Next Article in Journal
Revisiting Rare Earth Permanent Magnetic Alloys of Nd-Fe-C
Next Article in Special Issue
Fluidity of Pure Aluminum in a Narrow Channel Die Gap during Die Casting
Previous Article in Journal
Enhanced DC and AC Soft Magnetic Properties of Fe-Co-Ni-Al-Si High-Entropy Alloys via Texture and Iron Segregation
Previous Article in Special Issue
Mechanical Properties and Deformation Behavior of Open-Cell Type Aluminum Foams with Structural Conditions and Alloy Composition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Er on the Hot Deformation Behavior of the Crossover Al3Zn3Mg3Cu0.2Zr Alloy

by
Maria V. Glavatskikh
,
Leonid E. Gorlov
,
Irina S. Loginova
,
Ruslan Yu. Barkov
*,
Maxim G. Khomutov
,
Alexander Yu. Churyumov
and
Andrey V. Pozdniakov
Department of Physical Metallurgy of Non-Ferrous Metals, National University of Science and Technology “MISIS”, Leninskiy Prospekt 4, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Metals 2024, 14(10), 1114; https://doi.org/10.3390/met14101114
Submission received: 27 August 2024 / Revised: 25 September 2024 / Accepted: 25 September 2024 / Published: 29 September 2024
(This article belongs to the Special Issue Structure and Properties of Aluminium Alloys 2024)

Abstract

:
The effect of an erbium alloying on the hot deformation behavior of the crossover Al3Zn3Mg3Cu0.2Zr alloy was investigated in detail. First of all, Er increases the solidus temperature of the alloy. This allows hot deformation at a higher temperature. The precipitates resulting from the Er alloying of the Al3Zn3Mg3Cu0.2Zr alloy were analyzed using transmission electron microscopy. Erbium addition to the alloy produces the formation of more stable and fine L12-(Al3(Zr, Er)) precipitates with a size of 20–60 nm. True stress tends to increase with a decline in the temperature and an increase in the deformation rate. The addition of Er leads to decreases in true stress at the strain rates of 0.01–1 s−1 due to particle-stimulated nucleation softening mechanisms. The effective activation energy of the alloy with the Er addition has a lower value, enabling an easier hot deformation process in the alloy with an elevated volume fraction of the intermetallic particles. The addition of Er increases the strain rate sensitivity, which makes the failure during deformation less probable. The investigated alloys have a significant difference in the dependence of the activation volume on the temperature. The flow instability criterion allows better deformability of Er-doped alloys and enables the alloys to be formed more easily. The evenly distributed particles prevent the formation of shear bands with elevated storage energy and decrease the probability of crack initiation during the initial stages of hot deformation when only one softening mechanism (dynamic recovery) is working. The microstructure analysis proves that dynamic recovery is the main softening mechanism.

1. Introduction

Wrought Al-Zn-Mg-Cu alloys are widely used in the aviation and aerospace industries due to their high strength [1]. The highest content of alloying elements provides a great strengthening effect after aging treatment. The Zn/Mg ratio is more than 1 in the commercial Al-Zn-Mg-Cu alloys, ensuring a high yield strength (YS), but the casting properties, heat resistance, and corrosion resistance are low [1,2,3,4,5,6]. The Al-Zn-Mg-Cu alloys with a Zn/Mg ratio of 1 have medium strength and good castability, corrosion resistance, and heat resistance [1,3,4,5,6,7,8]. Another way to improve the casting properties is alloying using eutectic forming elements [7,8,9,10,11,12,13]. Scandium and zirconium are well-known for improving the mechanical properties of the Al-Zn-Mg-Cu alloysat room and elevated temperature, due to the nucleation of nano-sized precipitates of the L12-Al3(Sc,Zr) phase during solution treatment [14,15,16,17,18,19,20,21]. There is a similar effect on the microstructure and properties of Al that may render the combination of Er and Zr unnecessary [22,23,24,25,26]. However, Er is significantly cheaper than Sc. At the same time, Er is not only a precipitate-forming element but also a eutectic-forming element [27,28,29,30,31,32,33]. Er-rich eutectic origin phases have an improved resistance to growth with high-temperature solution treatment, which contributes to achieving a high strength of the alloy [27,28,29,30,31,32,33]. The novel cast and wrought Al-3Zn-3Mg-3Cu-Zr-Y(Er) alloy with improved heat resistance was developed based on these principles [34]. Developed alloys fit into the basic concept of the crossover alloys based on the use of aluminum alloy scraps [35]. The crossover alloys cover the structure and properties of combined alloys of the different groups [36,37,38,39,40,41,42,43,44]. The developed Al-Zn-Mg-Cu-Zr-Y(Er) alloys combined the best properties from the cast A-Zn-Mg alloys, high-strength Al-Zn-Mg-Cu alloys, and heat-resistant Al-Cu alloys [34,44]. The presented features of the new alloys make them very promising for deep investigations.
The main way to improve the mechanical properties of the Al-Zn-Mg-Cu alloys is to control the grain structure under thermo-mechanical processing [14,45,46,47,48,49,50,51,52,53]. Xu et al. showed that dynamic recovery is the main dynamic softening mechanism in the Al-5.6Zn-1.9Mg-0.3Cu-0.09Sc-0.09Zr alloy; at the same time, three dynamic recrystallization mechanisms (discontinuous, continuous, and geometric) lead to grain refinement [54]. A lower value of the effective activation energy was obtained in the as-forged alloy [55] compared to the as-cast alloy [50]. Tang et. al. investigated the influence of Zn content on the hot deformation behavior of Al-xZn-2Mg-2Cu alloys: the increase in Zn concentration from 6.3 to 10.1% decreases the rate of dynamic recovery and dynamic recrystallization and decelerates dynamic softening [56]. However, the increase in Zn content does not have a significant influence on the static softening after hot deformation [57]. Zirconium content has a more significant influence on the static softening [58]. An elevated volume fraction of Al3Zr dispersoids pinned dislocation and inhibited the dynamic recovery and/or recrystallization, resulting in a higher level of stored deformation energy in the hot deformed alloys. Consequently, more driving force for static softening is presented with increasing Zr additions. A similar influence of the hot deformation conditions (strain rate and temperature) on the post-deformation behavior was observed by Long et. al. [48]. The volume fraction of statically recrystallized grains is influenced by the stored deformation energy: at low deformation temperatures and high strain rates, the driving force for static recrystallization is higher. Hot deformation may also influence the phase composition of the alloys. Zhang et. al. has shown that hot rolling changes the morphology of Al3Zr particles in 7055 aluminum alloys and leads to an increase in the dislocation density in subgrains and, consequently, results in an increased number and variety of quenching-induced phases [59].
Previously, a power tool for the investigation and modeling of hot deformation behavior called “processing mapping” was developed. This approach lets us determine the optimal deformation conditions using minimal experimental data. This method was successfully applied to the 5A06 [60], 2195-O [61], and ECO-7175 [62] aluminum alloys and the TiC [63], CNT [64], and ZrB2 [65] aluminum-based composites.
The present investigation aims to determine the effect of Er alloying on parameters of the hot deformation of the Al-3Zn-3Mg-3Cu-0.2Zr alloy using the modeling of the 3D processing maps.

2. Materials and Methods

2.1. Alloy Melting

The Al-2.8Zn-2.7Mg-3Cu-0.2Zr-0.1Ti-0.15Si-0.15Fe (Al3Zn3Mg3Cu) and Al-2.9Zn-2.8Mg-3Cu-1.4Er-0.2Zr-0.1Ti-0.15Si-0.15Fe (Al3Zn3Mg3CuEr) alloys were melted in the graphite–chamotte crucibles in the resistance furnace. The melting and pouring processes are described in [34]. The samples for microstructure investigation and compression tests were cut out from the ingots that measured 20 mm in thickness, 40 mm in width, and 120 mm in height. The ingots were obtained in the water-cooling copper mold.

2.2. Sample Preparation and Structure Investigation

Evaluations of grain structure, microstructure, phase composition, and precipitates were conducted using a Neophot 21 optical microscope (OM) (Carl Zeiss AG, Oberkochen, Germany), a Tescan Vega 3 LMH scanning electron microscope (SEM) (Tescan, Brno, Kohoutovice, Czech Republic) with an XMax-80 electron diffraction X-ray (EDX) detector (Oxford Instruments Advanced AZtecEnergy, High Wycombe, UK), and a JEM2100 transmission electron microscope (TEM) (Jeol Ltd., Tokyo, Japan). The samples for microstructure investigation were grinded and polished using equipment from Struers Labopol (Struers APS, Ballerup, Denmark). The oxidation was applied to etch the grain structure. The oxidation proceeded in the Barker’s reagent [34]. The samples for TEM were electrochemically polished in the A2 electrolyte using equipment from Struers Tenupol (Struers APS, Ballerup, Denmark). The heat treatment was carried out in the Nabertherm furnace based on the differential calorimetry data obtained in [34]. The calculation of the volume fraction of the precipitates was performced in the Thermo-Calc software (https://thermocalc.com/, TCW5, Thermo-Calc Software AB, Stockholm, Sweden).

2.3. Hot Compression Testing

The hot compression was carried out using a hydraulic-type Gleeble 3800 (Dynamic Systems Inc., Poestenkill, NY, USA) thermomechanical simulator. The cylindrical samples with a diameter of 10 mm and a height of 15 mm were compressed at temperatures of 350 –500 °C and strain rates of 0.01, 0.1, 1, and 10 s−1. Graphite foil between the sample’s edges and dies was used to minimize the influence of friction. Control of the temperature was carried out using a chromel/alumel thermocouple welded to the center of the sample. The error in temperature determination was about 1 °C. The heating rate was 5 °C/s. The sample was held for 30 s at the compression temperature for unification of the temperature distribution before deformation. The overall strain was about 1. The primary curves that were obtained were corrected for friction and adiabatic heating [66,67] using the OriginLab Software (Version 9.1, Northampton, MA, USA).

3. Results and Discussion

3.1. Microstructure Analysis of As-Cast and Solution-Treated Alloys

The grain structure and phase composition (size and volume fraction of solidification origin phases and precipitates) affected the hot deformation behavior, as was noted earlier. The solution treatment was applied to the ingots of the as-cast alloys before hot deformation. The initial grain microstructure and as-solution-treated phase composition were investigated before the compression tests. The detailed investigation of the microstructure and phase composition of the as-cast and solution-treated Al3Zn3Mg3Cu and Al3Zn3Mg3CuEr alloys is presented in [34]. An average grain size in the Er-free alloy is 100 ± 15 µm (Figure 1a). A part of erbium atoms may substitute the titanium and/or zirconium atoms in the primary Al3(Zr,Ti) clusters. As a result, the fraction of the clusters may be increased. These clusters play the role of the substrates for the primary aluminum solid solution during solidification due to a similar lattice parameter. The erbium addition to the Al3Zn3Mg3CuEr alloy provides significant refining of the grain structure (Figure 1b). The average grain size of the Er-rich alloy is 45 ± 10 µm. The effectiveness of the erbium as a grain refiner has been demonstrated by many researchers [68,69,70,71,72,73].
The solidus temperature, which was determined by differential scanning calorimetry in [34], of the Al3Zn3Mg3Cu and Al3Zn3Mg3CuEr alloys is 493 and 477 °C, respectively. The homogenization annealing during solution treatment at 480 °C for 3 h provides for the dissolving of the non-equilibrium low-temperature melting phases and increases the solidus temperature in the Al3Zn3Mg3CuEr alloy only. The SEM microstructure of the alloys after solution treatment at 480 °C for 3 h is presented in Figure 2a,b. The solidus temperature of the Er-rich alloy increased to 533 °C after the first stage of homogenization annealing [34]. The second stage of homogenization annealing at 520 °C was performed for both alloys. Homogenization treatment at 520 °C is the burnout for the Er-free alloy. The low-temperature melting phases on the dendritic cell and grain boundaries were melted and solidified in the near-spherical particles (Figure 2c). The solidification origin intermetallic phases of the Al3Zn3Mg3CuEr alloy change the morphology to a more globular structure (Figure 2d). As a result, the final microstructure of the Al3Zn3Mg3Cu alloy after solution treatment at 480 °C for 3 h and before hot deformation was characterized by 10–11% of the solidification origin phases with a size of 0.5–3 µm (Figure 2a). The Er-rich alloy contains 11–12% of the solidification origin phases with a size of 0.5–4 µm (Figure 2d) after two-step solution treatment (480 °C for 3 h + 520 °C for 6 h).
An accompanying process during homogenization is heterogenization. Heterogenization is the decomposition of the supersaturated zirconium or zirconium and erbium aluminum solid solution. The decomposition of the aluminum solid solution under annealing of the as-cast alloy following with L12-Al3M (M = Zr,Er) precipitates nucleation. The measured EDX SEM content of the zirconium in the aluminum solid solution is 0.2–0.3%. The additional 0.3% fraction of erbium was determined in the aluminum solid solution in the as-cast Al3Zn3Mg3CuEr alloy. The higher content of the precipitates that form elements in the Al3Zn3Mg3CuEr alloy must provide the higher volume fraction of the L12 precipitates. Erbium substitutes the zirconium atoms in the lattice of the L12-structured Al3(Zr) precipitates [21,22,23,24,25,26,30,31]. The TEM microstructures of the alloys after solution treatment are presented in Figure 3. The Al3Zn3Mg3Cu alloy was solution-treated at 480 °C for 3 h, and the Al3Zn3Mg3CuEr alloy was solution-treated at (480 °C for 3 h + 520 °C for 6 h). The L12-(Al3Zr) precipitates with sizes of 40–100 nm were nucleated in the Er-free alloy (Figure 3a) after 3 h of solution treatment at 480 °C. The fast Fourier transformation (FFT) confirms the L12 structure of the precipitates (insert in the right image in Figure 3a). Erbium addition in the alloy provides the formation of more stable and fine L12-(Al3(Zr,Er)) precipitates after two-stage solution treatment with the higher temperature of the second stage. The L12-(Al3(Zr,Er)) precipitates have a size of 20–60 nm.
The characteristics of the microstructure components are summarized in Table 1. The grain size, volume fraction, and parameters of the intermetallic particles of the solidification origin, and precipitates were measured via the secant method using images of Figure 1, Figure 2 and Figure 3. The volume fraction of the precipitates was calculated using a multicomponent phase diagram of the Al-Zn-Mg-Cu-Zr system and binary Al-Er system. The fraction of the Al3(Zr,Er) precipitates was calculated as the sum of the volume fraction of the Al3Zr phase in the Al-3Zn-3Mg-3Cu-0.2Zr alloy and the Al3Er phase in the Al-0.3Er alloy. The microstructure after solution treatment is the initial one for hot deformation. The differences in the microstructure characteristics may affect the hot deformation behavior. The main difference between the two alloys is the solidus temperature. The erbium addition provides an increase in the alloy solidus and increases the possibility of increasing the temperature of hot deformation.

3.2. Hot Deformation Behavior

The investigated alloys show typical hot deformation behavior (Figure 4 and Figure 5). True stress tends to increase with a decline in the temperature and an increase in the deformation rate. The diffusion process mainly determines the non-conservative movement of the dislocations, which is the main mechanism of deformation at elevated temperatures. As a result, the increase in temperature accelerates diffusion and makes the deformation easier. At the same time, at low strain rates, the dislocations have more time for sliding and creeping to other planes, which leads to decreases in the stress. All curves show similar shapes with a peak in stress. Three main processes of the microstructure evaluation may proceed under hot deformation: strain hardening, dynamic recovery, and dynamic recrystallization. The strain hardening and dynamic recovery took place before peak stress was reached. Dynamic recrystallization is the main process that underlies the decrease in the stress. At the last stage of the deformation, all processes are mutually compensated, and the steady state has a constant stress value. The position of the peak depends on the temperature and strain rate. Following the general trend, the strain for peak stress is shifted to the larger values, with increases in the strain rate and decreases in the temperature. The addition of Er leads to decreases in true stress at strain rates of 0.01–1 s−1 due to particle-stimulated nucleation-softening mechanisms. At a high strain rate, the process of non-conservative dislocation movement is inhibited for both alloys.
The hot deformation behavior of the alloys may be compared in terms of the effective activation energy (Q). The Q-value is usually determined by the functional dependence of the stress (S) at a steady state and the Zener–Hollomon parameter (Z) [74]:
Z = ε ˙ e Q R T
where ε ˙ and T are strain rate (s−1) and temperature (K). The hyperbolic sine is the universal law that describes the dependence of the Z-parameter at all stress values:
Z = A 3 [ sinh ( α S ) ] n 2
where A3, n2, and α are the material’s constants. However, the approximate determination of the α constant needs the construction of particular dependencies between stress and the Z-parameter, such as the exponential law for high stresses:
Z = A 2 e β S
and the power law for a low level of stress:
Z   = A 1 S n 1
where the material’s constants A1, n1, A2, and β should be determined using the values of the flow stress.
The value of α may be found using the following formula:
α     β n P
The value of the Q-value was determined by the least squares method after the logarithmization of Equations (2)–(4). The values of the stress during the steady state were used for the determination of the material constants. Comparisons of the experimental and predicted true stress values using Equation (3), as well as constants of the constitutive equation, are presented in Figure 6. The value of Pearson’s coefficient for both collations is close to 1, indicating that the constructed model is accurate. The absolute average relative error has a value of less than 1%. The effective activation energy for the alloy with the Er addition has a lower value, suggesting that the hot deformation process is easier with the alloy with an elevated volume fraction of intermetallic particles.
The addition of Er to the alloys also influenced another hot deformation characteristic—the activation volume [75]:
V a = 3 3 k T ( ( l n ε ˙ ) S )
where k is the Boltzmann constant (k = 1.38·10−23 J/K).
This parameter quantifies the minimal volume that is necessary for the process of elemental deformation. The dependence of the activation volume on the temperature for both alloys at a steady state is shown in Figure 7. The activation volume increases with an increase in the temperature, which can be explained by a decrease in the dislocation density as the temperature increases. At low temperatures, Va values for the Er-free alloy are higher than for the Al3Zn3Mg3CuEr alloy, which can be explained by the additional accumulation of the dislocations at the intermetallic particles. However, at temperatures higher than 450 °C, the activation volume of the alloy with Er becomes significantly higher due to a decrease in the dislocation density caused by particle-stimulated nucleation dynamic recrystallization.
One more important parameter for hot deformation behavior analysis is the strain rate sensitivity coefficient:
m   = 1 n 1 ,
where n1 is the coefficient from Equation (4). The value of this coefficient is 0.174 ± 0.002 for the Al3Zn3Mg3Cu alloy and 0.181 ± 0.002 for the alloy with the Er addition. A larger value of the coefficient denotes a larger sensitivity to the strain rate increases and the prevention of deformation localization. The absence of the local large dislocation density (a larger strain) advances the deformation process and prevents premature failure. It shows an elevated technology adaptivity of the alloy with the Er addition.

3.3. Hot Processing Maps

For comparison of the hot formability of the alloys, the processing maps approach from [76] was also applied in this study. The approach involves the calculation of two criteria: the efficiency of power dissipation (η) and flow instability (ξ). The first criterion may be calculated as
η = 2 m m + 1 .
This criterion shows the part of the energy that dissipates in the material due to microstructural softening, such as dynamic recovery and dynamic recrystallization. The usual values of these parameters for aluminum alloys are in the range of 10–40% and are dependent on deformation temperature, strain rate, and strain.
The second criterion shows the instability of flow during the deformation and also may be calculated using the strain rate sensitivity coefficient value:
ξ ( ε ˙ ) = [ ln ( m m + 1 ) ln ε ˙ ] + m .
Its positive value shows that the deformation proceeds evenly without significant material flow localization and may provide potential places for local crack formation, stress concentration, and, consequently, material fracturing.
Comparisons of processing maps of different deformation parameters for both alloys is shown in Figure 8. Both alloys have a wide region of good energy dissipation (more than 30%) at high temperatures and low strain rates. However, the flow stability significantly differs. Compression is one of the softest deformation processes. The deformation of the Er-free alloy at strain rates higher than 0.1 s−1 may lead to flow localization. The flow localization may produce a fracture in deformation conditions that are harder than compression. At the same time, the ξ ( ε ˙ ) -value for the Er-doped alloy is lower than 0 only at high strain rates and temperatures below 420 °C, indicating that the alloy is better able to form under these conditions. The evenly distributed particles prevent the formation of shear bands with an elevated level of storage energy and decrease the probability of crack initiation during hot deformation during the initial stages when only one softening mechanism (dynamic recovery) is working.
Evaluations of the grain structure of the Al3Zn3Mg3Cu and Al3Zn3Mg3CuEr alloys under compression at different temperatures and strain rates are shown in Figure 9 and Figure 10. The polarized light in the OM makes the different grains appear as different colors. In general, the deformed non-recrystallized structures are visible in all conditions. The microstructure proves that dynamic recovery is the main softening mechanism. Figure 11 illustrates a comparison of the grain structure of the Al3Zn3Mg3Cu and Al3Zn3Mg3CuEr alloys after deformation at 450 °C and a strain rate of 0.01 s−1. The thin boundaries are seen as one deformed grain. The observation that the areas divided by these boundaries are the same color may indicate that these areas have very similar crystallographic orientations. In other words, the subgrain boundaries are clearly seen in the one deformed grain of both alloys. The slightly finer grain and substructure were formed in the Er-rich alloy due to finer precipitates with higher volume fractions (Figure 3 and Table 1). The L12-(Al3(Zr,Er) precipitates effectively suppress the recrystallization. By comparison, the static recrystallization in the cold rolled alloys proceeds at 350 °C after 1 h of annealing, and erbium provides a finer grain structure [34].
To summarize, we can provide three positive points of the Er alloying effect on the deformation behavior of the Al3Zn3Mg3Cu alloy:
-
Increasing the solidus temperature increases the hot deformation temperature;
-
Increasing the volume fraction of the solidification origin particles decreases the effective activation energy;
-
Increasing the volume fraction of the finer L12 precipitates suppresses recrystallization and provides stable hot deformation.

4. Conclusions

The effect of erbium as a eutectic and a precipitate-forming element on the hot deformation behavior of the novel crossover Al3Zn3Mg3Cu0.2Zr alloy was investigated in detail using optical, scanning, and transmission electron microscopy. The initial grain structure and as-solution-treated phase composition were investigated, and 3D processing maps were constructed after compression tests. The most important conclusions are presented below.
  • The formation of precipitates during the Er alloying of the Al3Zn3Mg3Cu0.2Zr alloy was investigated. Erbium addition to the alloy enabled the formation of more stable and fine L12-(Al3(Zr,Er)) precipitates, with sizes of 20–60 nm after two-stage solution treatment and higher temperatures during the second stage.
  • The investigated alloys demonstrated typical hot deformation behavior. True stress tended to increase with a decline in the temperature and an increase in the deformation rate. The addition of Er led to decreases in true stress at strain rates of 0.01–1 s−1 due to particle-stimulated nucleation-softening mechanisms.
  • The effective activation energy for the alloy with the Er addition had a lower value, allowing the hot deformation process to proceed more easily in the alloy with an elevated volume fraction of intermetallic particles.
  • The strain rate sensitivity coefficient was higher in the alloy with the Er addition due to elevated particle-stimulated softening.
  • The activation volume for deformation of the Er-free alloy was higher at low temperatures and lower at elevated temperatures, which can be explained by different rates of hardening and softening due to the presence of more intermetallic particles.
  • The investigated alloys had a wide region of good energy dissipation at high temperatures and low strain rates. The flow instability criterion for the Er-doped alloy had a negative value only at high strain rates and temperatures below 420 °C, indicating that the alloy was better able to form under these conditions. The evenly distributed particles prevented the formation of shear bands with elevated storage energy and decreased the probability of crack initiation during hot deformation during the initial stages when only one softening mechanism (dynamic recovery) was working. The microstructure proved that the main softening mechanism is dynamic recovery.

Author Contributions

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

Funding

This work was supported by the Russian Science Foundation (Project No. 22-79-10142), https://rscf.ru/project/22-79-10142/.

Data Availability Statement

The data can be available by request.

Conflicts of Interest

On behalf of all authors, the corresponding author states that there are no conflicts of interest.

References

  1. ASM International Handbook Committee. ASM Handbook, Vol. 2: Properties and Selection—Nonferrous Alloys and Special-Purpose Materials; ASM International: Almere, The Netherlands, 2001. [Google Scholar]
  2. Gerchikova, N.S.; Fridlyander, I.N.; Zaitseva, N.I.; Kirkina, N.N. Change in the structure and properties of Al-Zn-Mg alloys. Met. Sci. Heat Treat. 1972, 14, 233–236. [Google Scholar] [CrossRef]
  3. Zolotorevsky, V.S. Microstructure and mechanical properties of cast aluminum alloys. Diss. Doc. Tech. Sci. M. MISiS 1978. [Google Scholar]
  4. Zou, Y.; Wu, X.; Tang, S.; Zhu, Q.; Song, H.; Guo, M.; Cao, L. Investigation on microstructure and mechanical properties of Al-Zn-Mg-Cu alloys with various Zn/Mg ratios. J. Mater. Sci. Technol. 2021, 85, 106–117. [Google Scholar] [CrossRef]
  5. Novikov, I.I. Hot-Shortness of Non-Ferrous Metals and Alloys (Nauka, Moscow). 1966. Available online: https://ntrl.ntis.gov/NTRL/dashboard/searchResults/titleDetail/AD677354.xhtml (accessed on 26 September 2024).
  6. Pan, Y.; Zhang, D.; Liu, H.; Zhuang, L.; Zhang, J. Precipitation hardening and intergranular corrosion behavior of novel Al–Mg–Zn(-Cu) alloys. J. Alloys Compd. 2021, 853, 157199. [Google Scholar] [CrossRef]
  7. Zolotorevskii, V.S.; Pozdnyakov, A.V.; Churyumov, A.Y. Search for promising compositions for developing new multiphase casting alloys based on Al-Zn-Mg matrix using thermodynamic calculations and mathematic simulation. Phys. Met. Metallogr. 2014, 115, 286–294. [Google Scholar] [CrossRef]
  8. Pozdniakov, A.V.; Zolotorevskiy, V.S.; Mamzurina, O.I. Determining hot cracking index of Al–Mg–Zn casting alloys calculated using effective solidification range. Int. J. Cast Met. Res. 2015, 28, 318–321. [Google Scholar] [CrossRef]
  9. Shurkin, P.K.; Akopyan, T.K.; Galkin, S.P.; Aleshchenko, A.S. Effect of Radial Shear Rolling on the Structure and Mechanical Properties of a New-Generation High-Strength Aluminum Alloy Based on the Al–Zn–Mg–Ni–Fe System. Met. Sci. Heat Treat. 2019, 60, 764–769. [Google Scholar] [CrossRef]
  10. Akopyan, T.K.; Belov, N.A.; Alabin, A.N.; Zlobin, G.S. Calculation-experimental study of the aging of casting high-strength Al-Zn-Mg-(Cu)-Ni-Fe aluminum alloys. Russ. Metall. 2014, 2014, 60–65. [Google Scholar] [CrossRef]
  11. David, M.D.; Foley, R.D.; Griffin, J.A.; Monroe, C.A. Microstructural Characterization and Thermodynamic Simulation of Cast Al-Zn-Mg-Cu Alloys. Int. J. Met. 2016, 10, 2–20. [Google Scholar] [CrossRef]
  12. Petrova, A.N.; Brodova, I.G.; Razorenov, S.V.; Shorokhov, E.V.; Akopyan, T.K. Mechanical Properties of the Al-Zn-Mg-Fe-Ni Alloy of Eutectic Type at Different Strain Rates. Phys. Met. Metallogr. 2019, 120, 1221–1227. [Google Scholar] [CrossRef]
  13. Brodova, I.G.; Shirinkina, I.G.; Rasposienko, D.Y.; Akopyan, T.K. Structural Evolution in the Quenched Al-Zn-Mg-Fe-Ni Alloy during Severe Plastic Deformation and Annealing. Phys. Met. Metallogr. 2020, 121, 899–905. [Google Scholar] [CrossRef]
  14. Zuo, J.; Hou, L.; Shi, J.; Cui, H.; Zhuang, L.; Zhang, J. Enhanced plasticity and corrosion resistance of high strength Al-Zn-Mg-Cu alloy processed by an improved thermomechanical processing. J. Alloys Compd. 2017, 716, 220–230. [Google Scholar] [CrossRef]
  15. Chen, Z.; Yuan, Z.; Ren, J. The mechanism of comprehensive properties enhancement in Al-Zn-Mg-Cu alloy via novel thermomechanical treatment. J. Alloys Compd. 2020, 828, 154446. [Google Scholar] [CrossRef]
  16. Suh, D.W.; Lee, S.Y.; Lee, K.H.; Lim, S.K. Microstructural evolution of Al-Zn-Mg-Cu-(Sc) alloy during hot extrusion and heat treatments. J. Mater. Process. Technol. 2004, 155–156, 1330–1336. [Google Scholar] [CrossRef]
  17. Senkov, O.N.; Shagiev, M.R.; Senkova, S.V.; Miracle, D.B. Precipitation of Al3 (Sc, Zr) particles in an Al-Zn-Mg-Cu-Sc-Zr alloy during conventional solution heat treatment and its effect on tensile properties. Acta Mater. 2008, 56, 3723–3738. [Google Scholar] [CrossRef]
  18. Senkova, S.V.; Senkov, O.N.; Miracle, D.B. Cryogenic and elevated temperature strengths of an Al-Zn-Mg-Cu alloy modified with Sc and Zr. Metall. Mater. Trans. A 2006, 37, 3569–3575. [Google Scholar] [CrossRef]
  19. Kotov, A.D.; Mikhaylovskaya, A.V.; Kishchik, M.S.; Tsarkov, A.A. Superplasticity of high-strength Al-based alloys produced by thermomechanical treatment. J. Alloys Compd. 2016, 688, 336–344. [Google Scholar] [CrossRef]
  20. Wang, Y.; Xiong, B.; Li, Z.; Zhang, Y.; Teng, H. Precipitation Behavior of Al3 (Sc, Zr) Particles in High-Alloyed Al-Zn-Mg-Cu-Zr-Sc Alloy During Homogenization. Arab. J. Sci. Eng. 2021, 46, 6027–6037. [Google Scholar] [CrossRef]
  21. Wen, S.P.; Gao, K.Y.; Li, Y.; Huang, H.; Nie, Z.R. Synergetic effect of Er and Zr on the precipitation hardening of Al-Er-Zr alloy. Scr. Mater. 2011, 65, 592–595. [Google Scholar] [CrossRef]
  22. Li, H.; Bin, J.; Liu, J.; Gao, Z.; Lu, X. Precipitation evolution and coarsening resistance at 400 °C of Al microalloyed with Zr and Er. Scr. Mater. 2012, 67, 73–76. [Google Scholar] [CrossRef]
  23. Wen, S.P.; Gao, K.Y.; Huang, H.; Wang, W.; Nie, Z.R. Precipitation evolution in Al-Er-Zr alloys during aging at elevated temperature. J. Alloys Compd. 2013, 574, 92–97. [Google Scholar] [CrossRef]
  24. Gao, Z.; Li, H.; Lai, Y.; Ou, Y.; Li, D. Effects of minor Zr and Er on microstructure and mechanical properties of pure aluminum. Mater. Sci. Eng. A 2013, 580, 92–98. [Google Scholar] [CrossRef]
  25. Li, H.; Gao, Z.; Yin, H.; Jiang, H.; Su, X.; Bin, J. Effects of Er and Zr additions on precipitation and recrystallization of pure aluminum. Scr. Mater. 2013, 68, 59–62. [Google Scholar] [CrossRef]
  26. Huang, H.; Wen, S.P.; Gao, K.Y.; Wang, W.; Nie, Z.R. Age hardening behavior and corresponding microstructure of dilute Al-Er-Zr alloys. Metall. Mater. Trans. A 2013, 44, 2849–2856. [Google Scholar] [CrossRef]
  27. Amer, S.M.; Barkov, R.Y.; Yakovtseva, O.A.; Pozdniakov, A.V. Comparative Analysis of Structure and Properties of Quasibinary Al–6.5Cu–2.3Y and Al–6Cu–4.05Er Alloys. Phys. Met. Metallogr. 2020, 121, 476–482. [Google Scholar] [CrossRef]
  28. Pozdnyakov, A.V.; Barkov, R.Y.; Sarsenbaev, Z.; Amer, S.M.; Prosviryakov, A.S. Evolution of Microstructure and Mechanical Properties of a New Al–Cu–Er Wrought Alloy. Phys. Met. Metallogr. 2019, 120, 614–619. [Google Scholar] [CrossRef]
  29. Amer, S.M.; Barkov, R.Y.; 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]
  30. Amer, S.M.; Mikhaylovskaya, A.V.; Barkov, R.Y.; Kotov, A.D.; Mochugovskiy, A.G.; Yakovtseva, O.A.; Glavatskikh, M.V.; Loginova, I.S.; Medvedeva, S.V.; Pozdniakov, A.V. Effect of Homogenization Treatment Regime on Microstructure, Recrystallization Behavior, Mechanical Properties, and Superplasticity of Al-Cu-Er-Zr Alloy. JOM 2021, 73, 3092–3101. [Google Scholar] [CrossRef]
  31. Amer, S.; Yakovtseva, O.; Loginova, I.; Medvedeva, S.; Prosviryakov, A.; Bazlov, A.; Barkov, R.; Pozdniakov, A. The Phase Composition and Mechanical Properties of the Novel Precipitation-Strengthening Al-Cu-Er-Mn-Zr Alloyn. Appl. Sci. 2020, 10, 5345. [Google Scholar] [CrossRef]
  32. Amer, S.M.; Barkov, R.Y.; Prosviryakov, A.S.; Pozdniakov, A.V. 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]
  33. Amer, S.M.; Barkov, R.Y.; Prosviryakov, A.S.; Pozdniakov, A.V. Structure and Properties of New Wrought Al–Cu–Y- and Al–Cu–Er-Based Alloys. Phys. Met. Metallogr. 2021, 122, 915–922. [Google Scholar] [CrossRef]
  34. Glavatskikh, M.V.; Barkov, R.Y.; Gorlov, L.E.; Khomutov, M.G.; Pozdniakov, A.V. Novel Cast and Wrought Al-3Zn-3Mg-3Cu-Zr-Y(Er) Alloys with Improved Heat Resistance. Metals 2023, 13, 909. [Google Scholar] [CrossRef]
  35. Raabe, D.; Tasan, C.C.; Olivetti, E.A. Strategies for improving the sustainability of structural metals. Nature 2019, 575, 64–74. [Google Scholar] [CrossRef] [PubMed]
  36. Stemper, L.; Tunes, M.A.; Tosone, R.; Uggowitzer, P.J.; Pogatscher, S. On the potential of aluminum crossover alloys. Prog. Mater. Sci. 2022, 124, 100873. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Li, Y.; Li, H.; Zhang, D.; Zhang, J. Effect of high Cu concentration on the mechanical property and precipitation behavior of Al–Mg–Zn-(Cu) crossover alloys. J. Mater. Res. Technol. 2022, 20, 4585–4596. [Google Scholar] [CrossRef]
  38. Zhang, L.; Peng, G.; Gu, Y.; Chen, S.; Li, J. Composition optimization and mechanical properties of Al–Zn–Mg–Si–Mn crossover alloys by orthogonal design. Mater. Chem. Phys. 2023, 307, 128216. [Google Scholar] [CrossRef]
  39. Zhang, Z.; Li, Y.; Liu, Y.; Li, H.; Zhang, D.; Zhang, J. A novel Al-Mg-Zn(-Cu) crossover alloy with ultra-high strength. Mater. Lett. 2023, 347, 134640. [Google Scholar] [CrossRef]
  40. Zhang, X.; Yang, C.; Meng, L.; Chen, Z.; Gong, W.; Sun, B.; Zhao, S.; Zhang, D.; Li, Y.; Zhou, D. The influence of precipitation on plastic deformation in a high Mg-containing AlMgZn-based crossover alloy: Slip localization and strain hardening. Int. J. Plast. 2024, 173, 103896. [Google Scholar] [CrossRef]
  41. Trink, B.; Weißensteiner, I.; Uggowitzer, P.J.; Strobel, K.; Hofer-Roblyek, A.; Pogatscher, S. Processing and microstructure–property relations of Al-Mg-Si-Fe crossover alloys. Acta Mater. 2023, 257, 119160. [Google Scholar] [CrossRef]
  42. Zhang, Z.; Hao, Z.; Wang, H.; Zhang, D.; Zhang, J. Modifying the microstructure and stress distribution of crossover Al-Mg-Zn alloy for regulating stress corrosion cracking via retrogression and re-aging treatment. Mater. Sci. Eng. A 2023, 884, 145564. [Google Scholar] [CrossRef]
  43. Tan, P.; Liu, Z.; Qin, J.; Wei, Q.; Wang, B.; Yi, D. Enhanced corrosion performance by controlling grain boundary precipitates in a novel crossover Al-Cu-Zn-Mg alloy by optimizing Zn content. Mater. Charact. 2024, 208, 113615. [Google Scholar] [CrossRef]
  44. Glavatskikh, M.V.; Barkov, R.Y.; Gorlov, L.E.; Khomutov, M.G.; Pozdniakov, A.V. Microstructure and Phase Composition of Novel Crossover Al-Zn-Mg-Cu-Zr-Y(Er) Alloys with Equal Zn/Mg/Cu Ratio and Cr Addition. Metals 2024, 14, 547. [Google Scholar] [CrossRef]
  45. Huo, W.T.; Shi, J.T.; Hou, L.G.; Zhang, J.S. An improved thermo-mechanical treatment of high-strength Al-Zn-Mg-Cu alloy for effective grain refinement and ductility modification. J. Mater. Process. Technol. 2017, 239, 303–314. [Google Scholar] [CrossRef]
  46. Marlaud, T.; Deschamps, A.; Bley, F.; Lefebvre, W. Evolution of precipitate microstructures during the retrogression and re-ageing heat treatment of an Al-Zn-Mg-Cu alloy. Acta Mater. 2010, 58, 4814–4826. [Google Scholar] [CrossRef]
  47. Cheng, Z.; Zhang, C.; Chu, G.; Liu, Z.; Wang, K.; Meng, Z.; Chen, L.; Sun, L.; Zhao, G. Dynamic precipitation and recrystallization behavior during hot deformation of Al-Zn-Mg-Cu alloy: Experiment and modeling. Int. J. Plast. 2024, 178, 103995. [Google Scholar] [CrossRef]
  48. Long, S.; Xia, R.-Z.; Jiang, Y.-P.; Peng, P.; Wang, S.-Y.; Li, S.-S.; Zhou, Y.-T.; Dai, Q.-W. Recrystallization behavior of an Al-Zn-Mg-Cu alloy during hot deformation and subsequent heat treatment. Mater. Today Commun. 2023, 37, 107570. [Google Scholar] [CrossRef]
  49. Yang, Y.; Li, Y.; Yi, D.; Zheng, Q.; Fang, X.; Wang, Y.; Chang, T.; Chen, Y.; Gao, Y.; Huang, K. Revealing the microstructure evolution mechanism and mechanical responses of a novel Al-Zn-Mg-Cu alloy by hot deformation process. J. Mater. Res. Technol. 2024, 29, 3699–3710. [Google Scholar] [CrossRef]
  50. He, D.; Xie, H.; Lin, Y.C.; Yan, X.-T.; Xu, Z.; Xiao, G. Microstructure evolution mechanisms and a physically-based constitutive model for an Al-Zn-Mg-Cu–Zr aluminum alloy during hot deformation. J. Mater. Res. Technol. 2023, 26, 4739–4754. [Google Scholar] [CrossRef]
  51. Wang, Z.; Li, K.; Luo, T.; Xia, P.; Geng, J.; Li, Y.; Chen, D.; Li, X.; Sha, G.; Wang, H. Hot deformation mechanism of a rapidly-solidified Al-Zn-Mg-Cu–Zr alloy. J. Mater. Res. Technol. 2024, 31, 2424–2434. [Google Scholar] [CrossRef]
  52. He, J.; Hu, M.; Zhou, Z.; Li, C.; Sun, Y.; Zhu, X. Effect of initial grain size on hot deformation behavior and recrystallization mechanism of Al-Zn-Mg-Cu alloy. Mater. Charact. 2024, 212, 114012. [Google Scholar] [CrossRef]
  53. Wei, D.; Han, L.; Lv, Z.; Wang, G.; Zhao, G. Investigation of constitutive models and microstructure evolution during hot deformation and solution treatment of Al-Zn-Mg-Cu alloy. J. Mater. Res. Technol. 2024, 32, 1028–1045. [Google Scholar] [CrossRef]
  54. Xu, C.; Huang, J.; Jiang, F.; Jiang, Y. Dynamic recrystallization and precipitation behavior of a novel Sc, Zr alloyed Al-Zn-Mg-Cu alloy during hot deformation. Mater. Charact. 2022, 183, 111629. [Google Scholar] [CrossRef]
  55. Li, H.; Huang, Y.; Liu, Y. Dynamic recrystallization mechanisms of as-forged Al-Zn-Mg-(Cu) aluminum alloy during hot compression deformation. Mater. Sci. Eng. A 2023, 878, 145236. [Google Scholar] [CrossRef]
  56. Tang, J.; Wang, J.; Teng, J.; Wang, G. Effect of Zn content on the dynamic softening of Al-Zn-Mg-Cu alloys during hot compression deformation. Vacuum 2021, 184, 109941. [Google Scholar] [CrossRef]
  57. Tang, J.; Zhang, H.; Teng, J.; Fu, D.; Jiang, F. Effect of Zn content on the static softening behavior and kinetics of Al-Zn-Mg-Cu alloys during double-stage hot deformation. J. Alloys Compd. 2019, 806, 1081–1096. [Google Scholar] [CrossRef]
  58. Chen, K.; Tang, J.; Jiang, F.; Teng, J.; Fu, D.; Zhang, H. The role of various Zr additions in static softening behavior of Al-Zn-Mg-Cu alloys during interval holding of double-stage hot deformation. J. Alloys Compd. 2019, 792, 1112–1121. [Google Scholar] [CrossRef]
  59. Zhang, M.; Li, C.; Zhang, Y.; Liu, S. Effect of hot deformation on microstructure and quenching-induced precipitation behavior of Al-Zn-Mg-Cu alloy. Mater. Charact. 2021, 172, 110861. [Google Scholar] [CrossRef]
  60. Wang, J.; Xiao, G.; Zhang, J. A new constitutive model and hot processing map of 5A06 aluminum alloy based on high-temperature rheological behavior and higher-order gradients. Mater. Today Commun. 2023, 36, 106502. [Google Scholar] [CrossRef]
  61. Kou, L.; Zhao, Y.; Wang, G.; Wang, F.; Yu, Z. Damage mechanisms of 2195-O aluminum alloy sheet induced by microstructural evolution during tensile deformation at different temperatures. Mater. Sci. Eng. A 2024, 914, 147096. [Google Scholar] [CrossRef]
  62. Seyed-Salehi, M.; KIM, B.H.; YANG, S.Y.; KIM, S.K.; Bozchaloei, G.E. Effect of Cr elimination on flow behavior and processing map of newly developed ECO-7175 aluminum alloy during hot compression. Trans. Nonferrous Met. Soc. China 2022, 32, 1442–1459. [Google Scholar] [CrossRef]
  63. Huang, C.C.; Qi, L.; Chen, J.; Guan, R.; Ojo, O.A.; Wang, Z.G. Effect of TiC nanoparticles on the hot deformation behavior of AA7075 aluminum alloy. Mater. Charact. 2021, 181, 111508. [Google Scholar] [CrossRef]
  64. Bian, X.; Ma, K.; Liu, Z.Y.; Xiao, B.L.; Wang, Q.Z.; Ma, Z.Y. Hot deformation behavior analysis of CNT/2009Al composite with bimodal heterostructure via constructing a new processing map. Mater. Charact. 2024, 212, 113979. [Google Scholar] [CrossRef]
  65. Han, X.; Li, H.; He, W.; Wang, G.; Zhang, X.; Wang, X.; Volodymyr, S.; Shcheretskyi, O. Study on the microstructure and mechanical properties of hot rolled nano-ZrB2/AA6016 aluminum matrix composites by friction stir processing. Mater. Today Commun. 2024, 40, 109815. [Google Scholar] [CrossRef]
  66. Churyumov, A.Y.; Khomutov, M.G.; Tsarkov, A.A.; Pozdniakov, A.V. Study of the structure and mechanical properties of corrosion-resistant steel with a high concentration of boron at elevated temperatures. Phys. Met. Metallogr. 2014, 115, 809–813. [Google Scholar] [CrossRef]
  67. Tang, P.; Lu, J.; Yang, Y.; Mo, F.; Wei, L. Effect of various Er/Al-Ti-C ratios on microstructure and tensile properties of the As-cast Al-10Si-0.8Fe alloy. J. Alloys Compd. 2023, 968, 172237. [Google Scholar] [CrossRef]
  68. Gao, Z.; Ma, T.; Li, H.; Yang, H.; Wei, W.; Huang, H.; Nie, Z. Microstructure and mechanical properties of a novel Al-Mg-Er-Zr-Sc alloy fabricated by selective laser melting. J. Mater. Res. Technol. 2023, 27, 6880–6891. [Google Scholar] [CrossRef]
  69. Cao, P.; Zhu, D.; Wu, N.; Ouyang, L.; Li, C.; Liu, Y.; Zhu, Y.; Liu, J. Effect mechanism of Er content variations on the mechanical and corrosion properties of T6 state Al-Mg-Si-xEr sheets. J. Mater. Res. Technol. 2024, 29, 1359–1376. [Google Scholar] [CrossRef]
  70. Zhang, X.; Wang, H.; Yan, B.; Zou, C.; Wei, Z. The effect of grain refinement and precipitation strengthening induced by Sc or Er alloying on the mechanical properties of cast Al-Li-Cu-Mg alloys at elevated temperatures. Mater. Sci. Eng. A 2021, 822, 141641. [Google Scholar] [CrossRef]
  71. Wu, Z.; Shi, Z.; Yan, H.; Luo, C.; Li, H.; Xie, X.; Nie, Q.; Hu, Z. Modification and strengthening mechanism of high strength-toughness as-cast Al–11Si–3Cu alloy modified with minor Sr and Er content. Mater. Sci. Eng. A 2024, 903, 146670. [Google Scholar] [CrossRef]
  72. Li, C.; Zhou, Q.; Han, M.; Sha, S.; Luo, Y.; Xu, X. Effect of rare earth element Er on the microstructure and properties of highly alloyed Al-Zn-Mg-Cu-Zr-Ti alloy. J. Alloys Compd. 2023, 956, 170248. [Google Scholar] [CrossRef]
  73. Guo, Y.; Wei, W.; Shi, W.; Zhang, B.; Zhou, X.; Wen, S.; Wu, X.; Gao, K.; Rong, L.; Huang, H.; et al. Effect of Er and Zr additions and aging treatment on grain refinement of aluminum alloy fabricated by laser powder bed fusion. J. Alloys Compd. 2022, 912, 165237. [Google Scholar] [CrossRef]
  74. Zener, C.; Hollomon, J.H. Effect of strain rate upon plastic flow of steel. J. Appl. Phys. 1944, 15, 22–32. [Google Scholar] [CrossRef]
  75. Kappacher, J.; Leitner, A.; Kiener, D.; Clemens, H.; Maier-Kiener, V. Thermally Activated Deformation Mechanisms and Solid Solution Softening in W-Re Alloys Investigated via High Temperature Nanoindentation. Mater. Des. 2020, 189, 108499. [Google Scholar] [CrossRef]
  76. Prasad, Y.V.R.K.; Gegel, H.L.; Doraivelu, S.M.; Malas, J.C.; Morgan, J.T.; Lark, K.A.; Barker, D.R. Modeling of dynamic material behavior in hot deformation: Forging of Ti-6242. Metall. Trans. A 1984, 15, 1883–1892. [Google Scholar] [CrossRef]
Figure 1. As-cast grain structure of (a) Al3Zn3Mg3Cu and (b) Al3Zn3Mg3CuEr alloys (OM).
Figure 1. As-cast grain structure of (a) Al3Zn3Mg3Cu and (b) Al3Zn3Mg3CuEr alloys (OM).
Metals 14 01114 g001
Figure 2. Microstrucutre after solution treatment at (a,b) (480 °C, 3 h) and (c,d) (480 °C, 3 h + 520 °C, 6 h) of the (a,c) Al3Zn3Mg3Cu and (b,d) Al3Zn3Mg3CuEr alloys (SEM).
Figure 2. Microstrucutre after solution treatment at (a,b) (480 °C, 3 h) and (c,d) (480 °C, 3 h + 520 °C, 6 h) of the (a,c) Al3Zn3Mg3Cu and (b,d) Al3Zn3Mg3CuEr alloys (SEM).
Metals 14 01114 g002
Figure 3. Microstructure of the solution treated at (a) (480 °C, 3 h), which corresponds to the Al3Zn3Mg3Cu alloy, and at (b) (480 °C, 3 h + 520 °C, 6 h), which corresponds to the Al3Zn3Mg3CuEr alloy (TEM).
Figure 3. Microstructure of the solution treated at (a) (480 °C, 3 h), which corresponds to the Al3Zn3Mg3Cu alloy, and at (b) (480 °C, 3 h + 520 °C, 6 h), which corresponds to the Al3Zn3Mg3CuEr alloy (TEM).
Metals 14 01114 g003
Figure 4. Compression stress–strain curves of Al3Zn3Mg3Cu alloy at strain rates of 0.01 s−1 (a), 0.1 s−1 (b), 1 s−1 (c), and 10 s−1 (d).
Figure 4. Compression stress–strain curves of Al3Zn3Mg3Cu alloy at strain rates of 0.01 s−1 (a), 0.1 s−1 (b), 1 s−1 (c), and 10 s−1 (d).
Metals 14 01114 g004
Figure 5. Compression stress–strain curves of Al3Zn3Mg3CuEr alloy at strain rates of 0.01 s−1 (a), 0.1 s−1 (b), 1 s−1 (c), and 10 s−1 (d).
Figure 5. Compression stress–strain curves of Al3Zn3Mg3CuEr alloy at strain rates of 0.01 s−1 (a), 0.1 s−1 (b), 1 s−1 (c), and 10 s−1 (d).
Metals 14 01114 g005
Figure 6. Comparison of calculated and experimental steady state true stress of Al3Zn3Mg3Cu (a) and Al3Zn3Mg3CuEr (b) alloys.
Figure 6. Comparison of calculated and experimental steady state true stress of Al3Zn3Mg3Cu (a) and Al3Zn3Mg3CuEr (b) alloys.
Metals 14 01114 g006
Figure 7. Dependence of the activation volume on the temperature at a steady state.
Figure 7. Dependence of the activation volume on the temperature at a steady state.
Metals 14 01114 g007
Figure 8. Three-dimensional processing maps of the hot deformation behavior of Al3Zn3Mg3Cu (a) and Al3Zn3Mg3CuEr (b) alloys.
Figure 8. Three-dimensional processing maps of the hot deformation behavior of Al3Zn3Mg3Cu (a) and Al3Zn3Mg3CuEr (b) alloys.
Metals 14 01114 g008
Figure 9. Evaluation of the grain structure of the Al3Zn3Mg3Cu alloy under compression at temperatures of 350–475 °C and strain rates of 0.01–10 s−1 (OM).
Figure 9. Evaluation of the grain structure of the Al3Zn3Mg3Cu alloy under compression at temperatures of 350–475 °C and strain rates of 0.01–10 s−1 (OM).
Metals 14 01114 g009
Figure 10. Evaluation of the grain structure of the Al3Zn3Mg3CuEr alloy under compression at temperatures of 350–500 °C and strain rates of 0.01–10 s−1 (OM).
Figure 10. Evaluation of the grain structure of the Al3Zn3Mg3CuEr alloy under compression at temperatures of 350–500 °C and strain rates of 0.01–10 s−1 (OM).
Metals 14 01114 g010
Figure 11. The grain structure of the Al3Zn3Mg3Cu (a) and Al3Zn3Mg3CuEr (b) alloys after deformation at 450 °C and 0.01 s−1 (OM).
Figure 11. The grain structure of the Al3Zn3Mg3Cu (a) and Al3Zn3Mg3CuEr (b) alloys after deformation at 450 °C and 0.01 s−1 (OM).
Metals 14 01114 g011
Table 1. Parameters of the solution-treated microstructure.
Table 1. Parameters of the solution-treated microstructure.
AlloyGrain Size, µmIntermetallic Particles Size, µmIntermetallic Particle Volume Fraction, %Precipitate Size, nmPrecipitate Volume Fraction, %
Al3Zn3Mg3Cu1000.5–310–1140–1000.26
Al3Zn3Mg3CuEr450.5–411–1220–600.35
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

Glavatskikh, M.V.; Gorlov, L.E.; Loginova, I.S.; Barkov, R.Y.; Khomutov, M.G.; Churyumov, A.Y.; Pozdniakov, A.V. Effect of Er on the Hot Deformation Behavior of the Crossover Al3Zn3Mg3Cu0.2Zr Alloy. Metals 2024, 14, 1114. https://doi.org/10.3390/met14101114

AMA Style

Glavatskikh MV, Gorlov LE, Loginova IS, Barkov RY, Khomutov MG, Churyumov AY, Pozdniakov AV. Effect of Er on the Hot Deformation Behavior of the Crossover Al3Zn3Mg3Cu0.2Zr Alloy. Metals. 2024; 14(10):1114. https://doi.org/10.3390/met14101114

Chicago/Turabian Style

Glavatskikh, Maria V., Leonid E. Gorlov, Irina S. Loginova, Ruslan Yu. Barkov, Maxim G. Khomutov, Alexander Yu. Churyumov, and Andrey V. Pozdniakov. 2024. "Effect of Er on the Hot Deformation Behavior of the Crossover Al3Zn3Mg3Cu0.2Zr Alloy" Metals 14, no. 10: 1114. https://doi.org/10.3390/met14101114

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