Next Article in Journal
Hydrogen Damage Behavior of X80 Pipeline Steel Under AC Interference
Previous Article in Journal
C15-Structured Zr-Ti-Fe-Ni-V Alloys for High-Pressure Hydrogen Compression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystal-Plasticity-Based Micro-Mechanical Model for Simulating Plastic Deformation of TC4 Alloy

by
Huanhuan Chen
1,2,
Wei Li
1,
Zhengming Qian
1,
Dong Mi
1,
Yangyang Wu
1,
Siqi Zhang
3,
Can Wu
4,
Keke Li
1,
Tiezheng Tang
3,* and
Dongfeng Li
3,*
1
AECC Hunan Aviation Powerplant Research Institute, Zhuzhou 412002, China
2
Research Institute of Aero-Engine, Beihang University, Beijing 100191, China
3
School of Science, Harbin Institute of Technology, Shenzhen 518055, China
4
Experimental and Innovative Practice Education Center, Harbin Institute of Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(24), 5486; https://doi.org/10.3390/ma18245486
Submission received: 4 November 2025 / Revised: 28 November 2025 / Accepted: 2 December 2025 / Published: 5 December 2025
(This article belongs to the Section Metals and Alloys)

Abstract

Ti-6Al-4V (TC4) alloy is widely used in aerospace and biomedical applications due to its excellent strength-to-weight ratio and corrosion resistance. Its plastic deformation behavior is strongly influenced by its microstructural characteristics, particularly grain size. In this study, a crystal plasticity model incorporating a Hall–Petch relationship was developed to simulate the plastic deformation of TC4, with explicit consideration of the effect of grain size on slip resistance. The model employs a thermally activated flow rule to describe the kinetics of slip systems, enabling accurate prediction of flow stress and strain hardening across different microstructural conditions. The model is calibrated and validated using experimental stress–strain data from uniaxial tensile tests on specimens with varying grain sizes. Simulation results demonstrate that the model successfully captures the grain-size-strengthening effect and predicts the corresponding evolution of local strain heterogeneity. Furthermore, a critical local equivalent plastic strain criterion was established, which effectively predicts the dependence of macroscopic failure strain on grain size. This work provides a physically based computational tool for optimizing TC4 processing parameters and predicting deformation under service conditions.

1. Introduction

The Ti-6Al-4V alloy (TC4) is widely used in fields such as aerospace and medicine due to its advantages, namely, low density, high strength, excellent corrosion resistance, and good biocompatibility [1,2,3]. Microstructural features directly affect the mechanical properties of polycrystalline materials, and this knowledge enables better material design. In addition, investigations at the grain level are based on quantitative analysis of stress states at the microscale and anisotropic plastic flow [4].
Micromechanical models based on crystal plasticity constitutive theory have developed into a relatively complete analytical framework. The crystal plasticity finite-element method (CPFEM) captures the physical characteristics of metal deformation and has a more rigorous physical foundation compared to macroscopic phenomenological models [5]. By calibrating the crystal plasticity model through quasi-static tensile tests conducted on specimens with different orientations, Somlo et al. [6] investigated the anisotropic tensile behavior of additively manufactured TC4. Taking into account thermal stress, back stress, isotropic hardening, and temperature-dependent tensile stress, Gupta et al. [7] investigated the fatigue behavior of TC4 under both intraphase and interphase thermomechanical loading conditions. Arcidiacono et al. [8] adopted the Armstrong–Frederick nonlinear hardening formulation and proposed a microstructure-sensitive fatigue initiation criterion for TC4 to predict the locations of fatigue crack initiation. Based on a simulation framework integrating Cellular Automaton, Chen et al. [9] investigated the anisotropic behavior of TC4 fabricated via the laser-engineering net-splinting process.
It is worth noting that the majority of the crystal plasticity models employed, particularly those in the aforementioned works, are based on power-law formulations. Although widely used due to their simplicity and computational efficiency, power-law models have inherent limitations in accurately capturing thermally activated deformation mechanisms, especially under conditions involving varying temperatures and strain rates. In contrast, exponential-type thermally activated models provide a more physically grounded framework for describing slip system kinetics, making them better suited for high-fidelity simulations of temperature- and rate-dependent plasticity [10]. Yin and Umezawa [11] investigated the temperature-dependent dwell fatigue behavior of TC4 using thermally activated models. A representative volume element (RVE) model was established within a rate-dependent crystal plasticity framework to study the strain rate sensitivity of TC4. Liu and Dunne [12] investigated the role of macrozone crystallographic orientation and morphology in the dwell fatigue behavior of TC4 alloy using an exponential-type flow rule. Later, they further explored the microstructural evolution and fatigue failure behavior of TC4 titanium alloy under cold dwell fatigue conditions [13].
Compared to microstructural experimental characterization, crystal plasticity modeling is better able to isolate and control individual microstructural parameters while offering time-resolved, full-field deformation insights at the slip-system level (which would be experimentally inaccessible), all with significantly reduced costs and time investment [14]. Meanwhile, various constitutive models have also been developed for investigating polycrystalline micro-deformation. However, these models typically employ homogenization assumptions to bridge single-crystal responses with polycrystalline aggregate behavior. For instance, the Taylor model [15] assumes there is uniform strain across all grains, while self-consistent schemes [16,17,18] apply Eshelby’s method [19] to homogenize polycrystals containing ellipsoidal grains. These simplifications limit these methods’ ability to resolve critical microstructural details such as local slip system activation patterns, grain boundary stress concentrations, and deformation incompatibilities between neighboring grains.
Recent advancements in high-resolution experimental techniques such as in situ synchrotron X-ray diffraction and electron backscatter diffraction (EBSD) have enabled direct validation of CPFEM predictions at the grain scale [20,21,22]. These tools allow for more accurate calibration of model parameters and provide valuable insight into intragranular stress evolution and slip system activation. Furthermore, the integration of modeling with advanced experiments facilitates the study of grain boundary effects [23], heterogeneous deformation [24], and the role of crystallographic texture [25]. Thus, physically based CPFEM simulations can be more reliably applied to assess microstructural damage mechanisms.
In this study, we developed a novel thermally activated crystal plasticity framework that uniquely integrates temperature-dependent slip resistance with grain size effects through a Hall–Petch relationship. The proposed model enables accurate prediction of deformation behavior under varying thermomechanical conditions. Systematic experimental validation established quantitative correlations between grain size and strain localization, demonstrating the model’s capability to predict damage initiation in different microstructures. Comparative analysis revealed that the crystal plasticity simulations using EBSD-derived microstructures achieve high predictive accuracy in capturing the macroscopic stress–strain response when validated against experimental tensile curves. The integrated experimental and computational approach presented in this work advances our fundamental understanding of microstructure–property relationships in TC4 while providing a practical tool for deformation prediction in engineering applications.

