Next Article in Journal
Investigation of In-Flight Alumina Particle Interaction with a Plasma Jet in a Thermal Plasma Reactor
Previous Article in Journal
The Comparison of Microstructure, Phase Composition and Mechanical Properties of Inconel 625 Alloys Obtained by Wire Arc and Wire Electron Beam Additive Manufacturing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Investigation of the Performance of Equal Channel Angular Pressed Copper Electrodes in Electric Discharge Machining

1
Roketsan Missiles Industries Inc., Ankara 06780, Türkiye
2
Mechatronics Engineering Department, Çankaya University, Ankara 06800, Türkiye
*
Author to whom correspondence should be addressed.
Crystals 2025, 15(10), 849; https://doi.org/10.3390/cryst15100849 (registering DOI)
Submission received: 20 August 2025 / Revised: 15 September 2025 / Accepted: 25 September 2025 / Published: 29 September 2025
(This article belongs to the Section Crystalline Metals and Alloys)

Abstract

This study examines the mechanical, thermal, and electrical properties of copper tool electrodes processed via Equal Channel Angular Pressing (ECAP), with a specific focus on their performance in Electrical Discharge Machining (EDM) applications. A novel Crystal Plasticity Finite Element Method (CPFEM) framework is employed to model anisotropic slip behavior and microscale deformation mechanisms. The primary objective is to elucidate how initial crystallographic orientation influences hardness, thermal conductivity, and electrical conductivity. Simulations are performed on single-crystal copper for three representative Face Centered Cubic (FCC) orientations. Using an explicit CPFEM model, the study examines texture evolution and deformation heterogeneity during the ECAP process of single-crystal copper. The results indicate that the <100> single-crystal orientation exhibits the highest Taylor factor and the most homogeneous distribution of plastic equivalent strain (PEEQ), suggesting enhanced resistance to plastic flow. In contrast, the <111> single-crystal orientation displays localized deformation and reduced hardening. A decreasing Taylor factor correlates with more uniform slip, which improves both electrical and thermal conductivity, as well as machinability, by minimizing dislocation-related resistance. These findings make a novel contribution to the field by highlighting the critical role of crystallographic orientation in governing slip activity and deformation pathways, which directly impact thermal wear resistance and the fabrication efficiency of ECAP-processed copper electrodes in EDM.

1. Introduction

Electrical discharge machining (EDM) is classified among unconventional fabrication techniques [1,2]. During EDM, material removal occurs through electric discharges (sparks) generated in the gap between the tool electrode (shortly referred to as the electrode) and the workpiece. The energy released during each spark pulse causes localized melting and vaporization of the electrode and workpiece material surfaces. The total cost of a component fabricated via EDM is primarily influenced by the electrode, including material and fabrication costs, as well as the duration of the operation [3].
The electrode wear, one of the process’s limitations, is affected by several factors, including the thermal and electrical conductivity of the electrode material. These properties affect how efficiently electrical energy is transferred during discharge and how spark heat dissipates. Although EDM does not involve mechanical loading, the high-energy discharges generate microscopic stresses in the material. Therefore, the electrode’s durability, structural stability, and homogeneity are critical factors that directly influence EDM performance and electrode service life [4,5]. Furthermore, the machinability of electrode material is an important consideration, as it substantially affects electrode production costs. If the chosen material is difficult to shape, it becomes impractical for EDM applications and severely limits its usability [6].
Pure copper and its alloys are the most widely used material for EDM electrodes, primarily due to their excellent electrical and thermal conductivity and easy fabrication [7,8,9]. They exhibit desirable mechanical strength, high thermal conductivity, good oxidation resistance, and dimensional stability, in addition to being easy to machine. Copper electrodes, with higher thermal conductivity, dissipated heat more effectively, reducing localized melting and wear.
Equal-Channel Angular Pressing (ECAP) is a widely used severe-plastic-deformation (SPD) technique that improves hardness, extends wear life [10,11] and enhances corrosion resistance [12,13,14] in a range of alloys, particularly aluminum- and zinc-based systems. Magnesium and nickel alloys subjected to three ECAP passes exhibited a nearly 50% reduction in surface roughness and an over 50% decrease in abrasive wear, resulting in enhanced machining precision [15,16]. In the case of pure copper, optimized control of grain size and dislocation density has been shown to yield high electrical conductivity alongside increased tensile strength and elongation [17]. For Zn-Ag-Mg alloy, suppressing twin boundaries and reducing dislocation density further contribute to improved ductility with successive ECAP passes [18]. Texture evolution during deformation also plays a significant role; initial [111]<110> textures transform into stronger [110]<112> Brass textures, particularly after four passes, at which point electrical conductivity has been observed to reach up to 96.1% [19]. Similarly, for copper rods, Ciemiorek et al. [20] reported an initial drop in electrical conductivity after the first ECAP pass, followed by a recovery and eventual increase through subsequent passes up to eight. Shaban et al. [21] also confirmed that ECAP strengthens pure copper billets without significantly compromising their electrical performance. SPD has been demonstrated to significantly refine the microstructure of copper, thereby enhancing its mechanical performance. Yield and ultimate tensile strengths were reported to increase by approximately 85% and 80%, respectively. Despite these improvements, electrical conductivity was only marginally reduced (about 3% IACS), and hydrostatically extruded electrodes showed over 60% lower wear rates in EDM tests compared to their commercial counterparts [22]. The other study on EDM using ECAP-processed copper electrodes reported approximately 6% lower electrode wear compared to as-received samples, while maintaining a consistent material removal rate. This resulted in the lowest wear ratio observed after a single pass [23]. Additionally, Gopal et al. [24] observed a general trend of increasing hardness and decreasing electrical conductivity in copper electrodes subjected to EDM with each successive ECAP pass. Nevertheless, the magnitude of these changes was found to diminish progressively after the initial passes.
The literature review reveals a scarcity of studies investigating the use of ECAP-processed electrodes to improve the performance of the EDM process, particularly in terms of reducing electrode wear and facilitating electrode fabrication. Moreover, only a few studies [25,26] have systematically evaluated ECAP performance using numerical methods. Therefore, the current study aims to determine ECAP processing conditions that promote microstructural states conducive to improved mechanical strength, thermal transport, and electrical conductivity in pure copper electrodes, using a Crystal Plasticity Finite Element Method (CPFEM) approach. In this study, a novel rate-dependent crystal plasticity algorithm is developed and implemented as a user-defined material subroutine (VUMAT) in Abaqus/Explicit. The algorithm is utilized to investigate the influence of crystallographic deformation orientations, specifically <100>, <110>, and <111>, on the electrical and thermal conductivity, as well as the hardness of copper electrodes, which are critical factors affecting their wear behavior and machinability in EDM. The proposed model is employed to perform CPFEM simulations on single-crystal copper under ECAP-induced boundary conditions, enabling a detailed assessment of orientation-dependent responses. The effect of the ECAP process on the hardness, thermal conductivity, and electrical conductivity of copper electrodes is analyzed and compared with data from the literature. To the best of our knowledge, this is the first study to integrate CPFEM-based crystallographic analysis with EDM performance metrics for ECAP-processed copper electrodes.

