Next Article in Journal
Theoretical Advancements on a Few New Dependence Models Based on Copulas with an Original Ratio Form
Next Article in Special Issue
Molecular Dynamics Simulations Correlating Mechanical Property Changes of Alumina with Atomic Voids under Triaxial Tension Loading
Previous Article in Journal
Nonlinear Modeling of an Automotive Air Conditioning System Considering Active Grille Shutters
Previous Article in Special Issue
Empirical Modeling of Transverse Displacements of Single-Sided Transversely Cracked Prismatic Tension Beams
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid Finite-Discrete Element Modeling of the Mode I Tensile Response of an Alumina Ceramic

Department of Mechanical Engineering, University of Alberta, Edmonton, AB T6G 2R3, Canada
*
Author to whom correspondence should be addressed.
Modelling 2023, 4(1), 87-101; https://doi.org/10.3390/modelling4010007
Submission received: 2 February 2023 / Revised: 4 March 2023 / Accepted: 6 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Modeling Dynamic Fracture of Materials)

Abstract

:
We have developed a three-dimensional hybrid finite-discrete element model to investigate the mode I tensile opening failure of alumina ceramic. This model implicitly considers the flaw system in the material and explicitly shows the macroscopic failure patterns. A single main crack perpendicular to the loading direction is observed during the tensile loading simulation. Some fragments appear near the crack surfaces due to crack branching. The tensile strength obtained by our model is consistent with the experimental results from the literature. Once validated with the literature, the influences of the distribution of the flaw system on the tensile strength and elastic modulus are explored. The simulation results show that the material with more uniform flaw sizes and fewer big flaws has stronger tensile strength and higher elastic modulus.

1. Introduction

Advanced ceramics are often used as structural components in shielding applications [1,2,3,4] because of their desirable properties, such as low density [5], high hardness [6], and high wear resistance [7]. In ballistic applications, the role of the ceramic upon impact is to blunt the projectile [8,9,10] and to distribute the load [9]. During impact, pre-existing micro-cracks or pores in ceramics serve as stress concentration sites [11], which significantly affect fracture behavior [12,13,14,15]. The relationships between microstructure and failure processes have been widely investigated [14,16,17,18,19,20]. For example, Munro [16] and Nohut [17] found that the strength of alumina ceramics is limited by the distribution of flaws in the material, and the Weibull analysis could be used to characterize both the strength and the flaw system. Recently, Lo et al. [14] studied the microstructural and mechanical variability of AD85 alumina, and they found the flaw characterization (e.g., pore size, spatial distribution, orientation, and morphology) significantly influences the mechanical response of the alumina ceramic. In a separate study, Hogan et al. [18,19,20] identified the relationship between microstructure and fragment size by observing impact and compression experiments of ceramics. To better understand the failure mechanisms of advanced ceramics, controlled experiments such as uniaxial compression, beam bending tests, and Brazilian disk experiments are often coupled with advanced high-speed imaging [21,22,23,24,25,26,27,28,29]. In these experiments, the mode I tensile opening cracks are widely accepted as an important failure mode of advanced ceramics [30,31,32,33,34]. However, direct tension tests is a challenge for advanced ceramics because of their brittleness. In the current study, the mode I tensile behavior of an alumina ceramic is investigated. The widespread use of high-purity alumina (Al 2 O 3 ) as body armor material is due to its beneficial combination of favorable ballistic properties, affordability, and well-established manufacturing processes [35]. A significant number of studies have been conducted with the objective of enhancing the performance of alumina ceramics [35,36,37,38]. These ceramics are usually produced through various sintering techniques such as flash sintering (FS) [38], hot-pressing sintering (HPS) [36], and spark plasma sintering (SPS) [35]. The trend of 3D printing ceramics is gradually becoming more prevalent, offering multiple methods for production [35]. Some of the most common methods include stereo lithography, which involves curing photo-curable binder loaded ceramic pastes, and selective laser sintering, involving laser sintering of green powder beds [35]. Other techniques such as modified inkjet printing and binder jetting are also utilized in ceramic material production [37]. These different manufacturing methods would affect alumina ceramic’s microstructure (e.g., flaws). However, these experimental studies are limited by manufacturing technologies and testing methods for the tensile response of ceramics. For example, many factors (e.g., additives [38], debinding step [36], and temperature [35]) can affect the internal flaws and mechanical performance of ceramics in sintering manufacturing, which is complex [37]. In addition, exploring the influence of flaw systems on direct tensile performance is expensive and challenging [37,39]. To overcome these challenges, in the current study, the inherent microstructural flaws based on experimental studies [16,17,40] are incorporated to the numerical modelling method to more realistically explore the mode I tensile behavior of an alumina ceramic.
Numerous numerical models have been established to describe the failure behavior of brittle materials. These include the continuum damage mechanics (CDM) [41,42], the extended Finite Element Method (XFEM) [43], virtual crack closure technique (VCCT) [44,45,46], and the cohesive zone method (CZM) [47]. The CDM method utilizes damage parameters to explain the failure process but cannot capture crack-induced discontinuities [48]. VCCT, on the other hand, can simulate pre-defined crack propagation by imposing constraints on the nodes at crack edges but requires re-meshing [44,45,46]. XFEM avoids mesh refinement and reconstruction by modifying the displacement approximation function in conventional FEM with an enrichment function term [43]. While these methods can model progressive cracking behavior, they necessitate additional limitations, such as external criteria for discontinuous displacement enrichment [43], re-meshing requirements [44,45,46], and complex model pre-definitions [44,45,46]. Additionally, these methods cannot effectively address complex cracking problems, including crack intersection, coalescence, and branching. The CZM has several advantages over the other methods, including (1) creating new surfaces to generate cracks, (2) allowing for branched and intersecting cracks, and (3) eliminating the singularity present in linear elastic fracture mechanics [49]. In the literature [50,51], the CZM framework has been applied to study the strength of advanced ceramics, dynamic fracture events [52], and fragmentation of brittle materials [53]. In the CZM method, new surfaces are created when the fracture occurs, and these new faces require numerical contact algorithms, which makes CZM ideally suited for discretized methods [54]. To address the mutually interacting separate fragments in fracture processes, Munjiza [48] developed an innovative numerical approach called the hybrid finite-discrete element method (HFDEM). One distinct feature of the HFDEM is that it is able to capture the transition from a continuum (e.g., finite element method) to a discontinuous-based method (e.g., discrete element method) [55,56] to overcome the inability of these methods to capture progressive damage and failure processes in brittle materials (e.g., geomaterials [57,58,59] and ceramics [56]). In HFDEM, materials are often discretized as triangle elements (two-dimensional) or tetrahedral elements (three-dimensional), and cohesive elements are utilized to connect these discrete elements to represent the potential arbitrary crack path [60]. In addition, the use of tetrahedral elements offers the advantage of generating more potential fracture surfaces when compared with hexahedron elements, which make the results using tetrahedral element more reliable [48,60,61,62,63]. Then, an explicit finite difference time integration scheme is applied to solve the motion of the discretized system [48]. Recently, HFDEM has been used in modelling brittle materials under different loading conditions, as it can explicitly describe the process of fracture nucleation and growth, as well as the interaction of newly created discrete fragments [54,60,64]. In the current study, the HFDEM is applied to investigate the failure process of the alumina ceramic.
Motivated by these previous studies, this paper aims to develop a three-dimensional HFDEM model to investigate the mode I tensile opening failure of the alumina ceramic. This method is focused on the macroscopic failure process accounting for the flaw system with an implicit method. The material is assumed to have Weibull distributed initial flaws, making its failure mode stochastic. The failure of the specimen is simulated through the nucleation and propagation of these cracks generated by flaws. Limited experimental results are available for the direct tension problems, owing to the challenges posed by testing, such as specimen gripping and alignment [65]. The indirect tensile strength from our previous study is used [29], and the simulation results show reasonable agreement with the experimental results. The influence of flaw system on the tensile strength is also investigated. Overall, the current study provides new insight into the mode I tensile fracture behavior of alumina ceramic.