2. Experimental Procedures

Ti-6Al-4V (TC4) alloy specimens were extracted from mechanical components with an initial diameter of 6 mm and a length of 14 mm. The specimens were supplied by AECC Hunan Aviation Powerplant Research Institute, Zhuzhou, China. The chemical composition of the material is presented in Table 1.
Uniaxial tensile tests were conducted using a Zwick universal testing machine (100 kN capacity) at room temperature (20 °C) (Figure 1a). The tests were performed at a constant strain rate of 10−3 s−1. All tensile specimens with different grain sizes were machined from different locations of a single TC4 disk component. Figure 1a shows the true stress–true strain curves, while Figure 1b presents the relationship between yield strength and elongation at fracture.
Electron backscatter diffraction (EBSD) data were acquired with a step size of 0.25 μm at a resolution of 261 × 197 (Figure 2a). The results indicate that the α phase content exceeded 98%, and thus only the hexagonal close-packed (HCP) α phase was considered in the modeling. Representative samples have average equivalent circular diameters (D) of 7.4 μm (ranging from 0.9 to 33.9 μm), 4.8 μm (ranging from 0.9 to 27.8 μm), and 2.6 μm (ranging from 0.7 to 24.9 μm) (Figure 2b shows the D = 4.8 μm case).
As shown in Figure 1a, the three true stress–true strain curves reach their maximum true stress at strains of 0.028, 0.030, and 0.033, respectively; these points are marked with solid circles. All true strain values reported are dimensionless. A clear grain-size dependence is observed—the finer the grain size, the larger the strain corresponding to the maximum true stress. This trend suggests that finer-grained microstructures can accommodate greater uniform plastic deformation prior to the onset of damage, whereas coarser grains exhibit earlier strain localization and reduced strain-hardening capability.

3. Micromechanical Model for Polycrystals

Polycrystal models utilizing the finite-element method enable the discrete representation of grain morphology and continuum-scale material deformation while explicitly accounting for intergranular interactions to predict the aggregate behavior of polycrystalline materials. When subjected to macroscopic loading conditions, individual grains undergo complex deformation modes dictated by mechanical constraints imposed by their local grain neighborhood. Under the influence of specific macroscopic deformation paths, certain grains may exhibit pronounced strain localization arising from incompatibility with adjacent grains. Such localized finite deformations can induce substantial lattice distortions, thereby modifying crystallographic orientation distributions (texture evolution) and influencing subsequent mechanical behavior through geometric effects. Thus, the incorporation of finite-strain single-crystal constitutive relations becomes essential for accurate microscale representation within polycrystalline simulations.

