Next Article in Journal
Possible Approaches to Studying the Influence of Magnetic Fields and Mechanical Effects on the Physicochemical Properties of Aqueous IgG Colloids
Previous Article in Journal
Deep-Learning-Based Approach for Automated Detection of Irregular Walking Surfaces for Walkability Assessment with Wearable Sensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure Evolution and Constitutive Modelling of Deformation Behavior for Al-Mg-Si-Cu-Sc-Zr Alloy Processed with Isothermal Multidirectional Forging

by
Andrey G. Mochugovskiy
1,*,
Ludmila Yu. Kaplanskaya
1,
Ahmed O. Mosleh
2,
Valeria V. Palacheva
1 and
Anastasia V. Mikhaylovskaya
1
1
Physical Metallurgy of Non-Ferrous Metals, National University of Science and Technology “MISIS,” Leninsky Prospekt, 4, 119049 Moscow, Russia
2
Mechanical Engineering Department, Faculty of Engineering at Shoubra, Benha University, Cairo 11629, Egypt
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(24), 13054; https://doi.org/10.3390/app132413054
Submission received: 17 October 2023 / Revised: 14 November 2023 / Accepted: 29 November 2023 / Published: 7 December 2023
(This article belongs to the Section Materials Science and Engineering)

Abstract

:
This research is devoted to the microstructure evolution and deformation behavior of the Al-1.2Mg-0.7Si-1.0Cu-0.1Sc-0.2Zr alloy during the isothermal multidirectional forging (MDF) in a large cumulative strain and temperature range. The structure investigation of the studied alloy revealed several phases precipitated during solidification, among which θ(Al2Cu), Q(Al5Cu2Mg8Si6), Mg2Si, Sc-bearing W(AlScCu) and V(AlSi2Sc2) phases were observed. The MDF at 150–350 °C and a maximum cumulative strain of 14.4 significantly refined grain structure providing a mean grain size of 1.2–2.1 µm. The L12 structured Al3(Sc,Zr) dispersoids with a mean size of 10 ± 1 nm were formed during two-step homogenization annealing. Due to Zener pinning of the nanoscale dispersoids and fine-grained structure, the alloy exhibited near-superplastic behavior in a temperature range of 460–500 °C and strain rate range of 2 × 10−3–1 × 10−2 s−1 with the maximum elongation to failure of ~300%. After a strengthening heat treatment, the forged alloy exhibited the yield strength of 326 ± 5 MPa, ultimate tensile strength of 366 ± 5 MPa, and elongation of 10 ± 3%. The hot deformation behavior was described using the Arrhenius type model. The developed model demonstrated high predictability accuracy with a maximum average absolute relative error of 6.6%.

1. Introduction

The fine-grained Al-based alloys attract regular interest due to their combination of properties, among which are increased plasticity, formability, and strength [1]. Grain refinement in Al-based alloys is currently the main focus of multiple research studies that reflect relevant industrial needs [2]. The popular Al-based heat-treatable 6xxx-type Al-Mg-Si-based alloys exhibit good corrosion resistance and outstanding strength-to-weight ratio [3,4]. Moreover, the Al-Mg-Si-based alloys demonstrate good formability at low temperatures [5]. However, at elevated temperatures, owing to the low alloyed solid solution, the 6xxx-type alloys are prone to a grain growth that complicates the formation of a fine-grained structure [6]. Thus, the intense dynamic grain growth of these alloys leads to poor superplasticity [7,8]. The grain growth leads to structure inhomogeneity that provides inhomogeneous flow and results in a decrease in uniform elongation.
A fine-grained structure is commonly achieved through a severe plastic deformation (SPD) method that has been proven to be universal and effective for various metals and alloys [9]. The SPD is realized by various techniques that are united by the ultra-high plastic strain accumulated during material processing. Popular SPD techniques, such as the equal channel angular pressing (ECAP) [10], friction stir processing (FSP) [11] and high pression torsion (HPT) [12], have demonstrated high efficiency in grain refinement and mechanical properties improvement. However, the high energy consumption of the previously mentioned SPD approaches limits the sample size and industrial application. A promising SPD technique that allows for processing bulk samples while maintaining their geometry is isothermal multidirectional forging (MDF) [13,14,15,16,17]. MDF has been effectively used for achieving outstanding grain-refinement in the Al-based alloys [18,19], by which a grain size of 0.5–2 µm was achieved in 5xxx type alloys [20,21]. A similar grain size was observed in 6xxx-type alloys after multidirectional forging at cryogen temperatures [22,23].
Although the MDF treatment effectively refines the grains, the thermal stability of the refined grain structure requires special approach. To suppress grain growth, Al alloys are usually doped with small additions of transition metals (TM) and rare-earth elements (RE) [22,24,25], including Sc and Zr which have proven their significant effectiveness [26,27]. These metals exhibit low equilibrium solubility in aluminum. During rapid solidification, the solubility of Sc and Zr can be increased to allow the formation of a supersaturated solid solution. Subsequent heat treatment provides the decomposition of supersaturated solid solution with a precipitation of fine nanoscale dispersoids of Al3Sc or Al3Zr phases with an L12 crystal structure type [28,29,30]. The co-adding of Sc and Zr results in a formation of a complex Al3(Sc,Zr) phase with core-shell structure [31,32] that is distinguished by an increased potential resistance to coarsening at elevated temperatures [33]. The nanoscale dispersoids provide for the Zener pinning effect that suppress static and dynamic grain growth [34]. The positive effect of Sc and Zr co-addition on grain stability at elevated temperatures has been multiply reviewed for thermomechanical-treated Al alloys [35,36]. As a result, the improved superplastic behavior due to the Zener pinning effect has been previously demonstrated in 2xxx- [37], 5xxx- [38,39,40,41], 6xxx- [42,43], and 7xxx-type [44,45] Al alloys. The effectiveness of microalloying directly depends on the particle parameters. Following the Smith–Zener Equation (1), the pinning pressure increases with increasing particle volume fraction and decreasing particle radius [46].
P S Z = 3 f γ 2 r ,
where f is the volume fraction, r is the radius of particles, and γ is the grain boundary energy.
The parameters of heat and thermomechanical treatment play a principal role in dispersoid precipitation. Different researchers have observed that the high number density and nanoscale size of dispersoid in Sc- and Zr-bearing Al alloys can be achieved through a low temperature annealing [30,47] or two-step annealing with a low temperature first stage at 300–400 °C [48,49]. The first annealing step intensifies the nucleation of the dispersoids, whereas a high temperature second step accelerates the growth of them. Thus, multiple studies confirmed the effectiveness of the SPD method and microalloying for grain refinement in Al-based alloys. However, the effect of the MDF technique on the grain refinement and structural stability of low alloyed Al-Mg-Si-based alloys is still poorly studied.
In this context, the current study had two aims: (1) to achieve a refined grain structure of an Al-Mg-Si-Cu-based alloy through two simultaneous approaches, an MDF and an effective SPD approach and microalloying with Sc and Zr; and (2) to improve the hot deformation behavior of the studied alloy by suppressing the dynamic grain growth through activation of the Zener pinning effect. In this work, the structural evolution and near superplastic behavior of the Al-Mg-Si-Cu-Sc-Zr alloy processed by isothermal MDF in the wide temperature and cumulative strain ranges were investigated. In addition, the hot deformation behavior of the Al-Mg-Si-Cu-Sc-Zr alloy was constitutively modelled.

