Next Article in Journal
Study on Flow Characteristics of Francis Turbine Based on Large-Eddy Simulation
Next Article in Special Issue
Multi-Scale Polar Object Detection Based on Computer Vision
Previous Article in Journal
Study of the Process of Electrochemical Oxidation of Active Pharmaceutical Substances on the Example of Nitrofurazone ((2E)-2-[(5-Nitro-2-furyl)methylene]hydrazine Carboxamide)
Previous Article in Special Issue
Study on the Constitutive Equation and Mechanical Properties of Natural Snow under Step Loading
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigations on Flexural Strength of a Columnar Saline Model Ice under Circular Plate Central Loading

1
State Key Laboratory of Structural Analysis, Optimization and CAE Software for Industrial Equipment, Dalian University of Technology, Dalian 116024, China
2
China Ship Scientific Research Center, National Key Laboratory of Science and Technology on Hydrodynamics, Wuxi 214082, China
3
State Key Laboratory of Coastal and Offshore Engineering, Dalian University of Technology, Dalian 116024, China
4
School of Ocean Science and Technology, Dalian University of Technology, Panjin 124221, China
*
Author to whom correspondence should be addressed.
Water 2023, 15(19), 3371; https://doi.org/10.3390/w15193371
Submission received: 30 August 2023 / Revised: 19 September 2023 / Accepted: 21 September 2023 / Published: 26 September 2023
(This article belongs to the Special Issue Cold Regions Ice/Snow Actions in Hydrology, Ecology and Engineering)

Abstract

:
The properties of ice strength have a significant impact on the design and safety of structures in ice-infested waters. To analyze the flexural strength of columnar saline model ice, we conducted circular plate center loading tests at the Small Ice Model Basin of the China Ship Scientific Research Center (CSSRC SIMB) in China. The tests involved varying the loading rate and ice temperature, and a numerical model was developed using FEM and LS-DYNA for validation and comparison. The results of the tests revealed the crack propagation process, stress distribution, load response, and failure mode of the model ice. The model ice displayed typical brittle failure, and the flexural strength was linearly related to ice temperature but not significantly correlated with loading rate. The porosity of the model ice affected the load response and time of failure but not the failure mode. The model ice with 7% porosity had a 7.8% reduction in load response compared to the nonporous model ice. This study provides a reliable method for measuring and analyzing the flexural strength of model ice. It also serves as a foundation for further research on the interaction between structures and ice sheets.

1. Introduction

In regions where ice is present, such as beneath a floating object, the ice sheet may fracture under vertical forces. This circumstance has gained significance in the fields of polar exploration and engineering [1]. The durability of the ice plays a crucial role in the design and assessment of structures regarding safety. The upward pressure exerted by the floating entity can trigger a range of failures, including shearing, bending, and flexural failure, which is the most frequently observed [2,3]. Conducting tests to determine flexural strength is essential for accurately predicting ice loads on structures and ensuring their safety. This information is critical in developing and assessing structures for floating objects under ice.
There exist three primary methods to assess ice strength: in situ cantilever tests, simply supported beam tests, and circular plate center loading tests [4,5,6]. An in situ cantilever test involves obtaining an ice beam directly from the ice layer and conducting the test while the ice remains in its natural environment [7]. Frederking and Timco executed in situ cantilever tests on model ice at the National Research Council Hydraulics Laboratory, discovering no significant relationship between loading rate and flexural strength [8]. Krupina also conducted in situ cantilever tests on sea ice in the Barents Sea, obtaining the distribution of sea ice flexural strength in the region [9]. Gang of CSSRC conducted an in situ cantilever beam test on model ice, revealing that the flexural strength of model ice remains constant as the loading rate changes but decreases with increased rewarming time [10]. The simply supported beam test can be split into three-point bending and four-point bending tests [11]. Ji et al. executed an indoor three-point bending test on the flexural strength of Bohai sea ice and found that the sea ice flexural strength had a linear relationship with the loading rate [4]. Gagnon conducted a four-point bending test on iceberg ice and found that the flexural strength of ice decreases with increased ice temperature [12]. Barrette et al. performed a four-point bending test on iceberg ice and found that the larger the average grain size inside the ice, the smaller its flexural strength [13]. Lastly, in circular plate center loading tests, icicles are cut from the ice layer and then shaped into disks according to predetermined thickness–diameter ratios. Krupina conducted an extensive series of circular plate center loading tests on sea ice situated in both southeastern and northeastern regions of the Barents Sea. The study aimed to investigate the effects of different ice properties, such as temperature, salinity, and brine volume, on flexural strength. Furthermore, the plate test outcomes were juxtaposed with those of the cantilever test, and a suitable correlation was established between the two [9]. Marchenko conducted circular plate center loading tests on sea ice in the northwest Barents Sea and found that flexural strength depends on ice temperature [6].
In conclusion, ice is a complex crystalline material, and its flexural strength is influenced by factors such as the structure, temperature, and external loading conditions of ice crystals [14].
In addition to experimental studies, numerical simulations can also be utilized to investigate the flexural strength of ice. The discrete element method, the finite element method, and the peridynamics method are the most popular simulation methods used today. The discrete element method is suitable for the mechanics of discontinuous media [15], but its application for dynamic failure problems is limited due to result deviation in the continuum stage [16]. On the other hand, the finite element method has advantages in terms of fast calculation efficiency and high accuracy and is widely used to simulate ice–structure interaction [17]. However, the simulation accuracy of the finite element method depends on the material model chosen [18]. The peridynamics method analyzes the mechanics of a continuum with the help of the point method of matter and the idea of molecular dynamics. It constructs the object’s motion equation in the form of an integral [19]. However, compared with other numerical calculation methods, the peridynamics method also has the problem of low computational efficiency. When simulating the interaction between sea ice and structure, most scholars consider ice as a continuum, considering its elasticity and plasticity. For example, Li used an isotropic elastoplastic fracture model to simulate ice based on LS-DYNA [20]. Some scholars incorporated the effects of temperature and strain rate and constructed a numerical model of ice based on the multi-surface failure criteria of sea ice. Still, there were some differences between the test and simulation results [21]. Other scholars considered the existence of pores inside the ice, such as Von Bock und Polach et al., who applied LS−DYNA and selected the Lemaitre damage model to simulate ice and analyzed the dependence of the load response on the microstructure of ice [22]. To maintain the quality and energy of the model, some scholars used the node-splitting technique instead of the commonly used elemental erosion technique [23].
Despite the wealth of experimental and numerical data on in situ cantilever and simply supported beam tests, there has been limited research conducted on circular plate center loading tests. The International Towing Tank Conference’s mechanical properties test method for model ice also lacks relevant specifications for circular plate center loading tests, indicating the need for improvement in this area. As mentioned earlier, the flexural strength of ice is influenced by various factors, such as temperature and loading rate, and further experimental exploration is necessary. Additionally, there is a need for expanded research based on the material model, particularly regarding the influence of internal porosity on ice strength.
In this paper, the influence of loading rate and temperature on the flexural strength was explored by using a saline model ice under a circular plate center loading test, and the simulation of the porosity of ice was also realized to study the influence of the internal porosity of the model ice. The results of center loading on a circular plate of ice were discussed, and, in particular, the physical mechanisms of crack propagation were analyzed. This study provides a reference for the prediction of the ice’s mechanical characteristics.