3.1. Kinetic Formulation for a Single Crystal

The mechanical deformation of crystalline materials is typically described through a combination of elastic lattice distortions, crystalline rotation, and dislocation-induced slip across multiple slip systems [26,27]. A general description of plastic deformation starts typically from the multiplicative decomposition of the deformation gradient as proposed by Lee [28]:
F = F e F p ,
where F e is the elastic part of the deformation gradient ( F ), and F p is the plastic component. The constitutive relationship between stress and strain can be formulated using the second Piola–Kirchhoff stress tensor T * (defined as T * = F e F e 1 σ F e T ) and its work-conjugate measure, the Lagrangian Green strain tensor E e (given by E e = 1 2 F e T F e I ) as
T * = C : E e ,
where C is the fourth-order anisotropic elasticity tensor, and σ is the Cauchy stress tensor. In this formulation, the tensor exponent ( · ) T denotes the inverse transpose operation (i.e., A T = ( A 1 ) T ), and the operator: represents the inner product between two second-order tensors. To represent the plastic flow, the plastic velocity gradient, L p , is introduced as
L p = F ˙ p F p 1
A superimposed dot denotes the time derivative of the variable. The plastic component of the velocity gradient can be expressed as a linear superposition of slip rates across active crystallographic slip systems [29]:
L p = α = 1 N γ ˙ α m α n α
where γ ˙ α denotes the slip rate, m α and n α are the slip direction and slip plane normal of the slip system, α , and N is the total number of slip systems. The operator “⊗” is a tensor product. In this paper, a total of 12 slip systems, including 3 basal slip systems, 3 prismatic slip systems, and 6 pyramidal a slip systems, are taken into account. The specific slip systems and slip directions are listed in Table 2.
The constitutive model employs a thermally activated plasticity formulation originally developed by Busso and McClintock [30]:
γ ˙ α = γ ˙ 0 exp F k θ 1 | τ α | S α τ 0 p q sgn ( τ α )
where θ is the thermodynamic temperature; k is the Boltzmann constant; F represents the Helmholtz activation free energy, defined as the minimum energy required to surmount localized barriers in the absence of external stress; γ ˙ 0 , p, and q are the flow rule parameters; and the term S α represents the total slip resistance with respect to dislocation motion. The operator “ x ” assumes a value of x when x > 0 and is zero otherwise. τ α denotes the resolved shear stress defined by
τ a = F e T F e T * : m α n α
τ 0 is the lattice friction stress at 0 K. The slip resistance of the slip system S α is governed by
S ˙ α = β = 1 N h s α β S s a t S β S s a t S 0 γ ˙ β
where S 0 is the initial slip resistance, S s a t is the saturated slip resistance, and h s α β is the hardening matrix. In order to capture the grain size effect, a Hall–Petch relation [31,32] is applied for S 0 as follows:
S 0 = S s + K D ,
where S s corresponds to the slip resistance for a sufficiently large grain size, K is the Hall–Petch coefficient, and D is the average grain size. h s α β is defined as follows:
h s α β = h 0 ω 1 + 1 ω 2 δ α β
where δ α β is Kronecker’s Delta, assuming a value of 1 when α = β and zero otherwise. h 0 is the hardening constant, and ω 1 and ω 2 reflect the hardening behavior of the material. In this study, ω 1 = ω 2 = 1 , which corresponds to Taylor hardening. An effective scalar representation of the combined slip activity in all crystallographic systems is provided by the accumulated equivalent plastic strain:
ε ˜ p = 0 t 2 3 D p : D p d τ
where t denotes the current time, and τ is the integration variable. D p is the plastic deformation rate tensor given by
D p = 1 2 L p + L p T
The proposed constitutive model was incorporated into the commercial finite-element software ABAQUS 6.14 through a user-defined material subroutine, utilizing an implicit backward Euler integration scheme. Detailed formulations of the model can be found in Refs. [33,34].

3.2. Representative Volume Elements

Finite-element representative volume element (RVE) models of the microstructure were constructed based on the EBSD (Figure 3a), as shown in Figure 3b. The RVE model applies the grain orientations from the EBSD data for modeling and subsequent calculations.The EBSD-derived microstructure was reconstructed as an RVE with single-element discretization in the out-of-plane direction, maintaining full crystallographic orientation data while enabling efficient computation. In the simulated tests discussed later, the load is applied in the e 2 direction. The Euler angle ϕ 2 represents the final rotation about the crystal’s c-axis. The model employs 43,200 hexahedral elements (C3D8 type), containing 120 grains.