2. Computational Approach

In this section, a three-dimensional hybrid finite-discrete element method (HFDEM) is established to describe the failure process of brittle materials, and then applied to the alumina ceramic. First, we develop the main features of the cohesive law to represent the mode I tensile response of brittle materials. Second, the distribution of the flaw system [16,17,40] is considered in this model. Finally, the model is implemented with a FORTRAN vectorized user-material (VUMAT) subroutine in ABAQUS/Explicit to solve the model numerically.

2.1. The Cohesive Law

An extensive account of the HFDEM theories and their finite element implementation can be found in [48,55,56,57,58,59,66]. In this section, we summarize the main features of the cohesive law used in the current study:
σ = ( 1 D ) K δ
with
σ = σ 1 τ 1 τ 2 ,       K = K 1 0 0 0 K 2 0 0 0 K 3 ,       δ = δ 1 δ 2 δ 3
where σ represents the interface stress, σ 1 is the normal stress, and τ 1 and τ 2 are the shear stresses in the other two directions. D is a scalar damage parameter of the REA, where D = 0 is the intact state, and D = 1 represents a fully damaged state. K is the penalty stiffness of the interface where subscript 1 represents the normal direction, and subscripts 2 and 3 represent the two shear directions. δ represents relative displacements.
The linear irreversible cohesive law is widely used for the decaying response of brittle materials, such as rocks [54,60,64] and ceramics [51,66,67,68]. For a damage value in the range of 0∼1, the damage evolution can be expressed as:
D = max { 0 , min { 1 δ c δ e δ c δ m 0 , 1 } }
where δ e is the effective relative displacement ( δ e = δ 1 2 + δ 2 2 + δ 3 2 ), and δ c represents the critical displacement when interface failure occurs. δ m 0 is the relative displacement when the damage initiates under mixed mode loading, which is obtained by:
δ m 0 = δ 1 0 2 1 + β 2 δ 2 0 δ 3 0 δ 2 0 δ 3 0 + β 2 δ 1 0 2 and β = δ 2 2 + δ 3 2 δ 1
δ 1 0 = σ 1 0 K 1 , δ 2 0 = δ 3 0 = τ 1 0 K 2 = τ 2 0 K 3
where σ 1 0 , τ 1 0 , and τ 2 0 are the strengths in the normal and two shear directions under pure mode loading, and δ 1 0 , δ 2 0 and δ 3 0 are the corresponding displacements.
Until here, the only unknown parameter in Equation (2) is the critical displacement δ c . In the cohesive zone method, mode II and mode III fracture is often regarded as the same due to a lack of mode III mechanical property information [69]. Thus, the mixed mode energy-based failure function [70] can be expressed as:
G 1 G 1 c γ + G s h e a r G s h e a r c γ = 1
where G s h e a r c is the critical shearing energy release rate with G s h e a r c = G 2 c = G 3 c , and G 1 c , G 2 c and G 3 c terms are the fracture energies under pure mode loading. The G s h e a r is the energy dissipation rate by shearing, which is sum of the energy release by the mixed mode II and III crack, G s h e a r = G 2 + G 3 , and the G 2 and G 3 are given in Equation (6). A quadratic γ = 2 failure criteria is frequently chosen according to the mixed mode experimental results [70,71] and so it is used here. For the cohesive interface, the energy dissipation rates are:
G i = σ i d δ i ( i = 1 , 2 , 3 )
In HFDEM model proposed in this current study, the bonding stresses transferred by the material are functions of the relative displacements across the crack elements, and this is illustrated in Figure 1 for the mode I tensile response.

2.2. The Microscopic Stochastic Fracture Model

For brittle materials such as ceramics, the strength is limited by the distribution of flaws in the material specimen, and any flaw in the material can serve as an origin of a crack [16]. The Weibull strength distribution is widely used in ceramics to characterize the influence of flaws statistically [16,17,40].
P ( σ , V ) = 1 exp V V 0 σ σ 0 m
where P ( σ , V ) is the cumulative failure probability of an alumina ceramic, V is the volume of the investigated component, V 0 is the characteristic volume, σ is the applied stress, m is the Weibull modulus, and σ 0 is the Weibull characteristic strength.
In HDFEM, the cohesive elements represent the intrinsic and extrinsic flaws [60]. Thus, Weibull’s statistical strength theory [17] is applied to cohesive elements to show the stochastic properties of the alumina ceramics. In the recent study by Daphalapurkar et al. [68], a modified microscopic facet-strength probability distribution based on Equation (7) is applied to the dynamically introduced cohesive method.
f ( σ ) = m 0 σ 0 A A 0 m 0 m a σ σ 0 m 0 1 exp A A 0 m 0 m a · σ σ 0 m 0
where A is the facet area of the cohesive element, A 0 is the characteristic area; m 0 is the Weibull modulus of strength distribution, and m a is the Weibull modulus for the effective area modification. In the current study, we applied the microscopic facet-strength probability model to the pre-inserted cohesive method to represent the flaws in the material. Monte Carlo simulations generate the strength data affected by flaws according to Equation (8).
Figure 2 shows a Monte Carlo simulation for Weibull’s statistical strength distribution of cohesive elements with m 0 = 11.0 , m a = 11.0 , σ 0 = 440.0 , A = 0.01268 and A 0 = 0.013 . The percentage of low-strength cohesive elements (below 350 MPa) is around 7.9%, which is associated with the big flaws in the material. The rest of the cohesive elements (around 92.1 %) have strong strength (between 350 and 530 MPa), which corresponds to the material with smaller flaws. The Weibull statistical model for failure has been adapted to fit within the cohesive element framework by incorporating the following assumptions: (1) The pre-inserted cohesive elements in the finite element mesh are treated as potential locations for flaws. (2) When the tensile stress applied on a facet exceeds its strength, the flaw in that location becomes a microcrack. The cohesive elements with low strength will activate at nearly negligible loads and can be considered as intrinsic microcracks.