2. Methods

Both laboratory experiments and numerical simulations were conducted in the present study, and the implementation of the two methods is briefly summarized below.

2.1. Laboratory Experiments

2.1.1. The Columnar Saline Model Ice

The SIMB of the CSSRC is an ice water tank with 8 m in length, 2 m in width, and 1 m in depth, as shown in Figure 1 [24]. The ice-making process used in the SIMB is similar to but partly improved from the method by the Evers [25,26] of the Hamburg Tank (HSVA) in Germany. Cooling fans are suspended from the roof, ceiling panels with tiny holes are installed underneath to exhaust cooling air, and circulating fans with guide plates are set on the side walls to suck the air used for cooling; thus, forced cooling air circulation is formed [27].
Model ice is made of sodium chloride solution, and the preparation process is divided into precooling, crystallization, and ice making. To control the density and to lower the strength, an underwater air bubbling system is used to produce tiny bubbles that are trapped in the model ice during the freezing process. Its mechanical properties can be adjusted by rewarming and underwater microbubble generation systems [10]. After rewarming, the temperature of model ice is usually around −0.8 °C.
Through a series of experimental measurements, the mechanical properties of model ice in the SIMB were statistically analyzed. The fresh model ice’s main characteristic parameters (−0.8 °C) in the SIMB are summarized in Table 1. With the continuous exploration of the experimenters, it was found that after adding bubbles to the saline model ice, the brittleness of the ice was greatly improved. The flexural fracture mode of the ice and the distribution ratio of the crushed ice flakes of various scales were consistent with natural sea ice, and the similarity could be maintained in terms of physical and mechanical properties and fracture behavior patterns [28].

2.1.2. Circular Plate Center Loading Tests

In a circular ice plate indentation test, vertical loads are exerted on the central area of the disk ice specimen under the circumferentially uniform support, as shown on the schematic diagram in Figure 2.
When the plate is thin enough, it belongs to a typical plate–shell structure according to the theory of shell mechanics [29]. The maximum bending moment of the circular plate is at the center of the circular plate with radius R when the load q is uniformly distributed in a circle of radius r. By integrating the deflection caused by the load on the torus plane with radius b and width db (Figure 2b), the deflection of the center of the plate can be obtained, and then the bending moment Mmax at the center of the plate can be determined [28]:
M max = q 0 r 1 v 4 R 2 b 2 R 2 1 + v 2 ln b R b d b = F 4 π ( 1 + v ) ln R r + 1 ( 1 v ) r 2 4 R 2
where F represents the total load πr2q. ν = 0.33, is the Poisson’s ratio [30]. The relationship between the bending moment and stress is:
σ = 6 M h 2
In the above equation, h is the thickness of the plate, so the maximum tensile stress in the plate, that is, the flexural strength σf, is:
σ f = 3 F 8 π h 2 4 ( 1 v ) r R 2 4 ( 1 + v ) ln r R
The relationship between the central bending moment and curvature obtained from the pure bending hypothesis is the basis of the above derivation, and the effect of shear forces on the flexural strength in a plane parallel to the plate surface is not considered. The value of σf so far is only an approximate solution, and its accuracy depends on the ratio of the thickness of the disk to the outer diameter [31]. When the thickness–diameter ratio is less than 0.15, the effect of shear force on the bending strength is negligible [6].

2.1.3. Test Conditions

The test was carried out in a small cold laboratory adjacent to the ice water basin. The flexural strength of ice is affected by a variety of factors [1], and the loading rate and ice temperature were selected as the control factors of the present test. Different test conditions are listed in Table 2. The peak force and failure time of the circular plate indentation test are two important measurements. The average temperature of the newly made model ice was about −0.8 °C, the density was 0.91 g·cm−3, and the salinity was 2.75 ppt.
The test process was divided into three phases: the test preparation stage, the plate specimen sampling and storage freezing stage, and the measurement and data recording stage. In the test preparation stage, the small cold laboratory was first cooled to reach the target air temperature, and the high-speed camera, electronic universal testing machine, thermometer, and other related instruments were precooled. In the plate sampling and storage freezing stage, a total of 80 plate specimens were drilled from the flat model ice sheet. The diameter of the ice plate specimen was 140 mm, and the thickness was 20 mm (Figure 3). It produced a thickness–diameter ratio of less than 0.15, agreeing with the calculation requirements. After sampling, 20 newly made model ice specimens were placed in a small cold laboratory for test measurement, and the rest were stored in a refrigerator for subsequent tests. In the last test and measurement stage, the loading rate of the electronic testing machine was first adjusted according to the requirements of the test conditions. Then the force curve of the ice specimen from the initial bearing to the flexural failure was recorded. The high-speed camera was also used to capture the failure details of the ice specimen. The thickness and temperature of model ice were quickly measured and recorded after failure.
In the data recording stage, the force curve of the ice specimen from the initial loading to the flexural failure was recorded with a force measurement system, and the flexural strength was obtained according to the peak force. At the same time, the failure details of the ice specimen were documented with a high-speed video camera. The overall layout scheme of the instruments is shown in Figure 4, and the parameters of the test instrument are shown in Table 3.