2. Finite Element Model

The CPFEM simulation for a single-pass ECAP uses the commercial FE software Abaqus 2023. Figure 1a provides a 3D CAD model representation of the ECAP mold and the FE mesh configuration utilized for the analysis.The mesh properties employed in the CPFEM model (Figure 1b) are detailed in Table 1. The single-crystal copper electrode underwent meshing using 2560 linear hexahedral C3D8R elements, selected based on dilatational wave speed and the deformation speed ratio to prevent distortion issues. Meanwhile, the rigid punch is meshed using 504 linear hexahedral C3D8R elements. The ECAP mold is meshed utilizing 381,163 linear tetrahedral elements. The technical drawing of ECAP mold is given in Figure 1c. Increasing the ϕ angle (Figure 1d) leads to a decrease in both strain homogenization and hardness. To avoid these adverse effects and to ensure a more uniform strain distribution [27], a ϕ angle of 90° is selected in the present study. The ECAP mold and punch are treated as rigid bodies, representing non-deformable structures.
Encastre boundary conditions ( u 1 = u 2 = u 3 = R 1 = R 2 = R 3 = 0 ) are applied to the mold via a rigid body reference point definition. The square electrode boasts a cross-sectional area of 16 mm 2 ( 4 mm × 4 mm ) and spans a length of 50 mm. In the contact algorithm, explicit general contact is employed with a Coulomb friction coefficient ( μ ) of 0.05. CPFEM analyses of copper single crystals during ECAP report that μ = 0.05 yields better agreement with experimentally observed texture than μ = 0.1 [28,29]. In all cases, the simulations are conducted at a pressing speed of 3 mm/s [30], employing a mass scaling factor of sixteen. A dummy polycrystalline specimen with the same cross-sectional geometry as the electrode sample is modeled using von Mises plasticity to isolate and eliminate the influence of the punch force on the actual specimen during simulation.
Energy balance equations are commonly used to verify whether a simulation produces an appropriate quasi-static response. A widely accepted criterion in the literature involves comparing the kinetic energy history with the internal energy. In metal forming simulations, the majority of internal energy originates from plastic deformation. For a solution to be considered quasi-static, the kinetic energy should constitute only a small fraction of the internal energy, typically less than 1–5% [31]. As illustrated in Figure 2, the kinetic energy remains below 1% of the internal energy throughout the analysis, apart from the initial stage. This confirms that the criterion is satisfied, thereby indicating a valid quasi-static response. The figure also highlights the variations in strain energy (SE) and plastic dissipation energy (PD). Moreover, the total energy (Etotal) of the model remains nearly constant during the entire simulation.

3. Methods

3.1. Crystal Plasticity Framework

Metal forming simulations using the CPFEM can be performed with both implicit and explicit algorithms. Implicit CPFEM may encounter convergence issues when applied to highly non-linear and dynamic processes [32]. On the other hand, explicit algorithms are often preferred due to their stability under contact nonlinearity and frictional discontinuities. However, CPFEM simulations remain computationally demanding due to the stress update algorithms. Numerous studies have explored the pros and cons of implicit versus explicit CPFEM [33,34].
The crystallographic slip mechanism can effectively model the material’s constitutive response in single crystals [35,36]. The overall deformation gradient F = R * U , where R * represents rigid body rotation and U signifies the right stretch tensor, can be decomposed multiplicatively into elastic ( F e ) and plastic ( F P ) components [37] (Equation (1)). Furthermore, the elastic portion of deformation ( F e ) can be expressed as the product of the left stretch tensor ( V e ) and the rotation tensor ( R e ).
F = F e F p
In the explicit formulation within Abaqus, the total rotation tensor R * represents rigid body rotation, and it is formulated utilizing the Green-McInnis-Naghdi rate [38]. In the context of plastic deformation, the plastic component of the deformation gradient has no impact on the rigid body rotation of a single crystal. Consequently, the elastic portion of rotation R e can be equated to the total rotation R * of the deformation, as expressed in Equation (2).
F = R * U ,   R * U = V e R * F p
Resolved shear stress τ α acting on the α th slip system in the undeformed lattice can be defined as;
τ n α = k = 1 α C n e 0 S n e 0 s ˜ 0 α n ˜ 0 α
where C n e 0 in Equation (3) is the elastic right Cauchy-Green strain tensor. The S n e 0 term represents the Piola-Second Kirchhoff stress (PK2) on the intermediate state U 0 as illustrated in Figure 3. s ˜ 0 α n ˜ 0 α refers to the tensor product of the slip direction and nominal direction of the initial slip system directions. This product can be obtained by multiplying the slip system ( s α , n α ) with the Bunge-formulated Euler angles ( α , β , γ ) from s ˜ 0 α = Q 0 s α and n ˜ 0 α = Q 0 n α .
The power law type rate-dependent formulation, which establishes the relationship between the slip rate on each slip system, the current yield stress τ α , and the current strength g α , is presented in Equation (4) [39].
γ ˙ n α = γ ˙ 0 | τ n α g n α | m s i g n ( τ n α )
Here, γ ˙ 0 represents the reference shearing rate, and the exponent m denotes the strain rate sensitivity coefficient.
The expression for current strength g α is defined by Equation (5). Initially, the critical resolved shear stress τ 0 α is equal to g 0 . The hardening moduli h α β describe the rate of strain hardening on slip system α due to slip β . This phenomenon of self and latent hardening is phenomenologically characterized by h α β = q h ( γ ) [40].
g ˙ α = k = 1 12 h α β | γ ˙ β |
The secant-type model is introduced in Equation (6) [41].
h α α = h ( γ ) = h 0 s e c h 2 | h 0 γ τ s τ 0 |
Here, h 0 represents the initial hardening modulus, τ 0 stands for the yield stress, which is equal to the initial value of the current strength g 0 α , τ s denotes the stage I stress, and γ signifies the Taylor cumulative shear strain on all slip systems as presented in Equation (7).
γ = α 0 t | γ ˙ α | d t
Plastic deformation is induced by shear or slip γ α occurring on slip systems. These slip planes are characterized by s ˜ 0 and n ˜ 0 vectors, representing the unit vector in the slip direction and the unit normal vector to the slip plane in the intermediate configuration U 0 , as illustrated in Figure 3. The velocity gradient is defined by Equation (8);
L = F ˙ F 1
Substituting Equation (1) into the equation above and then applying straightforward differentiation through the product rule results in the subsequent additive decomposition of the velocity gradient: L = L e + F e L p ( F e ) 1 . Here, L e and L p represent the elastic and plastic components of velocity gradients, respectively. As defined in the previous expression, the plastic part of the velocity gradient pertains to the intermediate configuration. Consequently, the second term represents a push forward to the current configuration. As plastic deformation is attributed to dislocation slip along multiple slip systems, the plastic part of velocity gradient L p can be determined as the sum of shear rates associated with each slip system, denoted as γ ˙ α , according to Equation (9).
L p = k = 1 α γ ˙ n α s ˜ 0 α n ˜ 0 α
The plastic component of the deformation gradient can be expressed using Equation (10).
F n + 1 p = ( I + Δ t k = 1 α γ ˙ n α s ˜ 0 α n ˜ 0 α ) F n p
Thus, the elastic component of the deformation gradient F n + 1 e can be written as;
F n + 1 e = V n + 1 e R *
The rate of Cauchy stress is linked to the second Piola-Kirchhoff (PK2) stress, and the rate of PK2 stress can be determined by the multiplication of the rotated elastic modulus by transformed initial orientations [42]. This is represented as C ˜ = Q 0 c C Q 0 c T , where C is the material’s elastic stiffness tensor. The PK2 stress rate is calculated by combining the material’s strain rate tensor and the Green-Lagrange strain rate ( E ˙ n + 1 ), as outlined in Equations (12).
S ˙ n + 1 e 0 = C ˜ : E ˙ n + 1
The corotational stress rate can be expressed using Equation (13) and the term R * T σ ˙ R * can be calculated from R * T σ ˙ R * = 1 d e t F n + 1 e U n + 1 e S ˙ n + 1 e 0 U n + 1 e
σ ˙ = R ˙ * T σ R * + R * T σ ˙ R * + R * T σ R ˙ *
Stress rate can be defined as σ ˙ = σ + Ω σ n σ n Ω so the increment of Cauchy stress can be calculated as σ ˙ Δ t = Δ σ + R σ n R T in which σ is the rate of objective, corotational stress, or the rate associated with the constitutive response [43]. Skew symmetric angular velocity tensor Ω can be found from Ω = R ˙ * R * T . Hence, the update of Cauchy stress for the spatial configuration U s can be derived from σ n + 1 = σ n + σ ˙ Δ t .
Tensile test, simple shear test, and pure shear test simulations are conducted to verify the consistency of the explicit algorithm with experimental results [44] and the UMAT code developed by Huang [45]. The boundary conditions are treated as shown in Figure 4 [46,47]. Material parameters of single crystal copper, taken from the literature [48], are compared with [100] and [111] directed single copper crystal test data. However, under simple shear boundary conditions, an inconsistency is observed with the UMAT code, as shown in Figure 5b. One of the main reasons for this situation is as explained in detail in reference [49]. The material parameters of single copper for the VUMAT subroutine are illustrated in Table 2. The algorithm of the explicit crystal plasticity user subroutine (VUMAT) is illustrated in Figure 6.