2. Materials and Methods

The alloy with the following composition: Al-1.2Mg-0.7Si-1.0Cu-0.1Sc-0.2Zr wt.%, was prepared using a graphite–fireclay crucible in an inductive furnace (Interselt, Saint Petersburg, Russia). When the temperature reached 780 °C, the casting was carried out in a copper water-cooling mold with a cooling rate of ~15 K/s. The temperature of the melt was measured using a chromel-alumel thermocouple. To prepare the melt, the following components were co-added: 99.99Al, 99.95Mg, Al-12Si, Al-53.5Cu, and Al-5Zr, Al-2Sc. The master alloys were placed in the molten Al in decreasing order of their solidus temperatures.
The obtained ingots were annealed in two steps at 350 °C for 8 h and subsequently at 480 °C for 3 h. The annealing was carried out using a Nabertherm N30/65A air-forced convection furnace (Nabertherm GmbH, Lilienthal, Germany).
The multidirectional forging (MDF) was carried out using a hydraulic press machine (Nordberg N3650E, Changshu Tongrun Auto Acessory Co., Ltd., Changshu, China) providing a force of 50 tons. The scheme of the MDF process is demonstrated in Figure 1. The forging process included several circles. Each circle involved three passes in mutually perpendicular directions. During the MDF, the deformation occurred in two directions. The sample shape remained unchanged at the end of each pass. After each pass, the samples were rotated through 90° to maintain multiaxial deformation. One pass provided a strain of 0.8 and the cumulative strain for a full circle was 2.4. The initial crosshead speed for each pass was 5 mm/min. The samples exhibited parallelepipedal shape with a size l in one dimension and a similar size a in two other dimensions (Figure 1). The true stress-true strain curves and structure evolution during forging were studied using small samples with l = 16 mm and a = 8 mm. The superplastic behavior was analyzed using bulk samples with l = 56 mm and a = 28 mm. The MDF procedure is described in detail in ref. [50]. The maximum number of MDF circles for the studied alloy was 6 (18 passes) that provided for a maximum cumulative strain of 14.4.
The microstructure evolution was analyzed using a Tescan-VEGA3 LMH scanning electron microscope (Tescan Brno s.r.o., Kohoutovice, Czech Republic). The elemental distribution maps were obtained using an energy dispersive spectroscopy (EDS) detector (Oxford Instruments plc, Abingdon, UK). An electron backscatter diffraction (EBSD) maps were obtained using an HKL NordlysMax detector (Oxford Instruments plc, Abingdon, UK). The accelerating voltage and operating distance were 20 kV and 10–15 mm, respectively. The samples’ SEM were preliminarily prepared by grinding and polishing in a Struers LaboPol machine (Struers APS, Ballerup, Denmark).
The analysis of nanoscale dispersoids was performed using a JEOL JEM transmission electron microscope (JEOL, Tokyo, Japan). Disc-shaped specimens with a diameter of 3 mm and a thickness of 0.18–0.22 were examined. The specimens were preliminarily electrochemically polished at 20 ± 1 °C and 19 ± 2 V in a 30% HNO3 in methanol electrolyte using a Struers TenuPol-5 twinjet polishing machine (Struers APS, Ballerup, Denmark).
The EBSD data were obtained with a scanning step of 0.15 µm in the area of 250 × 250 μm, and the maps were analyzed using Aztec (EBSD) software (v.2.0, Oxford Instruments plc, Abingdon, UK). The high-angle grain boundaries (HAGBs) were identified as boundaries between the grains with a misorientation angle higher than 15°. The low-angle grain boundaries (LAGBs) were identified as boundaries between the grains with a misorientation angle of 2–15°, respectively.
The deformation behavior of the alloy was analyzed during a constant strain rate tensile test at 440–520 °C and 2 × 10−3–1 × 10−2 s−1 strain rate. The tensile test was carried out using a Walter-Bai LFM 100 universal testing machine (Walter + Bai AG, Löhningen, Switzerland). The samples for the tensile test had a gauge part size of 4 × 6 × 1 mm. The samples were cut from the as-forged specimens using a Charmilles Robofil 190 wire-electro discharge machine (Agie Charmilles, Geneva, Switzerland).
The room temperature mechanical properties were determined using a tensile test in the Zwick Z250 universal testing machine (Zwick Roell Group, Ulm, Germany). The gauge part of samples was 18 × 6 × 1 mm in size. Three samples were used per state.

3. Results

3.1. Microstructure in As-Cast and Homogenized States

Figure 2a demonstrates the BSE images of the studied alloy in the as-cast state. Several phases distinguished by contrast were revealed in the images. The aluminum solid solution (Al) was the predominant microstructural feature. The phases with dark and bright contrast were found on the periphery of dendrite cells. The bright areas predominantly corresponded to the Cu-rich phases (CuAl2 and Q(Al5Cu2Mg8Si6)) while the dark areas correspond to the Mg2Si phase. The volume fraction of Cu-rich phases and Mg2Si phase were 1.8 ± 0.2% and 2.1 ± 0.1%, respectively. The two-step homogenization annealing provided for near complete dissolution of the Cu-rich phases (Figure 2b). The Mg2Si particles were partially fragmented and spherodized. After homogenization, the volume fraction of the Mg2Si phase was 2.0 ± 0.1% that was similar to as-cast state. The small fraction of particles enriched with Sc were also defined after annealing (Figure 2b). The size of particles was varied from 0.2 to 2 µm.
The TEM studies of the as-homogenized alloy revealed nanoscale dispersoids (Figure 2a,b). Typical for the L12-Al3(Sc,Zr) phase superlattice reflections were observed in the selected area electron diffraction (SAED) (Figure 2d insert). The ordered position of the superlattice reflections against the reflections of Al confirm at least partial coherency of the dispersoids and the matrix. The mean dispersoid size was 10 ± 1 nm.

3.2. Deformation Behavior during MDF

The MDF was performed at 150, 250, and 350 °C. The true stress vs. true strain dependencies were corrected considering the increase in the strain rate during compression [51] (Figure 3). Relatively stable flow stage occurred in a strain interval of 0.1–0.4. Further deformation to a strain higher than ~0.4 led to intense strengthening, which was attributed to the increased friction between the die and the matrix due to the increase in processed surface during the pass.
At the steady stage of deformation, the stress-strain behavior depended on the strain and temperature. An insignificant softening was observed at 150 °C (Figure 3a). At 250 °C, the softening was observed at a cumulative strain below 7 (Figure 3b). At a higher cumulative strain, the flow was accompanied by dynamic strengthening effect. At 350 °C, dynamic softening occurred only during the first pass and subsequent passes were accompanied with strengthening (Figure 3c). The temperature increase provided for a stress decrease at the stable stage. The flow stress also decreased with cumulative strain. A significant decrease of flow stress was observed during the first 11 passes at 150 °C and 9 passes at 250–350 °C. During a subsequent increase in cumulative strain, the flow stress changed insignificantly.

3.3. The Microstructure of Alloy after MDF

The microstructure was analyzed after three and six circles of multidirectional forging (Figure 4). During the forging process, the Mg2Si particles were fragmentized. The particle parameters are presented in Table 1. The particle mean size and aspect ratio were insignificantly influenced by temperature and cumulative strain. However, the decrease in standard deviation of interparticle distance confirmed the increasing particle distribution homogeneity with increasing MDF temperature. The bright Sc-bearing particles which were observed after homogenization were also found for the sample subjected to MDF. The analysis of the EDS maps revealed the particles (marked as type-A) with bright contrast enriched with Cu, Sc, Si, and a minor amount of Zr (Figure 5a–c). However, the localization of Cu was not found in the particles marked as type-B. Higher magnification BSE images and EDS analysis also revealed fine precipitates enriched with Cu in the samples as-forged at 250 °C and 350 °C. The precipitates size increased with an increase in forging temperature and cumulative strain.
The EBSD IPF mapping of the studied alloy was carried out after six circles of forging (Figure 6). The grain and subgrain size increased with increasing temperature. The subgrain/grain sizes were 0.9 ± 0.1/1.2 ± 0.1 µm, 1.7 ± 0.1/2.0 ± 0.1 µm, and 1.8 ± 0.1/2.1 ± 0.1 µm for 150, 250, and 350 °C, respectively. The grain boundary misorientation histograms revealed that the value of LAGBs decreased with increasing temperature from 150 to 250 °C. The further increase in temperature to 350 °C insignificantly changed the fraction of LAGBs and HAGBs.

3.4. The Hot Deformation Behavior of Alloy after MDF

The hot deformation behavior was studied for the bulk sample subjected to six circles of MDF at 350 °C. The true stress vs. true strain curves for a constant strain rate tensile test in a temperature range of 460–500 °C and a strain rate of 2 × 10−3–1 × 10−2 s−1 are shown in Figure 7. The flow stress increased with the strain and decreased with the deformation temperature. The elongation decreased with increasing deformation temperature (Table 2). The maximum values of elongations to failure within the range of 290–310% were observed at 460 °C under studied strain rate conditions.

3.5. Room Temperature Mechanical Properties

The mechanical characteristics of the alloy studied at room temperature were examined after six MDF circles at 350 °C. The samples were preliminarily solid solution treated (SST) at 520 °C for 30 min and artificially aged at a temperature of 180 °C for 8 h. The values of yield strength (YS), ultimate tensile strength (UTS) and elongation to failure (El) were 326 ± 5 MPa, 366 ± 5 MPa, and 10 ± 3%, respectively.