2.2. Numerical Modeling

2.2.1. Numerical Model of Ice Material

A big advantage of the finite element method is that many contact algorithms allow the coupling of ice and structural models. Therefore, in this paper, the finite element solver LS-DYNA was used to establish a model ice failure model. The circular plate center loading tests were used to vertically load the central area of the disk ice specimen under the circumferentially uniform support. The contact between the indenter and the ice disk specimen occurred during loading, and the explicit nonlinear finite element method can be used to solve such contact problems. The calculation method used by the nonlinear finite element program is the explicit integration method [32].
When using the finite element method to simulate the model ice mechanical test, it is necessary to determine the material model parameters that match the macroscopic characteristics of the material. Standard model ice constitutive models include the elastoplastic model, elastic brittleness model, etc. According to Karr and Choi [33], model ice materials are considered isotropic in their undeformed state. This assumption was adopted in this paper, so the isotropic elastoplastic fracture model (*MAT_ISOTROPIC_ELASTIC_FAILURE) in LS-DYNA was selected to simulate the model ice. The failure criterion for the material is the Von Mises yield criterion [34].
The material parameters of the model ice are shown in Table 4. The specific parameters of density, plastic hardening modulus, and plastic failure strain were obtained by the model ice mechanical test [1]. According to the relevant ice mechanics numerical simulation, the ranges of other material parameters were obtained, and finally, the specific parameters were given by trial and error.
According to von Bock und Polach [22], the air pores inside the model ice can be simulated by deleting elements by using a random algorithm, and the number of deleted elements depends on the porosity. This is another advantage of numerical simulation as compared with the experiments in Section 2.1, because ice porosity is difficult to control according to predetermined values, although an underwater air bubbling system has been placed already. The so-called K file of the model ice numerical model was employed and set the random number function, which was used to delete a part of the model ice elements according to the input value of porosity. The flow chart is shown in Figure 5. There are several key points in this program. First, read the K file, defining each line of information in it as a string. Determine whether each line string represents the coordinate information of the cell; if so, proceed to the next step, and vice versa, and output to a K file. Secondly, enter the sampling rate, combine the random function to obtain a set of unit numbers that need to be deleted, correspond them one-to-one with the read unit string, replace the unit numbers that need to be deleted with spaces, and output them to a new K file.
The corrected ice model with a different porosity is shown in Figure 6. Since the elements used in this model were hexahedral, the resulting pore shape was also hexahedral. In fact, model ice with different porosities differs in the number of pores, but the size of the individual pores does not change, which is the same as Von Bock und Polach [22]. The intervals between pores are random in Figure 6 and may be different from the actual internal pore distribution of ice because the brine channels of actual sea ice tend to be vertical and continuous, but the pores in numerical models cannot guarantee continuity. In addition, the distribution of pores is different from the actual sea ice, the pores of the actual sea ice tend to exist only in the interior, and the surface is continuous. Still, the pores in this paper also exist on the surface. The resulting difference in loading force is hard to evaluate at present, but it is supposed to be ignorable because the given porosity is small (≤7%).

2.2.2. Numerical Simulation of Circular Plate Center Loading Tests

In the numerical calculation, the ice specimen was loaded similarly to that shown in Figure 2. During modeling, the annular support below the ice specimen was set as a fixed boundary, the cone indenter had only the degrees of freedom in the Z direction, and the load was applied downward on the upper surface of the ice specimen at a constant loading rate. The main parameters of the numerical model are shown in Table 5. To simulate the fragmentation phenomenon of the model ice, a fine mesh was used in the central area of the model ice, and a large gradient mesh was used in the outer part, as shown in Figure 7a. Through the circular plate center loading tests, the numerical calculation condition was determined. Combining this with the pore simulation principle of Section 2.2.1, a model ice with a porosity of 3% can also be obtained, as shown in Figure 7b.
A rigid body model was selected to simulate the annular support and cone indenter, and the material parameters were defined using the keyword *MAT_RIGID, as shown in Table 6.
LS-DYNA provides a variety of contact algorithms for explicit analysis, divided into single-sided contact, point-to-face contact, and face-to-face contact [35]. The contact between the subglacial surface of the model and the upper surface of the annular support had a large contact area and symmetrical shape, so the surface contact was selected. Since the ring support was modeled using shell elements, the Automatic Contact Algorithm in surface contact was chosen. It can consider the influence of element thickness, allowing contact to appear on both sides of the shell element, making it more accurate when calculating contact forces. The contact between the upper surface of the model ice and the lower surface of the cone table indenter also adopted surface contact, and the erosion contact algorithm in surface contact was selected because the ice breakage effect of the erosion contact simulation model is good. Erosion contact was employed to control time steps. It automatically invoked negative volume failure criteria for all solid elements in the model, which circumvents procedural errors due to negative volumes by removing solid elements that produce negative volumes.
Different test conditions in numerical simulations are listed in Table 7. Compared with the test conditions in laboratory experiments (Table 2), the impact of ice temperature was ignored. In fact, the influence of temperature on the properties of ice materials is difficult to achieve by numerical simulation. Therefore, to reduce the influence of temperature on the results, the input parameters of the ice material in the numerical model were the average mechanical parameters of the model ice at a specific temperature (−0.8 °C). Similarly, in numerical simulation, the flexural strength as an input quantity, and the influence of external parameters’ (porosity) changes on the flexural strength cannot be reflected, and the influence of external parameters on bending performance can only be analyzed through the change in load response.