3. The Hybrid Finite-Discrete Element Method

The hybrid finite-discrete element method is an advanced numerical method that combines continuum mechanics methods with the discrete element method (DEM) algorithms to solve complicated crack problems involving multiple interacting deformable bodies [48,55,56]. In HFDEM, the specimens are considered a collection of elastic bulk elements connected by cohesive elements [48]. The cohesive elements represent the inherent flaws in the specimens, which become the potential cracks during the loading process, and are also referred to as “crack elements” in HFDEM [60]. When D=1, the crack element is completely broken, and the cohesive element is deleted from the model, which generates new crack surfaces. In the HFDEM model, cohesive elements introduce a well-defined length scale into the material description, and are consequently sensitive to the size of the element [72]. In the current study, the physical and statistical model (Equation (8)) has introduced an internal microstructural length scale ( A 0 ) into our HFDEM model, and thus regularizes the problem in terms of mesh convergence. This technique (Equation (8)) has been applied to the dynamically introduced cohesive method and demonstrated the effectiveness of mesh convergence [67,68]. However, this technique (Equation (8)) has not been used in the pre-inserted cohesive method before, and we use the pre-inserted cohesive method considering the physical and statistical model (Equation (8)) to simulate the mode I tensile behavior in this paper.

3.1. Modeling Mode I Failure Considering Distributed Flaws

In this study, the mechanical properties of the CeramTec 98% are provided by the manufacturer [73] and evaluated in our previous studies [29,74,75]. This CeramTech ceramic has an alumina content of 98 mass percentage, a low porosity of less than 2%, a high hardness of 13.5 GPa, a low density of 3.8 g/cm 3 , Young’s modulus of E = 335 GPa, and Poisson’s ratio of v = 0.23 . For the physical and statistical model [16,17,29,40,74,75], the Weibull modulus of the strength distribution is m 0 = 11 , the Weibull modulus for the effective area modification is m a = 11 , the Weibull characteristic strength is σ 0 = 440 MPa, the characteristic area is A 0 = 0.013 mm 2 , and mode I fracture energy is G c = 0.04 N/mm, which are summarized in Table 1.
The simulation sample for the mesh sensitivity analysis is a ceramic block with dimensions L x = 1.5 mm, L y = 0.5 mm, and L z = 2.5 mm. In the simulations, a fixed-displacement boundary condition is implemented to the bottom, and the specimen is free to expand or contract freely in the lateral direction. The loading process is performed by imposing a velocity boundary condition in the z-direction to mimic the direct tension loading. The boundary velocity, v 0 , is fixed to 1 mm/s, which corresponds to quasi-static direct tension loading. This model has also been also employed by Zhou and Molinari [67] to study mesh sensitivity. In the literature [61,72], various studies have focused on the influence of mesh size on the mechanical response of brittle materials, and the upper limit of mesh size for advanced ceramics is suggested to be around 0.3 mm. In addition, experimentally determined fragment sizes for ceramics [18,19] can also guide in choosing the element size because each tetrahedral elastic element acts as a potential fragment generated in the post-fracture process in the HFDEM model. In their experimental study, Hogan et al. [18] measured more than 1500 ceramic fragments generated by quasi-static uniaxial compression. According to their measurements, most of the fragments are between 0.01 and 1 mm in size, and over 70% of the fragments are larger than 0.1 mm [19]. With this taken into account, the current study focuses on capturing the macroscopic failure (fracture and fragmentation process) numerically, therefore, we performed simulations with mesh sizes of 0.075, 0.1, 0.15, and 0.25 mm to confirm that there was minimal sensitivity. Monte Carlo simulations are carried out to simulate the variety of the specimen strength because of the different flaw distributions inherent to the samples. Three numerical tests are performed for each simulation case (fixed mesh and material parameters). The facet area of the cohesive element is A = 0.00336 mm 2 for mesh sizes of 0.075 mm, A = 0.00611 mm 2 for mesh sizes of 0.1 mm, A = 0.01268 mm 2 for mesh sizes of 0.15 mm, and A = 0.038 mm 2 for mesh sizes of 0.25 mm, which are summarized in Table 2.
The stable time step used in the current study satisfies the criteria given by [68]:
Δ t s t a b l e α l e c
where c is the wave speed in the alumina ceramic, l e is the smallest element size, and α is a factor whose value is 0.1 or less. This stable time step is able to resolve the two kinds of time-scales. The first time-scale is the response time associated with the cohesive law, and is given by [76]:
t 0 = E c G c σ 0 2
The other time-scale is associated with the time required for complete decohesion of the cohesive law [68]:
t c = δ c c ε ˙ 0.5
where ε ˙ is the applied strain rate, and δ c is the critical crack opening displacement.
Figure 3a illustrates the tensile stress and strain evolution of the alumina under quasi-static direct tension loading. The stress increases linearly during the loading process with a constant strain rate until catastrophic failure occurs. The variety of the tensile strength in the simulations with the same mesh size is due to the different flaws in the materials generated by Monte Carlo simulations. The peak stresses of the four kinds of simulation with different mesh sizes are consistent with the experimental results [29], and the difference of simulated stress-strain responses are within 7 % (see Figure 3b). Although the difference in stress (or strain) states between indirect and direct tension samples, the difficulties inherent in conducting conventional direct tensile tests on advanced ceramics have resulted in using indirect methods, such as Brazilian tests, for evaluating their tensile strength [23,27,29]. Our previous investigation [29] involved the application of an experimental approach coupled with a theoretical method to determine the tensile strength of alumina ceramic materials, which makes the obtained tensile strength more reliable.
Next, the failure pattern of the simulation with four different mesh sizes is shown in Figure 4. A horizontal crack perpendicular to the loading direction is observed during the loading process. The single main crack causes catastrophic failure at the low loading rate. Some fragments appear near the crack surfaces due to crack branching. The fracture process of simulation results under the tensile loading is consistent with the observation in brittle materials [77,78,79,80]. The origins of fractures can stem from either internal volume flaws (such as cracks, pores, uneven density, and composition variations) or surface flaws (like cracks from machining, surface pits, and voids) [80]. In the current study, we consider the tensile characteristics in the loading direction based on the statistical studies of flaws for alumina ceramics [16,17,40]. However, the flaws on the sample surfaces and the influence of the flaws on shear or compression directions are not considered. In brittle materials, fracture in brittle materials occurs due to the application of stress to the critical flaws in the material leading to unstable propagation of that cracks. When subjected to direct tension testing, the crack at the origin expands approximately perpendicular to the principal tensile stress and spreads symmetrically from the origin in a uniform stress field. The fracture processes include crack surface creation, fragment release, and the main crack splits into multiple branches [80]. In the simulation, the branching phenomenon is more pronounced for the samples with smaller mesh sizes, and this is because each cohesive element associated with flaws is a potential source of microcracking in the material. The smaller meshed sample has more cohesive elements (or crack elements) with critical strength in the material. The increase in the number of crack initiation sources naturally promotes branching behavior [77,78,79,80]. Overall, our model is mesh-independent based on qualitative (e.g., failure patterns) and quantitative (e.g., stress-strain curves) evaluations.