4. Results and Discussion

4.1. Calibrations of the Material Parameters in the Crystal Plasticity Model

The elastic matrix for HCP can be illustrated as follows:
C = C 11 C 12 C 13 0 0 0 C 12 C 11 C 13 0 0 0 C 13 C 13 C 33 0 0 0 0 0 0 C 44 0 0 0 0 0 0 C 66 0 0 0 0 0 0 C 66
where C 44 = 0.5 ( C 11 C 12 ) . The components C i j and the parameters for the CPFEM calibrated from the experimental data are listed in Table 3.
The initial slip resistance for the model whose equivalent circular diameter D = 7.4 μm was determined to be 293 MPa. Figure 4 shows the results simulated by the RVE model.
The numerical results demonstrate that the EBSD-based CPFEM exhibits excellent agreement with the experimental tensile data, albeit with subtle differences in predictive accuracy. It should be noted that the crystal plasticity model employed in this work does not account for damage initiation and subsequent evolution, and thus the comparison with experimental uniaxial tensile data becomes unreliable once the experimental curve reaches the UTS point.

4.2. Grain-Size Model Validation

To validate the integrated Hall–Petch formulation within the crystal plasticity framework, the calibrated model was extended to predict the mechanical response of TC4 specimens with finer grain sizes (D = 2.6 μm and 7.4 μm). Using the Hall–Petch coefficient, we calculated the initial slip resistance S 0 for each grain size: 311 MPa for 2.6 μm and 283 MPa for 7.4 μm.
Figure 5a compares the simulated stress–strain curves using the derived S 0 values against experimental data for all three grain sizes. Figure 5b illustrates the yield strength. The CPFEM simulations accurately reproduce the experimental trends across the full range of grain sizes, accurately capturing both the increased yield strength and strain-hardening behavior at reduced grain dimensions. This consistency validates two critical aspects: (1) the physical basis of the thermally activated crystal plasticity model for describing size-dependent plasticity and (2) the appropriateness of the selected Hall–Petch parameters for TC4 alloy.
The numerical results reveal that decreasing grain size from 7.4 μm to 2.6 μm increases yield strength. This agreement confirms that the Hall–Petch relationship, when coupled with the crystal plasticity framework, effectively captures the grain-boundary-strengthening mechanism in TC4. Furthermore, the model correctly predicts the influence of grain size on post-yield behavior—finer microstructures exhibit enhanced strain hardening due to greater dislocation accumulation at grain boundaries.

4.3. Effect of Grain Size

Figure 6a shows the simulated distribution of the accumulated equivalent plastic strain for the alloy with D = 4.8 μm at the yield point (in the RVE (Figure 3b)). A target line was strategically positioned to traverse the region of maximum local cumulative equivalent plastic strain concentration, allowing for quantitative comparison of strain localization patterns across different microstructures (Figure 6b).
Fine-grained specimens exhibited lower cumulative equivalent plastic strain values along the target path, suggesting a more uniform distribution of plastic deformation. In contrast, coarser-grained materials showed more pronounced strain localization, along with higher maximum cumulative equivalent plastic strain values. However, as the grain size increases, the rate of increase in the maximum local cumulative equivalent plastic strain gradually slows, indicating that the extent of strain localization tends to saturate with further grain coarsening.

4.4. Failure Prediction

Accurate prediction of failure initiation is crucial for the reliability assessment of metallic components under monotonic loading. While macroscopic tensile tests provide the overall stress–strain response, they lack the ability to directly characterize the local damage accumulation that precedes fracture. Classical criteria, such as the ultimate tensile strength, identify the onset of necking but do not necessarily correspond to the critical local material state for failure. To bridge this gap between global deformation and local failure, we established a novel failure prediction methodology by synergistically integrating experimental data with computational simulations. The core of this approach is to define a critical local failure criterion based on the maximum equivalent plastic strain (PEEQ) at the microstructural level and then calibrate it against the point of maximum true stress measured in uniaxial tensile tests.
The strain at the maximum stress point for the alloy with D = 4.8 μm ϵf was identified as 5.01% of the experimental true stress–strain curve in Figure 1a. A finite element simulation of the tensile test for this grain size was conducted up to this macroscopic strain level. The corresponding maximum local PEEQ value in the simulation at this instant was found to be 0.14. This value of 0.14 was subsequently defined as the critical failure criterion ε ˜ c r i t p . This criterion was then applied to predict the failure strains for the other grain sizes by determining the macroscopic strain at which the maximum local PEEQ in their respective simulations reached 0.14, i.e.,
ε ˜ f p = ε ˜ c r i t p
The correlation between the simulated and experimentally measured failure strains across all grain sizes is summarized in Figure 7.
The comparison between the simulation and experimental results demonstrates favorable agreement in predicting the grain size dependence of the failure strain. A clear trend was observed, where the failure strain decreases with increasing grain size. This inverse relationship can be attributed to the effect of grain size governed by strain gradients. In coarser-grained microstructures, stronger constraints at grain boundaries lead to higher concentrations of geometrically necessary dislocations, which accelerate local damage accumulation.