3.6. Arrhenius Constitutive Model

The stress–strain curves depicted in Figure 7 were utilized in the development of an Arrhenius constitutive model for the purpose of predicting and modeling the flow behavior of the alloy under investigation. The characterization of the deformation response of this particular alloy, taking into account the effects of strain rate and temperature, can be effectively clarified by the utilization of a Zener–Hollomon parameter (Equation (2)) and an exponent-type formula (Equation (3)) [52,53].
Z = ε ˙ . e x p Q R T
ε . ˙ = A f σ e x p Q R T
In the given Equation (2), A represents a material constant, ε ˙ denotes the strain rate measured in s−1, T signifies the temperature measured in Kelvin, Q represents the effective activation energy of deformation measured in kJ/mol, R is the gas constant equal to 8.314 J/(molK), and f σ is a function related to the flow stress (Equations (4)–(6)):
σ = { (4) σ n 1 ,                                                                               f o r   α σ < 0.8 (5) e x p β σ ,                                                                     f o r   α σ > 1.2   (6) sinh α σ n 2 ,                                               f o r   v a r i o u s   σ
The constants of the material are denoted as β, n 1 , n2  a n d   α , where α is defined as the ratio of β to n 1 . While Equations (4) and (5) are suitable for low stress conditions (ασ < 0.8) and high stress conditions (ασ > 1.2), respectively, the hyperbolic sine law (Equation (6)) can be regarded as universally applicable and can be employed across a broad range of stress levels. The Equations (4)–(6) can be expressed in the form provided below (Equations (7) and (8)), taking into account the Zener–Holomon parameter:
ε ˙ = { (7) A 1 σ n 1 e x p Q 1 R T P o w e r   l a w (8) A 2 e β σ e x p Q 2 R T E x p o n e n t i a l   l a w (9) A 3 sinh α σ n 2 e x p Q 3 R T H y p e r b o l i c   s i n e   l a w

3.6.1. Calculation of the Arrhenius Constitutive Model Constants

The material constants were determined by utilizing the true stress and true strain data. The calculating approach was demonstrated using a true strain of 0.4 (50%) as a pilot route. Equations (10) and (11) were derived through the natural logarithm of Equations (7) and (8), respectively.
ln ε ˙ = ln A 1 + n 1 ln σ Q 1 R T n 1 = ln ε ˙ ln σ T
ln ε ˙ = ln A 2 + β σ Q 2 R T β = ln ε ˙ σ T
By applying a logarithm to Equation (9), Equation (12) can be derived, which was subsequently utilized for the computation of Q3 and n2. Equations (13) and (14) were derived by differentiating Equation (12).
ln ε ˙ = ln A 3 + n 2 ln sinh α σ Q 3 R T
n 2 = ln ε ˙ ln sinh α σ T
Q 3 = R × ln ε ˙ ln sinh ( α σ ) T × ln sinh ( α σ ) ( 1 T ) ε ˙
The determined constants corresponding to a strain level of 0.4 (50%) are compiled in Table 3.
The pattern between T , ε ˙ , a n d σ was expressed as follows (Equation (15)):
σ = 1 α ln z A 3 1 n 2 + z A 3 2 n 2 + 1 1 2
where Z = ε ˙ e x p Q 3 R T (Equation (2)).

3.6.2. Compensation of Strain Effect

The magnitude of strain has a noticeable impact on the material constants. Hence, it is imperative to take into account the influence of strain while figuring out flow stress. The material constants were obtained for flow stress values in strain ranging from 0.05 to 0.6. The constants (α, n2, Q 3 and ln(A3)) strain dependances are illustrated in Figure 8. The values of the constants (α, n2, Q 3 and ln(A3)) were obtained through the utilization of polynomial functions of strain, as described in Equation (16). The coefficients of Equation (16) are listed in Table 4. The dependances of the constants on strain were fitted using third-order polynomials fit based on the Adj. R-Square values (Table 4); there were no significant differences in the Adj. R-Square values when using higher order polynomials, such as fifth-order polynomials (Table 4). The number of constants when using the fifth-order polynomials is more than that of third-order polynomials fitting. Thus, the third-order polynomials exhibited the required accuracy with a lower number of constants than in fifth-order polynomials.
α = Y 0 + B 1 ε 1 + B 2 ε 2 + B 3 ε 3 n 2 = Y 0 + B 1 ε 1 + B 2 ε 2 + B 3 ε 3 A 3 = Y 0 + B 1 ε 1 + B 2 ε 2 + B 3 ε 3 Q 3 = Y 0 + B 1 ε 1 + B 2 ε 2 + B 3 ε 3
In order to assess the appropriateness of the acquired constitutive model, the correlation coefficient (R), average absolute relative error (AARE), and root mean square error (RMSE) were calculated using Equations (17)–(19) and accounted for 0.93, 6.6%, and 1.15, respectively. The comparison between the experimental and predicted values of flow stress is illustrated in Figure 9.
The patterns collected from the data demonstrate a satisfactory level of consistency between the experimental results and support the model’s high level of prediction (Figure 9d).
R = i = 1 N E i E ¯ P i P ¯ i = 1 N E i E ¯ 2 i = 1 N P i P ¯ 2
A A R E = 1 N i = 1 N E i P i E i
R M S E = 1 N i = 1 N E i P i 2