3.2. The Effect of Flaw Distribution

In recent years, the influences of cohesive parameters (i.g., cohesive strength and the critical energy release rate) on the mechanical response of brittle materials have been widely investigated [54,60,62,64]. However, limited numerical studies focused on the influence of the distribution of the flaw systems. The influence of the flaw systems in materials on the strength at the quasi-static loading rate has traditionally been related to the size of the largest flaw in the material based on experimental studies [68,81]. Moreover, researchers found that the distribution of the flaw systems is also significant to the elasticity and strength of ceramic materials [82,83]. In the current study, we consider the different Weibull distributions of the flaw system with five choices of Weibull modulus m 0 = 9, 10, 11, 12, and 13 [16]. Figure 5 shows that the percentage of low-strength cohesive elements (below 350 MPa) is around 12.2% for m 0 = 9, 9.9% for m 0 = 10, 7.9% for m 0 = 11, 6.4% for m 0 = 12, and 5.1% for m 0 = 13. This means the higher value of m 0 corresponds to more uniform flaw sizes with fewer big flaws, and a smaller m 0 physically represents a more heterogeneous material with a higher amount of bigger flaws.
Figure 6a shows the tensile stress-strain curves during the constant low strain-rate loading, and Figure 6b summarizes the tensile strength and elastic modulus obtained by simulation with five different Weibull modulus. It is observed that material with smaller m 0 (i.e., more heterogeneous material with a higher number of bigger flaws) exhibits lower tensile strength and elastic modulus. This agrees with the experimental observation [82,84] that both the strength limit and elasticity moduli decrease with a higher percentage of big flaws. This is because the strength of a ceramic specimen is not only determined by the existing critical flaws in the material [85] but also by the flaw distribution [16,17,40]. Next, Figure 7 shows that the failure pattern is consistent with Figure 4, for which a single main crack perpendicular to the loading direction is generated during the loading process. Some fragments appear near the crack surfaces due to crack branching. The branching phenomenon is more pronounced for the material with a smaller m 0 value. This is because the material has more critical flaws with larger sizes, and these flaws are activated during the loading process. The increase in the number of crack initiation sources naturally stimulates the branching behavior.

4. Conclusions

We performed a three-dimensional HFDEM model to investigate the mode I tensile opening failure of the alumina ceramic. In this model, the ceramic material was divided into two parts: bulk material regions, represented by tetrahedral elements, and pre-inserted cohesive elements that appear at the interfaces (facets) between the tetrahedral elements. The bulk material was linear, homogeneous, and isotropically elasticit, while the behavior of the micro-cracks was described using cohesive elements. A microscopic stochastic fracture model is developed considering a random distribution of internal flaws in the material. The microscopic stochastic fracture model, including a Weibull strength distribution, is adapted to fit within the HFDEM model. Our model implicitly considered the tensile failure processes related to the flaw system in the material and explicitly showed the macroscopic failure patterns. This model is mesh-independent based on qualitative (e.g., failure patterns) and quantitative (e.g., stress-strain curves) evaluations. The tensile strength obtained by our model is consistent with the indirect tensile testing results from our previous study [29]. In the simulation, micro-cracks nucleated randomly within the sample and grew, eventually merging, leading to the catastrophic failure of the specimen. A single main crack perpendicular to the loading direction is observed during the tensile loading process. Some fragments appear near the crack surfaces due to crack branching. Furthermore, the influences of the flaw system distribution on the tensile strength and elastic modulus are explored. The simulation results show that the material with more uniform flaw sizes and fewer big flaws has stronger tensile strength and higher elastic modulus. Overall, applying this new model provides theoretical guidance for future material design and optimization.

Author Contributions

Conceptualization, J.Z., H.L. and J.D.H.; methodology, J.Z.; software, J.Z.; validation, J.Z., H.L. and J.D.H.; formal analysis, J.Z.; investigation, J.Z.; resources, J.D.H.; writing—original draft preparation, J.Z.; writing—review and editing, H.L. and J.D.H.; visualization, J.Z.; supervision, J.D.H.; project administration, H.L.; funding acquisition, J.D.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Defence Research and Development Canada, grant number NSERC ALLRP 560447-2020. This research was funded by General Dynamics Land Systems–Canada, grant number NSERC ALLRP 560447-2020. This research was funded by NP Aerospace, grant number NSERC ALLRP 560447-2020. The funds receiver is James D. Hogan.

Data Availability Statement

Data will be made available on request.

Acknowledgments

This work is supported by Defence Research and Development Canada (DRDC), General Dynamics Land Systems–Canada, and NP Aerospace through NSERC Alliance project ALLRP 560447-2020. The views and conclusions contained in this document are those of the authors and should not be interpreted as representing the official policies, either expressed or implied, of General Dynamics, NP Aerospace, DRDC or the Government of Canada. The Government of Canada is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation herein.

Conflicts of Interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: James D. Hogan reports financial support was provided by Defence Research and Development Canada. James D. Hogan reports financial support was provided by General Dynamics Land Systems Canada. James D. Hogan reports financial support was provided by NP Aerospace.