5. Summary and Conclusions

This work investigated the plastic deformation and failure behavior of Ti-6Al-4V (TC4) alloy of different grain sizes by integrating uniaxial tensile tests, electron backscatter diffraction (EBSD) characterization, and crystal plasticity finite element simulations. The primary findings and conclusions are as follows:
(1)
A thermally activated crystal plasticity model, which explicitly incorporates a Hall–Petch relationship for the initial slip resistance, was successfully developed and calibrated. The model demonstrates high fidelity in predicting the flow stress and strain-hardening behavior of TC4 alloy under uniaxial tension at room temperature.
(2)
The model quantitatively captures the grain-size-strengthening effect. Simulations and experiments consistently showed that a reduction in grain size from 7.4 µm to 2.6 µm leads to a significant increase in yield strength, validating the integrated Hall–Petch parameters.
(3)
Grain size has a profound influence on strain localization. The numerical results reveal that coarser-grained microstructures develop more intense and heterogeneous local plastic strain concentrations, which act as potential sites for damage initiation.
(4)
A local failure initiation criterion, defined by a critical value of accumulated equivalent plastic strain (PEEQcrit = 0.14), was established based on the experimental ultimate tensile strength. This criterion successfully predicts the macroscopic failure strain for different grain sizes.
(5)
The predicted failure strain decreases with an increasing grain size. This trend is associated with strain localization phenomena that lead to accelerated damage accumulation in coarser-grained microstructures.
(6)
It is important to note that the Hall–Petch relationship is typically valid for grain sizes ranging from approximately 1 to 30 µm in titanium alloys. The current model was calibrated and validated within a specific subset (2.6–7.4 µm) of this established range, where its predictive accuracy is high. Extrapolation to grain sizes significantly outside this validated range (particularly to ultra-fine grains < 1 µm or very coarse grains > 30 µm) may require the incorporation of additional physical mechanisms into the constitutive model.

