic waveforms measured at different temperatures. According to Figure 7, the first peak of the presented waveforms cannot be clearly identified. Under these circumstances, the first wave trough of the excitation cycle is selected as a reference to calculate the propagation velocities of longitudinal and shear waves in the ice. Correspondingly, the selected troughs of the longitudinal and transverse waves are annotated by the green circleabaquss in Figure 7.
Figure 8 presents the calculated wave velocities with temperatures ranging from −35 °C to −0.5 °C. Evidently, the velocities of longitudinal and shear waves are strongly related to temperature. To explicitly interpret the relationship between temperature and wave velocities, two quadratic functions are utilised to fit the velocity-temperature curves in Figure 8, respectively, which are:
V l = 0.01139   T 2 4.36647   T + 3783.56 ,   R = 0.99598 ,
V s = 0.00401   T 2 1.14069   T + 1797.705 ,   R = 0.99449 ,
where T represents the temperature and R is the correlation coefficient.
Table 2 shows the longitudinal wave velocity in ice obtained from existing research. The velocity values are in good agreement, which validates the accuracy of experimental results in this study. The wave velocity and Young’s modulus strongly depend on properties (such as salinity, impurity content, etc.) during the production of the ice, and the influences of temperature and time are not taken into account in these experiments. This may be the reason behind the deviation from the values, such as the distance of the transducers measured in the present experiment.
In the experiments, the density of the utilised ice specimen was 890 kg/m3, which is the average value determined by the drainage method [45,46,47]. Combined with Equations (3)–(6), the mechanical properties of ice at different temperatures can be achieved, including the Young’s modulus, shear modulus, Poisson’s ratio, and bulk modulus. The recognition results are exhibited in Figure 9, with the vertical axis representing the corresponding properties and the horizontal axis signifying the temperature variations. All the properties presented in Figure 9 decline continuously with increasing temperature.

3. Physics-Directed Numerical Simulations

In nature, ice contains pores, which are related to the growth process of ice. The formation of ice from water is a process from the outside to the inside. After the surface freezes, the air in the water cannot pass through the ice, forming pores inside the ice. These pores vary in size and are randomly located, so randomly distributed circular pores are used in the manuscript to simulate the pores inside the ice. The dimension, distribution, and amount of the entrapped air bubbles highly affect the dynamic mechanical properties of ice. Hou et al. [48] indicted that those bubbles in ice could neither be manufactured artificially nor obtained in the desired amount and distribution. Therefore, the influence of the pores in the ice on the mechanical properties can only be studied by numerical methods. In this study, a stochastic algorithm is presented to produce the sparsely distributed air bubbles in the numerical model of ice.

3.1. Material Properties of the Ice Model

In order to better compare the numerical and experimental results, four sets of experimental results were selected as the model material’s mechanical properties. The specific values are shown in Table 3.

3.2. Two-Dimensional Ice Random Pore Model

In the numerical model, ice is treated as a linear solid without considering its nonlinear behaviour or tensile or compressive damage. The dimensions of the tested specimens are 210 × 210 mm2, and the pores (air bubbles) are concentrated within a central region of 180 × 180 mm2. Additionally, the round pores in the 2D model are utilised. Before establishing the model, we observed and measured a large number of pores in natural and artificial ice. These pores, except for those larger than 5 mm on the surface of the ice, are very small in size inside the ice, some of which cannot be measured using conventional measurement methods. In order to restore the uniformly distributed pores accumulated in the ice, the porosity during simulation was fixed, the distribution was selected from the range of most pore sizes, and the appropriate pore distribution was calculated using MATLAB R2018a. They are classified into small pores having diameters of 0.5–1 mm and large pores having diameters of 1–1.5 mm, with the volume ratios being 6:4. This volume ratio was obtained through extensive testing, which not only meets the requirements of achieving porosity within the model but also achieves a uniform distribution of pores within the model. Moreover, the pores in the ice model are reciprocally independent, with distances between the pores greater than 1.5 mm. In the modelling process, random pores are first generated using MATLAB R2018a, and then the generated pore model is imported into the ice model built in ABAQUS 6.14. The final model is obtained by cutting out random pores from the original ice model. In the model, the excitation of the PZT actuator is simplified by applying a concentrated force at the same position as the PZT sensor [49]. The concentrated force is generated by the same 2-cycle 250 kHz tone burst signal as the experimental test (as shown in Figure 5). In addition, the distance between the actuator and receiver is 70 mm. In the model, the boundaries around the ice are the boundaries of wave propagation.
The porosity is obtained using the calculation method of Cox and Weeks [50]. The specific method can be expressed as:
v a = 1 ρ ρ i T i + ρ S i F 2 T i F 1 T i ,
The density of pure ice ρ i (g/cm3) is described as a function of temperature as [49]:
ρ i T i = 0.917 1.403 × 10 4 T i ,
where   v a   is the air volume ratio, ρ i is the pure ice density, S i is ice salinity, and T i is the ice temperature. In this study, distilled water is used to make ice samples, so in the numerical model S i   = 0 is adopted. Therefore, Equation (10) becomes:
v a = 1 ρ ρ i T i .
In the established numerical model, the density of the ice sample obtained from the experiment is employed to calculate the porosity at different temperatures following Equations (10)–(12).
Three 2D ice models (shown in Figure 10 and Figure 11) with different porosities (0.5%, 1%, and 3%) were established to investigate the influence of porosities on the wave propagations in ice. The white dots in Figure 10 and Figure 11 represent pores within the ice, while the green areas in Figure 10 represent the ice. The blue part in Figure 11 shows the grid inside ice. The mesh size is 1 mm, and the time step of numerical integration is 2 × 10−8 s. These small computational properties can ensure the accurate capture of the propagative behaviour of the ultrasonic wave.