4. Discussion

The analysis of the as-cast structure of the studied alloy revealed Mg2Si and Cu-enriched Al2Cu and Q phases. The non-equilibrium Cu-rich phases dissolved during homogenization annealing whereas the Mg2Si particles were fragmented and spherodized. The spherodization and fragmentation of phases of eutectic origin during annealing positively influenced the structural homogeneity and particles distribution after thermomechanical treatment [54,55]. The homogenization annealing of the alloy studied also provided for the formation of the nanoscale L12 phase dispersoids, which were homogeneously distributed in the Al-based matrix. The dispersoids provided a strong Zener pinning effect during MDF [56,57,58,59] and hot deformation inhibiting the static and dynamic grain growth. It should be noted that the Si atoms can also dissolve in the Al3(Sc,Zr) phase, increasing its thermal stability due to the formation of an (Al,Si)3(Sc,Zr) phase [60,61]. In the present study, dispersoids exhibited a size of ~10 nm after a second high temperature homogenization step that can be a result of partial Al substitution by Si atoms, which increase the thermal stability of dispersoids. This is in agreement with our earlier studies; dispersoids with similar parameters were found after two-step homogenization in 6xxx-type alloys [42,62].
The increase in cumulative strain and MDF temperature resulted in an increase in particle distribution homogeneity. The homogeneously distributed microscale particles can contribute to grain refinement during thermomechanical treatment and improve deformation behavior due to the particles’ stimulated nucleation effect (PSN) [46,51,63,64,65]. During MDF, the Mg2Si phase was noticeably fragmentized, the result of both the temperature and strain effect. The Mg2Si particle size after MDF was in the range of 1.3–1.5 µm. Despite a small volume fraction, these particles can provide for the PSN effect and additionally refine grain in the alloys studied. The positive effect of Mg2Si particles with similar parameters on grain refinement was previously discussed in refs. [7,42].
The Sc-rich type-A and type-B phases observed in the as-forged sample exhibited a solidification origin. It was suggested that the Cu-bearing type-A phase corresponded to the W(AlCuSc) phase [66,67]. The Si atoms may dissolve in W-phase by analogy with the (Al,Si)3(Sc,Zr) phase. The Cu-free type-B phase should correspond to the V(AlSc2Si2) phase discussed in refs. [60,66]. The Cu-enriched microscale particles with bright contrast observed in the as-forged at 250–350 °C samples were formed due to the ageing that occurred during MDF. These particles corresponded to the Q-(AlMgSiCu)-phase [68,69] or/and the Al2Cu phase.
The MDF significantly refined grain structure in the alloy studied. The grain-refinement effect accompanied by a decrease in stress value with increasing cumulative strain reflects recrystallization duringac MDF. The decrease in temperature of MDF to 150 °C resulted in a decrease in grain/subgrain size. A high-volume fraction of LAGBs was observed after forging at 150 °C. This phenomenon suggests the continuous dynamic recrystallization [46] as the dominant recrystallization mechanism occurred during MDF of the alloy studied [70,71]. The continuous dynamic recrystallization is the result of LAGBs transforming into HAGBs induced by the strain [46]. A similar process can be observed during the heating of a pre-strained alloy [71,72,73]. We suggest that the dynamic recrystallization is the dominant process at 150 °C, whereas at higher temperatures of 250–350 °C, post-dynamic and static recrystallization can also occur. The static recrystallization may occur when the sample is heated to a treatment temperature before starting the forging. The increase in grain size with MDF temperature can be explained using a Zener factor [46]. The L12 –phase dispersoids prevent the intense grain coarsening due to the Zener pinning effect. The dispersoids exhibited a mean size of ~10 nm. Taking into account that all atoms of Sc and Zr precipitated from the solid solution and formed an Al3(Sc,Zr) phase, the dispersoid volume fraction f is approximately ~0.005. The Zener factor allows for calculating the particles’ contribution to the grain growth suppression [46,74]:
Z H = 3 f d 4 r
where d is the mean grain size. The grain size after MDF at 150 °C was ~1.2 µm and at 250–350 °C it was ~2.0–2.1 µm. The corresponding Zener factor value was equal to 0.9 at 150 °C and ~1.5 at 250–350 °C. The Z = 1 is a threshold value above which the static grain growth is suppressed [74,75]. The Zener parameter value Z > 1 after MDF at both 250 and 350 °C. The threshold grain size value above which the grain growth is considered suppressed is ~1.3 µm. While the Zener factor does not consider the strain-induced grain growth, an insignificant difference in the grain size after MDF at 250 and 350 °C is most obviously associated with the Zener effect.
The fine-grained structure of the alloy after MDF and the Zener pinning effect provided a near superplastic behavior in a temperature range of 460–500 °C. The dynamic strengthening observed in the stress–strain curves is associated with dynamic grain growth that occurs at elevated temperatures [6]. This explains a decrease of elongation with increasing temperature during hot deformation. A hyperbolic Arrhenius constitutive model [76,77,78] is successfully constructed to estimate the flow stress behavior of this alloy in the studied temperature and strain rate range. The developed model demonstrated high predictability accuracy with the maximum error AARE of 6.6%.
The as-forged alloy demonstrated improved mechanical properties. The same level of mechanical characteristics was observed in the alloy with similar composition but increased Sc after thermomechanical treatment included hot and cold rolling [43]. The as-rolled and further annealed sheets with Sc exhibited a non-recrystallized structure that contributes significantly to strength due to dislocation strengthening mechanism [79,80]. Moreover, Sc and Zr provide for a high precipitation contribution to strength (Orowan strengthening) due to L12 dispersoids [81]. In the present study, the alloy contains low Sc of 0.1% and exhibits a recrystallized structure, whereas the strength was maintained at a similar level Sthat was achieved through both grain refinement Hall-Petch mechanism and Zener pinning effect. A near-superplastic behavior and high strength observed in the bulk samples offers additional opportunities for industrial application of the studied alloy and MDF technique.
The Arrhenius type constitutive model based on the experimental stress–strain results was constructed to model the hot flow behavior of this alloy. The constants of the Arrhenius type constitutive model were illustrated on the dependance of the deformation level. The stress exponent n is commonly employed in order to determine the primary mechanism governing the hot deformation process. Hence, the diverse methods of superplastic deformation can be correlated with independent values of n. Based on the findings of previous studies [82,83], it has been observed that the grain boundary sliding and the dislocation glide mechanism can result in values of n that are approximately two and three, respectively. On the other hand, processes associated with n values ranging from four to six are attributed to dislocation climb. According to the data presented in Figure 8b, the n values exhibit a range between 2.10 and 2.75 in the studied range of the strain and within the studied strain rates and temperatures, and this revealed that the grain boundary sliding and dislocation glide mechanisms controlled the hot deformation. The phenomenon of grain boundary sliding is observed to occur simultaneously with continuous dynamic recrystallization in both deformed region and substructured volume as the applied strain increases. The observed decrease in the n value at higher strains could potentially be attributed to the development of a recrystallized grain structure dominated in all samples. Furthermore, it is likely that the process of active grain boundary sliding is facilitated by dislocation mechanisms.
The material constant α, as shown in Figure 8d, exhibits a consistent decrease as the strain increases, and the same characteristic is observed for the activation energy Q. The specific deformation mechanisms and related mechanisms were also proposed by the effective activation energy of deformation [84]. The Q value exhibited an impressive decrease, reaching a strain value of 0.6 (Figure 8a). The observed reduction in Q as strain increases is consistent with the findings reported in previous studies [82,83,85,86]. The observed decrease in the Q value provides clear evidence of a reduction in the stored energy within the material as a result of strain [84,87].

5. Conclusions

The microstructure evolution and deformation behavior of the Al-1.2Mg-0.7Si-1.0Cu-0.1Sc-0.2Zr alloy after the homogenization annealing and multidirectional forging (MDF) to a cumulative strain of ∑e = 14.4 in a temperature range of 150–350 °C were studied. The main conclusions are drawn as follows:
  • Two Sc-bearing phases were formed in the studied alloy during solidification: the Cu-rich W(AlCuSc)-phase with dissolved Si atoms and the Cu-free V(AlSc2Si2) phase were suggested. A minor amount of particles of these phases with a size in a range of 0.2–2 µm were also observed after homogenization annealing and MDF. The dispersoids of the L12-Al3(Sc,Zr) phase with a mean size of 10 ± 1 nm were precipitated during homogenization annealing.
  • Due to the Zener pinning effect of nanoscale L12 phase dispersoids and a large cumulative strain of 14.4, the fine-grained structure with a mean grain size of 1.2–2.1 µm was formed after MDF.
  • Fine-grained as-forged at 350 °C alloy exhibited the maximum elongation to failure of ~300% at 460 °C in a strain-rate range of 2 × 10−3–1 × 10−2 s−1; at room temperature, the alloy exhibited yield strength of 326 ± 5 MPa, ultimate tensile strength of 366 ± 5 MPa, and elongation at fracture of 10 ± 3%.
  • The hot deformation behavior was described using the Arrhenius type constitutive model. The developed model demonstrated a high predictability accuracy with a maximum average absolute relative error of 6.6%.

Author Contributions

Conceptualization, A.G.M. and A.V.M.; methodology, L.Y.K. and V.V.P.; software, A.O.M.; validation, A.G.M. and V.V.P.; formal analysis, A.O.M.; investigation, A.G.M. and L.Y.K.; resources, A.G.M.; data curation, A.O.M.; writing—original draft preparation, A.G.M.; writing—review and editing, A.G.M. and V.V.P.; visualization, A.O.M.; supervision, A.V.M.; project administration A.G.M.; funding acquisition, A.G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Russian Science Foundation under grant number 22-79-00253.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented are available on request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