Author Contributions

Methodology, D.L.; Software, D.L.; Validation, H.C., W.L., Z.Q., D.M., Y.W. and K.L.; Investigation, H.C., S.Z. and C.W.; Writing—original draft, T.T.; Writing—review & editing, D.L.; Funding acquisition, H.C., W.L., Z.Q., D.M., Y.W. and K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bambach, M.; Sizova, I.; Szyndler, J.; Bennett, J.; Hyatt, G.; Cao, J.; Papke, T.; Merklein, M. On the hot deformation behavior of Ti-6Al-4V made by additive manufacturing. J. Mater. Process. Technol. 2021, 288, 116840. [Google Scholar] [CrossRef]
  2. Murr, L.E.; Esquivel, E.V.; Quinones, S.A.; Gaytan, S.M.; Lopez, M.I.; Martinez, E.Y.; Medina, F.; Hernandez, D.H.; Martinez, E.; Martinez, J.L.; et al. Microstructures mechanical properties of electron beam-rapid manufactured Ti-6Al-4V biomedical prototypes compared to wrought Ti-6Al-4V. Mater. Charact. 2009, 60, 96–105. [Google Scholar] [CrossRef]
  3. Zhao, R.; Chen, C.; Wang, W.; Cao, T.; Shuai, S.; Xu, S.; Hu, T.; Liao, H.; Wang, J.; Ren, Z. On the role of volumetric energy density in the microstructure and mechanical properties of laser powder bed fusion Ti-6Al-4V alloy. Addit. Manuf. 2022, 51, 102605. [Google Scholar] [CrossRef]
  4. Li, D.F.; O’Dowd, N.P. On the evolution of lattice deformation in austenitic stainless steels—The role of work hardening at finite strains. J. Mech. Phys. Solids 2011, 59, 2421–2441. [Google Scholar] [CrossRef]
  5. Ling, C.; Besson, J.; Forest, S.; Tanguy, B.; Latourte, F.; Bosso, E. An elastoviscoplastic model for porous single crystals at finite strains and its assessment based on unit cell simulations. Int. J. Plast. 2016, 84, 58–87. [Google Scholar] [CrossRef]
  6. Somlo, K.; Poulios, K.; Funch, C.V.; Niordson, C.F. Anisotropic tensile behaviour of additively manufactured Ti-6Al-4V simulated with crystal plasticity. Mech. Mater. 2021, 162, 104034. [Google Scholar] [CrossRef]
  7. Gupta, R.; Kazim, S.M.; Prasad, K.; Chakraborty, P. Crystal plasticity modeling of a titanium alloy under thermo-mechanical fatigue. Mech. Res. Commun. 2021, 111, 103647. [Google Scholar] [CrossRef]
  8. Arcidiacono, M.F.; Violatos, I.; Rahimi, S. Predicting Fatigue Crack Initiation in Milled Aerospace Grade Ti-6Al-4Vparts Using CPFEM. In Proceedings of the 9th Engineering Integrity Society International Conference on Durability & Fatigue, Cambridge, UK, 19–21 June 2024; pp. 115–126. [Google Scholar]
  9. Chen, X.; Zhang, J.; Hu, L.; Ji, D. Investigation on Anisotropic Behavior of Additively Manufactured Ti-6Al-4V Based on Cellular Automaton CPFEM. Met. Mater. Int. 2025, 31, 2578–2597. [Google Scholar] [CrossRef]
  10. Roters, F.; Eisenlohr, P.; Hantcherli, L.; Tjahjanto, D.D.; Bieler, T.R.; Raabe, D. Overview of constitutive laws, kinematics, homogenization and multiscale methods in crystal plasticity finite-element modeling: Theory, experiments, applications. Acta Mater. 2010, 58, 1152–1211. [Google Scholar] [CrossRef]
  11. Yin, L.; Umezawa, O. Crystal plasticity analysis of temperature-sensitive dwell fatigue in Ti-6Al-4V titanium alloy for an aero-engine fan disc. Int. J. Fatigue 2022, 156, 106688. [Google Scholar] [CrossRef]
  12. Liu, Y.; Dunne, F.P.E. The mechanistic link between macrozones and dwell fatigue in titanium alloys. Int. J. Fatigue 2021, 142, 105971. [Google Scholar] [CrossRef]
  13. Liu, Y.; Adande, S.; Britton, T.B.; Dunne, F.P.E. Cold dwell fatigue analyses integrating crystal-level strain rate sensitivity and microstructural heterogeneity. Int. J. Fatigue 2021, 151, 106398. [Google Scholar] [CrossRef]
  14. Roters, F. Advanced Material Models for the Crystal Plasticity Finite Element Method: Development of a General CPFEM Framework. Ph.D. Thesis, RWTH Aachen University, Aachen, Germany, 2011. [Google Scholar]
  15. Taylor, G.I. Plastic strain in metals. J. Inst. Met. 1938, 62, 307–324. [Google Scholar]
  16. Hutchinson, J.W.; Hill, R. Elastic-plastic behaviour of polycrystalline metals and composites. Proc. R. Soc. Lond. A Math. Phys. Sci. 1997, 319, 247–272. [Google Scholar]
  17. Neil, C.J.; Wollmershauser, J.A.; Clausen, B.; Tomé, C.N.; Agnew, S.R. Modeling lattice strain evolution at finite strains and experimental verification for copper and stainless steel using in situ neutron diffraction. Int. J. Plast. 2010, 26, 1772–1791. [Google Scholar] [CrossRef]
  18. Wang, H.; Wu, P.D.; Tomé, C.N.; Huang, Y. A finite strain elastic–viscoplastic self-consistent model for polycrystalline materials. J. Mech. Phys. Solids 2010, 58, 594–612. [Google Scholar] [CrossRef]
  19. Eshelby, J.D.; Peierls, R.E. The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1957, 241, 376–396. [Google Scholar]
  20. Cai, W.; Sun, C.; Wang, C.; Qian, L.; Li, Y.; Fu, M.W. Modelling of the intergranular fracture of TWIP steels working at high temperature by using CZM–CPFE method. Int. J. Plast. 2022, 156, 103366. [Google Scholar] [CrossRef]
  21. Song, M.; Geng, S.; Qiu, Y.; Xu, B.; Wang, Y.; Jiang, P.; Hu, Y.; Li, S. In-situ EBSD-DIC simulation of microstructure evolution of aluminum alloy welds. Int. J. Mech. Sci. 2024, 284, 109741. [Google Scholar] [CrossRef]
  22. Chakrabarty, A.; Chakraborty, P.; Jain, R.; Sahu, V.K.; Gurao, N.P.; Bar, H.N.; Khutia, N. Influence of Scanning and Building Strategies on the Deformation Behavior of Additively Manufactured AlSi10Mg: CPFEM and Finite Element Studies. Met. Mater. Int. 2023, 29, 2978–3008. [Google Scholar] [CrossRef]
  23. Han, W.Z.; Demkowicz, M.J.; Fu, E.G.; Wang, Y.Q.; Misra, A. Effect of grain boundary character on sink efficiency. Acta Mater. 2012, 60, 6341–6351. [Google Scholar] [CrossRef]
  24. Allain-Bonasso, N.; Wagner, F.; Berbenni, S.; Field, D.P. Astudy of the heterogeneity of plastic deformation in IF steel by EBSD. Mater. Sci. Eng. A 2012, 548, 56–63. [Google Scholar] [CrossRef]
  25. Lin, H.P.; Ng, T.-S.; Kuo, J.-C.; Chen, Y.-C.; Chen, C.-L.; Ding, S.-X. EBSD study on crystallographic texture and microstructure development of cold-rolled FePd alloy. Mater. Charact. 2014, 93, 163–172. [Google Scholar] [CrossRef]
  26. Balasubramanian, S.; Anand, L. Plasticity of initially textured hexagonal polycrystals at high homologous temperatures: Application to titanium. Acta Mater. 2002, 50, 133–148. [Google Scholar] [CrossRef]
  27. Kothari, M.; Anand, L. Elasto-viscoplastic constitutive equations for polycrystalline metals: Application to tantalum. J. Mech. Phys. Solids 1998, 46, 51–83. [Google Scholar] [CrossRef]