3.3. Numerical Simulations of 2D Ice Random Pore Model

3.3.1. Wave Propagation Analysis with 0.5% Porosity

Figure 12 shows the wave propagation process of the ice model with 0.5% porosity at different time values. Both the longitudinal and shear waves can be clearly distinguished within a short period of time after being excited. The red circle in Figure 12d shows that waves are superimposed on each other during propagation, and various waves cannot be separated. Figure 12e shows a horizontal rebound wave at the red circle. The whole specimen in Figure 12f is filled with mixed waveforms, and it is impossible to distinguish between shear waves and longitudinal waves.
The waveforms received by the response point at four temperatures are extracted, and the drawn waveform is shown in Figure 13. As can be seen in Figure 13, there is no significant change in the morphology of the waves at the four temperatures. If the image is not enlarged, it is difficult to detect the difference between the four curves. Therefore, we believe that the change in temperature has little effect on the waveform. However, in the red dashed circle in Figure 13, the results of amplifying the peaks show that the time of the peaks appearing in the four curves is different, which means that the wave propagation speeds corresponding to the four temperatures are different (because the distance between the sensors remains constant).
With the application of ice properties from experiment tests, the numerical results of longitudinal and shear wave velocities at different temperatures can be obtained, as shown in Figure 14.

3.3.2. Non-Porous 2D Ice Model

To reveal the existence of pores on the wave propagation in ice, this section compares the numerical results attained by a non-porous ice model and a model with 3% porosity. The result of −17.2 °C is randomly selected to demonstrate the process of wave propagation in ice in numerical simulations. Of course, other temperature results can also be selected for display. The selection of different temperatures has no effect on the propagation process of waves in ice. The comparison results of wave propagations are depicted in Figure 15 and Figure 16. In Figure 15b,c, the longitudinal and shear waves can be apparently inspected. According to the non-porous ice models in Figure 15b–f, no wave superpositions are observed near the excitation point, demonstrating the boundary of the critical rebound effect of the pore boundaries on the wave propagations in ice, even when the pore size is small.
Figure 17 presents the normalized waveforms of the response point. The circles represent the positions of the peaks used in velocity calculation, with the red circles indicating the peaks used for calculating the longitudinal wave velocity and the green circles used for calculating the shear wave velocity. According to Figure 17, the presence or absence of pores obviously affects the wave propagation speed. In Figure 17, the amplitude in the first wave packet of the non-porous ice model is greater than that of the porous ice model. However, in the range of 0.4–0.5 s, the amplitude of the porous ice model is greater than that of the non-porous ice model. This is because the pore size in the porous model is 0.5–1 mm, which is much smaller than the wavelength (about 1500 m). Therefore, waves will form multiple reflections, refractions, and diffraction superpositions under the rebound effect of the pore wall.
Figure 18 shows the normalised waveforms of the three porosities at −17.2 °C. The circles in Figure 18 indicate the values used for calculating the wave velocities, with the red circle is used for calculating the shear wave velocity and the green circles are used for calculating the longitudinal wave velocity. It can be seen from Figure 18 that the higher the porosity, the slower the wave propagation velocity.

3.4. Numerical Simulation Results of the 3D Ice Random Pore Model

In this section, the 3D random pore ice model is built by expanding the 2D model. The dimensions of the investigated 3D model are 150 × 150 × 50 mm3, and the pores are concentrated within a central cuboid range of 140 × 140 × 45 mm3. Similar to the 2D model, the random pores in the 3D model are spheres, which are classified into the small size with diameters of 0.5–1 mm and the large size with diameters of 1–1.5 mm. The ratio of small and large pores is 6:4. Each pair of pores has a distance larger than 3 mm. Additionally, the porosity of the 3D model can be designated in accordance with the porosity of the 2D model (shown in Equations (10)–(12)). Three 3D ice models with porosities of 0.5%, 1%, and 3% are counted below.
The coordinates of the centre and the diameter of the pores are randomly generated. When the diameter rotates 360 degrees around the centre, a spherical pore is developed. All pores of the entire model are shown in Figure 19a. By subtracting the pores from the whole ice, the final 3D ice model with random pores, as shown in Figure 19b, can be obtained. In addition, the three columns in Figure 19 represent three model diagrams with different porosities.
The excitation signal in the 3D model adopts the waveform signal of Figure 5 and acts with concentrated force on the excitation point. The distance between the excitation and the response points is 70 mm. The centre frequency of the sine wave is f c = 250 kHz. As the mesh size of the model is 1 mm, the time step is 2 × 10−8 s. The grid map is shown in Figure 20, where Figure 20b is a planned view of the model and Figure 20c is an enlarged picture of the red circle of Figure 20b, which is used to highlight the grid at the gap. The red circles in Figure 20c highlight the pores in the model.