3. Results and Analysis

The test results and the numerical simulation results were compared from the two aspects of the time history curve and damage phenomenon.

3.1. Time History Curve and Failure Mode of Model Ice

The numerically calculated time history curve was compared with the model ice mechanics test time history curve in Figure 8. During the bending process of the ice specimen, both time history curves showed linear changes without an obvious yield stage and increased sharply from 0, reaching the maximum value Pmax within only 0.8 s, and it can be seen that the failure mode of the ice specimen was an obvious brittle failure. The flexural strength σf of the model ice can be calculated from Pmax according to Equation (3). The peak force of the two was similar, the loading time difference was 0.01 s, and the absolute error did not exceed 5 N, so the rationality of the calculation model can be verified from the changing trend of the time history curve. However, due to differences in material properties and test conditions, the downward trend of the test curve and the numerical curve was not the same. The downward trend of the test curve was faster because in the model test, when the sensor read the maximum force peak, the indenter stopped pressing down and upward, and the numerical calculation did not simulate this behavior, but this did not affect the numerical model’s simulation of the mechanical properties of the model ice.
A high-speed camera was used to observe the state of the model ice at the four moments A, B, C, and D in Figure 8, and the flexural failure process of the model ice was captured, as shown in Figure 9. It is clear that the model ice broke along the diameter, and the crack propagation was rapid, with an expansion time of less than 0.06 s. The model ice had no prominent yield stage, and there was no obvious plastic deformation at the failure location. It can be deduced that the failure of the model ice belongs to a brittle failure. Compared with the experimental phenomenon in the breakthrough loads of floating ice sheets carried out by Sodhi [36], the model ice disk did not have holes in the middle area but just failed cracks along the radial direction. This may be because of the different constraints.
The experiment could not observe the failure pattern of the bottom during the model ice fracture failure, but numerical simulations supplemented this part of the study. By observing the bottom of the model ice at the three moments E, F, and G in Figure 8, it was found that the model ice started to crack from the center of the bottom (Figure 10). The crack extended firstly along the radius direction to the bottom boundary, and then extended along the thickness direction to the top surface until the model completely failed, as shown in Figure 10. The flexural failure of the model ice was essentially the tensile failure of the bottom of the model ice, consistent with the flexural failure of the sea ice proposed by Lainey [37]. It indicates that the numerical model of the model ice can better reflect the crack propagation during model ice breakage.
Meanwhile, the damage phenomenon of the model ice was compared between numerical simulations with the picture captured by the high-speed camera in Figure 11. When the ice specimen failed, under the action of the cone indenter, there was a significant deflection change at the center position of the model ice, and the crack propagation was more consistent. It can be considered that the numerical simulation results and the experimental results have good similarity, and the conclusions obtained can confirm each other’s results. The impact factors on model ice flexural strength are discussed in two different ways, and these conclusions are considered to apply to both methods.
The numerical simulation also studied the stress distribution of the model ice at the time of cracking, and the stress distribution is shown in Figure 12. The stress on the top was 249.6 kPa, and the stress on the bottom was 71.3 kPa. The stress on the top was about 3.5 times the stress on the bottom. In fact, under the vertical action of the indenter, the stress on the top was caused by extrusion, and the stress on the bottom was caused by tension. Figure 10 shows that the destruction of the model ice occurred firstly on the bottom surface because the model ice is a brittle material: its tensile strength is much lower than its compressive strength [38]. Thus, the failure usually occurs at the stretch of the bottom surface.

3.2. Impact Factors on Flexural Strength

3.2.1. The Effect of Loading Rate

Discussions in this section are based on model test results because the behavior of the indenter was not fully represented in numerical simulations, considering that this may interfere with the conclusions drawn for the variable loading rate. Figure 13 shows a typical time history curve for model ice at different loading rates (−0.8 °C). Under different loading rates, the loading force of the model ice failure did not change much. When the loading force increased to the peak point, the time history curve decreased rapidly, and the downward trend was basically the same, which further indicates that the failure of the model ice was a brittle failure. It can be seen from the figure that the time difference between the peaks of the time history curves at different loading rates was obvious, followed by the insignificant change in the peak of the time history curve.
Based on experimental data, the correlation between loading rate and flexural strength was analyzed using the Pearson correlation coefficient and the Spearman correlation coefficient. Among them, the Pearson correlation coefficient can reflect the degree of linear correlation between two random variables; the Spearman correlation coefficient measures the strength of monotonicity between variables. Through calculation, it was found that the Pearson correlation coefficient between the flexural strength and loading rate of the model ice was 0.33, and the p-value was greater than 0.05, indicating that there was no significant linear correlation between the loading rate and the flexural strength of the model ice. The Spearman correlation coefficient between the flexural strength and loading rate was 0.37, and the p-value was greater than 0.05, indicating that there was no significant monotonic correlation between the loading rate and the flexural strength of the model ice, as shown in Figure 14. Therefore, it can be judged that there was no obvious correlation between the ice flexural strength and the loading rate of the model. In fact, in the other four test groups at temperatures of −2 °C, −4 °C, −6 °C, and −8 °C, the relationship between loading rate and flexural strength also showed no significant correlation, and thus was not shown here. This agrees with the conclusion of Frederking et al. in an ice basin [8]. Similarly, in polar sea ice observations, it has been found that the loading rate has less influence on the flexural strength of the ice [39].

3.2.2. The Effect of Ice Temperature

As mentioned above, since it is difficult to show the influence of temperature on the mechanical properties of ice materials in numerical simulations, this section is also based on the experiments. This section quantifies the relationship between model ice flexural strength and model ice temperature. The temperature range of the model ice was −0.8 °C to −9 °C, as can be seen from Section 3.2.1.
According to the statistical test data, the Pearson correlation coefficient between ice temperature and flexural strength was −0.89, and the p-value was 3.15 × 10−24. The Spearman correlation coefficient was −0.78, and the p-value was 7.1 × 10−13. Curve fitting to the experimental data was conducted, with a 95% confidence band, as shown in Figure 15, and the regression equation is as follows:
σ f = 192.24 204.36 · T
where σf is the model ice flexural strength in kPa, and T is the model ice temperature. It is clear from Figure 15 that in the temperature range from −0.8 °C to −9 °C, the model ice flexural strength decreased linearly with ice temperature, consistent with previous results of sea ice. Ding et al. measured the flexural strength of Bohai sea ice, and the experimental data showed that the flexural strength increased with the decrease in ice temperature, and the relationship between the two was linear [40]. It is worth noting that in the −2 °C to −6 °C range, there are some data points that deviate from the 95% confidence band, possibly due to errors in the measurement of ice temperature. Marchenko [6] tested the flexural strength of sea ice in the northwestern part of the Barents Sea and obtained the relationship between flexural strength in MPa and temperature and salinity, as follows:
σ f = 0.236 0.095 · S 0.0134 · T
where S is the salinity of sea ice. The empirical formulas for the flexural strength of Barents Sea ice also show ice temperature has an approximate linear relationship with the flexural strength if a constant or average salinity is given. However, it should be especially noted that the model results and the results of the flexural strength measurement in marine conditions in absolute values cannot coincide, as well as other strength characteristics.

3.2.3. The Effect of Ice Porosity

Since ice crystal observation experiments are labor-intensive and time-consuming, and the expected porosity is difficult to control artificially, this section is based on numerical simulation results. The relationship between porosity, the model ice load response, and the model ice failure mode were studied through four sets of numerical simulations with different porosities, as listed in Table 7. As mentioned earlier, the flexural strength was used as a parameter input to the numerical model, so the influence of porosity on the bending properties of the model ice cannot be discussed in this section through the flexural strength but only by the load response. It was found that the load response of the model ice flexural failure was in the range of 24.44 N~26.5 N, as shown in Figure 16. The greater the porosity, the smaller the load response of the model ice, and the earlier the time for failure to occur. Statistics show that the load response of model ice with 7% porosity was 0.04 s earlier than model ice without porosity. In addition, the peak load decreases gradually with increasing porosity, and this conclusion is similar to Wang’s conclusions about sea ice at Prydz Bay, East Antarctic [41].
The final failures of model ice at different porosities are also shown in Figure 17. A crack extending along the diameter appeared on the surface of the model ice under different porosities. There was a significant deflection change at the center of the model ice, and there was a gap in the thickness direction. This indicates that within the range of porosity < 7%, porosity had little effect on the failure mode of the model ice. Compared with nonporous model ice, model ice with pores was more likely to develop microcracks around the pores or further expand the pores. There was a certain crush failure near the indenter, but the failure mode of the model ice was still dominated by flexural failure.

4. Conclusions

In this paper, both laboratory experiments and numerical modeling were carried out to investigate the flexural strength of a columnar saline model ice under circular plate central loading. The influence of different factors on the flexural strength of model ice was analyzed and the crack propagation law of the model ice flexural fracture failure was found. The main conclusions are as follows:
  • The experimental and numerical results were compared from two aspects including the time history curve and damage phenomenon, and their results agree well; these could reflect the flexural strength characteristics of the model ice and confirm each other’s results.
  • According to the time history curve of the ice specimen from the initial bearing to the flexural failure, it was found that the ice specimen had no obvious yield stage. The high-speed camera observed no obvious plastic deformation at the failure location. The model ice began to crack from the center of the bottom surface, and the crack extended along the radius direction to the lower surface boundary, then extended along the thickness direction to the top surface until complete failure. The failure process of the model ice was judged to be a typical brittle failure.
  • There was no significant correlation between the loading rate and the flexural strength. A significant linear correlation between the model ice temperature and the flexural strength was explored, and the flexural strength of the model ice increases continuously with the continuous decrease in the model ice temperature in the range from −0.8 °C to −9 °C. The larger the porosity, the smaller the load response of the model ice, and the earlier the time of failure. Compared with the nonporous model ice, the load response of model ice with 7% porosity was reduced by 7.8%, and the failure time was 0.04 s earlier. Within the range of 7%, ice porosity had little effect on the failure mode of the model ice.
This paper provided a feasible means to measure and predict the mechanical properties of model ice and realized the simulation of the internal pores of the model ice, forming a preliminary foundation for the research of the mechanism of ice loading under the vertical interaction between the structure and the ice cover. In the future, the results obtained by different methods such as circular plate center loading tests, in situ cantilever tests, and simply supported beam tests will be enriched and compared with each other, to establish a systematic ice flexural strength measurement and analysis technology. In addition, the numerical model will continue to be optimized and improved to reflect the influence of temperature and brine, especially the internal structure and anisotropic characteristics of the model ice.

Author Contributions

Conceptualization, Y.T.; methodology, Y.T. and W.Z.; software, W.Z.; validation, W.Z., C.Y. and X.G.; formal analysis, W.Z.; investigation, Y.T.; writing—original draft preparation, Y.T. and W.Z.; writing—review and editing, P.L. and Q.Y.; project administration, Y.T.; funding acquisition, Y.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fundamental Research and Development Project of China [Grant No. JCKY2020206B073], the National Natural Science Foundation of China [Grant Nos. 52192690, 52192694], and the Hi-Tech Ship Project of the Ministry of Industry and Information Technology [Grant No. [2021]-342].

Data Availability Statement

The data are not publicly available due to the confidentiality policy of our institution requires.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhao, W.; Tian, Y.; Yu, C. Experimental Investigation on Breaking Strength of Model Ice with Circular Ice Plate Indentation Tests. In Proceedings of the 32nd International Ocean and Polar Engineering Conference, Shanghai, China, 5–10 June 2022. [Google Scholar]
  2. Ye, L.; Guo, C.; Wang, C.; Wang, C.; Chang, X. Peridynamic solution for submarine surfacing through ice. Ships Offshore Struc. 2020, 15, 535–549. [Google Scholar] [CrossRef]
  3. Meng, X. Research on the Force of Submarine to Destory the Ice Layer Upwards; Harbin Institute of Technology: Harbin, China, 2019. [Google Scholar]
  4. Ji, S.; Wang, A.; Su, J.; Yue, S. Experimental studies and characteristic analysis of sea ice flexural strength in Bohai Sea. Adv. Water Sci. 2011, 22, 266–272. [Google Scholar] [CrossRef]
  5. Weeks, W.; Anderson, D. An experimental study of strength of young sea ice. Trans. Am. Geophys. Union 1958, 39, 641–647. [Google Scholar] [CrossRef]
  6. Marchenko, A.; Karulina, M.; Karulin, E. Flexural strength of ice reconstructed from field tests with cantilever beams and laboratory tests with beams and disks. In Proceedings of the International Conference on Port and Ocean Engineering under Arctic Conditions, Busan, Republic of Korea, 11–16 June 2017. [Google Scholar]
  7. Li, Z.; Wang, Y.; Li, G. Experimental analysis of flexural strength and elastic modulus of the DUT-1 model ice. Adv. Water Sci. 2002, 13, 292–297. [Google Scholar] [CrossRef]
  8. Frederking, R.; Timco, G. On measuring flexural properties of ice using cantilever beams. Ann. Glaciol. 1983, 4, 58–65. [Google Scholar] [CrossRef]
  9. Krupina, N.; Kubyshkin, N. Flexural strength of drifting level first-year ice in the Barents Sea. In Proceedings of the 17th International Offshore and Polar Engineering Conference, Lisbon, Portugal, 1–6 July 2007. [Google Scholar]
  10. Gang, X.; Tian, Y.; Ji, S.; Guo, W.; Kou, Y. Experimental analysis on flexural strength of columnar saline model ice. J. Ship Mech. 2021, 25, 336–341. [Google Scholar] [CrossRef]
  11. Xu, Y.; Hu, Z.; Chen, G.; Xu, Y. Overview of the investigation methods for ship-ice interaction analysis. J. Ship Mech. 2019, 23, 110–123. [Google Scholar]
  12. Gagnon, R.; Gammon, P. Characterization and flexural strength of iceberg and glacier ice. J. Glaciol. 1995, 41, 103–111. [Google Scholar] [CrossRef]
  13. Barrette, P.; Jordaan, I. Beam Bending and Fracture Behavior of Iceberg Ice; National Research Council Program on Energy Research and Development: Saskatoon, SK, Canada, 2001; pp. 4–18. [Google Scholar] [CrossRef]
  14. Timco, G.; O’Brien, S. Flexural strength equation for sea ice. Cold Reg. Sci. Technol. 1994, 22, 285–298. [Google Scholar] [CrossRef]
  15. Gang, X.; Tian, Y.; Yu, C.; Ji, S.; Kou, Y. Numerical simulation analysis of flexural strength of columnar saline ice model. Chin. J. Ship Res. 2021, 16, 143–149. [Google Scholar] [CrossRef]
  16. Yu, C.; Tian, Y.; Wang, W. Review of research on ice loading of offshore structures in level ice fields. Chin. J. Ship Res. 2021, 16, 39–53. [Google Scholar] [CrossRef]
  17. Sand, B. Nonlinear Finite Element Simulations of Ice Forces on Offshore Structures; University of Technology: Luleå, Sweden, 2008. [Google Scholar]
  18. Wang, C.; Wang, J.; Wang, C.; Guo, C.; Zhu, G. Research on movement of cylindrical structure out of water and breaking through ice layer based on S-ALE method. Chin. J. Theor. Appl. Mech. 2021, 53, 3110–3123. [Google Scholar] [CrossRef]
  19. Liu, R.; Yan, J.; Li, S. Modeling and simulation of ice-water interactions by coupling peridynamics with updated Lagrangian particle hydrodynamics. Comp. Part. Mech. 2020, 7, 241–255. [Google Scholar] [CrossRef]
  20. Li, J. Study on Hydrodynamic Effects on Ice-Structure Interaction; Dalian University of Technology: Dalian, China, 2020. [Google Scholar]
  21. Wang, J.; Derradji-Aouat, A. Implementation, verification, and validation of the multi-surface failure envelope for ice in explicit FEA. In Proceedings of the 20th International Conference on Port and Ocean Engineering Under Arctic Conditions, Luleå, Sweden, 9–12 June 2009. [Google Scholar]
  22. Von Bock und Polach, R.B.; Ehlers, S. Model scale ice-Part B: Numerical mode. Cold Reg. Sci. Technol. 2013, 94, 53–60. [Google Scholar] [CrossRef]
  23. Herrnring, H.; Ehlers, S. A finite element model for compressive ice loads based on a mohr-coulomb material and the node splitting technique. In Proceedings of the 40th ASME International Conference on Ocean, Offshore and Arctic Engineering (OMAE), Virtual, 21–30 June 2021. [Google Scholar]
  24. Tian, Y.; Ji, S.; Wang, Y.; Guo, W.; Kou, Y.; Gang, X. Research on sea ice simulation and measurement in small ice model basin of CSSRC. Mar. Environ. Sci. 2021, 40, 277–286. [Google Scholar] [CrossRef]