  28. Lee, E.H. Elastic-Plastic Deformation at Finite Strains. J. Appl. Mech. 1969, 36, 1–6. [Google Scholar] [CrossRef]
  29. Asaro, R.J.; Rice, J.R. Strain localization in ductile single crystals. J. Mech. Phys. Solids 1977, 25, 309–338. [Google Scholar] [CrossRef]
  30. Busso, E.P.; McClintock, F.A. A dislocation mechanics-based crystallographic model of a B2-type intermetallic alloy. Int. J. Plast. 1996, 12, 1–28. [Google Scholar] [CrossRef]
  31. Venkatramani, G.; Ghosh, S.; Mills, M. A size-dependent crystal plasticity finite-element model for creep and load shedding in polycrystalline titanium alloys. Acta Mater. 2007, 55, 3971–3986. [Google Scholar] [CrossRef]
  32. Ghosh, S.; Chakraborty, P. Microstructure and load sensitive fatigue crack nucleation in Ti-6242 using accelerated crystal plasticity FEM simulations. Int. J. Fatigue 2013, 48, 231–246. [Google Scholar] [CrossRef]
  33. Yin, J.-D.; Wang, L.; Zhu, B.-Y.; Jin, X.; Li, C.-F.; Busso, E.P.; Li, D.-F. Experimental and micromechanical investigation of precipitate size effects on the creep behaviour of a high chromium martensitic steel. Eur. J. Mech. A/Solids 2025, 111, 105591. [Google Scholar] [CrossRef]
  34. Guo, H.-J.; Ling, C.; Li, D.-F.; Li, C.-F.; Sun, Y.; Busso, E.P. A data-driven approach to predicting the anisotropic mechanical behaviour of voided single crystals. J. Mech. Phys. Solids 2022, 159, 104700. [Google Scholar] [CrossRef]
  35. Simmons, G. Single Crystal Elastic Constants and Caluculated Aggregate Properties: A Handbook; MIT Press: Cambridge, MA, USA, 1971. [Google Scholar]
  36. Benmessaoud, F.; Cheikh, M.; Velay, V.; Vidal, V.; Matsumoto, H. Role of grain size and crystallographic texture on tensile behavior induced by sliding mechanism in Ti-6Al-4V alloy. Mater. Sci. Eng. A 2020, 774, 138835. [Google Scholar] [CrossRef]
Figure 1. (a) Experimental uniaxial tensile true stress–strain curves, with solid circles marking the ultimate tensile strength points. (b) Yield strength versus elongation at fracture for different grain sizes.
Figure 1. (a) Experimental uniaxial tensile true stress–strain curves, with solid circles marking the ultimate tensile strength points. (b) Yield strength versus elongation at fracture for different grain sizes.
Materials 18 05486 g001
Figure 2. Microstructural of TC4 samples: (a) inverse pole figure map and phase distribution determined from electron backscatter diffraction, and (b) statistical distribution of equivalent circular diameter for the α -phase grains.
Figure 2. Microstructural of TC4 samples: (a) inverse pole figure map and phase distribution determined from electron backscatter diffraction, and (b) statistical distribution of equivalent circular diameter for the α -phase grains.
Materials 18 05486 g002
Figure 3. Construction of the representative volume element model: (a) experimentally obtained electron backscatter diffraction inverse-pole figure map, and (b) corresponding finite-element RVE model with crystallographic orientations assigned from the EBSD data.
Figure 3. Construction of the representative volume element model: (a) experimentally obtained electron backscatter diffraction inverse-pole figure map, and (b) corresponding finite-element RVE model with crystallographic orientations assigned from the EBSD data.
Materials 18 05486 g003
Figure 4. Validation of the crystal plasticity model against experimental data for the sample with an average grain size of 4.8 µm: Comparison between the experimental tensile curve and the simulation results from the EBSD-based representative volume element (RVE) model.
Figure 4. Validation of the crystal plasticity model against experimental data for the sample with an average grain size of 4.8 µm: Comparison between the experimental tensile curve and the simulation results from the EBSD-based representative volume element (RVE) model.
Materials 18 05486 g004
Figure 5. Mechanical response of TC4 alloy with varying grain sizes: (a) experimental and simulated true stress–strain curves, and (b) Comparison of yield strength between experiments and model predictions as a function of grain size.
Figure 5. Mechanical response of TC4 alloy with varying grain sizes: (a) experimental and simulated true stress–strain curves, and (b) Comparison of yield strength between experiments and model predictions as a function of grain size.
Materials 18 05486 g005
Figure 6. (a) Full field distribution of accumulated equivalent plastic strain (PEEQ) at yield across the entire RVE (D = 4.8 µm). (b) Local PEEQ values along a target line for different grain sizes, showing the effect of grain size on local strain concentration.
Figure 6. (a) Full field distribution of accumulated equivalent plastic strain (PEEQ) at yield across the entire RVE (D = 4.8 µm). (b) Local PEEQ values along a target line for different grain sizes, showing the effect of grain size on local strain concentration.
Materials 18 05486 g006
Figure 7. The macroscopic strain required for the local maximum cumulative equivalent plastic strain (PEEQ) to reach the critical value of 5.01%, predicted by finite element simulations for alloys with different grain sizes, and the comparison with the experimental data.
Figure 7. The macroscopic strain required for the local maximum cumulative equivalent plastic strain (PEEQ) to reach the critical value of 5.01%, predicted by finite element simulations for alloys with different grain sizes, and the comparison with the experimental data.
Materials 18 05486 g007
Table 1. Chemical composition of TC4 alloy (wt.%).
Table 1. Chemical composition of TC4 alloy (wt.%).
ElementAlVFeOCNTi
Content5.5–6.53.5–4.5≤0.25≤0.20≤0.08≤0.05Bal.
Table 2. Slip systems in HCP crystals.
Table 2. Slip systems in HCP crystals.
Slip SystemsSlip PlaneSlip Direction
BasalMaterials 18 05486 i001 { 0001 } 11 2 ¯ 0
{ 0001 } 1 2 ¯ 10
{ 0001 } 2 1 ¯ 1 ¯ 0
PrismaticMaterials 18 05486 i002 { 1 1 ¯ 00 } 11 2 ¯ 0
{ 10 1 ¯ 0 } 1 2 ¯ 10
{ 01 1 ¯ 0 } 2 1 ¯ 1 ¯ 0
Pyramidal a Materials 18 05486 i003 { 10 1 ¯ 1 } 1 ¯ 2 1 ¯ 0
{ 01 1 ¯ 1 } 2 ¯ 110
{ 1 ¯ 101 } 1 ¯ 1 ¯ 20
{ 1 ¯ 011 } 1 2 ¯ 10
{ 0 1 ¯ 11 } 2 1 ¯ 1 ¯ 0
{ 1 1 ¯ 01 } 11 2 ¯ 0
Table 3. Material parameters used in the model developed in this study.
Table 3. Material parameters used in the model developed in this study.
ParametersValuesUnitsSources
C 11 162.4GPa[35]
C 12 92GPa
C 13 69GPa
C 33 180.7GPa
C 66 46.7GPa
p0.45The present work
q1.9
F160kJ/mol
τ 0 300MPa
γ ˙ 0 1500s−1
S s a t 350MPa
h s 100MPa
S s 242.7MPa
K109.3MPa µ m [36]
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