3.2. Von Mises Plasticity Framework

The von Mises plasticity model is the constitutive model used to simulate severe plastic deformations of the dummy workpiece. The true stress-strain curve of the copper specimen is illustrated in Figure 7.

4. Results and Discussion

4.1. Plastic Strain and Taylor Factor Evolution in ECAP-Processed Copper Electrodes

Figure 8 shows that the plastic equivalent strain (PEEQ) values near the top of the single-crystal copper electrode exhibit homogeneity, with PEEQ maintaining an approximately constant value across all its elements. Conversely, the strain distribution in the lower part of the electrode appears heterogeneous, with each region showing a consistent value of effective strain. The strain distribution along the axial direction remains nearly uniform except for the initial billet segment, where steady-state deformation hasn’t yet been achieved. The distribution of effective strain along the height of the billet is the characteristic of ECAP processes employing dies with external curvature. The visible gap between the electrode and the die represents an external curvature during the process, resulting in reduced shear strain levels in the outer region of the deformation zone. The electrode near the inner corner experiences more significant strain and strain gradients than the material near the outer corner. The intense deformation near the inner corner imparts a higher compressive force to the material neighboring the right-hand side interface of the entry channel. As a result, this leads to increased frictional stress at the interface on the right-hand side and a notably significant plastic strain rate.
Experimental measurement of deformation modes is not directly feasible due to the complex, multiscale nature of plastic deformation, which involves heterogeneous slip activity, internal lattice rotations, and evolving microstructural features that are difficult to observe in situ and at sufficient resolution. The CPFEM approach offers the advantage of capturing such local heterogeneities in strain and stress distribution, enabling a more detailed and mechanistic understanding of orientation-dependent plastic behavior. The extent of strain inhomogeneity can be evaluated utilizing the equation proposed as S H i = A c c u m a x A c c u m i n A c c u A v g [52]. In this equation, S H i represents the strain inhomogeneity index, A c c u m i n stands for the minimum accumulated plastic strain, A c c u m a x denotes the maximum accumulated plastic strain, and A c c u A v g signifies the average accumulated plastic strain. The strain inhomogeneity index and the coefficient of variation of accumulated plastic strain for different initial orientations are shown in Figure 9. The black color corresponds to the initially oriented <100> direction. The orange and blue colors represent initially oriented <110> and <111> directions, respectively. Lower values of the index indicate better homogeneity. Although the <111> initially oriented single crystal electrode exhibits more inhomogeneous deformation compared to the <100> and <110> orientations after one pass of ECAP, the resulting inhomogeneity values remain relatively close to each other. The results suggest that the heterogeneity of through-thickness accumulated plastic strain in the Z-direction is also orientation-dependent.
The analysis of deformation behavior and texture evolution in the single copper crystal involves examining specific regions and elements of the specimen, as shown in Figure 10. Initially <100> oriented single crystal copper demonstrates the highest Taylor factor ( T a y l o r f a c t o r ( M ) = α | γ ˙ α | t ϵ ˙ e f f p t ) when it is deformed in A deformation path within the plastic deformation zone. Conversely, initially, a <111>-oriented single-crystal copper exhibits the lowest values throughout the process. Toward the end of the process, initially <100> oriented single-crystal copper again displays higher Taylor factors than initially <111> oriented single-crystal copper. It’s important to clarify that reducing the Taylor or Schmid factors does not necessarily imply material softening. Instead, grains with higher Taylor factors experience more significant stress reduction during deformation. A decrease in the Taylor factor suggests that the grain’s orientation has shifted towards more favorable slip conditions. Smaller Taylor factor values indicate orientations more conducive to deformation, as they require lower stress levels for plastic deformation to occur. The crystals with smaller Taylor factors tend to deform more quickly than those with larger values [53]. Conversely, an increase in the Taylor factor is associated with higher energy accumulation due to deformation [54]. The presence of low Taylor factor values with a narrow distribution in microstructural orientations facilitates more uniform deformation among the grains. This homogeneous deformation can limit local increases in dislocation density, reduce electron scattering, and thereby improve electrical conductivity [55]. High electrical conductivity ensures efficient spark generation and minimizes the electrode’s energy loss, contributing to lower wear rates [56]. Moreover, metals with high electrical conductivity also tend to exhibit high thermal conductivity [57]. The thermal conductivity of the electrode material influences the transfer of spark energy within the electrode body during EDM, thereby affecting electrode wear rates. Electrodes with higher thermal conductivity dissipate heat more effectively, which reduces localized melting and subsequent wear [58]. A decrease in the Taylor factor generally enhances the machinability of the material, as it offers lower resistance to deformation during cutting, resulting in reduced tool loading and facilitated plastic flow. However, secondary factors such as chip morphology and surface integrity must be carefully considered due to their sensitivity to local deformation behavior [59]. Initially, <111>-oriented single-crystal copper electrodes exhibit a reduction in Taylor factor values, particularly under deformation routes A and B, resulting in a concurrent decrease in dislocation density. This microstructural refinement, achieved through SPD methods such as ECAP, results in lower machining forces due to enhanced plastic deformability [60]. Furthermore, the elevated electrical conductivity observed following ECAP processing along routes A and B has also been reported in the literature, correlating with the lower Taylor factor and dislocation density in the <111> orientation [61,62].