  25. Evers, K. Model tests with ships and offshore structures in HSVA’s ice tanks. In Proceedings of the 24th International Conference on Port and Ocean Engineering under Arctic Conditions, Busan, Republic of Korea, 11–16 June 2017. [Google Scholar]
  26. Evers, K.; Jochmann, P. An advanced technique to improve the mechanical properties of model ice developed at the HSVA ice tank. In Proceedings of the 12th International Conference on Port and Ocean Engineering under Arctic Conditions, Hamburg, Germany, 17–20 August 1993. [Google Scholar]
  27. Tian, Y.; Ji, S.; Kou, Y.; Guo, W.; Chen, Z.; Gang, X. Characterization of Uniaxial Compression Strength for Columnar Saline Model Ice in CSSRC Small Ice Model Basin. J. Ship Mech. 2020, 24, 1647–1656. [Google Scholar] [CrossRef]
  28. Yu, C. Research on Ice Loading of Typical Offshore Ice-Resistant Structure; China Ship Scientific Research Center: Wuxi, China, 2021. [Google Scholar]
  29. Timoshenko, S.; Woinowsky-Krieger, S. Theory of Plates and Shells, 2nd ed.; McGraw-Hill: New York, NY, USA, 1959; pp. 135–146. [Google Scholar]
  30. Timco, G.W.; Weeks, W.F. A review of the engineering properties of sea ice. Cold Reg. Sci. Technol. 2010, 60, 107–129. [Google Scholar] [CrossRef]
  31. Bassali, W. The transverse flexure of thin elastic plates supported at several points. Math. Proc. Camb. 1957, 53, 728–743. [Google Scholar] [CrossRef]
  32. Guo, T.; Zhang, A.; Yu, B. Underwater collision simulation of the pressure hull with spherical head based on LS-DYNA. J. Ship Mech. 2021, 25, 210–219. [Google Scholar]
  33. Karr, D.; Choi, K. A 3-dimensional constitutive damage model for polycrystalline ice. Mech. Mater. 1989, 8, 55–66. [Google Scholar] [CrossRef]
  34. Qian, Y. Study on the Calculation Method for the Ice-Load of Vertical Ice-Breaking; Harbin Engineering University: Harbin, China, 2020. [Google Scholar]
  35. Zhao, H. LS-DYNA Dynamic Analysis Guide; The Publishing House of Ordnance Industry: Beijing, China, 2003; pp. 64–71. [Google Scholar]
  36. Sodhi, D.S. Breakthrough loads of Floating Ice Sheet. J. Cold Reg. Eng. 1995, 9, 4–22. [Google Scholar] [CrossRef]
  37. Lainey, L.; Tinawi, R. The mechanical properties of sea ice a compilation of available data. Can. J. Civil. Eng. 1984, 11, 884–923. [Google Scholar] [CrossRef]
  38. Ni, B.Y.; Tan, H.; Di, S.-C.; Zhang, C.X.; Li, Z.; Huang, L.; Xue, Y.Z. When does a light sphere break ice plate most by using its net buoyance? J. Mar. Sci. Eng. 2023, 11, 289. [Google Scholar] [CrossRef]
  39. Barrette, P.; Phillips, R.; Clark, J.; Crocker, G.; Jones, S. Flexural behavior of model sea ice in a centriguge. J. Cold Reg. Eng. 1999, 13, 122–138. [Google Scholar] [CrossRef]
  40. Ding, D. An Introduction to Engineering Sea Ice; Ocean Press: Beijing, China, 2000; pp. 84–88. [Google Scholar]
  41. Wang, Q.; Li, Z.; Lu, P.; Xu, Y.; Li, Z. Flexural and compressive strength of the landfast sea ice in the Prydz Bay, East Antarctic. Cryosphere 2022, 16, 1941–1961. [Google Scholar] [CrossRef]
Figure 1. Interior scene of CSSRC SIMB.
Figure 1. Interior scene of CSSRC SIMB.
Water 15 03371 g001
Figure 2. Circular plate center loading test: (a) schematic diagram; (b) partial schematic diagram.
Figure 2. Circular plate center loading test: (a) schematic diagram; (b) partial schematic diagram.
Water 15 03371 g002
Figure 3. Photo of model ice sample during sampling stage.
Figure 3. Photo of model ice sample during sampling stage.
Water 15 03371 g003
Figure 4. Test device layout.
Figure 4. Test device layout.
Water 15 03371 g004
Figure 5. Flow chart of internal defects in construction of model ice.
Figure 5. Flow chart of internal defects in construction of model ice.
Water 15 03371 g005
Figure 6. Numerical model of model ice with different porosities of (a) 0%, (b) 3%, (c) 5%, and (d) 7%.
Figure 6. Numerical model of model ice with different porosities of (a) 0%, (b) 3%, (c) 5%, and (d) 7%.
Water 15 03371 g006
Figure 7. Model ice meshing with different porosities of (a) 0% and (b) 3%.
Figure 7. Model ice meshing with different porosities of (a) 0% and (b) 3%.
Water 15 03371 g007
Figure 8. Comparison of time history curves between numerical result and test result (−0.8 °C, 150 mm/min). A–G denote the different loading pictures shown in Figure 9 and Figure 10.
Figure 8. Comparison of time history curves between numerical result and test result (−0.8 °C, 150 mm/min). A–G denote the different loading pictures shown in Figure 9 and Figure 10.
Water 15 03371 g008
Figure 9. The loading process for model ice at (a) 0.11 s, (b) 0.45 s, (c) 0.79 s, and (d) 0.85 s (−0.8 °C, 150 mm/min), corresponding to points A, B, C, and D in Figure 8.
Figure 9. The loading process for model ice at (a) 0.11 s, (b) 0.45 s, (c) 0.79 s, and (d) 0.85 s (−0.8 °C, 150 mm/min), corresponding to points A, B, C, and D in Figure 8.
Water 15 03371 g009
Figure 10. Crack propagation during model ice breakage (a) 0.77 s, (b) 0.78 s, and (c) 0.79 s (−0.8 °C, 150 mm/min), corresponding to points E, F, and G in Figure 8.
