1. Introduction
The 2D or 3D finite element simulations of Representative Volume Elements (RVEs) are increasingly popular as they help the researchers to understand the strength mechanisms within the loaded materials. A careful design of the RVE is crucial to ensure an accurate and reliable representation of material microstructures, capable of effectively predicting macro-scale properties, as demonstrated by significant validation with experimental data.
As pointed in reference [
1], the RVE methodology encompasses two primary approaches: statistically modeling the microstructure as a virtual representative volume [
2], or discretizing a statically representative real volume [
3]. Based on microstructure features such as spherical particles, randomly distributed needle-like precipitates in a matrix, isotropic polycrystalline material, or strong anisotropy due to texture or grain shape, the choice between 2D and 3D RVEs is required to investigate the mechanical response under various loading conditions. The main focus of this work is to explore an alternative to 3D RVEs, also called 2.5D RVEs, in order to save the CPU time, particularly in the context of isotropic materials featuring a soft matrix containing hard particles. The 2.5D RVE approach is defined by a one-plane layer of 3D elements and relies on specific constraints in the out-of-plane direction to model any isotropic material presenting a matrix with particles. Subsequently, this introduction discusses the range of simulation goals achievable with RVEs, exemplified through several metal-based examples.
The quality of RVE results explains their use in virtual material and process design, as RVEs can predict material properties based on a given microstructure. They are now integrated into materials engineering to enhance material design. For instance, Maity et al. [
3] use 2D RVEs to understand the effect of the addition of Mn in the Al-12.6 Si alloy on bulk hardness, yield stress, and fracture propagation. Indeed, Mn presence varies the microstructure morphology and the micromechanical response of the alloy. Using 3D RVEs, Shalimov and Tashkinov [
4] demonstrate that for cell porous gold crystals, the random morphology can play an as important role in the mechanical tensile curve as the pore fraction. Akbari et al. [
5] study the link between the yield surface of a polycrystalline brittle material and its microstructure by using 2D RVEs, including cohesive elements to model the grain interface behavior. Sun and Jain [
6] simulate the AA7075 elastic behavior based on its complex microstructure involving irregularly shaped Al
3Fe particles, elliptical MgZn
2 ones, and needle-like CuAl
2 ones. Their 3D RVE models accurately estimate the effective elastic properties of the AA7075-O sheet. Reis et al. [
7] use 2D RVEs to study heterogeneous ductile materials, employing non-local formulations to predict the damage localization path. As pointed out by Gillner and Münstermann [
8] while working on 2D RVEs of a ferrite pearlite steel, the RVE results can decrease the cost of experimental campaigns by identifying the target microstructure able to increase fatigue lifetime. The above examples confirm the use and the interest in both 2D and 3D RVEs. Based on statistical data analysis, a systematic comparison of local results has been carried out by Qayyum et al. [
9] to conclude that the 3D RVEs provide better quantitative results than experiments, whereas the 2D RVEs already provide appropriate qualitative information about the damage initiation sites as well as an accurate macroscopic stress–strain response for a small and medium plastic range, which is often sufficient for many practical applications. In a pragmatic way, based on [
10] and their own previous study [
11], Qayyum et al. [
9] suggest to generalize that damage initiation in 2D RVEs occurs for a global strain of 6% earlier than the damage initiation in 3D RVEs for Dual-Phase (DP) steel. Indeed, 2D RVE models can definitively be exploited to understand hardening or damage mechanisms, as shown in [
12], as well as study the effect of a ferrite–martensite interface.
In materials science, the machine learning approach is currently applied for material design. See, for instance, the DP steel design through 3D RVE simulations [
13], where an accurate use of 2D RVEs could significantly reduce computational costs, which is crucial when generating the large datasets needed to train machine learning models. This efficiency comes from the simpler geometry and numerical requirements of 2D simulations compared to 3D ones, allowing for a quicker exploration of a wider range of material behaviors and parameters. Furthermore, 2D RVEs make it easier to visualize and interpret the material microstructure and its properties, which is beneficial during the early stages of a material design (process parameter optimization as well as post-processing operations). These observations explain our efforts to quantify the accuracy of a 2D RVE versus a 3D one for matrix–precipitate material and to develop a specific 2.5D RVE.
Laser Powder Bed Fusion (L-PBF) produces materials with high strength [
14,
15], whereas Friction Stir Processing (FSP) reduces the out-of-equilibrium microstructural state without causing excessive softening [
16]. A new FSP microstructure has the added benefit of closing porosities, which positively impacts ductility and fatigue behavior within the single FSP pass post-processed material [
17]. In the present article, an L-PBF material post-processed by FSP was chosen as the case study. The as-built L-PBF AlSi10Mg material investigated has undergone testing in two orthogonal tensile directions [
18] and has been extensively characterized in previous research [
14,
18]. Its microstructure is defined by an interconnected Si network forming cells of different sizes according to their position within each melt pool zone. Post-processed by FSP [
15,
17], the Si network is globularized into Si particles. The alloy forms an aluminum matrix called the α phase (α-Al) containing Si-rich precipitates which exhibit macroscopic isotropic behavior. As pointed out by different authors [
19,
20], combining FSP with additive manufacturing (AM) on an industrial scale presents both opportunities and challenges. FSP can significantly improve the mechanical properties of AM parts by refining microstructures and addressing defects like porosity and anisotropy, making it valuable for industries, such as aerospace and automotive, where high material performance is crucial. However, large-scale implementation still requires overcoming practical challenges, particularly the integration of FSP into highly automated AM workflows. This would involve developing efficient, high-throughput systems to keep up with production rates, and addressing issues like tool wear and process control, which can be influenced by part geometries. Recent research has been increasingly focused on combining AM and FSP to leverage the strengths of both processes [
19,
20,
21].
Hereafter, we investigate the ability of 2D and 3D RVE simulations to accurately predict the measured tensile hardening curve of a typical Al alloy microstructure, formed by L-PBF and post-processed by FSP. In our simulations, currently the absence of cohesive elements limits the detailed analysis of fracture mechanisms, a topic we intend to address in future research as well as the fatigue behavior. As demonstrated by Yuan et al. [
22], who validated their plane strain RVE results with a tensile test before addressing fatigue prediction, the accuracy of 2D and 3D RVEs is hereafter checked for the macroscopic tensile stress–strain curve.
To define an RVE for this L-PBF FSP AlSi10Mg material, it is necessary to determine its size, its finite element mesh, and the applied boundary conditions [
23]. These choices are not independent as periodic boundary conditions allow reducing the size of the RVE volume. The selected mesh generator has to handle both geometric and boundary condition periodicity [
24,
25]. Generating RVE meshes with periodic conditions and flexible refinement for the matrix and particles is a complex task. Specific tools like Neper [
26], Digimat [
27], and GMSH v 4.11.0 [
28] have been investigated for this purpose. Hereafter, it has been decided to develop scripts based on GMSH v 4.11.0 to build the desired microstructure mesh, since GMSH v 4.11.0 is quite flexible and comes with a well-documented Python 3.8 interface. This capability enables the creation of intricate scripts, including algorithms that can reconstruct meshes using only quadrangle elements [
29,
30]. Additionally, features such as imposing periodicity conditions on the mesh and handling complex geometric operations such as cuts, fusions, and intersections between geometrical entities make it straightforward to use with support from the OpenCascade library [
31]. In the ULiege Lagamine FEM software [
32], to keep accuracy, it was chosen to mesh the 2D cut of the L-PBF FSP AlSi10Mg microstructure using only quadrangle elements to avoid any locking.
Figure 1 summarizes the procedure described in this article to build a 2.5D RVE: a one-plane layer of 3D finite elements relying on specific constraints in the out-of-plane direction. This approach is able to model any isotropic material presenting a matrix with particles. It keeps the low computation time of a 2D RVE compared with a 3D RVE.
The structure of this article is as follows:
Section 2 presents the microstructure data collected for designing the RVEs, along with the experimental tensile curve.
Section 3 summarizes the various 2.5D RVE models tested under tension, including their mesh size, specific boundary conditions tailored to mimic a 3D behavior, and the material constitutive laws employed.
Section 4 analyzes the predictions obtained from the 2.5D RVE simulations versus experimental results, highlighting their superiority over simpler 2D cases such as a simple membrane in a plane strain for instance.
Section 5 compares all the 2D and 3D RVE results with experimental data, while conclusions and future perspectives are discussed in
Section 6.
2. Material Description
The material under investigation is an AlSi10Mg alloy manufactured by L-PBF and post-processed with a single pass of FSP. The process parameters employed to fabricate the samples in this study are detailed in Dedry et al. [
33].
Figure 2a shows a typical microstructure observed by SEM [
15] while its post-processing by “ImageJ 1.52a” allows a clear observation of the microstructure composed of globularized Si-rich particles within an α-Al solid solution (see
Figure 2b). The studied zone of 11.4 × 7.7 = 87.78 µm
2 contains 157 Si hard particles embedded in the α-Al matrix. The smallest Si particles, i.e., area less than 0.011 µm
2 (
Figure 2b), were not considered in this amount, as they do not affect the mechanical behavior of the RVE. This conclusion was derived from the sensitivity analysis of FE simulations [
34] and from more than 100 experimental nanoindentations [
33].
Each particle is described by several parameters: the equivalent diameter of a circle with the same area Ø
eq, the equivalent ellipse with its aspect ratio
AR defined by Equation (1) varying between 0.247 and 0.941, and the particle orientation defined by the angle
α between the horizontal axis of the image and the major axis of the ellipse. The
AR is given as follows:
where 2
b and 2
a are the lengths of the minor and major axes of the equivalent ellipse. The distribution of the particles with respect to their Ø
eq is in the range of [0.118; 0.695] µm with a higher rate of small particles. The
AR analysis shows few highly elongated equivalent ellipses (
AR < 0.4). The angles
α show no strong dominant direction. A detailed description is available in the
Supplementary Material. The well-known Pearson correlation coefficient
r used to determine the degree of linear correlation between two variables is recalled in Equation (2):
If
m and
n are the variables,
cov(
m,
n) is their covariance, and
σm and
σn are their standard deviations, respectively. No correlation exists when
rm,n = 0, while a perfect negative or positive correlation is found when
rm,n = −1 or +1, respectively. Here, the
r coefficients (
Table 1), computed for each pair of geometric parameters describing the Si particles, indicate a moderate (almost weak) correlation between
AR and Ø
eq (
Figure 3a) and a weak correlation between both
α and Ø
eq (
Figure 3b) and
α and
AR (
Figure 4a). A corrected Pearson coefficient value is provided for the angle correlation, due to the periodic context. These microstructural statistical observations are used to generate the RVE models. If the blue dots in
Figure 3 and
Figure 4 illustrate the low relationships between the particle size, shape, and orientation observed in the SEM image, the black bullets correspond to the 10 particles of the medium-B RVE model (see
Section 3.1).
In
Figure 4b, the tensile test results of specimens, cut perpendicularly to the L-PBF building direction and parallel to the FSP tool displacement direction, are depicted (see process and test details in Zhao et al. [
15] and Santos Macias et al. [
17]). Very low experimental result scattering is observed, and the average curve will be henceforth used as a reference for the macroscopic material tensile behavior.
4. Law Identification of 2.5D RVE Model and 2D RVE Versus 2.5D RVE Results
The constitutive laws used for each phase, the material dataset, and the interface behavior between the Al and Si phases are key ingredients for the quality of the RVE results. In previous studies [
33,
34], the Young modulus of particles and the material behavior for each phase were separately measured by nanoindentation tests. Both Poisson ratios were estimated to be equal to 0.3, consistent with the values found in the literature. The numerical simulation of a macro tensile test with the medium-B RVE model and a fine mesh (model shown in
Figure 7c with 1881 elements/µm
2 and a total of 10,519 brick elements) is used to adjust the Voce material parameter and the numerical parameters
,
of the model by the inverse method.
More specifically, the data of the constitutive law of the matrix (
,
K, and
n in
Table 5) are identified by minimizing the difference between the numerical prediction and the average measured macro tensile curve (see curve “Exp-aver” in
Figure 9a). These parameters, close to those determined from the indentation measurements [
33,
34], allow recovering the experimental stress–strain curve (see
Figure 9a). The spring stiffnesses values (
,
) of the 2.5D model are adjusted by reducing the deviation between the deformations
and
over all the elements of the model (
Table 6). This result is used to compute the macro stiffnesses values
and
, through Equations (4) and (5). Those generic values are exploited to compute the spring stiffnesses of all the other models (small, medium, large) and all the mesh densities. Note that the applied α-Al matrix behavior identification methodology required less experimental data than the former approach based on indentation experiments.
Figure 9a also shows the behavior of the two phases present in the composite material (α-Al matrix and Si-rich particles). The 2.5D RVE developed in this research is compared in
Figure 9b with two 2D RVEs: a classical membrane one, in the plane strain state, and a 2D RVE called “Free in Z” where the out-of-plane springs are disabled, which behaves like a thin membrane composed of 3D elements. These results show the effect of the out-of-plane boundary conditions on the predicted stress–strain curves. The plane strain state is far too stiff to correctly predict the experimental behavior of the material, while the 2.5D RVE and 2D Free in Z predictions and experiments overlap. Although the difference between the 2.5D and “Free in Z” for macro predictions seems negligible, the local distribution of stress still needs to be analyzed. Looking at the AA cross-section passing through several Si particles and passing through a zone where the deformations are high (
Figure 10a), one can easily verify that the distributions of strains and stresses are indeed close in the X and in Z directions in the 2.5D RVE model (
Figure 10b,c) for Si particles and for the α-Al matrix. The predictions the local stress components XX and ZZ,
and
respectively, along cut AA of the two models, “Free in Z” and 2.5D RVEs, are further investigated in
Figure 11 and
Figure 12. These plots underline the effect of the specific boundary conditions on these stress components. The stress distribution of the 2.5D RVE model is closer to the expected physical isotropic behavior.
The decrease in the difference between
and
in the 2.5D RVE model compared to the 2D RVE “free in Z” case is further quantified in absolute value in
Figure 12a for cut AA, while the stress evolution in the E element during the tensile simulation (see element E position in
Figure 10a) confirms the improvement of the “isotropic” behavior in an average sense in
Figure 12b for the 2.5D RVE compared to the “Free in Z” RVE model. Indeed, due to the material macroscopic isotropy, the Y tensile load should not affect the Z and X stress and strain fields differently on average. To quantify the improvement of the stress state in the 2.5D RVE model, Equation (6) computes the average difference Δ between the two stress components
and
, for a macro strain of 10% in the Y direction, considering all the elements along cut AA and throughout the entire model. Three models are compared: the membrane in the plane strain state, the “Free in Z” model, and the 2.5D RVE. All the values pertain to the medium-B model and use the same fine mesh.
Table 7 shows that the 2.5D model better describes the behavior of the material thanks to the decrease in the Δ value.
5. Result Analysis and 2D/3D Validation
All the 2.5D RVE models (5, 10, and 15 elliptical Si-rich particles with different particle distributions, different mesh sizes, and node densities) and 3D RVE models (5, 10, 15, and 20 elliptical Si-rich particles and different mesh sizes) presented in
Table 2 and
Table 3 are used for a sensitivity and convergence analysis. A single input dataset (out-of-plane macro stiffness parameters
KAl and
KSi (
Table 6) and material data of
Table 5) is applied within all these simulations. A uniaxial tensile test up to a macro strain of 10% is modeled, and the FEM simulation results of a “20 particles_3D RVE refined mesh” as well as a “10 particles_2.5D RVE Medium B mesh” are compared, with the experiment in
Figure 13a showing good accuracy.
The macro-(true) stresses, calculated as the applied force associated with an imposed displacement corresponding to a true strain of 10%, divided by the average cross-sectional area relative to each RVE case and mesh density, are compared in
Figure 13b. From a macroscopic point of view, all the 2.5D RVE models (small, medium-A, medium-B, large) are quite similar. They converge to a single value, slightly underestimating the experimental average stress, while the scattering between the 3D RVE results is a little larger, and the 3D RVE results present an overestimation of the experiment.
For the 2.5D RVE and microstructure studied, a fine mesh with around 1900 elements per µm
2 can accurately predict the macro stress behavior. A logic curve convergence based on mesh densities and RVE sizes is observed, and the impact of particle distribution is quantified between medium-A and -B simulations, confirming that 10 particles is enough in 2D RVE simulations to have a reliable answer. Note that the medium-B model with 10 particles, 1881 elements per µm
2, and named “Ref. case” in
Figure 13 (hollow circle dot) was the one used to identify the behavior of the α-Al matrix material (
Table 5) and to analyze the local stress and strain fields.
In 3D simulations, a complete convergence analysis on mesh densities and RVE sizes was not performed; however, refining the mesh decreases the macro stress computed at a 10% strain and increases the accuracy. This indicates that a coarse mesh cannot adequately handle the high strain and stress gradients near the matrix–particle interface due to the material stiffness differences.
One might argue that the relative position of the results for the different number of particles (also defining the RVE size) in 2D or 3D RVEs, as shown in
Figure 13b, appears erratic beyond the positive effect of mesh refinement. However, the authors believe that this just confirms that all these RVE sizes are reliable for predicting the macroscopic curve, provided the mesh refinement is correct (see
Figure 14). The results also show a difference between a 2.5D RVE and 3D RVE, justifying the effort to consider boundary conditions in the transverse direction.
Figure 14 indicates that using 5 to 20 particles in a 3D RVE still results in some scattering like the physical experiments. While larger RVEs would reduce this scattering, this investigation was not performed as the main goal of the article is to save computational costs and develop a 2.5D RVE approach.
Figure 14 shows that the average responses of 2.5D RVE and 3D RVE predictions fall within or close to the scattering observed in experimental tests (4%). Even though the 3D RVE demonstrates a better accuracy, the 2.5D RVE maintains an error of less than 4% and remains within the standard experimental deviation.
The result scattering between the 2.5D and 3D RVE models is due to the high sensitivity of the maximum local stress to the relative position of the particles. The small model with five particles predicts a lower maximum stress due to its low number of particles unable to represent all the possible interactions between the particles and matrix as in the real material. On the contrary, when the number of particles increases up to 10, like in the medium-A, the medium-B, or for the 15 particle models in the 2.5 D RVE, the gap between the maximum stresses in the simulations is reduced. The same observation is also found for the maximum stresses in the transverse and in the out-of-plane directions or in the 3D RVE models.
Experimental material isotropy is validated in the 2.5D RVE for 10 particles and for 3D RVE for 15 and 20 particle cases. A minimum number of particles is indeed required to correctly model the material behavior. For the 2.5D RVE,
Figure 15 and
Figure 16 plot the strain and the stress fields, respectively, according to the directions X (a) and Z (b) at the end of the simulation for a macro strain of 10% and for the “Ref. case”. These figures confirm that similar stress and strain states are observed in both the X and Z directions and that, of course for a Y tensile direction, the internal transversal stress is very low only due to the differences in matrix and particle strengths. The internal total transversal strains have to account for the elastic volume changes and the plastic deformation heterogeneity due to the particles. The equal local strain fields in the X and Z directions in any cut of the 2.5D RVE are far from being exactly reached, as the macroscopic equality is applied as a light constraint, which just helps to define a consistent behavior in the Z direction within the 2.5D RVE simulation. Moreover,
Figure 15 and
Figure 16 confirm the efficient implementation of the periodic boundary conditions.
Figure 17 presents the 2.5D RVE distribution of strains (a) and stresses (b) in the loading direction (Y). The critical areas are at the matrix–particle interface with risks of decohesion and where the experiment predicts rupture initiation [
15]. One could even see in
Figure 17a the beginning of a strain localization between some particles. The current model without cohesive elements at a matrix–particle interface does not include the damage evolution present at the interface in the real material for a macro strain larger than 0.10. The degradation of the interface is probably already beginning earlier, so no quantitative interpretation was completed here within the current results, assuming continuous interfaces. However, the predicted trends seem consistent with the damage mechanisms already experimentally identified in [
15] (decohesion, localization, rupture of particles starting around 10% strain).
The local stress fields (
Figure 15,
Figure 16,
Figure 17,
Figure 18 and
Figure 19) computed by the Lagamine (2.5D RVE) and METAFOR v 3494 (3D RVE) codes show a close agreement for the matrix stress value, even if the distribution of particles differs between the 2.5D square and the 3D cube, and if these stresses are computed by different element types. These RVEs are built for the same particle statistics (see Image J post-processing in
Section 2), so indeed both models should provide similar results on average.
A direct comparison of the local fields between 2.5D and 3D RVE simulations is not straightforward, as even if built for the same particle statistics, the particle distribution methodology used does not impose a similar choice of particles. For the 3D RVE (
Figure 18) and 2.5D RVE (
Figure 16 and
Figure 17), simulations with a similar mesh density and edge size are presented. The computed local stress fields are in close range. In agreement with the predicted global macro stress shown in
Figure 14, the 2.5D RVE predicts local stresses as being slightly lower than the 3D RVE results. The level of the 2.5D RVE results is sensitive to the transversal boundary condition optimized to recover the macroscopic stress–strain curve. As shown in
Figure 9, the plane strain is definitively too stiff and the Free Z condition does not allow having a similar stress state in the X and Z directions (
Figure 12), as expected and indeed predicted by the 3D RVE simulations (
Figure 18a,b).
The 3D RVE simulation result in the Y direction (
Figure 19) confirms the strong heterogeneity of the stress and strain fields close to the particles. The 2.5D RVE model presents, versus a 3D RVE one, a high CPU advantage as the associated simulation times are decreased by a factor 417 (real time) or 2581 (CPU time). The parallel computing distribution is only interesting for the 3D RVE. This computing time comparison was performed on an AMD Thread Ripper 3970X 32-Core Processor (PRIMINFO, Liège, Belgium), exploiting16 threads for similar mesh densities.
6. Conclusions
A methodology for constructing a 2.5D RVE that improves the FEM prediction of local stress and strain fields in an isotropic two-phase material (matrix and particles) has been described. Less accurate than a 3D RVE simulation but improved compared to classical 2D RVE approaches, a 2.5D RVE provides a quick alternative for the identification of particle shapes and sizes, generating an optimal tensile behavior.
The tensile behavior of the L-PBF AlSi10Mg material post-processed by FSP and presenting a soft matrix compared to stiff particles has been predicted by RVEs. The high local gradients near the matrix–particle interfaces, which corresponds to the experimentally observed damage modes, were computed. The simulation results show that all RVE simulations could predict stress–strain curves. The plane strain 2D RVE evaluates a tensile curve with an excessive stiffness compared to the macroscopic experiment. The “Free in Z” 2D RVE and 2.5D RVE predictions are quantitatively close to the experimental values. By analyzing the strain fields according to the X and Z directions for a tensile test in the Y direction, the isotropic behavior is better recovered in an average way for a 2.5D RVE than for a “Free in Z” 2D RVE assumption. Cohesive elements would be necessary to quantitatively simulate static failure (strain above 10% not investigated here).
The key advantage of the 2.5D model is its computational efficiency. The identified microstructure by a 2.5D RVE could speed up an accurate 3D RVE optimization. Indeed, the CPU time of a 3D RVE becomes an issue if optimization loops, deep learning training, or FE2 computation are foreseen. A long-term goal could be the extension of the 2.5D RVE with matrix–particle decohesion and advanced cyclic damage constitutive law to address the prediction of Wöhler curves. It should speed up material design for improved behavior in fatigue.
While the 2.5D RVE approach offers substantial computational efficiency and maintains accuracy in many cases, there are situations where the 3D RVE may be more suitable. For example, when dealing with materials that exhibit strong anisotropy or complex 3D microstructures—such as intricate grain orientations, strong anisotropy behavior of each particle, void distribution with a non-representative planar pattern, or fiber reinforcements—the 2.5D approximation will not fully capture the material behavior. In those cases, where out-of-plane stresses present strong heterogeneity between particles or where full 3D representation is necessary to characterize the microstructure geometry, the prediction of mechanical responses even under a uniaxial tensile or compression load with a 2.5D RVE will be inaccurate. Therefore, for materials with complex 3D microstructures, the 3D RVE remains the preferred choice to ensure accuracy across all spatial dimensions.
The uniaxial target loading is also a limitation, even for material presenting an isotropic behavior. Any macroscopic load where a relation between some average strain components can be found thanks to the material isotropic property could bring a methodology extension; however, complex loading will be excluded. While the current model successfully estimates the monotonic tensile curve, it may benefit from further extensions, such as incorporating kinematic hardening into the matrix constitutive model, to enhance its accuracy and applicability into cyclic scenarios.
In conclusion, the 2.5D method classifies different microstructures, speeds up material design, and saves resources compared with the 3D RVE method. The simplicity of the 2D RVE mesh operation as well as the short direct link from 2D SEM images to define a representative set of particles without complex tomography experiments or image reconstruction is of interest. Our qualitative comparison of a 2.5D RVE versus a 3D one has not pointed out a huge accuracy decrease in the local interface matrix–particle stress and strain field. So, future work should include a statistical comparison of 2.5D and 3D RVE results to identify if a factor relating 2.5D RVE local interface information to a 3D one for matrix–particle microstructure is required, as suggested by Qayyum et al. [
9]. Cyclic loading simulations are also foreseen to address fatigue behavior. Purely numerical perspectives are numerous, such as an easy study of the effects of the proportions, sizes, shapes, orientations, and distributions of the particles, and of the strength ratio matrix/particle on mechanical properties, as long as the macroscopic behavior stays isotropic.