3.5. Numerical Results from the 3D Ice Model

3.5.1. Wave Propagation in Ice with 1% Porosity

Figure 21 presents the propagation process of waves in the 3D model. According to Figure 21a–c, within a short period of time after excitation, the longitudinal and shear waves can be clearly distinguished. Nevertheless, as time goes on, the longitudinal and shear waves become bounced and superimposed on each other. Compared with the 2D model, these bounces and superpositions of ultrasonic waves in the 3D model are more serious and complex.
The mechanical properties of ice at various temperatures obtained from the experiment are input into the 3D numerical model as the material properties, and the shear wave together with the longitudinal wave velocities at different temperatures can be obtained. Finally, all wave velocities are plotted as a scatter diagram, as shown in Figure 22.
Four groups of data are selected from all these numerical outcomes. The waveforms of wave propagation in ice at these four temperatures are shown in Figure 23. It can be seen that the shape of the waveforms at each temperature is similar, and there are obvious differences in the speed of wave propagation at different temperatures. The positions of the red circles in Figure 23 correspond to the first troughs of the longitudinal wave and the shear wave, respectively.

3.5.2. Comparisons

This part investigates the influence of porosity on wave propagation in ice using a 3D model. The normalised waveforms at different porosities are shown in Figure 24. Similar to the results in 2D, the porosity only affects the speed of wave propagation, and the influence on the waveform can be ignored. The circles represent the peak positions used for velocity calculation. Among them, the red circle and the green circle represent the peak values used for calculating longitudinal wave velocity and shear wave velocity, respectively. The cluttered waveforms at the tail (the wave propagation time   t 5.5 × 10 5 ) in Figure 24 indicate the complex superposition and rebound of waves in the 3D model.

4. Discussion of Experimental and Numerical Results

This section compares the numerical and experimental results to explicitly verify the reliability of the established 2D and 3D numerical ice models. Taking the case of temperature −17.2 °C as an example, the corresponding numerical and experimental results are presented in Figure 25. The figure shows that the waveform is normalised. The circles represent the peak positions used for velocity calculation. Among them, the red circle and the green circle represent the peak values used for calculating longitudinal wave velocity and shear wave velocity, respectively. According to Figure 25, both experimental and numerical results can distinguish the transverse wave from the longitudinal wave. Moreover, the waveforms are similar except for a certain difference in magnitude. Table 4 presents the simulation errors of the established 2D and 3D ice models at various temperatures. The relative errors are within the range of 0.15–5.7% according to the comparison of numerical and experimental results. The errors in the 3D model are less than those in the 2D model, which indicates that the 3D numerical ice model is more appropriate for wave propagation analysis of ice.
Table 4 shows that the small deviation range demonstrates the reliability of both the experimental tests and the numerical simulations. Moreover, the mechanical properties of ice obtained from the experimental results are reasonable and applicable. In addition, the 2D and 3D numerical ice models can prove that this method of simulating ice with a random pore model is feasible. The ice model provides a reliable and feasible way for further investigations of ice-related engineering problems.

5. Conclusions

In this study, a sophisticated embedded ultrasonic system is proposed to inspect the mechanical properties of ice in real time and online. This embedded ultrasonic system provides a platform to continuously obtain the response of the ice. With this system, the ice properties under specific temperature conditions are identified based on the intrinsic relationships between the wave propagation velocities and the mechanical properties of ice. Furthermore, the feasibility and effectiveness of the proposed embedded ultrasonic system-based method are validated via ultrasonic experiments and numerical simulations. Both numerical and experimental results demonstrate the effectiveness of the proposed embedded ultrasonic system-based method to inspect the mechanical properties of ice. The conclusions can be summarised as follows:
(1)
With the artificial ice specimen made of distilled water, the experiment results indicate the velocities of wave propagation in the ice decrease gradually with the elevated temperatures. Moreover, the velocity of the longitudinal wave is within the range of 3954.8–3787.88 m/s, while the velocity of the shear wave is within the range of 1831.54–1797.64 m/s;
(2)
The experimentally obtained wave velocities have been brought into known formulas to calculate the material properties of ice at different temperatures, including Young’s module, Poisson’s ratio, shear modulus, and bulk modulus. The results have shown that the values of the four properties have a downward trend with the increase in temperature;
(3)
With the help of the numerical models, the wave propagation in ice has been investigated, and the influence of varying temperatures on the wave propagation has been discussed in detail. The relative errors are within the range of 0.15–5.7% according to the comparison of numerical and experimental results, which proves the validity of the experimental results and reasonability of the calculated ice mechanics properties;
(4)
By comparing the waveform diagrams of the non-porous model and different porosity models, it can be found that the porosity affects the wave propagation velocity. By which, the greater the porosity, the slower the wave propagation velocity;
(5)
This research proves that the random pore model can simulate ice well. Additionally, by modifying the properties of the input model, more different ice models at different temperatures can be obtained, which provides a reliable modelling method for later research on ice.

Author Contributions

Methodology, software, formal analysis, writing—original draft, H.H.; data curation, validation, writing—original draft, L.W.; data curation, validation, writing—original draft, N.F.A.; conceptualisation, supervision, writing—review and editing, funding acquisition, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the Fundamental Research Funds for the Central Universities: No.B220204002; the 2022 National Young Foreign Talents Program of China: No. QN2022143002L; Jiangsu International Joint Research and Development Program: No. BZ2022010; and the Nanjing International Joint Research and Development Program: No. 202112003.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets used and/or analysed during the current study are available from the corresponding author on reasonable request.

Conflicts of Interest

All authors certify that they have no affiliations with or involvement in any organisation or entity with any financial or non-financial interest in the subject matter or materials discussed in this manuscript.