4.2. Texture Evolution of ECAP-Processed Copper Electrode

The crystallographic texture of polycrystalline materials significantly influences their thermal conductivity due to the anisotropic nature of phonon and electron transport. Specific texture components, such as cube ([100]<001>), Goss ([110]<001>), and Brass ([011]<211>), are commonly observed in FCC metals and can be identified through characteristic intensity peaks in pole figures [63]. The cube component, for example, aligns heat-conducting directions <100> with the measurement axis, often resulting in higher thermal conductivity in textured materials. Conversely, orientations such as Brass or Goss may lead to increased phonon scattering due to higher dislocation density or less favorable transport pathways. Studies have shown that a firm Cube texture can enhance thermal conductivity in rolled or annealed metals by minimizing grain boundary resistance and maximizing aligned conduction pathways [64]. Therefore, pole figure analysis not only aids in texture quantification but also provides insights into the anisotropic thermal performance of engineered materials.
The pole figures after a single ECAP pass of single-crystal copper are shown in Figure 11. According to these pole figures, Goss and Brass orientations emerge after one ECAP pass for initially <100> and <110> oriented single crystals under deformation paths A, B, and C. Similarly, Goss and Brass components are also observed in initially <111> oriented crystals under paths A and B. Therefore, the electrical conductivity of initially <100> oriented copper electrodes is expected to decrease due to the loss of the favorable <100> alignment. In contrast, for initially <110> oriented electrodes, the conductivity is expected to remain relatively stable due to the retention of suitable transport orientations. Moreover, for initially <111> oriented copper, the transformation to Goss and Brass textures under deformation paths A and B is expected to enhance electrical conductivity. Besides that, under deformation path C, shear texture components become dominant after a single pass, which may negatively impact conductivity. The obtained pole figures are in good agreement with those reported in previous studies [65,66]. Shear deformation paths, such as those imposed by SPD techniques like ECAP, significantly influence the microstructure and, consequently, the electrical conductivity of metallic materials. The applied shear strain leads to grain refinement, dislocation multiplication, and texture evolution, all of which affect electron scattering mechanisms. In particular, high-angle grain boundaries and increased dislocation density associated with shear deformation typically reduce electrical conductivity [67,68]. However, under controlled deformation paths such as route A or B, it is possible to promote favorable crystallographic textures that align low-resistivity directions with the measurement axis. This alignment can partially offset conductivity losses by facilitating electron transport along preferred orientations [69]. Thus, while SPD generally introduces microstructural defects that hinder conductivity, carefully selected deformation paths can optimize texture to mitigate these effects and even enhance electrical conductivity. The Taylor factor and pole figure results indicate that both electrical and thermal conductivities can be improved through the ECAP process, particularly under deformation paths A and B for initially <111> oriented copper electrodes.

4.3. Thermal Effects on Texture Analysis of ECAP-Processed Copper Electrodes

During the ECAP process, the imposed deformation path has a significant influence on the temperature evolution within the material, which in turn affects its microstructural development. Different ECAP routes (such as A, B, and C) result in varying shear strain distributions and heat generation patterns due to differences in strain path repetition and crystal lattice reorientation. These temperature variations, particularly under adiabatic or near-adiabatic conditions, can accelerate recovery or dynamic recrystallization in localized regions, thereby modifying grain size, dislocation density, and texture evolution [70]. The recrystallization temperature for pure copper is typically expected to be in the range of 30–40% of its melting point (1085 °C), corresponding to approximately 350–450 °C. However, severe plastic deformation techniques such as ECAP can significantly reduce this temperature to below 200 °C [71]. Therefore, to accurately evaluate the temperature rise during the ECAP process, it is necessary to perform a detailed thermal analysis under appropriately defined boundary conditions. The present study, the temperature analysis is initially carried out under adiabatic assumptions, considering only the contribution from shear deformation. However, it should be noted that frictional dissipation contributes directly to additional heat generation at the die–specimen interface, which can further accelerate recrystallization processes in localized regions. In order to illustrate the temperature rise independently of die parameters, only the contribution from plastic strain has been presented in this work. In this context, the temperature rise due to both plastic deformation work and interfacial frictional heating should be explicitly taken into account. A general equation for the temperature rise in ECAP is derived, and, as also demonstrated in previous studies in the literature, the temperature increase scales with the material strength, ram speed, and channel angle, whereas it decreases with higher density, heat capacity, and die corner angle [72,73].
Assuming adiabatic conditions, the temperature change is calculated using ρ c θ ˙ = ξ k = 1 N α τ α γ ˙ α ( k ) [74]. Where ρ denotes the material density, and c is the specific heat capacity. The left-hand side represents the rate of internal energy accumulation per unit volume, attributed to the temperature change under adiabatic conditions. On the right-hand side, ξ is the Taylor–Quinney coefficient, which quantifies the fraction of plastic work converted into heat (typically ranging between 0.85 and 1). The double summation accounts for the plastic dissipation within all crystallographic slip systems α , evaluated at each material integration point k. Here, τ α is the resolved shear stress, and γ ˙ α is the shear strain rate associated with the α -th slip system. The expression thus captures the total rate of plastic dissipation transformed into heat and subsequently manifested as a temperature increment in the material during the plastic deformation. Figure 12 indicates that deformation route A exhibits the highest temperature increase compared to routes B and C. This observation is consistent with the results of accumulated plastic strain. A route can also be reached through the recrystallization temperature during the ECAP process. Upon reaching the recrystallization temperature, the material undergoes significant microstructural transformations driven by the stored deformation energy, primarily in the form of dislocations. The high dislocation density accumulated during severe plastic deformation is progressively reduced through recovery and recrystallization mechanisms. During this process, dislocation rearrangement and annihilation occur, followed by the nucleation and growth of new, defect-free grains. As a result, the heterogeneity in dislocation distribution is eliminated, and internal residual stresses are substantially relieved. Simultaneously, the crystallographic texture of the material is also altered: deformation-induced texture components, such as Brass, Copper, and S-type orientations (common in FCC metals), are weakened or replaced by recrystallization textures. In the case of copper, a strong Cube texture ([001]<100>) often develops due to the preferential growth of favorably oriented grains. The extent and nature of these changes are strongly influenced by factors such as prior strain, annealing temperature, and the mobility of high-angle grain boundaries [75,76,77].