References

  1. Swab, J.J.; Lasalvia, J.C.; Gilde, G.A.; Patel, P.J.; Motyka, M.J. Transparent armor ceramics: Alon and spinel. In Proceedings of the 23rd Annual Conference on Composites, Advanced Ceramics, Materials, and Structures: B: Ceramic Engineering and Science Proceedings; Wiley Online Library: Hoboken, NJ, USA, 1999; pp. 79–84. [Google Scholar]
  2. Willmann, G. Ceramic femoral head retrieval data. Clin. Orthop. AndRelated Res. 2000, 379, 22–28. [Google Scholar] [CrossRef]
  3. Subhash, G.; Maiti, S.; Geubelle, P.H.; Ghosh, D. Recent advances in dynamic indentation fracture, impact damage and fragmentation of ceramics. J. Am. Ceram. Soc. 2008, 91, 2777–2791. [Google Scholar] [CrossRef]
  4. Medvedovski, E. Ballistic performance of armour ceramics: Influence of design and structure. part 2. Ceram. Int. 2010, 36, 2117–2127. [Google Scholar] [CrossRef]
  5. Medvedovski, E. Ballistic performance of armour ceramics: Influence of design and structure. part 1. Ceram. Int. 2010, 36, 2103–2115. [Google Scholar] [CrossRef]
  6. Anglada, M. Assessment of Mechanical Properties of Ceramic Materials. In Advances in Ceramic Biomaterials; Elsevier: Amsterdam, The Netherlands, 2017; pp. 83–109. [Google Scholar]
  7. Suh, M.-S.; Chae, Y.-H.; Kim, S.-S. Friction and wear behavior of structural ceramics sliding against zirconia. Wear 2008, 264, 800–806. [Google Scholar] [CrossRef]
  8. Woodward, R.; Gooch Jr, W.; O’donnell, R.; Perciballi, W.; Baxter, B.; Pattie, S. A study of fragmentation in the ballistic impact of ceramics. Int. Impact Eng. 1994, 15, 605–618. [Google Scholar] [CrossRef]
  9. Kaufmann, C.; Cronin, D.; Worswick, M.; Pageau, G.; Beth, A. Influence of material properties on the ballistic performance of ceramics for personal body armour. Shock Vib. 2003, 10, 51–58. [Google Scholar] [CrossRef]
  10. Toussaint, G.; Martinez-Rubi, Y.; Noormohammed, S.; Li, C. Novel alternative multi-hit ballistic testing method developed to evaluate bonding technologies performance. Tech. rep., DRDC-RDDC-2020-R045.
  11. Arrowood, R.; Lankford, J. Compressive fracture processes in an alumina-glass composite. J. Mater. Sci. 1987, 22, 3737–3744. [Google Scholar] [CrossRef]
  12. Swab, J.J.; Wereszczak, A.A.; Strong, K.T., Jr.; Danna, D.; LaSalvia, J.C.; Ragan, M.E.; Ritt, P.J. Knoop hardness–apparent yield stress relationship in ceramics. Int. J. Appl. Ceram. Technol. 2012, 9, 650–655. [Google Scholar] [CrossRef]
  13. Hogan, J.D.; Farbaniec, L.; Sano, T.; Shaeffer, M.; Ramesh, K. The effects of defects on the uniaxial compressive strength and failure of an advanced ceramic. Acta Mater. 2016, 102, 263–272. [Google Scholar] [CrossRef]
  14. Lo, C.; Sano, T.; Hogan, J.D. Microstructural and mechanical characterization of variability in porous advanced ceramics using x-ray computed tomography and digital image correlation. Mater. Charact. 2019, 158, 109929. [Google Scholar] [CrossRef]
  15. DeVries, M.; Subhash, G. Influence of carbon nanotubes as secondary phase addition on the mechanical properties and amorphization of boron carbide. J. Eur. Ceram. Soc. 2019, 39, 1974–1983. [Google Scholar] [CrossRef]
  16. MUNRO, M. Evaluated material properties for a sintered alpha-alumina. J. Am. Ceram. Soc. 1997, 80, 1919–1928. [Google Scholar] [CrossRef]
  17. Nohut, S. Influence of sample size on strength distribution of advanced ceramics. Ceram. Int. 2014, 40, 4285–4295. [Google Scholar] [CrossRef]
  18. Hogan, J.D.; Farbaniec, L.; Shaeffer, M.; Ramesh, K. The effects of microstructure and confinement on the compressive fragmentation of an advanced ceramic. J. Am. Ceram. Soc. 2015, 98, 902–912. [Google Scholar] [CrossRef]
  19. Hogan, J.D.; Farbaniec, L.; Daphalapurkar, N.; Ramesh, K. On compressive brittle fragmentation. J. Am. Ceram. Soc. 2016, 99, 2159–2169. [Google Scholar] [CrossRef]
  20. Hogan, J.D.; Farbaniec, L.; Mallick, D.; Domnich, V.; Kuwelkar, K.; Sano, T.; McCauley, J.W.; Ramesh, K.T. Fragmentation of an advanced ceramic under ballistic impact: Mechanisms and microstructure. Int. J. Impactengineering 2017, 102, 47–54. [Google Scholar] [CrossRef]
  21. Langford, J.; Perras, M.A. Obtaining reliable estimates of intact tensile strength. In Proceedings of the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, MA, USA, 1–4 June 2014. [Google Scholar]
  22. Swab, J.J.; Tice, J.; Wereszczak, A.A.; Kraft, R.H. Fracture toughness of advanced structural ceramics: Applying astm c1421. J. Am. Soc. 2015, 98, 607–615. [Google Scholar] [CrossRef]
  23. Dai, Y.; Li, Y.; Xu, X.; Zhu, Q.; Yan, W.; Jin, S.; Harmuth, H. Fracture behaviour of magnesia refractory materials in tension with the brazilian test. J. Eur. Ceram. Soc. 2019, 39, 5433–5441. [Google Scholar] [CrossRef]
  24. Lo, C.; Li, H.; Toussaint, G.; Hogan, J.D. On the evaluation of mechanical properties and ballistic performance of two variants of boron carbide. Int. J. Impact Eng. 2021, 152, 103846. [Google Scholar] [CrossRef]
  25. Swab, J.; Chen, W.; Hogan, J.; Liao, H.; Lo, C.; Mates, S.; Meredith, C.; Pittari, J.; Rhorer, R.; Quinn, G. Dynamic compression strength of ceramics: What was learned from an interlaboratory round robin exercise? J. Dyn. Behav. Mater. 2021, 7, 34–47. [Google Scholar] [CrossRef]
  26. Subhash, G.; Awasthi, A.; Ghosh, D. Dynamic Response of Advanced Ceramics; John Wiley & Sons: Hoboken, NJ, USA, 2021. [Google Scholar]
  27. Zhang, L.; Townsend, D.; Petrinic, N.; Pellegrino, A. Loading mode and lateral confinement dependent dynamic fracture of a glass ceramic macor. J. Dynamic Behav. Mater. 2022, 8, 255–272. [Google Scholar] [CrossRef]
  28. Bavdekar, S.; Subhash, G. Failure mechanisms of ceramics under quasi-static and dynamic loads: Overview. In Handbook of Damage Mechanics: Nano to Macro Scale for Materials and Structures; Springer: Berlin/Heidelberg, Germany, 2022; pp. 579–607. [Google Scholar]
  29. Zheng, J.; Li, H.; Hogan, J.D. Strain-rate-dependent tensile response of an alumina ceramic: Experiments and modelling. Int. J. Impact Eng. 2022, 173, 104487. [Google Scholar] [CrossRef]
  30. Brace, W.; Bombolakis, E. A note on brittle crack growth in compression. J. Geophys. Res. 1963, 68, 3709–3713. [Google Scholar] [CrossRef]