References

  1. Screen, J.A.; Simmonds, I. The central role of diminishing sea ice in recent Arctic temperature amplification. Nature 2010, 464, 1334–1337. [Google Scholar] [CrossRef] [Green Version]
  2. Kumar, A.; Perlwitz, J.; Eischeid, J.; Quan, X.; Xu, T.; Zhang, T.; Hoerling, M.; Jha, B.; Wang, W. Contribution of sea ice loss to Arctic amplification. Geophys. Res. Lett. 2010, 37, 1–17. [Google Scholar] [CrossRef] [Green Version]
  3. Gebre, S.; Boissy, T.; Alfredsen, K. Sensitivity to climate change of the thermal structure and ice cover regime of three hydropower reservoirs. J. Hydrol. 2014, 510, 208–227. [Google Scholar] [CrossRef] [Green Version]
  4. Zhou, L.; Riska, K.; Ji, C. Simulating transverse icebreaking process considering both crushing and bending failures. Mar. Struct. 2017, 54, 167–187. [Google Scholar] [CrossRef]
  5. Balkhanov, V.K.; Bashkuev, Y.B.; Khaptanov, V.B. Deformation of freshwater ice cover measured by a horizontal electric antenna. Tech. Phys. 2007, 52, 120–122. [Google Scholar] [CrossRef]
  6. Liu, Z.; Amdahl, J.; Løset, S. Plasticity based material modelling of ice and its application to ship–iceberg impacts. Cold Reg. Sci. Technol. 2011, 65, 326–334. [Google Scholar] [CrossRef]
  7. Iliescu, D.; Baker, I. The structure and mechanical properties of river and lake ice. Cold Reg. Sci. Technol. 2007, 48, 202–217. [Google Scholar] [CrossRef]
  8. Wang, Y.; Xu, Y.; Huang, Q. Progress on ultrasonic guided waves de-icing techniques in improving aviation energy efficiency. Renew. Sustain. Energy Rev. 2017, 79, 638–645. [Google Scholar] [CrossRef]
  9. Sun, X.; Li, S.; Dun, X.; Li, D.; Li, T.; Guo, R.; Yang, M. A novel characterization method of piezoelectric composite material based on particle swarm optimization algorithm. Appl. Math. Model. 2019, 66, 322–331. [Google Scholar] [CrossRef]
  10. Wang, Q.; Li, Z.; Lei, R.; Lu, P.; Han, H. Estimation of the uniaxial compressive strength of Arctic sea ice during melt season. Cold Reg. Sci. Technol. 2018, 151, 9–18. [Google Scholar] [CrossRef]
  11. Han, H.; Li, Z.; Huang, W.; Lu, P.; Lei, R. The uniaxial compressive strength of the Arctic summer sea ice. Acta Oceanol. Sin. 2015, 34, 129–136. [Google Scholar] [CrossRef]
  12. Karulina, M.; Marchenko, A.; Karulin, E.; Sodhi, D.; Sakharov, A.; Chistyakov, P. Full-scale flexural strength of sea ice and freshwater ice in Spitsbergen Fjords and North-West Barents Sea. Appl. Ocean Res. 2019, 90, 101853. [Google Scholar] [CrossRef]
  13. Aksenov, V.I.; Gevorkyan, S.G.; Iospa, A.V. Temperature Dependence of Stress-Strain Properties of Freshwater Ice. Soil Mech. Found. Eng. 2019, 56, 366–370. [Google Scholar] [CrossRef]
  14. Moslet, P. Field testing of uniaxial compression strength of columnar sea ice. Cold Reg. Sci. Technol. 2007, 48, 1–14. [Google Scholar] [CrossRef]
  15. Qiu, W.-L.; Peng, R.-X. Research on the numerical simulation for plastic model of ice as building materials under triaxial compression. Constr. Build. Mater. 2020, 268, 121183. [Google Scholar] [CrossRef]
  16. Tinard, V.; François, P.; Fond, C. The Potential Scope of the Ultrasonic Surface Reflection Method Towards Mechanical Characterisation of Isotropic Materials. Part 2. Experimental Results. Exp. Mech. 2021, 61, 1161–1170. [Google Scholar] [CrossRef]
  17. Tinard, V.; François, P.; Fond, C. The Potential Scope of the Ultrasonic Surface Reflection Method Towards Mechanical Characterisation of Isotropic Materials. Part 1. A Theoretical Analysis. Exp. Mech. 2021, 61, 1153–1160. [Google Scholar] [CrossRef]
  18. Zhu, Z.H.; Post, M.A.; Meguid, S.A. The Potential of Ultrasonic Non-Destructive Measurement of Residual Stresses by Modal Frequency Spacing using Leaky Lamb Waves. Exp. Mech. 2012, 52, 1329–1339. [Google Scholar] [CrossRef]
  19. Mustansar, Z.; Shaukat, A.; Margetts, L. Effect of using different types of methods for the derivation of elastic modulus of bone—A critical survey. MATEC Web Conf. 2017, 108, 13002. [Google Scholar] [CrossRef]
  20. Lin, S.; Shams, S.; Choi, H.; Azari, H. Ultrasonic imaging of multi-layer concrete structures. NDT E Int. 2018, 98, 101–109. [Google Scholar] [CrossRef]
  21. Yang, X.; Verboven, E.; Ju, B.-F.; Kersemans, M. Comparative study of ultrasonic techniques for reconstructing the multilayer structure of composites. NDT E Int. 2021, 121, 102460. [Google Scholar] [CrossRef]
  22. Cao-Rial, M.; Moreno, C.; Quintela, P. Determination of Young modulus by using Rayleigh waves. Appl. Math. Model. 2019, 77, 439–455. [Google Scholar] [CrossRef]