Figure 10. Crack propagation during model ice breakage (a) 0.77 s, (b) 0.78 s, and (c) 0.79 s (−0.8 °C, 150 mm/min), corresponding to points E, F, and G in Figure 8.
Water 15 03371 g010
Figure 11. Comparison of failure phenomenon: (a) numerical simulation, (b) test result (−0.8 °C, 150 mm/min).
Figure 11. Comparison of failure phenomenon: (a) numerical simulation, (b) test result (−0.8 °C, 150 mm/min).
Water 15 03371 g011
Figure 12. Stress distribution of disk model in failure: (a) stress on the top—249.6 kPa; (b) stress on the bottom—71.3 kPa (−0.8 °C, 150 mm/min).
Figure 12. Stress distribution of disk model in failure: (a) stress on the top—249.6 kPa; (b) stress on the bottom—71.3 kPa (−0.8 °C, 150 mm/min).
Water 15 03371 g012
Figure 13. Time history curves of flexural strength experiments (−0.8 °C).
Figure 13. Time history curves of flexural strength experiments (−0.8 °C).
Water 15 03371 g013
Figure 14. Relationship between flexural strength and loading rate (−0.8 °C).
Figure 14. Relationship between flexural strength and loading rate (−0.8 °C).
Water 15 03371 g014
Figure 15. Relationship between flexural strength and ice temperature.
Figure 15. Relationship between flexural strength and ice temperature.
Water 15 03371 g015
Figure 16. Load response of model ice with different porosities: (a) time history curve (b) force–porosity curve (−0.8 °C, 150 mm/min).
Figure 16. Load response of model ice with different porosities: (a) time history curve (b) force–porosity curve (−0.8 °C, 150 mm/min).
Water 15 03371 g016
Figure 17. Failure results of model ice with different porosities of (a) 0%, (b) 3%, (c) 5%, and (d) 7% (−0.8 °C, 150 mm/min).
Figure 17. Failure results of model ice with different porosities of (a) 0%, (b) 3%, (c) 5%, and (d) 7% (−0.8 °C, 150 mm/min).
Water 15 03371 g017
Table 1. The main physical and mechanical performance parameters of model ice in SIMB [24].
Table 1. The main physical and mechanical performance parameters of model ice in SIMB [24].
ParameterValue
Density/g·cm−30.89~0.91
Thickness/mm10~100
Flexural strength/kPa94.7
Compressive strength/kPa80~150
Elastic modulus/MPa250~450
Elastic modulus/flexural strength800~1500
Table 2. Summary of plate specimen size and testing conditions in laboratory experiments.
Table 2. Summary of plate specimen size and testing conditions in laboratory experiments.
Diameter/Thickness
(mm/mm)
Specimen NumberLoading Rate
(mm/min)
Ice Average Temperature (°C) Laboratory
Temperature (°C)
Measurements
140/2020100, 150, 200, 250, 300−0.8−0.8Peak force
Failure time
140/2015−2.0−2.0
140/2015−4.0−4.0
140/2015−6.0−6.0
140/2015−8.0−8.0
Table 3. Test instrument parameter table.
Table 3. Test instrument parameter table.
InstrumentAccuracy
Thermometer0.01 °C
Electronic testing machine1 mm/min
High-speed camerasFrame rate: 2800
Sensor0.01 N
Vernier calipers0.01 mm
Ice density measurement instruments
Salinity meter
0.01 g/cm3
±3% (FS)
Table 4. Model ice material parameters.
Table 4. Model ice material parameters.
Material PropertiesValue
Density/g·cm−30.92
Shear modulus /MPa76.9
Plastic hardening modulus /MPa94.7
Yield stress/kPa83
Bulk modulus/MPa75
Plastic failure strain
Failure pressure/kPa
0.05
−110
Table 5. Main parameters of the numerical model of the circular plate center loading test.
Table 5. Main parameters of the numerical model of the circular plate center loading test.
Ice Specimen
Radius
(mm)
Ice Specimen Thickness
(mm)
Ring Support Outer Diameter
(mm)
Ring Support
Inner Diameter
(mm)
The Radius of the Lower Surface of the Tapered Indenter
(mm)
702070605
Table 6. Ring support and indenter material parameters.
Table 6. Ring support and indenter material parameters.
Material PropertiesValue
Density/g·cm−37.83
Elastic modulus/GPa207
Poisson’s ratio0.33
Table 7. Summary of plate specimen size and testing conditions in numerical simulations.
Table 7. Summary of plate specimen size and testing conditions in numerical simulations.
Diameter/Thickness
(mm)
Specimen NumberLoading Rate
(mm/min)
Ice Porosity (%) Measurements
140/2041500Peak force
Failure time
140/2043
140/2045
140/2047
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

Tian, Y.; Zhao, W.; Yu, C.; Gang, X.; Lu, P.; Yue, Q. Investigations on Flexural Strength of a Columnar Saline Model Ice under Circular Plate Central Loading. Water 2023, 15, 3371. https://doi.org/10.3390/w15193371

AMA Style

Tian Y, Zhao W, Yu C, Gang X, Lu P, Yue Q. Investigations on Flexural Strength of a Columnar Saline Model Ice under Circular Plate Central Loading. Water. 2023; 15(19):3371. https://doi.org/10.3390/w15193371

Chicago/Turabian Style

Tian, Yukui, Weihang Zhao, Chaoge Yu, Xuhao Gang, Peng Lu, and Qianjin Yue. 2023. "Investigations on Flexural Strength of a Columnar Saline Model Ice under Circular Plate Central Loading" Water 15, no. 19: 3371. https://doi.org/10.3390/w15193371

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