  31. Bradley, W.; Kobayashi, A. An investigation of propagating cracks by dynamic photoelasticity. Exp. Mech. 1970, 10, 106–113. [Google Scholar] [CrossRef]
  32. Nemat-Nasser, S.; Horii, H. Compression-induced nonplanar crack extension with application to splitting, exfoliation, and rockburst. J. Geophys. Solid Earth 1982, 87, 6805–6821. [Google Scholar] [CrossRef]
  33. Nemat-Nasser, S.; Horii, H. Rock failure in compression. Int. J. Eng. Sci. 1984, 22, 999–1011. [Google Scholar] [CrossRef] [Green Version]
  34. Horii, H.; Nemat-Nasser, S. Compression-induced microcrack growth in brittle solids: Axial splitting and shear failure. J. Geophys. Res. Earth 1985, 90, 3105–3125. [Google Scholar] [CrossRef]
  35. Appleby-Thomas, G.J.; Jaansalu, K.; Hameed, A.; Painter, J.; Shackel, J.; Rowley, J. A comparison of the ballistic behaviour of conventionally sintered and additively manufactured alumina. Def. Technol. 2020, 16, 275–282. [Google Scholar] [CrossRef]
  36. Nieto, A.; Huang, L.; Han, Y.-H.; Schoenung, J.M. Sintering behavior of spark plasma sintered alumina with graphene nanoplatelet reinforcement. Ceram. Int. 2015, 41, 5926–5936. [Google Scholar] [CrossRef]
  37. Schwentenwein, M.; Homa, J. Additive manufacturing of dense alumina ceramics. Int. J. Appl. Ceram. Technol. 2015, 12, 1–7. [Google Scholar] [CrossRef]
  38. Biesuz, M.; Sglavo, V.M. Flash sintering of alumina: Effect of different operating conditions on densification. J. Eur. Ceram. 2016, 36, 2535–2542. [Google Scholar] [CrossRef]
  39. Andreev, G. A review of the brazilian test for rock tensile strength determination. part i: Calculation formula. Min. Sci. Technol. 1991, 13, 445–456. [Google Scholar] [CrossRef]
  40. Graham-Brady, L. Statistical characterization of meso-scale uniaxial compressive strength in brittle materials with randomly occurring flaws. Int. J. Solids Struct. 2010, 47, 2398–2413. [Google Scholar] [CrossRef] [Green Version]
  41. Hu, G.; Liu, J.; Graham-Brady, L.; Ramesh, K. A 3D mechanistic model for brittle materials containing evolving flaw distributions under dynamic multiaxial loading. J. Mech. Phys. Solids 2015, 78, 269–297. [Google Scholar] [CrossRef]
  42. Ahmadi, M.H.; Molladavoodi, H. A micromechanical sliding-damage model under dynamic compressive loading. Period. Polytech. Civ. Eng. 2019, 63, 168–183. [Google Scholar] [CrossRef]
  43. Moës, N.; Gravouil, A.; Belytschko, T. Non-planar 3d crack growth by the extended finite element and level sets—Part i: Mechanical model. Int. J. Numer. Methods Eng. 2002, 53, 2549–2568. [Google Scholar] [CrossRef]
  44. Sun, L.; Ma, D.; Wang, L.; Shi, X.; Wang, J.; Chen, W. Determining indentation fracture toughness of ceramics by finite element method using virtual crack closure technique. Eng. Fract. Mech. 2018, 197, 151–159. [Google Scholar] [CrossRef]
  45. Wang, H.; Dong, H.; Cai, Z.; Li, Y.; Wang, W.; Liu, Y. Peridynamic-based investigation of the cracking behavior of multilayer thermal barrier coatings. Ceram. Int. 2022, 48, 23543–23553. [Google Scholar] [CrossRef]
  46. Cao, P.; Tang, C.; Liu, Z.; Shi, F.; Wang, Z.; Wang, X.; Zhou, C. Study on fracture behaviour of basalt fibre reinforced asphalt concrete with plastic coupled cohesive model and enhanced virtual crack closure technique model. Int. J. Pavement Eng. 2022, 384, 135585. [Google Scholar] [CrossRef]
  47. Falk, M.L.; Needleman, A.; Rice, J.R. A critical evaluation of cohesive zone models of dynamic fractur. Le J. Phys. IV 2001, 11, Pr5-43–Pr5-50. [Google Scholar] [CrossRef]
  48. Munjiza, A.A. The Combined Finite-Discrete Element Method; John Wiley & Sons: Hoboken, NJ, USA, 2004. [Google Scholar]
  49. Williams, C.; Love, B. Dynamic Failure of Materials: A Review. Tech. rep., Army Research Lab Aberdeen Proving Ground Md Weapons and Materials Research 2010-ADA528764. Available online: https://apps.dtic.mil/sti/pdfs/ADA528764.pdf (accessed on 1 February 2023).
  50. Rena, C.Y.; Ruiz, G.; Pandolfi, A. Numerical investigation on the dynamic behavior of advanced ceramics. Eng. Fract. Mech. 2004, 71, 897–911. [Google Scholar]
  51. Zhou, F.; Molinari, J.-F.; Ramesh, K. A cohesive model based fragmentation analysis: Effects of strain rate and initial defects distribution. Int. J. Solids Struct. 2005, 42, 5181–5207. [Google Scholar] [CrossRef]
  52. Arias, I.; Knap, J.; Chalivendra, V.B.; Hong, S.; Ortiz, M.; Rosakis, A.J. Numerical modelling and experimental validation of dynamic fracture events along weak planes. Comput. Methods Appl. Mech. Eng. 2007, 196, 3833–3840. [Google Scholar] [CrossRef] [Green Version]
  53. Espinosa, H.D.; Zavattieri, P.D.; Dwivedi, S.K. A finite deformation continuum∖discrete model for the description of fragmentation and damage in brittle materials. J. Mech. Phys. Solids 1998, 46, 1909–1942. [Google Scholar] [CrossRef]
  54. Zhou, W.; Yuan, W.; Chang, X. Combined finite-discrete element method modeling of rock failure problems. In International Conference on Discrete Element Methods; Springer: Berlin/Heidelberg, Germany, 2016; pp. 301–309. [Google Scholar]
  55. An, H.; Shi, J.; Zheng, X.; Wang, X. Hybrid finite-discrete element method modelling of brazilian disc tests. In Proceedings of the 2016 International Conference on Civil, Transportation and Environment, Guangzhou, China, 30–31 January 2016; Atlantis Press: Paris, France, 2016; pp. 371–377. [Google Scholar]
  56. Jaini, Z.M.; Feng, Y.T.; Mokhatar, S.N.; Seman, M.A. Numerical homogenization of protective ceramic composite layers using the hybrid finite-discrete element methods. Int. J. Integr. Eng. 2013, 5, 15–22. [Google Scholar]
  57. Mahabadi, O.; Lisjak, A.; Munjiza, A.; Grasselli, G. Y-geo: New combined finite-discrete element numerical code for geomechanical applications. Int. J. Geomech. 2012, 12, 676–688. [Google Scholar] [CrossRef]
  58. Liu, H.; Kang, Y.; Lin, P. Hybrid finite–discrete element modeling of geomaterials fracture and fragment muck-piling. Int. J. Geotech. Eng. 2015, 9, 115–131. [Google Scholar] [CrossRef]
  59. Miglietta, P.C. Application of the Hybrid Finite/Discrete Element Method to the Numerical Simulation of Masonry Structures. Ph.D. Thesis, University of Toronto, Toronto, ON, Canada, 2017. [Google Scholar]