  23. Maguire, R.R.; Schmerr, N.C.; Lekić, V.; Hurford, T.A.; Dai, L.; Rhoden, A.R. Constraining Europa’s ice shell thickness with fundamental mode surface wave dispersion. Icarus 2021, 369, 114617. [Google Scholar] [CrossRef]
  24. Bayón, A.; Gascón, F.; Nieves, F.J. Estimation of dynamic elastic constants from the amplitude and velocity of Rayleigh waves. J. Acoust. Soc. Am. 2005, 117, 3469–3477. [Google Scholar] [CrossRef]
  25. Medina, R.; Bayón, A. Elastic constants of a plate from impact-echo resonance and Rayleigh wave velocity. J. Sound Vib. 2010, 329, 2114–2126. [Google Scholar] [CrossRef]
  26. Polach, R.U.F.V.B.U.; Ettema, R.; Gralher, S.; Kellner, L.; Stender, M. The non-linear behavior of aqueous model ice in downward flexure. Cold Reg. Sci. Technol. 2019, 165, 102775. [Google Scholar] [CrossRef] [Green Version]
  27. Gagnon, R. Results of numerical simulations of growler impact tests. Cold Reg. Sci. Technol. 2007, 49, 206–214. [Google Scholar] [CrossRef]
  28. Wang, Y.; Ni, Y. Bayesian dynamic forecasting of structural strain response using structural health monitoring data. Struct. Control. Health Monit. 2020, 27, e2575. [Google Scholar] [CrossRef]
  29. Wang, Q.-A.; Zhang, C.; Ma, Z.-G.; Ni, Y.-Q. Modelling and forecasting of SHM strain measurement for a large-scale suspension bridge during typhoon events using variational heteroscedastic Gaussian process. Eng. Struct. 2022, 251, 113554. [Google Scholar] [CrossRef]
  30. Invernizzi, S.; Lacidogna, G.; Carpinteri, A. Particle-based numerical modeling of AE statistics in disordered materials. Meccanica 2012, 48, 211–220. [Google Scholar] [CrossRef]
  31. Markovic, N.; Nestorovic, T.; Stojic, D. Numerical modeling of damage detection in concrete beams using piezoelectric patches. Mech. Res. Commun. 2015, 64, 15–22. [Google Scholar] [CrossRef]
  32. Cao, M.; Ren, Q.; Wang, H. A method of detecting seismic singularities using combined wavelet with fractal. Chin. J. Geophys. 2005, 48, 672–679. [Google Scholar] [CrossRef]
  33. Alkayem, N.F.; Cao, M. Damage identification in three-dimensional structures using single-objective evolutionary algorithms and finite element model updating: Evaluation and comparison. Eng. Optim. 2018, 50, 1695–1714. [Google Scholar] [CrossRef]
  34. Bonath, V.; Edeskär, T.; Lintzén, N.; Fransson, L.; Cwirzen, A. Properties of ice from first-year ridges in the Barents Sea and Fram Strait. Cold Reg. Sci. Technol. 2019, 168, 102890. [Google Scholar] [CrossRef]
  35. Gharamti, I.; Dempsey, J.; Polojärvi, A.; Tuhkuri, J. Fracture of warm S2 columnar freshwater ice: Size and rate effects. Acta Mater. 2020, 202, 22–34. [Google Scholar] [CrossRef]
  36. Ji, S.; Yue, Q.; Bi, X. Probability distribution of sea ice fatigue parameters in JZ20-2 sea area of the Liaodong Bay. Ocean Eng. 2002, 20, 6. [Google Scholar]
  37. Cai, W.; Zhu, L.; Yu, T.; Li, Y. Numerical simulations for plates under ice impact based on a concrete constitutive ice model. Int. J. Impact Eng. 2020, 143, 103594. [Google Scholar] [CrossRef]
  38. Zhang, J.-P.; Zhou, D. Numerical modeling for strain rate effect and size effect of ice under uniaxial tension and compression. Commun. Nonlinear Sci. Numer. Simul. 2021, 96, 105614. [Google Scholar] [CrossRef]
  39. Wang, Z.; Wei, L.; Cao, M. Damage Quantification with Embedded Piezoelectric Aggregates Based on Wavelet Packet Energy Analysis. Sensors 2019, 19, 425. [Google Scholar] [CrossRef] [Green Version]
  40. Li, W.; Wang, Z.; Cao, M.; Fu, R. Concrete strength monitoring based on the variation of ultrasonic waveform acquired by piezoelectric aggregates. Struct. Eng. Mech. 2020, 76, 591–598. [Google Scholar]
  41. Zhang, W.; Hao, H.; Wu, J.; Li, J.; Ma, H.; Li, C. Detection of minor damage in structures with guided wave signals and nonlinear oscillator. Measurement 2018, 122, 532–544. [Google Scholar] [CrossRef]
  42. Kupperman, D.S.; Reimann, K.J.; Abrego-Lopez, J. Ultrasonic NDE of cast stainless steel. NDT Int. 1989, 22, 309. [Google Scholar] [CrossRef]
  43. Gammon, P.H.; Klefte, H.; Clouter, M.J. Elastic Constants of Ice Samples by Brillouin Elastic Constants of Ice Samples by Brillouin Spectroscopy. J. Phys. Chem. 1983, 87, 4025–4029. [Google Scholar] [CrossRef]
  44. Randhawa, K.S. The Measurement of the Young’s Modulus of Ice with Ultrasonic Waves. Ph.D. Thesis, Technische Universität Hamburg, Hamburg, Germany, 2018. [Google Scholar]
  45. Pustogvar, A.; Kulyakhtin, A. Sea ice density measurements. Methods and uncertainties. Cold Reg. Sci. Technol. 2016, 131, 46–52. [Google Scholar] [CrossRef]
  46. Timco, R.F.G.W. A review of sea ice density. Cold Reg. Sci. Technol. 1996, 24, 1–6. [Google Scholar] [CrossRef]