The authors are also grateful for TEM studies carried out by the Collective Use Equipment Center “Material Science and Metallurgy” working due to the equipment modernization program No. 075-15-2021-696 represented by the Ministry of Higher Education and Science of the Russian Federation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guan, R.-G.; Tie, D. A Review on Grain Refinement of Aluminum Alloys: Progresses, Challenges and Prospects. Acta Metall. Sin. Engl. Lett. 2017, 30, 409–432. [Google Scholar] [CrossRef]
  2. Huang, Y.; Langdon, T.G. Advances in Ultrafine-Grained Materials. Mater. Today 2013, 16, 85–93. [Google Scholar] [CrossRef]
  3. Miller, W.; Zhuang, L.; Bottema, J.; Wittebrood, A.; De Smet, P.; Haszler, A.; Vieregge, A. Recent Development in Aluminium Alloys for the Automotive Industry. Mater. Sci. Eng. A 2000, 280, 37–49. [Google Scholar] [CrossRef]
  4. Heinz, A.; Haszler, A.; Keidel, C.; Moldenhauer, S.; Benedictus, R.; Miller, W. Recent Development in Aluminium Alloys for Aerospace Applications. Mater. Sci. Eng. A 2000, 280, 102–107. [Google Scholar] [CrossRef]
  5. Hirsch, J. Recent Development in Aluminium for Automotive Applications. Trans. Nonferrous Met. Soc. China 2014, 24, 1995–2002. [Google Scholar] [CrossRef]
  6. Sotoudeh, K.; Bate, P.S. Diffusion Creep and Superplasticity in Aluminium Alloys. Acta Mater. 2010, 58, 1909–1920. [Google Scholar] [CrossRef]
  7. Troeger, L.P.; Starke, E.A. Particle-Stimulated Nucleation of Recrystallization for Grain-Size Control and Superplasticity in an Al–Mg–Si–Cu Alloy. Mater. Sci. Eng. A 2000, 293, 19–29. [Google Scholar] [CrossRef]
  8. Troeger, L.P.; Starke, E.A. Microstructural and Mechanical Characterization of a Superplastic 6xxx Aluminum Alloy. Mater. Sci. Eng. A 2000, 277, 102–113. [Google Scholar] [CrossRef]
  9. Furukawa, M.; Horita, Z.; Nemoto, M.; Langdon, T.G. The Use of Severe Plastic Deformation for Microstructural Control. Mater. Sci. Eng. A 2002, 324, 82–89. [Google Scholar] [CrossRef]
  10. Werenskiold, J.C.; Roven, H.J. Microstructure and Texture Evolution during ECAP of an AlMgSi Alloy: Observations, Mechanisms and Modeling. Mater. Sci. Eng. A 2005, 410–411, 174–177. [Google Scholar] [CrossRef]
  11. Orozco-Caballero, A.; Álvarez-Leal, M.; Hidalgo-Manrique, P.; Cepeda-Jiménez, C.M.; Ruano, O.A.; Carreño, F. Grain Size versus Microstructural Stability in the High Strain Rate Superplastic Response of a Severely Friction Stir Processed Al-Zn-Mg-Cu Alloy. Mater. Sci. Eng. A 2017, 680, 329–337. [Google Scholar] [CrossRef]
  12. Han, J.-K.; Jang, J.; Langdon, T.G.; Kawasaki, M. Bulk-State Reactions and Improving the Mechanical Properties of Metals through High-Pressure Torsion. Mater. Trans. 2019, 60, 1131–1138. [Google Scholar] [CrossRef]
  13. Mulyukov, R.R.; Imayev, R.M.; Nazarov, A.A. Production, Properties and Application Prospects of Bulk Nanostructured Materials. J. Mater. Sci. 2008, 43, 7257–7263. [Google Scholar] [CrossRef]
  14. Alemdağ, Y.; Karabiyik, S.; Pürçek, G. Microstructure and Mechanical Properties of Severely Deformed Eutectic Al–Si–Cu–Mn Alloy. Met. Mater. Int. 2023, 29, 1181–1194. [Google Scholar] [CrossRef]
  15. Sitdikov, O.S.; Avtokratova, E.V.; Atanov, B.I.; Markushev, M.V. Effect of Multidirectional Isothermal Forging on the Formation of Ultrafine-Grained Structure in Alloy 1570 C. Inorg. Mater. 2022, 58, 544–554. [Google Scholar] [CrossRef]
  16. Kumara, B.; Preetham Kumar, G.V. Influence of Multi-Directional Room-Temperature Forging Process on Microstructure and Mechanical Behaviour of Eutectic Al-12Si Alloy. Metallogr. Microstruct. Anal. 2022, 11, 175–182. [Google Scholar] [CrossRef]
  17. He, H.; Chen, K.; Yi, Y.; You, W.; Guo, Y.; Wang, B.; Tang, J.; Guo, W.; Huang, S. Influence of Forging Temperature on the Microstructures and Mechanical Properties of a Multi-Directionally Forged Al–Cu–Li Alloy. Met. Mater. Int. 2022, 28, 433–447. [Google Scholar] [CrossRef]
  18. Zhu, Q.; Li, L.; Ban, C.; Zhao, Z.; Zuo, Y.; Cui, J. Structure Uniformity and Limits of Grain Refinement of High Purity Aluminum during Multi-Directional Forging Process at Room Temperature. Trans. Nonferrous Met. Soc. China 2014, 24, 1301–1306. [Google Scholar] [CrossRef]
  19. Armstrong, P.E.; Hockett, J.E.; Sherby, O.D. Large Strain Multidirectional Deformation of 1100 Aluminum at 300 K. J. Mech. Phys. Solids 1982, 30, 37–58. [Google Scholar] [CrossRef]
  20. Sitdikov, O.; Garipova, R.; Avtokratova, E.; Mukhametdinova, O.; Markushev, M. Effect of Temperature of Isothermal Multidirectional Forging on Microstructure Development in the Al-Mg Alloy with Nano-Size Aluminides of Sc and Zr. J. Alloys Compd. 2018, 746, 520–531. [Google Scholar] [CrossRef]
  21. Sitdikov, O.S.; Avtokratova, E.V.; Murzinova, M.A.; Markushev, M.V. Effect of Multidirectional Forging with Decreasing Temperature on the Microstructure and Microhardness of the Aluminum-Magnesium-Scandium-Zirconium Alloy. Mater. Perform. Charact. 2020, 9, 20190006. [Google Scholar] [CrossRef]
  22. Rao, P.N.; Singh, D.; Jayaganthan, R. Mechanical Properties and Microstructural Evolution of Al 6061 Alloy Processed by Multidirectional Forging at Liquid Nitrogen Temperature. Mater. Des. 2014, 56, 97–104. [Google Scholar] [CrossRef]
  23. Hussain, M.; Rao, P.N.; Singh, D.; Jayaganthan, R.; Singh, S. Comparative Study of Microstructure and Mechanical Properties of Al 6063 Alloy Processed by Multi Axial Forging at 77K and Cryorolling. Procedia Eng. 2014, 75, 129–133. [Google Scholar] [CrossRef]
  24. Amer, S.M.; Barkov, R.Y.; Prosviryakov, A.S.; Pozdniakov, A.V. Pozdniakov. Structure and Properties of New Heat-Resistant Cast Alloys Based on the Al–Cu–Y and Al–Cu–Er Systems. Phys. Met. Metallogr. 2021, 122, 908–914. [Google Scholar] [CrossRef]
  25. Liu, F.; Zhu, X.; Qin, J.; Zhou, W.; Ling, J.; Dong, Q.; Yu, J.; Nagaumi, H.; Zhang, B. Effect of Mn/Cr Ratio on Precipitation Behaviors of α-Al(FeMnCr)Si Dispersoids and Mechanical Properties of Al–Mg–Si–Cu Alloys. Mater. Sci. Eng. A 2022, 860, 144269. [Google Scholar] [CrossRef]
  26. Liu, Y.; Zhang, C.-C.; Zhang, X.-Y.; Huang, Y.-C. Understanding Grain Refinement of Sc Addition in a Zr Containing Al-Zn-Mg-Cu Aluminum Alloy from Experiments and First-Principles. Intermetallics 2020, 123, 106823. [Google Scholar] [CrossRef]
  27. Jiang, H.; Xing, H.; Xu, Z.; Yang, B.; Liang, E.; Zhang, J.; Sun, B. Effect of Zn Content and Sc,Zr Addition on Microstructure and Mechanical Properties of Al-Zn-Mg-Cu Alloys. J. Alloys Compd. 2023, 947, 169246. [Google Scholar] [CrossRef]
  28. Dorin, T.; Jiang, L.; Langan, T. Formation of Al3Sc Dispersoids and Associated Strengthening. In TMS Annual Meeting & Exhibition; Springer Nature Switzerland: Cham, Switzerland, 2023; pp. 1207–1212. [Google Scholar]
  29. Croteau, J.R.; Jung, J.-G.; Whalen, S.A.; Darsell, J.; Mello, A.; Holstine, D.; Lay, K.; Hansen, M.; Dunand, D.C.; Vo, N.Q. Ultrafine-Grained Al-Mg-Zr Alloy Processed by Shear-Assisted Extrusion with High Thermal Stability. Scr. Mater. 2020, 186, 326–330. [Google Scholar] [CrossRef]
  30. Mikhaylovskaya, A.V.; Mochugovskiy, A.G.; Levchenko, V.S.; Tabachkova, N.Y.; Mufalo, W.; Portnoy, V.K. Precipitation Behavior of L12 Al3Zr Phase in Al-Mg-Zr Alloy. Mater. Charact. 2018, 139, 30–37. [Google Scholar] [CrossRef]
  31. Monastyrska, T.O.; Berezina, A.L.; Molebny, O.A.; Kotko, A.V. Effect of Alloying with Transition Metals on the Aging of Anomalously Supersaturated Solid Solution of Al–Mg Alloys. Appl. Nanosci. 2022, 12, 513–521. [Google Scholar] [CrossRef]
  32. Xue, C.; Wang, S.; Zhang, Y.; Tian, G.; Yang, X.; Chang, X.; Ke, Y.; Xie, Z.; Wang, J. Uncovering the Kinetics of Li-Rich Clusters and Monodisperse Core–Shell Al3(Zr, Sc) Structures in Al–Li–Cu Alloys. Mater. Sci. Eng. A 2023, 881, 145393. [Google Scholar] [CrossRef]
  33. Wang, Y.; Zhang, S.; Wu, R.; Turakhodjaev, N.; Hou, L.; Zhang, J.; Betsofen, S. Coarsening Kinetics and Strengthening Mechanisms of Core-Shell Nanoscale Precipitates in Al-Li-Yb-Er-Sc-Zr Alloy. J. Mater. Sci. Technol. 2021, 61, 197–203. [Google Scholar] [CrossRef]
  34. Manohar, P.A.; Ferry, M.; Chandra, T. Five Decades of the Zener Equation. ISIJ Int. 1998, 38, 913–924. [Google Scholar] [CrossRef]
  35. Kaibyshev, R.; Avtokratova, E.; Apollonov, A.; Davies, R. High Strain Rate Superplasticity in an Al–Mg–Sc–Zr Alloy Subjected to Simple Thermomechanical Processing. Scr. Mater. 2006, 54, 2119–2124. [Google Scholar] [CrossRef]
  36. Sitdikov, O.; Avtokratova, E.; Markushev, M. Influence of Strain Rate on Grain Refinement in the Al-Mg-Sc-Zr Alloy during High-Temperature Multidirectional Isothermal Forging. Mater. Charact. 2019, 157, 109885. [Google Scholar] [CrossRef]