Chen, H.; Li, W.; Qian, Z.; Mi, D.; Wu, Y.; Zhang, S.; Wu, C.; Li, K.; Tang, T.; Li, D. Crystal-Plasticity-Based Micro-Mechanical Model for Simulating Plastic Deformation of TC4 Alloy. Materials 2025, 18, 5486. https://doi.org/10.3390/ma18245486

AMA Style

Chen H, Li W, Qian Z, Mi D, Wu Y, Zhang S, Wu C, Li K, Tang T, Li D. Crystal-Plasticity-Based Micro-Mechanical Model for Simulating Plastic Deformation of TC4 Alloy. Materials. 2025; 18(24):5486. https://doi.org/10.3390/ma18245486

Chicago/Turabian Style

Chen, Huanhuan, Wei Li, Zhengming Qian, Dong Mi, Yangyang Wu, Siqi Zhang, Can Wu, Keke Li, Tiezheng Tang, and Dongfeng Li. 2025. "Crystal-Plasticity-Based Micro-Mechanical Model for Simulating Plastic Deformation of TC4 Alloy" Materials 18, no. 24: 5486. https://doi.org/10.3390/ma18245486

APA Style

Chen, H., Li, W., Qian, Z., Mi, D., Wu, Y., Zhang, S., Wu, C., Li, K., Tang, T., & Li, D. (2025). Crystal-Plasticity-Based Micro-Mechanical Model for Simulating Plastic Deformation of TC4 Alloy. Materials, 18(24), 5486. https://doi.org/10.3390/ma18245486

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