  47. Timco, G.W.; Frederking, R.M.W. Compressive strength of sea ice sheets. Cold Reg. Sci. Technol. 1990, 17, 227–240. [Google Scholar] [CrossRef]
  48. Song, Z.; Hou, L.; Whisler, D.; Gao, G. Mesoscopic numerical investigation of dynamic mechanical properties of ice with entrapped air bubbles based on a stochastic sparse distribution mechanism. Compos. Struct. 2020, 236, 111834. [Google Scholar] [CrossRef]
  49. Chen, H.; Xu, B.; Zhou, T.; Mo, Y.-L. Debonding detection for rectangular CFST using surface wave measurement: Test and multi-physical fields numerical simulation. Mech. Syst. Signal Process. 2018, 117, 238–254. [Google Scholar] [CrossRef]
  50. Cox, G.F.N.; Weeks, W.F. Equations for Determining the Gas and Brine Volumes in Sea-Ice Samples. J. Glaciol. 1983, 29, 306–316. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Schematic diagram of ice sample structure; (b) ice specimen with sensors (scale 1:3); (c) the PZT; and (d) the temperature sensor.
Figure 1. (a) Schematic diagram of ice sample structure; (b) ice specimen with sensors (scale 1:3); (c) the PZT; and (d) the temperature sensor.
Sensors 23 06045 g001
Figure 2. The experimental setup.
Figure 2. The experimental setup.
Sensors 23 06045 g002
Figure 3. Measured ultrasonic signals with the same centre frequency (250 kHz) and various number of cycles: (a) n = 2, (b) n = 3, (c) n = 4, and (d) n = 5.
Figure 3. Measured ultrasonic signals with the same centre frequency (250 kHz) and various number of cycles: (a) n = 2, (b) n = 3, (c) n = 4, and (d) n = 5.
Sensors 23 06045 g003
Figure 4. Measured ultrasonic signals with the same cycles (2T) and various central frequencies: (a) fc = 150 kHz, (b) fc = 200 kHz, (c) fc = 250 kHz, (d) fc = 300 kHz, (e) fc = 350 kHz, (f) fc = 400 kHz, (g) fc = 450 kHz, and (h) fc = 500 kHz.
Figure 4. Measured ultrasonic signals with the same cycles (2T) and various central frequencies: (a) fc = 150 kHz, (b) fc = 200 kHz, (c) fc = 250 kHz, (d) fc = 300 kHz, (e) fc = 350 kHz, (f) fc = 400 kHz, (g) fc = 450 kHz, and (h) fc = 500 kHz.
Sensors 23 06045 g004
Figure 5. Waveform of the normalised 2-cycle 250 kHz tone burst signal.
Figure 5. Waveform of the normalised 2-cycle 250 kHz tone burst signal.
Sensors 23 06045 g005
Figure 6. Time domain and spectrogram of the response signal: (a) original response signal, (b) signal after denoising, (c) spectrogram of the original signal, and (d) spectrogram of the signal after denoising.
Figure 6. Time domain and spectrogram of the response signal: (a) original response signal, (b) signal after denoising, (c) spectrogram of the original signal, and (d) spectrogram of the signal after denoising.
Sensors 23 06045 g006
Figure 7. Four sets of measured waveforms.
Figure 7. Four sets of measured waveforms.
Sensors 23 06045 g007
Figure 8. The relationship between wave velocity and temperature: (a) the relationship between longitudinal wave velocity and temperature and (b) the relationship between shear wave velocity and temperature.
Figure 8. The relationship between wave velocity and temperature: (a) the relationship between longitudinal wave velocity and temperature and (b) the relationship between shear wave velocity and temperature.
Sensors 23 06045 g008
Figure 9. The changes of various properties with ice samples’ temperatures: (a) the Poisson’s factor of ice versus temperature of ice; (b) the Young’s module versus temperature of ice; (c) the module of volume versus temperature of ice; and (d) the module of shear versus temperature of ice.
Figure 9. The changes of various properties with ice samples’ temperatures: (a) the Poisson’s factor of ice versus temperature of ice; (b) the Young’s module versus temperature of ice; (c) the module of volume versus temperature of ice; and (d) the module of shear versus temperature of ice.
Sensors 23 06045 g009
Figure 10. Numerical model of the ice specimen with different porosities: (a) 0.5%; (b) 1%; and (c) 3%.
Figure 10. Numerical model of the ice specimen with different porosities: (a) 0.5%; (b) 1%; and (c) 3%.
Sensors 23 06045 g010
Figure 11. FEM meshes of the ice specimen with different porosities: (a) 0.5%; (b) 1%; and (c) 3%.
Figure 11. FEM meshes of the ice specimen with different porosities: (a) 0.5%; (b) 1%; and (c) 3%.
Sensors 23 06045 g011
Figure 12. Wave propagation in a 2D ice model: (af) are the time slices of the wave propagation process: (a) time = 1.0 × 10−5 s, (b) time = 2.0 × 10−5 s, (c) time = 3.0 × 10−5 s, (d) time = 4.0 × 10−5 s, (e) time = 6 × 10−5 s, and (f) time = 7 × 10−5 s.
Figure 12. Wave propagation in a 2D ice model: (af) are the time slices of the wave propagation process: (a) time = 1.0 × 10−5 s, (b) time = 2.0 × 10−5 s, (c) time = 3.0 × 10−5 s, (d) time = 4.0 × 10−5 s, (e) time = 6 × 10−5 s, and (f) time = 7 × 10−5 s.