  37. Higashi, K.; Nagai, S.; Maeda, M.; Ohnishi, T. The Superplastic Behavior and Structural Changes during Hot Deformation of Al-Cu-Zr Alloys. J. Jpn. Inst. Light Met. 1986, 36, 361–371. [Google Scholar] [CrossRef]
  38. Avtokratova, E.; Sitdikov, O.; Markushev, M.; Mulyukov, R. Extraordinary High-Strain Rate Superplasticity of Severely Deformed Al–Mg–Sc–Zr Alloy. Mater. Sci. Eng. A 2012, 538, 386–390. [Google Scholar] [CrossRef]
  39. Smolej, A.; Klobčar, D.; Skaza, B.; Nagode, A.; Slaček, E.; Dragojević, V.; Smolej, S. Superplasticity of the Rolled and Friction Stir Processed Al–4.5 Mg–0.35Sc–0.15Zr Alloy. Mater. Sci. Eng. A 2014, 590, 239–245. [Google Scholar] [CrossRef]
  40. Duan, Y.L.; Tang, L.; Deng, Y.; Cao, X.W.; Xu, G.F.; Yin, Z.M. Superplastic Behavior and Microstructure Evolution of a New Al-Mg-Sc-Zr Alloy Subjected to a Simple Thermomechanical Processing. Mater. Sci. Eng. A 2016, 669, 205–217. [Google Scholar] [CrossRef]
  41. Cao, X.; Xu, G.; Duan, Y.; Yin, Z.; Lu, L.; Wang, Y. Achieving High Superplasticity of a New Al–Mg–Sc–Zr Alloy Sheet Prepared by a Simple Thermal–Mechanical Process. Mater. Sci. Eng. A 2015, 647, 333–343. [Google Scholar] [CrossRef]
  42. Mochugovskiy, A.G.; Mosleh, A.O.; Kotov, A.D.; Khokhlov, A.V.; Kaplanskaya, L.Y.; Mikhaylovskaya, A.V. Microstructure Evolution, Constitutive Modelling, and Superplastic Forming of Experimental 6XXX-Type Alloys Processed with Different Thermomechanical Treatments. Materials 2023, 16, 445. [Google Scholar] [CrossRef]
  43. Esmaeili Ghayoumabadi, M.; Mochugovskiy, A.G.; Tabachkova, N.Y.; Mikhaylovskaya, A.V. The Influence of Minor Additions of Y, Sc, and Zr on the Microstructural Evolution, Superplastic Behavior, and Mechanical Properties of AA6013 Alloy. J. Alloys Compd. 2022, 900, 163477. [Google Scholar] [CrossRef]
  44. Fridlyander, I.N.; Senatorova, O.G.; Ryazanova, N.A.; Nikiforov, A.O. Grain Structure and Superplasticity of High Strength AL-ZN-MG-CU Alloys with Different Minor Additions. Mater. Sci. Forum 1994, 170–172, 345–350. [Google Scholar] [CrossRef]
  45. Kumar, A.; Mukhopadhyay, A.K.; Prasad, K.S. Superplastic Behaviour of Al–Zn–Mg–Cu–Zr Alloy AA7010 Containing Sc. Mater. Sci. Eng. A 2010, 527, 854–857. [Google Scholar] [CrossRef]
  46. Humphreys, F.J.; Hatherly, M. Recrystallization and Related Annealing Phenomena, 2nd ed.; Pergamon Press: Oxford, UK, 1995. [Google Scholar]
  47. Belov, N.A.; Alabin, A.N.; Eskin, D.G.; Istomin-Kastrovskii, V.V. Optimization of Hardening of Al–Zr–Sc Cast Alloys. J. Mater. Sci. 2006, 41, 5890–5899. [Google Scholar] [CrossRef]
  48. Mousavi Anijdan, S.H.; Kang, D.; Singh, N.; Gallerneault, M. Precipitation Behavior of Strip Cast Al–Mg–0.4Sc–0.15Zr Alloy under Single and Multiple-Stage Aging Processes. Mater. Sci. Eng. A 2015, 640, 275–279. [Google Scholar] [CrossRef]
  49. Guo, Z.; Zhao, G.; Chen, X.-G. Effects of Two-Step Homogenization on Precipitation Behavior of Al3Zr Dispersoids and Recrystallization Resistance in 7150 Aluminum Alloy. Mater. Charact. 2015, 102, 122–130. [Google Scholar] [CrossRef]
  50. Mikhaylovskaya, A.V.; Kotov, A.D.; Kishchik, M.S.; Prosviryakov, A.S.; Portnoy, V.K. The Effect of Isothermal Multi-Directional Forging on the Grain Structure, Superplasticity, and Mechanical Properties of the Conventional Al–Mg-Based Alloy. Metals 2019, 9, 33. [Google Scholar] [CrossRef]
  51. Kishchik, M.S.; Mochugovskiy, A.G.; Cuda, M.; Kishchik, A.A.; Mikhaylovskaya, A.V. Particle Stimulated Nucleation Effect for Al-Mg-Zr-Sc Alloys with Ni Addition during Multidirectional Forging. Metals 2023, 13, 1499. [Google Scholar] [CrossRef]
  52. Kai, X.; Zhao, Y.; Wang, A.; Wang, C.; Mao, Z. Hot Deformation Behavior of in Situ Nano ZrB2 Reinforced 2024Al Matrix Composite. Compos. Sci. Technol. 2015, 116, 1–8. [Google Scholar] [CrossRef]
  53. Kai, X.; Chen, C.; Sun, X.; Wang, C.; Zhao, Y. Hot Deformation Behavior and Optimization of Processing Parameters of a Typical High-Strength Al–Mg–Si Alloy. Mater. Des. 2016, 90, 1151–1158. [Google Scholar] [CrossRef]
  54. Amer, S.M.; Yu. Barkov, R.; Yakovtseva, O.A.; Loginova, I.S.; Pozdniakov, A.V. Effect of Zr on Microstructure and Mechanical Properties of the Al–Cu–Er Alloy. Mater. Sci. Technol. 2020, 36, 453–459. [Google Scholar] [CrossRef]
  55. Mamzurina, O.I.; Amer, S.M.; Glavatskikh, M.V.; Barkov, R.Y.; Loginova, I.S.; Pozdniakov, A.V. Microstructure and Mechanical Properties of Novel Heat Resistant Cast Al-Cu-Yb(Gd)-Mg-Mn-Zr Alloys. Metals 2022, 12, 2079. [Google Scholar] [CrossRef]
  56. Huang, H.; Jiang, M.; Wei, L.; Liu, C.; Liu, S. Microstructural and Mechanical Behavior of Al-Mg-Mn-Sc-Zr Alloy Subjected to Multi-Axial Forging at Room Temperature. Mater. Charact. 2021, 171, 110734. [Google Scholar] [CrossRef]
  57. Sitdikov, O.; Avtokratova, E.; Markushev, M. Development of Ultrafine Grain Structure in an Al–Mg–Mn–Sc–Zr Alloy During High-Temperature Multidirectional Isothermal Forging. Met. Mater. Int. 2021, 27, 2743–2755. [Google Scholar] [CrossRef]
  58. Ringeval, S.; Piot, D.; Desrayaud, C.; Driver, J.H. Texture and Microtexture Development in an Al–3Mg–Sc(Zr) Alloy Deformed by Triaxial Forging. Acta Mater. 2006, 54, 3095–3105. [Google Scholar] [CrossRef]
  59. Sitdikov, O.S.; Avtokratova, E.V.; Krymskyi, S.V.; Ilyasov, R.R.; Markushev, M.V. Effect of Severe Forging and Rolling on the Microstructure and Mechanical Properties of Al–Mg–Sc–Zr Alloy. Inorg. Mater. 2021, 57, 101–111. [Google Scholar] [CrossRef]
  60. Dumbre, J.; Kairy, S.K.; Anber, E.; Langan, T.; Taheri, M.L.; Dorin, T.; Birbilis, N. Understanding the Formation of (Al,Si)3Sc and V-Phase (AlSc2Si2) in Al-Si-Sc Alloys via Ex Situ Heat Treatments and in Situ Transmission Electron Microscopy Studies. J. Alloys Compd. 2021, 861, 158511. [Google Scholar] [CrossRef]
  61. Vo, N.Q.; Dunand, D.C.; Seidman, D.N. Improving Aging and Creep Resistance in a Dilute Al–Sc Alloy by Microalloying with Si, Zr and Er. Acta Mater. 2014, 63, 73–85. [Google Scholar] [CrossRef]
  62. Mochugovskiy, A.G.; Prosviryakov, A.S.; Tabachkova, N.Y.; Mikhaylovskaya, A.V. The Effect of Ce on the Microstructure, Superplasticity, and Mechanical Properties of Al-Mg-Si-Cu Alloy. Metals 2022, 12, 512. [Google Scholar] [CrossRef]
  63. Humphreys, F.J. The Nucleation of Recrystallization at Second Phase Particles in Deformed Aluminium. Acta Metall. 1977, 25, 1323–1344. [Google Scholar] [CrossRef]
  64. Mikhaylovskaya, A.V.; Ryazantseva, M.A.; Portnoy, V.K. Effect of Eutectic Particles on the Grain Size Control and the Superplasticity of Aluminium Alloys. Mater. Sci. Eng. A 2011, 528, 7306–7309. [Google Scholar] [CrossRef]
  65. Loginova, I.S.; Sazerat, M.V.; Loginov, P.A.; Pozdniakov, A.V.; Popov, N.A.; Solonin, A.N. Evaluation of Microstructure and Hardness of Novel Al-Fe-Ni Alloys with High Thermal Stability for Laser Additive Manufacturing. JOM 2020, 72, 3744–3752. [Google Scholar] [CrossRef]
  66. Gazizov, M.; Teleshov, V.; Zakharov, V.; Kaibyshev, R. Solidification Behaviour and the Effects of Homogenisation on the Structure of an Al–Cu–Mg–Ag–Sc Alloy. J. Alloys Compd. 2011, 509, 9497–9507. [Google Scholar] [CrossRef]
  67. Kharakterova, M.L.; Eskin, D.G.; Toropova, L.S. Precipitation Hardening in Ternary Alloys of the Al-Sc-Cu and Al-Sc-Si Systems. Acta Metall. Mater. 1994, 42, 2285–2290. [Google Scholar] [CrossRef]
  68. Chakrabarti, D.J.; Peng, Y.; Laughlin, D.E. Precipitation in Al-Mg-Si Alloys with Cu Additions and the Role of the Q’ and Related Phases. Mater. Sci. Forum 2002, 396–402, 857–862. [Google Scholar] [CrossRef]
  69. Staab, T.E.M.; Krause-Rehberg, R.; Hornauer, U.; Zschech, E. Study of Artificial Aging in AlMgSi (6061) and AlMgSiCu (6013) Alloys by Positron Annihilation. J. Mater. Sci. 2006, 41, 1059–1066. [Google Scholar] [CrossRef]
  70. Kishchik, A.A.; Kishchik, M.S.; Kotov, A.D.; Mikhaylovskaya, A.V. Effect of Multidirectional Forging on the Microstructure and Mechanical Properties of the Al–Mg–Mn–Cr Alloy. Phys. Met. Metallogr. 2020, 121, 489–494. [Google Scholar] [CrossRef]
  71. Kishchik, M.S.; Mikhaylovskaya, A.V.; Kotov, A.D.; Mosleh, A.O.; AbuShanab, W.S.; Portnoy, V.K. Effect of Multidirectional Forging on the Grain Structure and Mechanical Properties of the Al-Mg-Mn Alloy. Materials 2018, 11, 2166. [Google Scholar] [CrossRef]
  72. Sakai, T.; Belyakov, A.; Kaibyshev, R.; Miura, H.; Jonas, J.J. Dynamic and Post-Dynamic Recrystallization under Hot, Cold and Severe Plastic Deformation Conditions. Prog. Mater. Sci. 2014, 60, 130–207. [Google Scholar] [CrossRef]
  73. Sakai, T.; Jonas, J.J. Plastic Deformation: Role of Recovery and Recrystallization. In Encyclopedia of Materials: Science and Technology; Elsevier: Amsterdam, The Netherlands, 2001; pp. 7079–7084. [Google Scholar]
  74. Humphreys, F.J. A Unified Theory of Recovery, Recrystallization and Grain Growth, Based on the Stability and Growth of Cellular Microstructures—II. The Effect of Second-Phase Particles. Acta Mater. 1997, 45, 5031–5039. [Google Scholar] [CrossRef]
  75. Humphreys, F.J. A Unified Theory of Recovery, Recrystallization and Grain Growth, Based on the Stability and Growth of Cellular Microstructures—I. The Basic Model. Acta Mater. 1997, 45, 4231–4240. [Google Scholar] [CrossRef]
  76. Xia, Y.; Shu, X.; Zhang, Q.; Pater, Z.; Li, Z.; Xu, H.; Ma, Z.; Xu, C. Modified Arrhenius Constitutive Model and Simulation Verification of 2A12-T4 Aluminum Alloy during Hot Compression. J. Mater. Res. Technol. 2023, 26, 1325–1340. [Google Scholar] [CrossRef]
  77. Li, J.; Yang, X.; Zhu, Y.; Zhang, Y.; Qiu, Y.; Sanders, R.E. Study of Thermal Compression Constitutive Relation for 5182-Sc-Zr Alloy Based on Arrhenius-Type and ANN Model. Crystals 2022, 12, 611. [Google Scholar] [CrossRef]
  78. Xie, Q.; Wu, Y.; Peng, S.; Yuan, Z. Prediction Flow Behaviour of a Magnesium Alloy in Hot Deformation and a Comparative Study by Using Hyperbolic Sine Model and ANN Model. Phys. Met. Metallogr. 2022, 123, 1471–1478. [Google Scholar] [CrossRef]
  79. Kalish, D.; LeFevre, B.G. Subgrain Strengthening of Aluminum Conductor Wires. Metall. Trans. A 1975, 6, 1319–1324. [Google Scholar] [CrossRef]
  80. Mochugovskiy, A.G.; Mikhaylovskaya, A.V.; Zadorognyy, M.Y.; Golovin, I.S. Effect of Heat Treatment on the Grain Size Control, Superplasticity, Internal Friction, and Mechanical Properties of Zirconium-Bearing Aluminum-Based Alloy. J. Alloys Compd. 2021, 856, 157455. [Google Scholar] [CrossRef]
  81. Knipling, K.E.; Karnesky, R.A.; Lee, C.P.; Dunand, D.C.; Seidman, D.N. Precipitation Evolution in Al–0.1Sc, Al–0.1Zr and Al–0.1Sc–0.1Zr (at.%) Alloys during Isochronal Aging. Acta Mater. 2010, 58, 5184–5195. [Google Scholar] [CrossRef]
  82. Mahmudi, R.; Rezaee-Bazzaz, A.; Banaie-Fard, H.R. Investigation of Stress Exponent in the Room-Temperature Creep of Sn–40Pb–2.5Sb Solder Alloy. J. Alloys Compd. 2007, 429, 192–197. [Google Scholar] [CrossRef]
  83. Langdon, T.G. Identifiying Creep Mechanisms at Low Stresses. Mater. Sci. Eng. A 2000, 283, 266–273. [Google Scholar] [CrossRef]
  84. Yang, X.; Guo, H.; Liang, H.; Yao, Z.; Yuan, S. Flow Behavior and Constitutive Equation of Ti-6.5Al-2Sn-4Zr-4Mo-1W-0.2Si Titanium Alloy. J. Mater. Eng. Perform. 2016, 25, 1347–1359. [Google Scholar] [CrossRef]
  85. Mandal, S.; Rakesh, V.; Sivaprasad, P.V.; Venugopal, S.; Kasiviswanathan, K.V. Constitutive Equations to Predict High Temperature Flow Stress in a Ti-Modified Austenitic Stainless Steel. Mater. Sci. Eng. A 2009, 500, 114–121. [Google Scholar] [CrossRef]
  86. Pu, Z.J.; Wu, K.H.; Shi, J.; Zou, D. Development of Constitutive Relationships for the Hot Deformation of Boron Microalloying Ti-Al-Cr-V Alloys. Mater. Sci. Eng. A 1995, 192–193, 780–787. [Google Scholar] [CrossRef]
  87. Rao, K.P.; Hawbolt, E.B. Development of Constitutive Relationships Using Compression Testing of a Medium Carbon Steel. J. Eng. Mater. Technol. 1992, 114, 116–123. [Google Scholar] [CrossRef]
Figure 1. The scheme of the multidirectional forging procedure.
Figure 1. The scheme of the multidirectional forging procedure.
Applsci 13 13054 g001
Figure 2. (a,b,e) SEM-BSE and (c,d) TEM micrographs of the (a) as-cast and (be) homogenized alloy; (c) bright field image; (d) dark filed image for (c) obtained in the reflection marked-up with a green circle in the SAED (insert in (d)); (e) SEM-EDS elemental distribution maps for the alloying elements.
Figure 2. (a,b,e) SEM-BSE and (c,d) TEM micrographs of the (a) as-cast and (be) homogenized alloy; (c) bright field image; (d) dark filed image for (c) obtained in the reflection marked-up with a green circle in the SAED (insert in (d)); (e) SEM-EDS elemental distribution maps for the alloying elements.
Applsci 13 13054 g002
Figure 3. The stress vs. true strain dependencies for each pass of MDF at (a) 150 °C, (b) 250 °C, and (c) 350 °C.
Figure 3. The stress vs. true strain dependencies for each pass of MDF at (a) 150 °C, (b) 250 °C, and (c) 350 °C.
Applsci 13 13054 g003
Figure 4. SEM-BSE micrographs of the studied alloy after (ac) three and (df) six circles of MDF at (a,d) 150 °C, (b,e) 250 °C, and (c,f) 350 °C.
Figure 4. SEM-BSE micrographs of the studied alloy after (ac) three and (df) six circles of MDF at (a,d) 150 °C, (b,e) 250 °C, and (c,f) 350 °C.
Applsci 13 13054 g004
Figure 5. SEM-BSE images and EDS elemental distribution maps for the studied alloy after (a) 6 MDF circles at 150 °C, (b) 3 MDF circles at 250 °C, and (c) 6 MDF circles at 350 °C. Red rectangular areas show the areas of EDS maps.
Figure 5. SEM-BSE images and EDS elemental distribution maps for the studied alloy after (a) 6 MDF circles at 150 °C, (b) 3 MDF circles at 250 °C, and (c) 6 MDF circles at 350 °C. Red rectangular areas show the areas of EDS maps.
Applsci 13 13054 g005
Figure 6. EBSD maps and corresponding grain misorientation angle histograms for the studied alloys after six circles of MDF at (a) 150 °C, (b) 250 °C, and (c) 350 °C.
Figure 6. EBSD maps and corresponding grain misorientation angle histograms for the studied alloys after six circles of MDF at (a) 150 °C, (b) 250 °C, and (c) 350 °C.
Applsci 13 13054 g006
Figure 7. The true stress vs. true strain curves corresponding to a tensile test in a constant strain rate range of 2 × 10−3–1 × 10−2 s−1 at (a) 460 °C, (b) 480 °C, and (c) 500 °C for the alloy after six MDF circles at 350 °C.
Figure 7. The true stress vs. true strain curves corresponding to a tensile test in a constant strain rate range of 2 × 10−3–1 × 10−2 s−1 at (a) 460 °C, (b) 480 °C, and (c) 500 °C for the alloy after six MDF circles at 350 °C.
Applsci 13 13054 g007
Figure 8. The dependence of (a) Q, (b) n, (c) ln(A3), and (d) α constants vs. true strain (ε). The square box symbols are the calculated constants, and the red lines are the third-order polynomials fits.
Figure 8. The dependence of (a) Q, (b) n, (c) ln(A3), and (d) α constants vs. true strain (ε). The square box symbols are the calculated constants, and the red lines are the third-order polynomials fits.
Applsci 13 13054 g008
Figure 9. The experimental and the predicted stress–strain from the constructed model at different temperatures: (a) 460 °C, (b) 480 °C, and (c) 500 °C; and (d) the verification of the constitutive equation (through points) between experimental and theoretical data. The square box symbols in (d) are the experimental data, and the red lines in (d) are the linear fits.
Figure 9. The experimental and the predicted stress–strain from the constructed model at different temperatures: (a) 460 °C, (b) 480 °C, and (c) 500 °C; and (d) the verification of the constitutive equation (through points) between experimental and theoretical data. The square box symbols in (d) are the experimental data, and the red lines in (d) are the linear fits.
Applsci 13 13054 g009
Table 1. Parameters of the Mg2Si particles after MDF.
Table 1. Parameters of the Mg2Si particles after MDF.
Particle ParameterNumber of CirclesTemperature of MDF
150 °C250 °C350 °C
Mean particle size, µm31.3 ± 0.11.3 ± 0.21.5 ± 0.2
61.3 ± 0.11.4 ± 0.11.5 ± 0.2
Aspect ratio30.62 ± 0.040.68 ± 0.030.65 ± 0.04
60.63 ± 0.030.67 ± 0.030.66 ± 0.03
Mean interparticle distance, µm353 ± 1147 ± 1053 ± 8
649 ± 1244 ± 843 ± 8
Standard deviation of interparticle distance342.936.834.9
649.235.136.7
Table 2. Elongation to failure (%) after tensile tests at various temperatures and strain rates.
Table 2. Elongation to failure (%) after tensile tests at various temperatures and strain rates.
Strain Rate (s−1)Temperature
460 °C480 °C500 °C
1 × 10−2310 ± 20130 ± 10110 ± 20
5 × 10−3290 ± 10210 ± 10120 ± 20
2 × 10−3310 ± 10240 ± 20200 ± 20
Table 3. The computed values of A1, A2, A3, n1, n2, β, Q 1,2 , 3 and α constants at strain level of 0.4 (50%).
Table 3. The computed values of A1, A2, A3, n1, n2, β, Q 1,2 , 3 and α constants at strain level of 0.4 (50%).
ln(A1)n1Q1 [KJ/mol]ln(A2)β [MPa−1]Q2 [KJ/mol]αln(A3)n2Q3 [KJ/mol]
233.551 ± 5170.2250 ± 50.062142.6751 ± 6
Table 4. Coefficients in Equation (20).
Table 4. Coefficients in Equation (20).
ParameterY0B1B2B3Adj. R-Square (Third Order)Adj. R-Square (Fifth Order)
α 0.161−0.7291.792−1.4780.960.99
n 2 2.176−2.21016.755−20.1770.980.98
l n ( A 3 ) 22.178−64.456190.923−200.3690.970.99
Q 103.819−394.3071143.484−1192.0500.980.99
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

Mochugovskiy, A.G.; Kaplanskaya, L.Y.; Mosleh, A.O.; Palacheva, V.V.; Mikhaylovskaya, A.V. Microstructure Evolution and Constitutive Modelling of Deformation Behavior for Al-Mg-Si-Cu-Sc-Zr Alloy Processed with Isothermal Multidirectional Forging. Appl. Sci. 2023, 13, 13054. https://doi.org/10.3390/app132413054

AMA Style

Mochugovskiy AG, Kaplanskaya LY, Mosleh AO, Palacheva VV, Mikhaylovskaya AV. Microstructure Evolution and Constitutive Modelling of Deformation Behavior for Al-Mg-Si-Cu-Sc-Zr Alloy Processed with Isothermal Multidirectional Forging. Applied Sciences. 2023; 13(24):13054. https://doi.org/10.3390/app132413054

Chicago/Turabian Style

Mochugovskiy, Andrey G., Ludmila Yu. Kaplanskaya, Ahmed O. Mosleh, Valeria V. Palacheva, and Anastasia V. Mikhaylovskaya. 2023. "Microstructure Evolution and Constitutive Modelling of Deformation Behavior for Al-Mg-Si-Cu-Sc-Zr Alloy Processed with Isothermal Multidirectional Forging" Applied Sciences 13, no. 24: 13054. https://doi.org/10.3390/app132413054

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