  60. Mahabadi, O.; Kaifosh, P.; Marschall, P.; Vietor, T. Three-dimensional fdem numerical simulation of failure processes observed in opalinus clay laboratory samples. J. Rock Mech. Geotech. Eng. 2014, 6, 591–606. [Google Scholar] [CrossRef] [Green Version]
  61. Camacho, G.T.; Ortiz, M. Computational modelling of impact damage in brittle materials. Int. J. Solids Struct. 1996, 33, 2899–2938. [Google Scholar] [CrossRef]
  62. Warner, D.; Molinari, J. Micromechanical finite element modeling of compressive fracture in confined alumina ceramic. Acta Mater. 2006, 54, 5135–5145. [Google Scholar] [CrossRef]
  63. Lisjak, A.; Figi, D.; Grasselli, G. Fracture development around deep underground excavations: Insights from fdem modelling. J. Rock Mech. And Geotech. Eng. 2014, 6, 493–505. [Google Scholar] [CrossRef] [Green Version]
  64. Lisjak, A.; Liu, Q.; Zhao, Q.; Mahabadi, O.; Grasselli, G. Numerical simulation of acoustic emission in brittle rocks by two-dimensional finite-discrete element analysis. Geophys. J. Int. 2013, 195, 423–443. [Google Scholar] [CrossRef] [Green Version]
  65. Walter, M.; Owen, D.; Ravichandran, G. Static and dynamic tension testing of ceramics and ceramic composites. Asme Appl. Mech.-Publ. 1994, 197, 13. [Google Scholar]
  66. Wang, Z.; Li, P. Voronoi cell finite element modelling of the intergranular fracture mechanism in polycrystalline alumina. Ceram. Int. 2017, 43, 6967–6975. [Google Scholar] [CrossRef] [Green Version]
  67. Zhou, F.; Molinari, J.-F. Stochastic fracture of ceramics under dynamic tensile loading. Int. J. Solids Struct. 2004, 41, 6573–6596. [Google Scholar] [CrossRef]
  68. Daphalapurkar, N.P.; Ramesh, K.; Graham-Brady, L.; Molinari, J.-F. Predicting variability in the dynamic failure strength of brittle materials considering pre-existing flaws. J. Mech. Phys. Solids 2011, 59, 297–319. [Google Scholar] [CrossRef]
  69. Zou, Z.; Reid, S.; Li, S. A continuum damage model for delaminations in laminated composites. J. Mech. Phys. Solids 2003, 51, 333–356. [Google Scholar] [CrossRef]
  70. Tikare, V.; Choi, S.R. Combined mode i and mode ii fracture of monolithic ceramics. J. Am. Ceram. Soc. 1993, 76, 2265–2272. [Google Scholar] [CrossRef]
  71. Shetty, D.; Rosenfield, A.; Duckworth, W. Mixed-mode fracture of ceramics in diametral compression. J. Am. Ceram. Soc. 1986, 69, 437–443. [Google Scholar] [CrossRef]
  72. Bazant, Z.P.; Planas, J. Fracture and Size Effect in Concrete and Other Quasibrittle Materials; CRC Press: Boca Raton, FL, USA, 1997; Volume 16. [Google Scholar]
  73. KMW, K. Ceramic Materials for Light-Weight Ceramic Polymer Armor Systems. Available online: https://www.ceramtec.com/files/et_armor_systems.pdf (accessed on 1 February 2023).
  74. Zheng, J.; Ji, M.; Zaiemyekeh, Z.; Li, H.; Hogan, J.D. Strain-rate-dependent compressive and compression-shear response of an alumina ceramic. J. Eur. Ceram. Soc. 2022, 42, 7516–7527. [Google Scholar] [CrossRef]
  75. Ji, M.; Li, H.; Zheng, J.; Yang, S.; Zaiemyekeh, Z.; Hogan, J.D. An experimental study on the strain-rate-dependent compressive and tensile response of an alumina ceramic. Ceram. Int. 2022, 48, 28121–28134. [Google Scholar] [CrossRef]
  76. Molinari, J.-F.; Gazonas, G.; Raghupathy, R.; Rusinek, A.; Zhou, F. The cohesive element approach to dynamic fragmentation: The question of energy convergence. Int. J. Numer. Methods Eng. 2007, 69, 484–503. [Google Scholar] [CrossRef] [Green Version]
  77. Zhou, F.; Molinari, J.-F. Dynamic crack propagation with cohesive elements: A methodology to address mesh dependency. Int. J. Numer. Eng. 2004, 59, 1–24. [Google Scholar] [CrossRef]
  78. Heizler, S.I.; Kessler, D.A.; Levine, H. Propagating mode-i fracture in amorphous materials using the continuous random network model. Phys. Rev. E 2011, 84, 026102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Ren, X.; Li, J. Dynamic fracture in irregularly structured systems. Phys. Rev. E 2012, 85, 055102. [Google Scholar] [CrossRef] [PubMed]
  80. Freiman, S.W.; Mecholsky, J.J., Jr. The Fracture of Brittle Materials: Testing and Analysis; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  81. Weibull, W. A Statistical Theory of Strength of Materials; Generalstabens Litografiska Anstalts Förlag: Stockholm, Sweden, 1939. [Google Scholar]
  82. Savchenko, N.; Sevostyanova, I.; Sablina, T.; Gömze, L.; Kulkov, S. The influence of porosity on the elasticity and strength of alumina and zirconia ceramics. In AIP Conference Proceedings; American Institute of Physics: College Park, MD, USA, 2014; Volume 1623, pp. 547–550. [Google Scholar]
  83. Sapozhnikov, S.; Kudryavtsev, O.; Dolganina, N.Y. Experimental and numerical estimation of strength and fragmentation of different porosity alumina ceramics. Mater. Des. 2015, 88, 1042–1048. [Google Scholar] [CrossRef]
  84. Hristopulos, D.T.; Demertzi, M. A semi-analytical equation for the young’s modulus of isotropic ceramic materials. J. Eur. Ceram. 2008, 28, 1111–1120. [Google Scholar] [CrossRef]
  85. Paliwal, B.; Ramesh, K. An interacting micro-crack damage model for failure of brittle materials under compression. J. Mech. Phys. Solids 2008, 56, 896–923. [Google Scholar] [CrossRef]
Figure 1. The mode I constitutive behavior of the cohesive element with K 1 = 4.6 × 10 7 N/mm 3 , σ 0 = 440 MPa, and G 1 c = 0.04 N/mm. In the inset figure, a cohesive (crack) element is interspersed throughout two tetrahedral elements.
Figure 1. The mode I constitutive behavior of the cohesive element with K 1 = 4.6 × 10 7 N/mm 3 , σ 0 = 440 MPa, and G 1 c = 0.04 N/mm. In the inset figure, a cohesive (crack) element is interspersed throughout two tetrahedral elements.
Modelling 04 00007 g001
Figure 2. The green line is Weibull’s statistical strength distribution of cohesive elements obtained from Equation (8) with m 0 = 11.0 , m a = 11.0 , σ 0 = 440.0 , A = 0.01268 and A 0 = 0.013 . The orange bar is the statistics of the facet strength of the cohesive element with random flaws generated by Monte Carlo simulations. The percentage of low-strength cohesive elements (below 350 MPa) is around 7.9%, which is associated with the big flaws in the material. The rest of the cohesive elements (around 92.1%) have strong strength (between 350 and 530 MPa), which corresponds to the material with smaller flaws.