Sensors 23 06045 g012
Figure 13. Four sets of numerical simulation waveforms.
Figure 13. Four sets of numerical simulation waveforms.
Sensors 23 06045 g013
Figure 14. The relationship between wave velocities and temperatures: (a) the relationship between longitudinal wave velocity and temperature and (b) the relationship between shear wave velocity and temperature.
Figure 14. The relationship between wave velocities and temperatures: (a) the relationship between longitudinal wave velocity and temperature and (b) the relationship between shear wave velocity and temperature.
Sensors 23 06045 g014
Figure 15. Wave propagation in a non-porous ice model and a 3% porosity ice model. (af) are the time slices of the wave propagation process: non-porous ice model (a) time = 2.0 × 10−5 s, (b) time = 3.0 × 10−5 s, (c) time = 4.0 × 10−5 s, (d) time = 4.5 × 10−5 s, (e) time = 6 × 10−5 s, and (f) time = 8 × 10−5 s.
Figure 15. Wave propagation in a non-porous ice model and a 3% porosity ice model. (af) are the time slices of the wave propagation process: non-porous ice model (a) time = 2.0 × 10−5 s, (b) time = 3.0 × 10−5 s, (c) time = 4.0 × 10−5 s, (d) time = 4.5 × 10−5 s, (e) time = 6 × 10−5 s, and (f) time = 8 × 10−5 s.
Sensors 23 06045 g015
Figure 16. Wave propagation in a 3% porosity ice model, (af) are the time slices of the wave propagation process: (a) time = 2.2 × 10−5 s, (b) time = 3.2 × 10−5 s, (c) time = 4.0 × 10−5 s, (d) time = 5.0 × 10−5 s, (e) time = 5.8 × 10−5 s, and (f) time = 8 × 10−5 s.
Figure 16. Wave propagation in a 3% porosity ice model, (af) are the time slices of the wave propagation process: (a) time = 2.2 × 10−5 s, (b) time = 3.2 × 10−5 s, (c) time = 4.0 × 10−5 s, (d) time = 5.0 × 10−5 s, (e) time = 5.8 × 10−5 s, and (f) time = 8 × 10−5 s.
Sensors 23 06045 g016
Figure 17. Comparison of the waveforms of the two models.
Figure 17. Comparison of the waveforms of the two models.
Sensors 23 06045 g017
Figure 18. Comparison of three porosity waveforms.
Figure 18. Comparison of three porosity waveforms.
Sensors 23 06045 g018
Figure 19. Three-dimensional numerical model: (ac) the spherical pore of the three different porosities: (a) 0.5% porosity, (b) 1% porosity, and (c) 3% porosity; (df) three-dimensional ice random pore model of the three different porosities: (d) 0.5% porosity, (e) 1% porosity, and (f) 3% porosity.
Figure 19. Three-dimensional numerical model: (ac) the spherical pore of the three different porosities: (a) 0.5% porosity, (b) 1% porosity, and (c) 3% porosity; (df) three-dimensional ice random pore model of the three different porosities: (d) 0.5% porosity, (e) 1% porosity, and (f) 3% porosity.
Sensors 23 06045 g019
Figure 20. FEM meshes of the 3D ice model: (a) global view; (b) internal view; and (c) zoomed view of the red circle in (b).
Figure 20. FEM meshes of the 3D ice model: (a) global view; (b) internal view; and (c) zoomed view of the red circle in (b).
Sensors 23 06045 g020
Figure 21. Wave propagation in the 3D ice model: (af) are the time slices of the wave propagation process: (a) time = 1.0 × 10−5 s, (b) time = 1.7 × 10−5 s, (c) time = 2.0 × 10−5 s, (d) time = 3.1 × 10−5 s, (e) time = 5.2 × 10−5 s, and (f) time = 6.3 × 10−5 s.
Figure 21. Wave propagation in the 3D ice model: (af) are the time slices of the wave propagation process: (a) time = 1.0 × 10−5 s, (b) time = 1.7 × 10−5 s, (c) time = 2.0 × 10−5 s, (d) time = 3.1 × 10−5 s, (e) time = 5.2 × 10−5 s, and (f) time = 6.3 × 10−5 s.
Sensors 23 06045 g021
Figure 22. Three-dimensional numerical simulation of the relationship between wave velocities and temperatures: (a) the relationship between longitudinal wave velocity and temperature and (b) the relationship between shear wave velocity and temperature.
Figure 22. Three-dimensional numerical simulation of the relationship between wave velocities and temperatures: (a) the relationship between longitudinal wave velocity and temperature and (b) the relationship between shear wave velocity and temperature.
Sensors 23 06045 g022
Figure 23. Combination chart of four sets of 3D numerical simulation.
Figure 23. Combination chart of four sets of 3D numerical simulation.
Sensors 23 06045 g023
Figure 24. Comparison of three porosity waveforms of 3D models.
Figure 24. Comparison of three porosity waveforms of 3D models.
Sensors 23 06045 g024
Figure 25. Comparison of experimental data and simulation.
Figure 25. Comparison of experimental data and simulation.
Sensors 23 06045 g025
Table 1. Dimensions and material properties of PZT.