5. Conclusions

This study applies an explicit CPFEM framework to assess the effect of crystallographic orientation on the deformation and performance of ECAP-processed copper electrodes for EDM. Simulations are performed for three FCC single-crystal orientations, <111>, <100>, and <110>, under one pass ECAP conditions. The <100> orientation displays the most homogeneous deformation and the highest Taylor factor, indicating strong resistance to plastic flow. In contrast, the <111> orientation shows more localized plasticity and lower hardening capacity, although with relatively uniform fields. Temperature variations are linked to accumulated plastic strain, and the recrystallization temperature is found to significantly influence texture evolution. Texture analysis and Taylor factor evaluations indicate that ECAP, particularly along paths A and B in <111> oriented electrodes, can improve electrical and thermal conductivity through favorable texture development. The reduction in Taylor factor, dependent on both orientation and path, promotes plastic softening and enhances machinability. <111> crystals also exhibit potential for slip band formation and structural homogeneity. While deformation twinning is likely in high-shear regions, particularly along <111> directions, it is not explicitly included in the present framework. The present findings underscore the pivotal role of crystallographic orientation and deformation path in enhancing the manufacturability, durability, and thermoelectric efficiency of copper electrodes employed in EDM. The computational framework developed herein establishes a direct link between orientation-dependent deformation mechanisms and the resulting functional performance, offering a solid basis for process optimization. Future investigations will focus on experimental validation to ensure the practical applicability and reliability of the predicted trends.

Author Contributions

Both authors contributed to the study’s conception and design. Ü.Ş. performed material preparation, data collection, and analysis. Ü.Ş. wrote the first draft of the manuscript, and C.Ç. supervised the study and provided continuous guidance during the preparation of the manuscript. Both authors commented on previous versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Ülke Şimşek was employed by the company Roketsan Missiles Industries. The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Çogun, C.; Esen, Z.; Genc, A.; Cogun, F.; Akturk, N. Effect of powder metallurgy Cu-B4C electrodes on workpiece surface characteristics and machining performance of electric discharge machining. Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 2016, 230, 2190–2203. [Google Scholar] [CrossRef]
  2. Benedict, G.F. Nontraditional Manufacturing Processes, 1st ed.; CRC Press: Boca Raton, FL, USA, 1987. [Google Scholar] [CrossRef]
  3. Kanagarajan, K.; Arunachalam, R.; Sathiya, P. Modeling of machining parameters in electrical discharge machining of Al 7075 composite using graphite electrode. J. Mater. Process. Technol. 2008, 205, 442–450. [Google Scholar] [CrossRef]
  4. Çoğun, C.; Ünses, E. Improvement of Electric Discharge Machining (EDM) Performance of Ti-6Al-4V Alloy with Added Graphite Powder to Dielectric. Stroj. Vestn.—J. Mech. Eng. 2015, 61, 409–418. [Google Scholar] [CrossRef]
  5. Torres-Salcedo, A.; Luis-Pérez, C.J.; Puertas-Arbizu, I.; Corres-Sanz, J.M. A Study on the EDM Drilling of Reaction-Bonded Silicon Carbide Using Different Electrode Materials. Met. Mater. Int. 2023, 126, 5139–5162. [Google Scholar] [CrossRef]
  6. Voigt, O.; Wendler, M.; Siddique, A.; Stöcker, H.; Quitzke, C.; Peuker, U.A. Characterization and Microstructure of Recycled Eroded Particles from Die-Sink Electro Discharge Machining of H11 Alloy for Applicability in Additive Manufacturing. Met. Mater. Int. 2024, 30, 1209–1226. [Google Scholar] [CrossRef]
  7. Şimşek, Ü.; Çoġun, C.; Esen, Z. Effects of electrolytic copper and copper alloy electrodes on machining performance in electrical discharge machining (EDM). Mach. Sci. Technol. 2022, 26, 229–244. [Google Scholar] [CrossRef]
  8. Raj, S.; Kumar, A. Copper and Copper Alloys: Applications and Properties. Mater. Today Proc. 2016, 3, 3580–3585. [Google Scholar] [CrossRef]
  9. Liu, Y.; Zhang, X. Recent Developments in Copper Alloys for Heat Transfer and Electronics Applications. J. Mater. Sci. 2021, 56, 11432–11458. [Google Scholar]
  10. Azadkoli, G.; Azadkoli, P.; Moazami-Goudarzi, M. Enhanced Strength and Ductility of Biodegradable Zn-1Mg Alloy Through EECAP Processing at Different Temperatures. Met. Mater. Int. 2025, 31, 2811–2822. [Google Scholar] [CrossRef]
  11. Vafaeenezhad, H.; Chegini, M.; Kalaki, A.; Serajian, H. Micromechanical Finite Element Simulation of Low Cycle Fatigue Damage Occurring During Sliding Wear Test of ECAP-Processed AA7075 Alloy. Met. Mater. Int. 2024, 30, 143–166. [Google Scholar] [CrossRef]
  12. Li, J.; He, T.; Du, X.; Vereschaka, A. Enhancing the corrosion resistance of high-strength Al-Zn-Mg-Cu alloys after equal channel angular pressing by developing retrogression and re-aging strategies. Corros. Sci. 2025, 246, 112736. [Google Scholar] [CrossRef]
  13. Orlov, D.; Ralston, K.; Birbilis, N.; Estrin, Y. Enhanced corrosion resistance of Mg alloy ZK60 after processing by integrated extrusion and equal channel angular pressing. Acta Mater. 2011, 59, 6176–6186. [Google Scholar] [CrossRef]
  14. Alipour, S.; Vafaeenezhad, H.; Ezatirad, F.; Fesahat, M.; Mahmoudi, M.; Mousavi-Khoshdel, S. Effect of grain refinement on the mechanical and corrosion behavior of ECAP-processed Sn-5Sb alloy studied by positron annihilation lifetime and Doppler broadening spectroscopy. J. Alloys Compd. 2025, 1011, 178175. [Google Scholar] [CrossRef]
  15. Zhang, S.; Zhang, P.; Xu, W.; Yan, H. Effect of ECAP on Mechanical Properties and Corrosion Resistance of 2024-CNTs@Ni Composite. Met. Mater. Int. 2023, 30, 1407–1423. [Google Scholar] [CrossRef]
  16. Khoubrou, I.; Nami, B.; Miresmaeili, S.M.; Yazdani, M. Evolution of Microstructure, Texture and Mechanical Properties of ECAP-Processed ZK60 Magnesium Alloy. Met. Mater. Int. 2025, 31, 227–236. [Google Scholar] [CrossRef]
  17. Zheng, Y.; Niu, J.; Lu, Y.; Xu, S.; Li, Z.; Zhao, H.; La, P. High strength and high conductivity pure copper prepared by ECAP-Conform combined with cryogenic rolling and recrystallization annealing. J. Mater. Res. Technol. 2025, 36, 3192–3204. [Google Scholar] [CrossRef]
  18. Zhao, L.; Zhuo, X.; Liu, H.; Jiang, J.; Ma, A. Effect of ECAP on the microstructure and mechanical properties of Zn–0.5Ag–0.08Mg alloy. J. Mater. Res. Technol. 2023, 25, 2619–2630. [Google Scholar] [CrossRef]
  19. Guo, T.; Gao, Y.; Feng, R.; Qian, D.; Zhang, G.; Ling, D.; Ding, Y. High Strength High Conductivity Copper Prepared by C-ECAP and Cryo-Rolling. Mater. Charact. 2024, 208, 113665. [Google Scholar] [CrossRef]
  20. Ciemiorek, M.; Pawliszak, L.; Chromiński, W.; Olejnik, L.; Lewandowska, M. Enhancing the Electrical Conductivity of Electrolytic Tough Pitch Copper Rods Processed by Incremental Equal Channel Angular Pressing. Metall. Mater. Trans. A 2020, 51, 3749–3753. [Google Scholar] [CrossRef]
  21. Shaban, M.; Alsharekh, M.F.; Alsunaydih, F.N.; Alateyah, A.I.; Alawad, M.O.; BaQais, A.; Kamel, M.; Nassef, A.; El-Hadek, M.A.; El-Garaihy, W.H. Investigation of the Effect of ECAP Parameters on Hardness, Tensile Properties, Impact Toughness, and Electrical Conductivity of Pure Cu through Machine Learning Predictive Models. Materials 2022, 15, 9032. [Google Scholar] [CrossRef]
  22. Skiba, J.; Kulczyk, M.; Przybysz-Gloc, S.; Skorupska, M.; Kobus, M.; Nowak, K. Effect of microstructure refinement of pure copper on improving the performance of electrodes in electro discharge machining (EDM). Sci. Rep. 2023, 13, 16686. [Google Scholar] [CrossRef]
  23. Marashi, H.; Jafarlou, D.M.; Sarahan, A.A.D.; Mardi, N.A. Employing severe plastic deformation to the processing of electrical discharge machining electrodes. Precis. Eng. 2016, 46, 309–322. [Google Scholar] [CrossRef]
  24. Gopal, R.; Thangadurai, K.; Thirunavukkarasu, K. Behavior of ECAP Processed Copper Electrodes in Electrical Discharge Machining of AISI H13 Steel. Mater. Today Proc. 2020, 21, 295–298. [Google Scholar] [CrossRef]
  25. Berndt, N.; Reiser, N.A.; Wagner, M.F.X. Evolution of plastic deformation during multi-pass ECAP of an AA6060 aluminum alloy—An experimental flow line analysis. J. Mater. Res. Technol. 2025, 34, 359–371. [Google Scholar] [CrossRef]
  26. Lu, C.; Deng, G.; Tieu, A.; Su, L.; Zhu, H.; Liu, X. Crystal plasticity modeling of texture evolution and heterogeneity in equal channel angular pressing of aluminum single crystal. Acta Mater. 2011, 59, 3581–3592. [Google Scholar] [CrossRef]
  27. Alateyah, A.; Ahmed, M.M.Z.; Alawad, M.O.; Elkatatny, S.; Zedan, Y.; Nassef, A.; El-Garaihy, W. Effect of ECAP die angle on the strain homogeneity, microstructural evolution, crystallographic texture and mechanical properties of pure magnesium: Numerical simulation and experimental approach. J. Mater. Res. Technol. 2022, 17, 1491–1511. [Google Scholar] [CrossRef]
  28. Deng, G.; Lu, C.; Su, L.; Liu, X.; Tieu, A. Modeling texture evolution during ECAP of copper single crystal by crystal plasticity FEM. Mater. Sci. Eng. A 2012, 534, 68–74. [Google Scholar] [CrossRef]
  29. Deng, G.; Tieu, K.A.; Su, L.; Lu, C.; Zhu, H.; Wei, P.; Liu, X. Crystal Plasticity FEM Study on the Influence of Crystallographic Orientation in Copper Single Crystals Subjected to Equal Channel Angular Pressing. Steel Res. Int. 2013, 84, 1258–1266. [Google Scholar] [CrossRef]
  30. Wongsa-Ngam, J.; Noraphaiphipaksa, N.; Kanchanomai, C.; Langdon, T.G. Numerical Investigation of Plastic Strain Homogeneity during Equal-Channel Angular Pressing of a Cu-Zr Alloy. Crystals 2021, 11, 1505. [Google Scholar] [CrossRef]
  31. Li, S.; Chen, W.; Kosimov, N.; Bhandari, K.S.; Jung, D. Reduction of Explicit Incremental Forming Simulation Time by Mass Scaling Parameters Adjusting and Symmetrical Setting. Int. J. Technol. Eng. Stud. 2021, 7, 27–34. [Google Scholar] [CrossRef]
  32. Şimşek, Ü.; Miyamoto, H.; Yalçınkaya, T. Crystal plasticity finite element analysis of linear and non-linear extrusion processes. Mater. Res. Proc. 2023, 28, 553–562. [Google Scholar] [CrossRef]
  33. Harewood, F.; McHugh, P. Comparison of the implicit and explicit finite element methods using crystal plasticity. Comput. Mater. Sci. 2007, 39, 481–494. [Google Scholar] [CrossRef]
  34. Wei, P.; Lu, C.; Tieu, K.; Su, L.; Deng, G.; Huang, W. A study on the texture evolution mechanism of nickel single crystal deformed by high pressure torsion. Mater. Sci. Eng. A 2017, 684, 239–248. [Google Scholar] [CrossRef]
  35. Hill, R.; Rice, J. Constitutive analysis of elastic-plastic crystals at arbitrary strain. J. Mech. Phys. Solids 1972, 20, 401–413. [Google Scholar] [CrossRef]
  36. Asaro, R.; Rice, J. Strain localization in ductile single crystals. J. Mech. Phys. Solids 1977, 25, 309–338. [Google Scholar] [CrossRef]
  37. Lee, E. Elastic-Plastic deformation at finite strains. J. Appl. Mech. 1969, 36, 243–245. [Google Scholar] [CrossRef]
  38. Abaqus Documentation. Solid (continuum) elements. In Theory Manual; Dassault Systemes Simulia, Inc.: Johnston, RI, USA, 2023; Section 14; pp. 1–2. [Google Scholar]
  39. Hutchinson, J.W. Bounds and self-consistent estimates for creep of pollycrystalline materials. Proc. R. Soc. Lond. 1976, 348, 101–127. [Google Scholar]
  40. Anand, L.; Kothari, M. A computational procedure for rate-independent crystal plasticity. J. Mech. Phys. Solids 1996, 44, 525–558. [Google Scholar] [CrossRef]
  41. Peirce, D.; Asoro, R.; Needleman, A. Material rate dependence and localized deformation in crystalline solids. Acta Metall. 1983, 31, 1951–1976. [Google Scholar] [CrossRef]
  42. Cowin, S.C. Continuum Mechanics of Anisotropic Materials; Springer: New York, NY, USA, 2013. [Google Scholar]
  43. Dunne, F.; Petrinic, N. Introduction to Computational Plasticity, 2nd ed.; Oxford University Press: Oxford, UK, 2006. [Google Scholar]
  44. Takeuchi, T. Work hardening of copper single crystals with multiple glide orientation. Trans. Jpn. Inst. Met. 1975, 16, 629–640. [Google Scholar] [CrossRef]
  45. Huang, Y. A User-Material Subroutine Incorporating Single Crystal Plasticity in the Abaqus Finite Element Program; Technical Report; Division of Applied Sciences, Harvard University: Boston, MA, USA, 1991. [Google Scholar]
  46. Wulfinghoff, S.; Böhlke, T. Equivalent plastic strain gradient crystal plasticity enhanced power law subroutine. GAMM-Mitteilungen 2013, 36, 134–138. [Google Scholar] [CrossRef]
  47. Ha, C.S.; Plesha, M.E.; Lakes, R.S. Chiral three-dimensional isotropic lattices with negative Poisson’s ratio. Phys. Status Solidi B 2016, 253, 1243–1251. [Google Scholar] [CrossRef]
  48. Lee, M.G.; Wagoner, R.H.; Kim, S.J. Comparative study of single crystal constitutive equations for crystal plasticity finite element analysis. Int. J. Mod. Phys. B 2008, 22, 5388–5393. [Google Scholar] [CrossRef]
  49. Şimşek, Ü. Crystal Plasticity Analysis of Severe Plastic Deformation Processes. Ph.D. Dissertation, Middle East Technical University, Ankara, Turkey, 2024. [Google Scholar]
  50. Şimşek, Ü.; Davut, K.; Miyamoto, H.; Yalçınkaya, T. Comparison of Linear and Nonlinear Twist Extrusion Processes with Crystal Plasticity Finite Element Analysis. Materials 2024, 17, 1139. [Google Scholar] [CrossRef] [PubMed]
  51. Yalçınkaya, T.; Şimşek, Ü.; Miyamoto, H.; Yuasa, M. Numerical analysis of a new nonlinear twist extrusion process. Metals 2019, 9, 513. [Google Scholar] [CrossRef]
  52. Basavaraj, V.; Chakkingal, U.; Kumar, T. Study of channel angle influence on material flow and strain inhomogeneity in equal channel angular pressing using 3D finite element simulation. J. Mater. Process. Technol. 2009, 209, 89–95. [Google Scholar] [CrossRef]
  53. Sarma, G.B.; Dawson, P.R. Effects of interactions among crystals on the inhomogeneous deformations of polycrystals. Acta Metall. 1996, 44, 1937–1953. [Google Scholar] [CrossRef]
  54. Takayama, Y.; Szpunar, J.A. Stored energy and Taylor factor relation in an Al-Mg-Mn alloy sheet worked by continuous cyclic bending. Mater. Trans. 2004, 45, 2316–2325. [Google Scholar] [CrossRef]
  55. Cho, C.H.; Cho, H. Effect of dislocation characteristics on electrical conductivity and mechanical properties of AA 6201 wires. J. Mater. Sci. Technol. 2021, 85, 13–22. [Google Scholar] [CrossRef]
  56. Kumar, S.; Beri, N. Study the effect of process parameters on MRR and TWR in EDM of aluminium and mild steel using copper and brass electrodes. Int. J. Adv. Eng. Res. Stud. 2012, 1, 28–30. [Google Scholar]
  57. Powell, R.W. Correlation of metallic thermal and electrical conductivities for both solid and liquid phases. Int. J. Heat Mass Transf. 1965, 8, 1033–1045. [Google Scholar] [CrossRef]
  58. Ho, K.H.; Newman, S.T. State of the art in electrical discharge machining. Int. J. Mach. Tools Manuf. 2003, 43, 1287–1300. [Google Scholar] [CrossRef]
  59. Demir, E. A Taylor-based plasticity model for orthogonal machining of single crystal FCC materials including frictional effects. Int. J. Adv. Manuf. Technol. 2009, 40, 847–856. [Google Scholar] [CrossRef]
  60. Dolatabadi, J.; Rafiee, M.; Hadad, M.H. Experimental investigation of the effects of cutting parameters on the machinability of ECAP-processed ultrafine-grained copper using tungsten carbide cutting tools. Energy Equip. Syst. 2022, 10, 241–254. [Google Scholar]
  61. Yang, P.; Zhang, H.; Luan, T.; Jin, Y. Molecular dynamics analysis of microstructural deformation mechanisms in single crystal copper undergoing equal channel angular pressing. AIP Adv. 2025, 15, 015110. [Google Scholar] [CrossRef]
  62. Dalan, F.C.; Andreani, G.F.D.L.; Travessa, D.N.; Faizov, I.A.; Faizova, S.; Cardoso, K.R. Effect of ECAP processing on distribution of second phase particles, hardness and electrical conductivity of Cu–0.81Cr–0.07Zr alloy. Trans. Nonferr. Met. Soc. China 2022, 32, 217–232. [Google Scholar] [CrossRef]
  63. Raabe, D.; Roters, F. Using texture components in crystal plasticity finite element simulations. Int. J. Plast. 2004, 20, 339–361. [Google Scholar] [CrossRef]
  64. Shi, L. Thermal and thermoelectric transport in nanostructures and low-dimensional systems. Nanoscale Microscale Thermophys. Eng. 2018, 22, 163–197. [Google Scholar] [CrossRef]
  65. Guo, T.; Tai, X.; Wei, S.; Wang, J.; Jia, Z.; Ding, Y. Microstructure and properties of bulk ultrafine grained Cu1.5Cr0.1Si alloy through ECAP by route C and aging treatment. Crystals 2020, 10, 207. [Google Scholar] [CrossRef]
  66. Fukuda, Y.; Ohishi, K.; Furukawa, M.; Horita, Z.; Langdon, T.G. Influence of crystal orientation on the processing of copper single crystals by ECAP. J. Mater. Sci. 2007, 42, 1501–1511. [Google Scholar] [CrossRef]
  67. Zhao, Z.; Godfrey, A.; Liu, Q.; Hansen, N. Relation between dislocation density and electrical conductivity in deformed copper. Acta Mater. 2002, 50, 2901–2912. [Google Scholar]
  68. Liu, Q.; Lin, J.; Huang, Y.; Zheng, Y. Effect of dislocation structures on electrical conductivity of cold-deformed aluminum. Mater. Sci. Eng. A 2014, 609, 154–160. [Google Scholar]
  69. Belyakov, A.; Kaibyshev, R.; Sakai, T. Texture evolution and its influence on electrical conductivity in copper processed by ECAP. Mater. Charact. 2015, 107, 310–316. [Google Scholar]
  70. Mishra, R.S.; Ma, Z.Y. Severe plastic deformation and thermomechanical effects during ECAP. Mater. Sci. Eng. A 2007, 466, 59–64. [Google Scholar] [CrossRef]
  71. Zhang, Y.; Horita, Z.; Langdon, T.G. Stored energy and recrystallization temperature in high purity copper after ECAP. J. Mater. Sci. 2008, 43, 7363–7369. [Google Scholar] [CrossRef]
  72. Skrotzki, W.; Scheerbaum, N.; Oertel, C.; Brokmeier, H.; Suwas, S.; Toth, L. Recrystallization of high-purity aluminium during equal channel angular pressing. Acta Mater. 2007, 55, 2211–2218. [Google Scholar] [CrossRef]
  73. Kim, H.S. Prediction of Temperature Rise in Equal Channel Angular Pressing. Mater. Trans. 2001, 42, 536–538. [Google Scholar] [CrossRef]
  74. Balasubramanian, S.; Anand, L. Elasto-viscoplastic constitutive equations for polycrystalline FCC materials at low homologous temperatures. J. Mech. Phys. Solids 2002, 50, 427–462. [Google Scholar] [CrossRef]
  75. Humphreys, F.J.; Hatherly, M. Recrystallization and Related Annealing Phenomena; Elsevier: Oxford, UK, 2004. [Google Scholar]
  76. Doherty, R.D.; Hughes, D.A.; Humphreys, F.J.; Jonas, J.J.; Jensen, D.J.; Kassner, M.E.; King, W.E.; McNelley, T.R.; McQueen, H.J.; Rollett, A.D. Current issues in recrystallization: A review. Mater. Sci. Eng. A 1997, 238, 219–274. [Google Scholar] [CrossRef]
  77. Rollett, A.D.; Srolovitz, D.J.; Anderson, M.P.; Grest, G.S. Recrystallization in metallic materials. Acta Mater. 2004, 52, 575–607. [Google Scholar] [CrossRef]
Figure 1. FEA model of the ECAP process and technical drawings of ECAP mold. (a) FEA model of ECAP process, (b) Mesh model of ECAP process, (c) Technical drawing of ECAP mold, (d) ECAP mold parameters.
Figure 1. FEA model of the ECAP process and technical drawings of ECAP mold. (a) FEA model of ECAP process, (b) Mesh model of ECAP process, (c) Technical drawing of ECAP mold, (d) ECAP mold parameters.
Crystals 15 00849 g001
Figure 2. Comparison of mechanical energies during ECAP process.
Figure 2. Comparison of mechanical energies during ECAP process.
Crystals 15 00849 g002
Figure 3. Kinematics of elastic and plastic deformation of single crystals.
Figure 3. Kinematics of elastic and plastic deformation of single crystals.
Crystals 15 00849 g003
Figure 4. (a) Tension boundary conditions, (b) simple shear boundary conditions and (c) pure shear boundary conditions.
Figure 4. (a) Tension boundary conditions, (b) simple shear boundary conditions and (c) pure shear boundary conditions.
Crystals 15 00849 g004
Figure 5. Single copper crystal UMAT and VUMAT CPFEM simulation results vs. experiment results. Under (a) Tension, (b) Simple shear, and (c) Pure shear boundary conditions.
Figure 5. Single copper crystal UMAT and VUMAT CPFEM simulation results vs. experiment results. Under (a) Tension, (b) Simple shear, and (c) Pure shear boundary conditions.
Crystals 15 00849 g005
Figure 6. Algorithm of explicit subroutine user-defined material (VUMAT).
Figure 6. Algorithm of explicit subroutine user-defined material (VUMAT).
Crystals 15 00849 g006
Figure 7. True stress vs true strain curve of copper specimen [51].
Figure 7. True stress vs true strain curve of copper specimen [51].
Crystals 15 00849 g007
Figure 8. The PEEQ distribution after one ECAP pass of initially <100> (left), <110> (right) and <111> (bottom) directed single crystal copper according to the deformation path (A, B and C).
Figure 8. The PEEQ distribution after one ECAP pass of initially <100> (left), <110> (right) and <111> (bottom) directed single crystal copper according to the deformation path (A, B and C).
Crystals 15 00849 g008
Figure 9. The accumulated plastic strain distribution after one ECAP pass of single crystal copper according to initial orientations <100> (up left), <110>(up right) and <111> (bottom left) according to the deformation path (A, B and C) and strain inhomogeneity (bottom right) according to the initial orientation (<100>, <110> and <111>).
Figure 9. The accumulated plastic strain distribution after one ECAP pass of single crystal copper according to initial orientations <100> (up left), <110>(up right) and <111> (bottom left) according to the deformation path (A, B and C) and strain inhomogeneity (bottom right) according to the initial orientation (<100>, <110> and <111>).
Crystals 15 00849 g009
Figure 10. Taylor factor distribution after one ECAP pass of single crystal copper according to <100> (left), <110> (right) and <111> (bottom) directed single crystal copper according to the deformation path (A, B and C).
Figure 10. Taylor factor distribution after one ECAP pass of single crystal copper according to <100> (left), <110> (right) and <111> (bottom) directed single crystal copper according to the deformation path (A, B and C).
Crystals 15 00849 g010
Figure 11. The [111] section Pole Figures after one ECAP pass of single crystal copper according to Texture components (Cube, Goss, Brass, and Shear) (up left), initial orientations <100> (up right), <110> (bottom left) and <111> (bottom right) according to the deformation path (A, B and C).
Figure 11. The [111] section Pole Figures after one ECAP pass of single crystal copper according to Texture components (Cube, Goss, Brass, and Shear) (up left), initial orientations <100> (up right), <110> (bottom left) and <111> (bottom right) according to the deformation path (A, B and C).
Crystals 15 00849 g011
Figure 12. Temperature distribution after one ECAP pass of single crystal copper according to the deformation path (A, B, and C) versus the pole figures of texture components and the after one pass ECAP process of initially <111> oriented single crystal copper.
Figure 12. Temperature distribution after one ECAP pass of single crystal copper according to the deformation path (A, B, and C) versus the pole figures of texture components and the after one pass ECAP process of initially <111> oriented single crystal copper.
Crystals 15 00849 g012
Table 1. CPFEM mesh model parameters of simulated ECAP process.
Table 1. CPFEM mesh model parameters of simulated ECAP process.
ModelElement TypeElement NumberModel Type
MoldLinear tet-C3D4381,163Rigid
PunchExplicit-Hex-C3D8R735Rigid
Polycrystal DummyExplicit-Hex-C3D8R758,462Deformable (VM)
ElectrodeExplicit-Hex-C3D8R2560Deformable (CP)
Table 2. Material parameters for the VUMAT single crystal copper model [50].
Table 2. Material parameters for the VUMAT single crystal copper model [50].
Model C 11 C 12 C 44 τ 0 τ s h 0 γ 0 ˙ nq
(MPa)(MPa)(MPa)(MPa)(MPa)(MPa)( s 1 )
VUMAT CP168,000121,40075,400251151200.001171.4
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

Şimşek, Ü.; Çoğun, C. An Investigation of the Performance of Equal Channel Angular Pressed Copper Electrodes in Electric Discharge Machining. Crystals 2025, 15, 849. https://doi.org/10.3390/cryst15100849

AMA Style

Şimşek Ü, Çoğun C. An Investigation of the Performance of Equal Channel Angular Pressed Copper Electrodes in Electric Discharge Machining. Crystals. 2025; 15(10):849. https://doi.org/10.3390/cryst15100849

Chicago/Turabian Style

Şimşek, Ülke, and Can Çoğun. 2025. "An Investigation of the Performance of Equal Channel Angular Pressed Copper Electrodes in Electric Discharge Machining" Crystals 15, no. 10: 849. https://doi.org/10.3390/cryst15100849

APA Style

Şimşek, Ü., & Çoğun, C. (2025). An Investigation of the Performance of Equal Channel Angular Pressed Copper Electrodes in Electric Discharge Machining. Crystals, 15(10), 849. https://doi.org/10.3390/cryst15100849

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