Figure 2. The green line is Weibull’s statistical strength distribution of cohesive elements obtained from Equation (8) with m 0 = 11.0 , m a = 11.0 , σ 0 = 440.0 , A = 0.01268 and A 0 = 0.013 . The orange bar is the statistics of the facet strength of the cohesive element with random flaws generated by Monte Carlo simulations. The percentage of low-strength cohesive elements (below 350 MPa) is around 7.9%, which is associated with the big flaws in the material. The rest of the cohesive elements (around 92.1%) have strong strength (between 350 and 530 MPa), which corresponds to the material with smaller flaws.
Modelling 04 00007 g002
Figure 3. (a) Variation of engineering stress-strain response of the CeramTec 98% alumina during the direct tension simulations with four kinds of mesh sizes. (b) The shaded region is the tensile strength of the CeramTec 98% obtained by Brazilian disk experiments [29], and the dots are the simulation results with four kinds of mesh sizes.
Figure 3. (a) Variation of engineering stress-strain response of the CeramTec 98% alumina during the direct tension simulations with four kinds of mesh sizes. (b) The shaded region is the tensile strength of the CeramTec 98% obtained by Brazilian disk experiments [29], and the dots are the simulation results with four kinds of mesh sizes.
Modelling 04 00007 g003
Figure 4. The failure pattern of the samples with different mesh sizes. The legends in the figure correspond to the displacement of the samples in the z-direction (U 3 ). It is observed that a horizontal crack perpendicular to the loading direction is generated during the loading process. The single main crack causes catastrophic failure. In some cases, the main crack may split into multiple branches. Some fragments appear near the crack surfaces due to crack branching.
Figure 4. The failure pattern of the samples with different mesh sizes. The legends in the figure correspond to the displacement of the samples in the z-direction (U 3 ). It is observed that a horizontal crack perpendicular to the loading direction is generated during the loading process. The single main crack causes catastrophic failure. In some cases, the main crack may split into multiple branches. Some fragments appear near the crack surfaces due to crack branching.
Modelling 04 00007 g004
Figure 5. Microscopic strength distributions with different Weibull modulus ( m 0 = 9, 10, 11, 12, and 13).
Figure 5. Microscopic strength distributions with different Weibull modulus ( m 0 = 9, 10, 11, 12, and 13).
Modelling 04 00007 g005
Figure 6. (a) The engineering stress-strain response of the CeramTec 98% alumina during the direct tension simulations with different Weibull modulus ( m 0 = 9, 10, 11, 12, and 13). (b) The full dots are the tensile strength and the hollow dots are the elastic modulus obtained by simulation with five different Weibull modulus.
Figure 6. (a) The engineering stress-strain response of the CeramTec 98% alumina during the direct tension simulations with different Weibull modulus ( m 0 = 9, 10, 11, 12, and 13). (b) The full dots are the tensile strength and the hollow dots are the elastic modulus obtained by simulation with five different Weibull modulus.
Modelling 04 00007 g006
Figure 7. The failure pattern of the simulation results with different Weibull modulus ( m 0 = 9, 10, 11, 12, and 13). The U 3 legend in the figure indicates the displacement of the samples in the z-direction. The failure pattern is consistent with Figure 4, a horizontal crack forms perpendicularly to the loading direction during the loading process, ultimately leading to a catastrophic failure caused by the single main crack. In certain instances, the main crack can divide into multiple branches, resulting in fragments near the crack surfaces due to crack branching.
Figure 7. The failure pattern of the simulation results with different Weibull modulus ( m 0 = 9, 10, 11, 12, and 13). The U 3 legend in the figure indicates the displacement of the samples in the z-direction. The failure pattern is consistent with Figure 4, a horizontal crack forms perpendicularly to the loading direction during the loading process, ultimately leading to a catastrophic failure caused by the single main crack. In certain instances, the main crack can divide into multiple branches, resulting in fragments near the crack surfaces due to crack branching.
Modelling 04 00007 g007
Table 1. The properties of CeramTec 98% alumina for the HFDEM model.
Table 1. The properties of CeramTec 98% alumina for the HFDEM model.
The Mechanical Properties of the CeramTec 98% Alumina
Porosity<2%
Hardness13.5 (GPa)
Density3.8 (g/cm 3 )
Young’s modulus335 (GPa)
Poisson’s ratio0.23
The Properties for the Microscopic Stochastic Fracture Model
Weibull modulus of the strength distribution11
Weibull modulus for the effective area modification−11
Weibull characteristic strength440 (MPa)
Characteristic area0.013 (mm 2 )
Mode I fracture energy0.04 (N/mm)
Table 2. Statistics of four FEM meshes.
Table 2. Statistics of four FEM meshes.
Validation ModelsAverage Mesh Size (mm)Average Facet Area (mm 2 )
Case 10.250.038
Case 20.150.01268
Case 30.10.00611
Case 40.0750.00336
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

Zheng, J.; Li, H.; Hogan, J.D. Hybrid Finite-Discrete Element Modeling of the Mode I Tensile Response of an Alumina Ceramic. Modelling 2023, 4, 87-101. https://doi.org/10.3390/modelling4010007

AMA Style

Zheng J, Li H, Hogan JD. Hybrid Finite-Discrete Element Modeling of the Mode I Tensile Response of an Alumina Ceramic. Modelling. 2023; 4(1):87-101. https://doi.org/10.3390/modelling4010007

Chicago/Turabian Style

Zheng, Jie, Haoyang Li, and James D. Hogan. 2023. "Hybrid Finite-Discrete Element Modeling of the Mode I Tensile Response of an Alumina Ceramic" Modelling 4, no. 1: 87-101. https://doi.org/10.3390/modelling4010007

Article Metrics

Back to TopTop