Table 1. Dimensions and material properties of PZT.
DiameterThicknessd33KT3
20 mm1 mm460 × 10−12 m/V1700
Table 2. Values of the longitudinal sound velocity in ice by different authors.
Table 2. Values of the longitudinal sound velocity in ice by different authors.
StudiesKupperman et al. [42]Gammon et al. [43]Randhawa [44]Present
Sound velocity (m/s)39503892–404039803790–3960
Table 3. Material mechanical properties of ice.
Table 3. Material mechanical properties of ice.
Temperature
T (°C)
Density
ρ (kg·m3)
Poisson’s Factor
μ
Young’s Module
E (GPa)
Shear Module G (GPa)Volume Module
K (GPa)
−24.58900.3608.03752.9549.581
−17.28900.3587.9822.9399.363
−8.88900.3567.8892.9089.143
−0.88900.3557.7922.8768.935
Table 4. Wave velocity of tests and simulations and the errors between them.
Table 4. Wave velocity of tests and simulations and the errors between them.
Temperature
T (°C)
Longitudinal Wave Velocity of the Test
Vl (m/s)
Two-Dimensional Simulated Longitudinal Wave Velocity
Vln2 (m/s)
Errors of 2D Simulation Results (%)Three-Dimensional Simulated Longitudinal Wave Velocity
Vln3 (m/s)
Deviation of 3D Simulation Results (%)Two-Dimensional Shear Wave Velocity of the Test
Vs (m/s)
Two-Dimensional Simulated Shear Wave Velocity
Vsn3 (m/s)
Deviation of 2D Simulation Results (%)Three-Dimensional Simulated Shear Wave Velocity
Vsn3 (m/s)
Errors
of 2D
Simulation
Results
(%)
−34.43954.803791.984.123736.215.531825.771928.375.621902.834.22
−32.43937.013787.883.793733.695.161830.541927.315.291901.533.88
−31.83937.013787.883.793730.155.251830.541927.315.291896.713.61
−30.53923.773783.783.573728.054.981829.591925.725.251893.213.48
−29.73919.373798.163.093725.754.941827.681924.135.281889.563.39
−28.63914.993794.043.093721.984.931827.681923.085.221885.243.15
−27.73910.613789.933.093718.664.911826.721922.555.251881.362.99
−26.93906.253783.783.143715.924.871825.771920.975.211878.642.90
−26.13906.253775.623.343713.454.941823.871919.915.271875.642.84
−25.43897.553769.523.283712.084.761823.871919.395.241872.222.65
−24.53897.553761.403.503711.564.771821.971918.905.301869.662.62
−23.73893.213757.383.493706.314.801821.971917.815.261867.542.50
−22.93893.213755.363.543703.334.881821.021916.765.261864.262.37
−223884.573751.343.433695.854.861820.071915.715.251862.742.34
−21.33884.573751.343.433691.334.971820.071915.715.251859.692.18
−20.53875.973749.333.273685.464.921819.131914.665.251856.242.04
−19.83871.683747.323.213679.514.961819.131914.145.221853.961.91
−18.73871.683747.323.213671.695.171818.181914.145.281849.541.72
−183863.133745.323.053667.525.061818.181913.095.221846.331.55
−17.23863.133743.324.003661.095.231817.241912.054.941843.081.42
−16.33854.633739.322.993657.745.111816.291910.485.191839.851.30
−15.63858.883735.333.203652.865.341815.351909.445.181836.341.16
−14.83854.633733.333.153648.695.341813.471907.885.211833.851.12
−13.93846.153729.363.043645.215.221810.661906.845.311831.021.12
−13.13846.153725.393.143640.985.331811.591905.805.201826.980.85
−12.43841.933723.403.093636.545.351810.661904.245.171823.870.73
−11.63837.723719.453.083632.235.351809.721903.735.191820.330.59
−10.63829.323713.533.023628.965.231809.721902.695.141817.520.43
−9.93825.143709.593.023625.435.221808.791902.175.161815.650.38
−8.83825.143703.703.173623.195.281807.851901.665.1018130.28
−83820.963701.753.123618.655.291806.921900.115.161809.890.16
−7.33816.793697.833.123612.125.361805.991898.565.131807.350.08
−6.63808.493693.933.013608.945.241805.051897.025.101804.220.05
−5.63808.493693.933.013602.545.411804.121895.995.091801.360.15
−4.73804.353690.043.003596.435.471803.191894.455.061797.950.29
−3.83800.223688.092.953591.235.501802.271893.435.061793.430.49
−2.83796.103684.212.953586.545.521801.341891.895.031789.530.66
−1.73787.883684.212.743581.935.441799.491890.365.051784.630.83
−0.83787.883682.272.703576.95.571797.641887.815.021782.080.87
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Han, H.; Wei, L.; Alkayem, N.F.; Cao, M. Embedded Ultrasonic Inspection on the Mechanical Properties of Cold Region Ice under Varying Temperatures. Sensors 2023, 23, 6045. https://doi.org/10.3390/s23136045

AMA Style

Han H, Wei L, Alkayem NF, Cao M. Embedded Ultrasonic Inspection on the Mechanical Properties of Cold Region Ice under Varying Temperatures. Sensors. 2023; 23(13):6045. https://doi.org/10.3390/s23136045

Chicago/Turabian Style

Han, Huimin, Li Wei, Nizar Faisal Alkayem, and Maosen Cao. 2023. "Embedded Ultrasonic Inspection on the Mechanical Properties of Cold Region Ice under Varying Temperatures" Sensors 23, no. 13: 6045. https://doi.org/10.3390/s23136045

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop