Next Article in Journal
Surface Post-Treatment Induced by Nanosecond Pulsed Laser Processing of HVAF-Sprayed Fe-Based Metallic Glass Coating
Next Article in Special Issue
Fabrication and Characterization of In Situ Ti-6Al-4V/TiB Composites by the Hot-Pressing Method using Recycled Metal Chips
Previous Article in Journal
Thermal Performance of the Thin Heat Pipe for Cooling of Solid-State Drives
Previous Article in Special Issue
The Microstructure, Mechanical Properties, and Corrosion Resistance of a Novel Extruded Titanium Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Modified Johnson–Cook Model and Microstructure Evolution of As-Extruded AA 2055 Alloy during Isothermal Compression

1
Mechanical Staff Room of Fundamental Department, Aviation University of Air Force, Changchun 130022, China
2
Physical and Chemical Staff Room of Fundamental Department, Aviation University of Air Force, Changchun 130022, China
3
State Key Laboratory of Solidification Processing, School of Materials Science & Engineering, Northwestern Polytechnical University, Xi’an 710072, China
4
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences (CAS), Changchun 130022, China
5
Heavy Machinery Engineering Research Center of the Ministry of Education, Taiyuan University of Science and Technology, Taiyuan 030024, China
6
Ganjiang Innovation Academy, Chinese Academy of Sciences (CAS), Ganzhou 341119, China
*
Authors to whom correspondence should be addressed.
Metals 2022, 12(11), 1787; https://doi.org/10.3390/met12111787
Submission received: 30 September 2022 / Revised: 19 October 2022 / Accepted: 20 October 2022 / Published: 23 October 2022
(This article belongs to the Special Issue Hot Forming/Processing of Metallic Materials)

Abstract

:
Isothermal compression behaviors of as-extruded AA 2055 alloy (T6 state) were studied at temperature of 320, 380, 440 and 500 °C with strain rate of 0.001, 0.01, 0.1 and 1 s−1 by a Gleeble-3800 testing machine. A modified Johnson–Cook model fitted by polynomial and power-exponential functions were established to describe the flow stress of the alloy. The constitutive models fitted by higher-order polynomials were more accurate than the ones fitted by second-order polynomial and power-exponential functions. The constitutive model fitted by a fourth-order polynomial was chosen for the optimal constitutive model in order to balance the prediction accuracy and model complexity. The modified Johnson–Cook constitutive model could predict the flow stress well, especially in high-temperature zone (around 500 °C) and low-temperature zone (around 320 °C). The dynamic precipitation and dissolution of the T1 phase during hot compression were discussed. The unusual dynamic precipitation of the T2 phase was investigated during hot compression by XRD and TEM. The massive dense fine precipitates effectively pinned dislocations or subgrain boundaries to accelerate DRV but suppressed DRX, leading to a low frequency of HAGBs in compressed samples.

1. Introduction

Lithium is the lightest metallic element on earth. Former research studies have revealed that an addition of 1 wt.% of lithium into an aluminum matrix can decrease the density of the aluminum alloy by 3% and increase the elastic modulus of the aluminum alloy by 5~6% [1,2,3,4,5]. Compared to the traditional 2xxx and 7xxx aluminum alloys, an Al-Li alloy has a low density, high elastic modulus, high specific strength, high specific elastic modulus and low fatigue crack growth rate [4,5,6,7]. An Al-Li alloy also has good mechanical properties at elevated temperature or at low temperature [6,8]. Compared with carbon-fiber-reinforced plastic (CFRP), an Al-Li alloy has irreplaceable advantages in impact resistance, plasticity and repairability. The AA 2055 alloy belongs to the third generation Al-Li alloys and was developed by Aluminum Company of America (Alcoa) in 2012 [8]. The alloy exhibits a higher fracture strength and fracture toughness than other third-generation Al-Li alloys through tweaking the alloying elements without increasing cost [9].
The Al-Li alloy components are always processed by plastic deformation [9]. In some case, the wrought Al alloys require further deformation, such as large-sized as-extruded bars or forged billet, while the reports about isothermal compression of wrought Al alloys are few, therefore an AA 2055 as-extruded bar with a diameter of 100 mm, a length of about 500 mm was chosen as the initial material in this work. The flow stress and constitutive model are important factors to describe the hot plastic deformation of an Al-Li alloy. During the hot deformation, a work hardening and softening of dynamic recovery (DRV) and dynamic recrystallization (DRX) both take place, under the joint impact of several thermal deformation parameters, such as the composition of the alloy, deformation temperature, strain rate and initial state [10,11,12,13]. A constitutive model establishes the relationship among flow stress, flow strain, deformation temperature and strain rate. Constructing an accurate and efficient constitutive model is the key to determine process route and optimize process parameters [14].
According to theoretical derivation methods, constitutive models can be divided into phenomenological models, physical models and neural network models [14,15]. A phenomenological constitutive model mainly reflects the macroscopic factors of hot deformation [15], without including the microscopic mechanism of hot deformation, such as the Arrhenius type model [16,17,18]. A physical model is built on a certain physical background, considering the influence of relevant physical parameters, such as the Zerilli–Armstrong (ZA) model [19,20,21]. However, in order to improve prediction accuracy, the physical constitutive model always introduces numerous complex variables which are hardly obtained. A neural network model is used to predict the flow stress of hot deformation by manual training because of its efficient nonlinear mapping ability [22,23]. Although a neural network model has a higher prediction accuracy resulting from its nonlinear mapping ability, its convergence speed is slow, and the problem of local minimization is common [14]. Therefore, a phenomenological model is a relatively appropriate model for an FEM analysis and industrial production.
The Johnson–Cook constitutive model is a classical phenomenological model which can predict the constitutive relationship of metal materials including the effect of strain, strain rate and temperature on the flow stress [17,24]. It is widely used in the numerical simulation of metal forming because the model is simple in form and the parameters are easy to obtain. However, since the Johnson–Cook constitutive model assumes that strain, strain rate and temperature have independent effects on the hot deformation behavior, the accuracy of prediction of the original model is inadequate. Our former report established the Arrhenius model of the alloy, which only described the relationship between stress and temperature or strain rate without reflecting the effect of strain on the flow stress [25]. Therefore, a modified Johnson–Cook constitutive model was established considering the coupling effect of temperature, strain rate and strain to improve the accuracy of model prediction and build the direct relationship between flow stress and strain.
The T1 phase (Al2CuLi), β’ (Al3Zr), the T2 phase (Al6CuLi3), δ’(Al3Li) and θ’(Al2Cu) are the precipitates in the Al-Li-Cu-Zr alloys forming according to the process of hot deformation and heat treatment [2,9,26,27,28]. The dissolution or precipitation of the second-phase particles in Al-Li alloys may take place during every stage of the hot deformation. The influence of variation of the second-phase particles on the microstructure evolution during hot deformation is non-negligible. The second-phase particles may retard or accelerate dynamic recovery (DRV) and dynamic recrystallization (DRX) depending on their size, spatial distribution and deformation conditions in Al-Li alloys [10,13,29]. According to former research studies [10,11,13,30,31], the dispersed fine second-phase particles can pin dislocations or subgrain boundaries to accelerate DRV but suppress DRX in Al-Li alloys. While the large stable particles (>1 μm in diameter) can promote DRX by producing regions of high lattice misorientation to offer preferential sites for recrystallization, i.e., particle-stimulated nucleation (PSN) [29,32], the second-phase particles also affect the deformation structure and texture depending on thermomechanical processing and their diameters and deformability. In general, Al-Li alloys are strengthened by precipitates of ordered atomic structure. Moreover, the fine deformable precipitates are sheared by moving dislocations leading to a strong slip localization on a few slip planes which can puncture adjacent grains to form shear bands [29].
The modified Johnson–Cook constitutive model fitted by a second-order polynomial could predict the flow curves well in former reports, in the case where the stress of alloy changed monotonically with the increase of strain on the whole after the stress reached the peak value [33,34,35]. However, the experimental results in this paper suggest that the overall stress–strain relationship of the Al-Li alloy cannot be considered monotonic after the peak stress at different temperatures and strain rates, which is quite different from previous studies. In order to improve the prediction accuracy, a modified Johnson–Cook model fitted by a power-exponential function and a higher-order polynomial is also established in this work. In addition, the dynamic precipitation and dissolution and the effect of the second phase on DRV and DRX in as-extruded AA 2055 alloy during hot compression are also discussed in this work.

2. Materials and Methods

The as-extruded AA 2055 alloy (T6) with a composition of 3.5% Cu, 1.1% Li, 0.5% Zn, 0.3% Mn, 0.4% Mg, 0.1% Zr, 0.3% Ag, 0.1% Ti, 0.15% Si, 0.2% Fe, <0.15% other elements except the above elements and Al (all in wt.%) was used. The composition of the alloy was analyzed by inductively coupled plasma spectrometry (ICP, ThermoScientific iCAP6300, Waltham, MA, USA). The hot compression experiment was carried out by a Gleeble-3800 thermo-force simulation system (DSI, New York, NY, USA). The heating rate was 5 °C/s with a holding time of 3 min. The temperatures of hot compression were 320 °C, 380 °C, 440 °C and 500 °C, and the strain rates were 0.001 s−1, 0.01 s−1, 0.1 s−1 and 1 s−1. The experiment with a higher strain rate of 5 s−1 at 380 °C was also carried out, but a deformation instability took place during hot compression; the result can be obtained from our former report [25]. The samples for hot compression were cut by wire electrical discharge machining from the as-extruded AA 2055 alloy with a diameter size of 10 mm × length 15 mm, as shown in Figure 1. The compressed samples were quenched by water with a height reduction of 60%. The detailed experimental procedures and the flow curves can be obtained from our previous report [25], as shown in Figure 1 and Figure 2.
In order to observe the precipitation behavior during hot compression, the X ray diffraction (XRD) analysis of the original sample and compressed samples were carried out (D8 Advance Bruker AXS, Karlsruhe, Germany) with a scan speed of 4°/min from 20° to 90°. The samples for the electron backscattered diffraction (EBSD) were sectioned parallel to the compression direction (CD) and microstructure observations were taken in the center of the sample, as shown in Figure 1. The samples were mechanically polished carefully, then electrolytically polished in a solution consisting of 10 mL of HClO4 and 90 mL of C2H5OH with a voltage of 25 V at room temperature. The EBSD analysis was conducted on a field emission scanning electron microscope (Qunta 250 FEI, Hillsboro, USA) equipped with an Oxford EBSD detector operating at 30 kV. The foils for transmission electron microscopy (TEM) were also taken from the compressed samples parallel to the compression direction as shown in Figure 1. The samples were polished to 30 μm in thickness followed by ion milling (Leica RES101, Wetzlar, Germany). The TEM analysis was carried out on a transmission electron microscopy (JEOL-F200, Tokyo, Japan) with an accelerating voltage of 200 kV.

3. Results and Discussion

3.1. The Modified Johnson–Cook Model

The Johnson–Cook model is suitable for describing hot deformation with a wide range of strains and temperatures [36]. The flow stress is the function of temperature, strain rate and strain, as shown in Equation (1),
σ = ( γ 1 + γ 2 ε n ) ( 1 + K 1 ln ε ˙ * ) ( 1 T * m )
where σ , ε , ε ˙ , and ε ˙ * are the flow stress, the effective plastic strain, the strain rate and the ratio of strain rate to reference strain rate. γ 1 , γ 2 and n denote the yield strength, the strain hardening coefficient and the strain hardening exponent at a reference strain rate and reference temperature. K1 is the strain-rate hardening coefficient and m is the thermal softening exponent. The homologous temperature T * can be calculated by Equation (2), where T r , T m and T represent the reference temperature, melting point of the material and the current temperature, respectively [36]:
T * = T T r T m T r
The relationship between stress and strain can be fitted by Equations (3) and (4) according to the compression curves with the effect of the variation of temperature and strain rate on stress ignored.
σ = ( a 1 + a 2 ε 1 + a 3 ε 2 )
σ = ( a 1 + a 2 ε a 3 )
Considering the effect of the strain rate on the flow stress, the relationship between σ and ln ( ε ˙ / ε r ) with the compression data at temperature of 593.15 K (320 °C) and strain rate of 0.001 s1 can be expressed as Equation (5), as shown in Figure 3.
σ = b 1 + b 2 ln ( ε ˙ / ε r ) ε r = 0.001   s 1
When considering the effect of temperature variation on stress, it can be found from Figure 4a that the logarithm of the stress is linear with the difference between the current temperature and the reference temperature ( T T r ) , when the strain rate is 0.01 s−1, and the relationship between the stress and the temperature can be expressed by Equation (6).
σ = exp ( M ( T T r ) )
where Tr = 593.15 K (320 °C), M is a parameter reflecting the coupling effect of temperature and strain rate, which can be expressed as Equation (7).
M = c 1 + c 2 ln ( ε ˙ / ε r ) + c 3 ( ln ( ε ˙ / ε r ) ) 2
The values of c 1 , c 2 and c 3 can be determined by the values of M with strain rates of 0.001 s−1, 0.01 s−1, 0.1 s−1 and 1 s−1, respectively.
In summary, the modified Johnson-Cook constitutive model can be expressed as the following equation, which considers the coupling effect among temperature, strain rate and strain. The model fitted by a polynomial of order two is shown below:
σ = ( a 1 + a 2 ε 1 + a 3 ε 2 ) [ 1 + a 4 ln ( ε ˙ / ε r ) ] exp [ [ a 5 + a 6 ln ( ε ˙ / ε r ) + a 7 ( ln ( ε ˙ / ε r ) ) 2 ] ( T T r ) ]
The model fitted by the power function is shown below:
σ = ( a 1 + a 2 ε a 3 ) [ 1 + a 4 ln ( ε ˙ / ε r ) ] exp [ [ a 5 + a 6 ln ( ε ˙ / ε r ) + a 7 ( ln ( ε ˙ / ε r ) ) 2 ] ( T T r ) ]
When T = 593.15   K and ε ˙ = 0.001   s 1 , Equations (8) and (9) can be simplified as Equations (3) and (4), respectively. According to the corresponding data from compressive stress–strain curves, the values of a 1 , a 2 and a 3 in Equations (8) and (9) were 74.19562, −30.20233, 17.79643 and 122.49779, −62.66139 and 0.09106, respectively.
When T = 593.15   K , Equations (8) and (9) can be simplified as follows:
σ = ( a 1 + a 2 ε + a 3 ε 2 ) [ 1 + a 4 ln ( ε ˙ / ε r ) ]
σ = ( a 1 + a 2 ε a 3 ) [ 1 + a 4 ln ( ε ˙ / ε r ) ]
Based on the data of strain rates of 0.001 s−1, 0.01 s−1, 0.1 s−1 and 1 s−1, the parameter a 4 in Equations (10) and (11) were 0.156564 and 0.156746, respectively. Moreover, the parameters a 5 , a 6 and a 7 in Equations (8) and (9) were −0.01255, 0.0021 and −0.000158905, respectively. The detailed parameters of the modified Johnson–Cook constitutive model are listed in Table 1.
The prediction accuracy of the constitutive model can be judged by the correlation coefficient R and average relative error A A R E , as shown below.
R = i = 1 N ( σ e i σ ¯ e ) ( σ p i σ ¯ p ) i = 1 N ( σ e i σ ¯ e ) 2 i = 1 N ( σ p i σ ¯ p ) 2
A A R E ( % ) = 1 N i = 1 N | σ e i σ p i σ e i | × 100 %
The closer the value of R is to 1, the smaller the value of A A R E , the more accurate the prediction of the model. Comparing the predicted data of the constitutive model with the actual compression stress–strain curve, the values of R for the model fitted by a second-order polynomial and a power-exponential function were 0.9639 and 0.9835, respectively, and the corresponding values of A A R E were 0.106722 and 0.10648, respectively. That is, the prediction of the constitutive model fitted by a power-exponential function was more accurate than that fitted by a second-order polynomial. Figure 5 and Figure 6 show the comparison between the prediction of the constitutive model fitted by a second-order polynomial and the actual compression stress–strain curves and the comparison between the prediction of the constitutive model fitted by a power-exponential function and the actual compression stress–strain curves.
In order to improve the prediction accuracy, the constitutive model fitted by a higher-order polynomial was developed. The stress was expressed as a polynomial function of the third, fourth and fifth order of the strain, and the corresponding constitutive models are shown in Equations (14)–(16), respectively.
σ = ( a 1 + a 2 ε 1 + a 3 ε 2 + a 4 ε 3 ) [ 1 + a 5 ln ( ε ˙ / ε r ) ] exp [ [ a 6 + a 7 ln ( ε ˙ / ε r ) + a 8 ( ln ( ε ˙ / ε r ) ) 2 ] ( T T r ) ]
σ = ( a 1 + a 2 ε 1 + a 3 ε 2 + a 4 ε 3 + a 5 ε 4 ) [ 1 + a 6 ln ( ε ˙ / ε r ) ] exp [ [ a 7 + a 8 ln ( ε ˙ / ε r ) + a 9 ( ln ( ε ˙ / ε r ) ) 2 ] ( T T r ) ]
σ = ( a 1 + a 2 ε 1 + a 3 ε 2 + a 4 ε 3 + a 5 ε 4 + a 6 ε 5 ) [ 1 + a 7 ln ( ε ˙ / ε r ) ] exp [ [ a 8 + a 9 ln ( ε ˙ / ε r ) + a 10 ( ln ( ε ˙ / ε r ) ) 2 ] ( T T r ) ]
The values of R and AARE of the constitutive models fitted by different higher-order polynomials are shown in Table 2. It can be concluded that the constitutive models fitted by higher-order polynomials were more accurate than the models fitted by a second-order polynomial and a power-exponential function. Although the prediction accuracy of the constitutive model can be improved by increasing the order of the fitting polynomial, the improvement of prediction accuracy is limited when the order of the polynomial exceeds four. Moreover, the model becomes more complex as the number the parameters of the model increases. Therefore, the constitutive model fitted by a fourth order polynomial was chosen in this work.
Figure 7 shows the comparison between the predictive value of the constitutive model fitted by a fourth-order polynomial and the experimental compression stress–strain curve. Figure 8 shows the correlation between the predicted flow stress value of the model fitted by a fourth-order polynomial and the experimental value. According to the comparison between the predicted and experimental values, the modified Johnson–Cook constitutive model could predict the flow stress during the isothermal compression and behaved better in the high-temperature zone (around 500 °C) and in the low-temperature zone (around 320 °C) than that in the medium-temperature zone (380~440 °C).

3.2. Microstructures of the Compressed Samples

Figure 9 shows the XRD patterns of the original sample and the compressed samples with different deformation conditions. The selected area electron diffraction (SAED) patterns of the T1 phase and β’ phase were obtained in the compressed sample at 440 °C/0.001 s−1 from a previous report [25]. The original sample contained the T1 phase (Al2CuLi) and β’ (Al3Zr) phase, while all the compressed samples contained the T2 phase (Al6CuLi3) besides the T1 phase and the β’ phase. It is known that the T1 phase is the predominant precipitate in Al-Cu-Li alloys, which has been widely investigated and forms as thin hexagonal plates with a {111}Al matrix habit plane [1,2,3,4]. The metastable Al3Zr phase always presents spherical particles in third-generation Al-Li alloys and is an important precipitate in the Al-Cu-Zr system, which has a LI2→D023 structural transformation with a Cu addition [26]. The T2 phase is a stable intermetallic (icosahedral, point group m 3 ¯ 5 ¯ ) forming via solid-state phase transformation on HAGBs, which may be detrimental to the toughness of the Al-Li alloy [37]. Figure 10a shows the schematic precipitate microstructure of the precipitates in compressed samples [9].
When the deformation temperature is low (320 °C), plenty of dislocations are introduced into the grains and the strong interaction between the T1 phase and dislocations are also observed, as shown in Figure 10b. The green arrows indicate dislocations, and the yellow ones indicate the T1 phase. The dislocation motion is impeded by precipitates distributing like a network. On the other hand, part of the T1 precipitates are distorted or broken by the dislocation movement, as shown in Figure 10c. The report from Sha et al. [38] reveals that plastic deformation could change the orientation between the precipitates and the matrix slightly, as well as the atomic structure of the interface, resulting in a growth of the interface mismatch energy. In addition, the plastic deformation can increase the distortion energy of the alloy. The surface energy of the precipitates also increases with precipitates being broken. The combined effects of the above factors result in the rise of free energy and redissolution of the T1 phase, which induces the transformation of the T1 phase from rodlike to spherical, as shown in Figure 10c. The redissolution of the T1 phase make the Al matrix become a supersaturated solid solution again followed by a secondary precipitation of the precipitates.
Moreover, the dislocations induced by deformation provide heterogeneous nucleation positions for the T1 phase nucleation. The stress field caused by the interaction between the dislocation and T1 phase also provides the required energy for the T1 phase nucleation. The short-circuit diffusion channel provided by the dislocations reduces the atomic diffusion activation energy, which is beneficial to increase the precipitates’ nucleation rate [39]. A former research [40] showed that further deformation of the precipitates during the process of redissolution might lead to a secondary precipitation of the precipitates. It can be observed that the fine T1 phase reprecipitated around the dislocation and subgrain boundary as indicated by the yellow arrows in Figure 10d. In addition, former investigations [41] suggested that Al3Zr particles and the dislocation loops around the Al3Zr phase could provide heterogeneous nucleation positions for the T1 phase nucleation, which can be observed in Figure 10d.
The T2 phase was found to be stable to about 420 °C and might reprecipitate at the high-angle grain boundaries (HAGBs), preferentially during the aging of the 8090 alloy [37]. However, in this work, the T2 phase only existed in the compressed sample but not in the original sample according to the results of XRD. The temperature of deformation ranged from 320 °C to 440 °C, which was suitable for the reprecipitation of the T2 phase. High-angle grain boundaries generated by dynamic recrystallization provided positions for the T2 phase nucleation. The distortion energy caused by the deformation provided the driving force for the T2 phase nucleation. Therefore, it was a reasonable result that the dynamic precipitation of the T2 phase took place during hot compression. As shown in Figure 10d, the coarse T2 particles can be observed at the HAGBs indicated by blue arrows.
Figure 11 shows the EBSD IPF maps of the compressed samples with deformation conditions of 320 °C/1 s−1, 320 °C/0.001 s−1, 440 °C/1 s−1 and 440 °C/0.001 s1. Figure 12 shows the misorientation angle distribution histograms of the samples. It can be observed that a large fraction of low-angle grain boundaries (<15°, LAGBs) existed in all the samples. With the deformation temperature increasing and strain rate decreasing, the frequency of HAGBs increased obviously. Furthermore, the sample of 440 °C/0.001 s−1 had a maximum frequency of HAGBs of 21.6% and a maximum average of misorientation of 9.91°. It could effectively promote DRX to increase the temperature or decrease the strain rate by providing sufficient energy or time to facilitate dislocation rearrangement and forming DRXed grains.
As mentioned above, the second-phase particles may retard or accelerate dynamic recovery (DRV) and dynamic recrystallization (DRX) depending on their size and spatial distribution. The large particles (>1 μm in diameter) can promote DRX via PSN. A recent report from Wang et al. [41] revealed that a fraction of recrystallization of 67% in an Al-Cu-Li alloy was triggered by a sparse coarsened T1 phase via PSN during hot compression at 380 °C/0.01 s−1. There were massive dense and fine T1 phase precipitates in the Al-Cu-Li alloys in this work, which effectively pinned dislocations (as shown in Figure 10b) or subgrain boundaries to accelerate DRV but suppress DRX. Therefore, a very small fraction of HAGBs existed in all the compressed samples.

4. Conclusions

In this work, the isothermal compression behaviors of as-extruded AA 2055 alloy were studied. A modified Johnson–Cook model was established to describe the flow stress of the alloy. The microstructures of compressed samples were investigated via XRD, EBSD and TEM. Several conclusions could be obtained, as shown below.
(1)
The constitutive model fitted by a fourth-order polynomial was chosen for the optimal modified Johnson–Cook constitutive model in order to balance the prediction accuracy and model complexity.
(2)
The modified Johnson–Cook constitutive model could predict the flow stress of the as-extruded AA 2055 alloy during isothermal compression well, especially in the high-temperature zone (around 500 °C) and the low-temperature zone (around 320 °C).
(3)
The massive dense fine precipitates effectively pinned dislocations or subgrain boundaries to accelerate DRV but suppress DRX. An unusual dynamic precipitation of the T2 phase was observed in the compressed samples.

Author Contributions

Conceptualization, D.J. and L.C.; methodology, J.Z.; validation, D.J., T.L. and Z.W.; formal analysis, T.H.; investigation, D.J.; resources, L.C., T.H. and C.C.; data curation, C.C.; writing—original draft preparation, D.J.; writing—review and editing, D.J.; visualization, W.L.; funding acquisition, L.C., T.H. and C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2020YFE0204500, key deployment projects of the Chinese Academy of Sciences, grant number ZDRW-CN-2021-3-4-15, National Natural Science Foundation of China, grant numbers 51905366, 51901149, 52275382 and U1910213, China Postdoctoral Science Foundation, grant number 2021M692626, Changchun Science and Technology Development Plan Project, grant number 21GD03, Youth Science and Technology Research Fund in Shanxi, grant number 201901D211307, Scientific and Technological Innovation Programs of Higher Education Institutions in Shanxi (STIP), grant number 2019L0625.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, P.; Chen, M. Progress in characterization methods for thermoplastic deforming constitutive models of Al-Li alloys: A review. J. Mater. Sci. 2020, 55, 9828–9847. [Google Scholar] [CrossRef]
  2. Zhang, Y.; Ling, J.; Li, H.-G.; Luo, X.-Y.; Ba, Z.-X. Effect of pre-deformation and artificial aging on fatigue life of 2198 Al-Li alloy. Mater. Res. Express 2020, 7, 046509. [Google Scholar] [CrossRef]
  3. Wang, X.; Xiao, D.; Wu, M.; Liu, W. Deformation behavior and microstructure evolution of Al–5.06Mg–1.67Li–0.51Zn alloy under hot compression. J. Mater. Res. Technol. 2021, 15, 4516–4528. [Google Scholar] [CrossRef]
  4. Zhao, T.; Sato, Y.S.; Xiao, R.; Huang, T.; Zhang, J. Hardness distribution and aging response associated with precipitation behavior in a laser pressure welded Al-Li alloy 2198. Mater. Sci. Eng. A Struct. Mater. Prop. Microstruct. Process. 2021, 808, 140946. [Google Scholar] [CrossRef]
  5. Zhang, C.; Liu, M.; Meng, Z.; Zhang, Q.; Zhao, G.; Chen, L.; Zhang, H.; Wang, J. Microstructure evolution and precipitation characteristics of spray-formed and subsequently extruded 2195 Al-Li alloy plate during solution and aging process. J. Mater. Process. Technol. 2020, 283, 116718. [Google Scholar] [CrossRef]
  6. Sun, J.; Wu, G.; Zhang, L.; Zhang, X.; Liu, L.; Zhang, J. Microstructure characteristics of an ultra-high strength extruded Al-4.7Cu–1Li-0.5Mg-0.1Zr–1Zn alloy during heat treatment. J. Alloys Compd. 2020, 813, 152216. [Google Scholar] [CrossRef]
  7. Li, S.; Wei, B.; Yu, C.; Li, Y.; Xu, G.; Li, Y. Evolution of microstructure and properties during homogenization of the novel Al-Li alloy fabricated by electromagnetic oscillation twin-roll casting. J. Mater. Res. Technol. 2020, 9, 3304–3317. [Google Scholar] [CrossRef]
  8. Balducci, E.; Ceschini, L.; Messieri, S. High Temperature Tensile Tests of the Lightweight 2099 and 2055 Al-Cu-Li Alloy: A Comparison. JOM 2018, 70, 2716–2725. [Google Scholar] [CrossRef]
  9. Rioja, R.J.; Liu, J. The Evolution of Al-Li Base Products for Aerospace and Space Applications. Metall. Mater. Trans. A 2012, 43, 3325–3337. [Google Scholar] [CrossRef]
  10. Xiang, S.; Liu, D.-y.; Zhu, R.-h.; Li, J.-f.; Chen, Y.-l.; Zhang, X.-h. Hot deformation behavior and microstructure evolution of 1460 Al–Li alloy. Trans. Nonferrous Met. Soc. China 2015, 25, 3855–3864. [Google Scholar] [CrossRef]
  11. Zhu, R.; Liu, Q.; Li, J.; Xiang, S.; Chen, Y.; Zhang, X. Dynamic restoration mechanism and physically based constitutive model of 2050 Al–Li alloy during hot compression. J. Alloys Compd. 2015, 650, 75–85. [Google Scholar] [CrossRef]
  12. Yang, Q.; Wang, X.; Li, X.; Deng, Z.; Jia, Z.; Zhang, Z.; Huang, G.; Liu, Q. Hot deformation behavior and microstructure of AA2195 alloy under plane strain compression. Mater. Charact. 2017, 131, 500–507. [Google Scholar] [CrossRef]
  13. Li, S.Q.; Deng, Z.H.; Yang, Q.Y.; Lin, L.; Guo, Y.L.; Zhang, Z.Q. Hot Deformation Behavior of 2195 Al-Li Alloy Under Plane Strain Compression. Rare Met. Mater. Eng. 2018, 47, 553–559. [Google Scholar]
  14. Jiao, M.Y.; Ma, L.F.; Jia, W.T.; Zhu, Y.C.; Lin, J.B.; Le, Q.C. A new phenomenological model describing the compressive thermal deformation flow stress of cast-rolled AZ31B Mg alloy. Mater. Res. Express 2019, 6, 096597. [Google Scholar] [CrossRef]
  15. Rezaei Ashtiani, H.R.; Shayanpoor, A.A. Prediction of the Hot flow Behavior of AA1070 Aluminum Using the Phenomenological and Physically-Based Models. Phys. Met. Metallogr. 2021, 122, 1436–1445. [Google Scholar] [CrossRef]
  16. Li, H.-Y.; Li, Y.-H.; Wang, X.-F.; Liu, J.-J.; Wu, Y. A comparative study on modified Johnson Cook, modified Zerilli–Armstrong and Arrhenius-type constitutive models to predict the hot deformation behavior in 28CrMnMoV steel. Mater. Des. 2013, 49, 493–501. [Google Scholar] [CrossRef]
  17. Chen, X.; Liao, Q.; Niu, Y.; Jia, W.; Le, Q.; Cheng, C.; Yu, F.; Cui, J. A constitutive relation of AZ80 magnesium alloy during hot deformation based on Arrhenius and Johnson–Cook model. J. Mater. Res. Technol. 2019, 8, 1859–1869. [Google Scholar] [CrossRef]
  18. Bu, H.Y.; Li, Q.; Li, S.H.; Li, M.N. Comparison of Modified Johnson-Cook Model and Strain-Compensated Arrhenius Constitutive Model for 5CrNiMoV Steel during Compression around Austenitic Temperature. Metals 2022, 12, 1270. [Google Scholar] [CrossRef]
  19. Zerilli, F.J.; Armstrong, R.W. Dislocation-mechanics-based constitutive relations for material dynamics calculations. J. Appl. Phys. 1987, 61, 1816–1825. [Google Scholar] [CrossRef] [Green Version]
  20. Shayanpoor, A.A.; Ashtiani, H.R.R. The phenomenological and physical constitutive analysis of hot flow behavior of Al/Cu bimetal composite. Appl. Phys. A Mater. 2022, 128, 636. [Google Scholar] [CrossRef]
  21. Venkata Ramana, A.; Balasundar, I.; Davidson, M.J.; Balamuralikrishnan, R.; Raghu, T. Constitutive Modelling of a New High-Strength Low-Alloy Steel Using Modified Zerilli–Armstrong and Arrhenius Model. Trans. Indian Inst. Met. 2019, 72, 2869–2876. [Google Scholar] [CrossRef]
  22. Sani, S.A.; Ebrahimi, G.R.; Vafaeenezhad, H.; Kiani-Rashid, A.R. Modeling of hot deformation behavior and prediction of flow stress in a magnesium alloy using constitutive equation and artificial neural network (ANN) model. J. Magnes. Alloy. 2018, 6, 134–144. [Google Scholar] [CrossRef]
  23. Ping, L.; Kemin, X.; Yan, L.; Jianrong, T. Neural network prediction of flow stress of Ti–15–3 alloy under hot compression. J. Mater. Process. Technol. 2004, 148, 235–238. [Google Scholar] [CrossRef]
  24. He, J.; Chen, F.; Wang, B.; Zhu, L.B. A modified Johnson-Cook model for 10%Cr steel at elevated temperatures and a wide range of strain rates. Mater. Sci. Eng. A 2018, 715, 1–9. [Google Scholar] [CrossRef]
  25. Jiang, D.M.; Yuan, C.Q.; Li, X.; Yu, H. Hot deformation behavior and microstructure characteristic of 2055 Al-Li alloy during uniaxial compression. Mater. Res. Express 2020, 7, 14. [Google Scholar] [CrossRef]
  26. Jia, Z.H.; Couzinie, J.P.; Cherdoudi, N.; Guillot, I.; Arnberg, L.; Asholt, P.; Brusethaug, S.; Barlas, B.; Massinon, D. Precipitation behaviour of Al3Zr precipitate in Al-Cu-Zr and Al-Cu-Zr-Ti-V alloys. Trans. Nonferrous Met. Soc. China 2012, 22, 1860–1865. [Google Scholar] [CrossRef]
  27. Decreus, B.; Deschamps, A.; De Geuser, F.; Donnadieu, P.; Sigli, C.; Weyland, M. The influence of Cu/Li ratio on precipitation in Al-Cu-Li-x alloys. Acta Mater. 2013, 61, 2207–2218. [Google Scholar] [CrossRef]
  28. Deschamps, A.; Decreus, B.; De Geuser, F.; Dorin, T.; Weyland, M. The influence of precipitation on plastic deformation of Al–Cu–Li alloys. Acta Mater. 2013, 61, 4010–4021. [Google Scholar] [CrossRef]
  29. Huang, K.; Marthinsen, K.; Zhao, Q.; Logé, R.E. The double-edge effect of second-phase particles on the recrystallization behaviour and associated mechanical properties of metallic materials. Prog. Mater. Sci. 2018, 92, 284–359. [Google Scholar] [CrossRef]
  30. Ou, L.; Zheng, Z.; Nie, Y.; Jian, H. Hot deformation behavior of 2060 alloy. J. Alloys Compd. 2015, 648, 681–689. [Google Scholar] [CrossRef]
  31. Yu, W.; Li, H.; Du, R.; You, W.; Zhao, M.; Wang, Z.-a. Characteristic constitution model and microstructure of an Al-3.5Cu-1.5Li alloy subjected to thermal deformation. Mater. Charact. 2018, 145, 53–64. [Google Scholar] [CrossRef]
  32. Niu, L.Q.; Cao, M.; Liang, Z.L.; Han, B.; Zhang, Q. A modified Johnson-Cook model considering strain softening of A356 alloy. Mater. Sci. Eng. A Struct. Mater. Prop. Microstruct. Process. 2020, 789, 139612. [Google Scholar] [CrossRef]
  33. Chen, X.M.; Lin, Y.C.; Hu, H.W.; Luo, S.C.; Zhou, X.J.; Huang, Y. An Enhanced Johnson-Cook Model for Hot Compressed A356 Aluminum Alloy. Adv. Eng. Mater. 2021, 23, 2000704. [Google Scholar] [CrossRef]
  34. Chao, Z.L.; Jiang, L.T.; Chen, G.Q.; Zhang, Q.; Zhang, N.B.; Zhao, Q.Q.; Pang, B.J.; Wu, G.H. A modified Johnson-Cook model with damage degradation for B4Cp/Al composites. Compos. Struct. 2022, 282, 115029. [Google Scholar] [CrossRef]
  35. Wang, X.Y.; Jiang, J.T.; Li, G.A.; Wang, X.M.; Sun, J.; Zhen, L. Effects of coarse Al2CuLi phase on the hot deformation behavior of Al-Li alloy. J. Alloys Compd. 2020, 815, 152469. [Google Scholar] [CrossRef]
  36. Wang, S.C.; Starink, M.J. Precipitates and intermetallic phases in precipitation hardening Al–Cu–Mg–(Li) based alloys. Int. Mater. Rev. 2005, 50, 193–215. [Google Scholar] [CrossRef] [Green Version]
  37. Sha, G.; Wang, Y.B.; Liao, X.Z.; Duan, Z.C.; Ringer, S.P.; Langdon, T.G. Influence of equal-channel angular pressing on precipitation in an Al–Zn–Mg–Cu alloy. Acta Mater. 2009, 57, 3123–3132. [Google Scholar] [CrossRef]
  38. Yang, S.L.; Shen, J.; Li, X.W.; Yan, X.D.; Zhang, Y.G.; Li, Z.H.; Huang, S.H.; Xiong, B.Q. Mechanical Behavior and Microstructure Evolution of Al-Cu-Li Alloy during Multi-Pass Hot Deformation. Rare Met. Mater. Eng. 2017, 46, 2825–2830. [Google Scholar]
  39. Garcia-Infanta, J.M.; Swaminathan, S.; Cepeda-Jimenez, C.M.; McNelley, T.R.; Ruano, O.A.; Carreno, F. Enhanced grain refinement due to deformation-induced precipitation during ambient-temperature severe plastic deformation of an Al-7%Si alloy. J. Alloys Compd. 2009, 478, 139–143. [Google Scholar] [CrossRef]
  40. Tsivoulas, D.; Robson, J.D.; Sigli, C.; Prangnell, P.B. Interactions between zirconium and manganese dispersoid-forming elements on their combined addition in Al–Cu–Li alloys. Acta Mater. 2012, 60, 5245–5259. [Google Scholar] [CrossRef]
  41. Wang, X.Y.; Jiang, J.T.; Li, G.A.; Wang, X.M.; Shao, W.-Z.; Zhen, L. Particle-stimulated nucleation and recrystallization texture initiated by coarsened Al2CuLi phase in Al–Cu–Li alloy. J. Mater. Res. Technol. 2021, 10, 643–650. [Google Scholar] [CrossRef]
Figure 1. The diagrammatic sketch of the experimental procedure. Adapted from ref. [25].
Figure 1. The diagrammatic sketch of the experimental procedure. Adapted from ref. [25].
Metals 12 01787 g001
Figure 2. The flow curves of 2055 Al-Li alloy at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, with strain rates ranging from 0.001 s−1 to 1 s−1. Reprinted from ref. [25].
Figure 2. The flow curves of 2055 Al-Li alloy at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, with strain rates ranging from 0.001 s−1 to 1 s−1. Reprinted from ref. [25].
Metals 12 01787 g002
Figure 3. The relationship between σ and ln ( ε ˙ / ε r ) at T = 593.15 K (320 °C).
Figure 3. The relationship between σ and ln ( ε ˙ / ε r ) at T = 593.15 K (320 °C).
Metals 12 01787 g003
Figure 4. (a) The relationship between ( T T r ) and ln σ ; (b) the relationship between ln ( ε ˙ / ε r ) and M .
Figure 4. (a) The relationship between ( T T r ) and ln σ ; (b) the relationship between ln ( ε ˙ / ε r ) and M .
Metals 12 01787 g004
Figure 5. Comparison between the predictive value of the constitutive model fitted by a second-order polynomial and the experimental stress-strain curves at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, respectively.
Figure 5. Comparison between the predictive value of the constitutive model fitted by a second-order polynomial and the experimental stress-strain curves at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, respectively.
Metals 12 01787 g005
Figure 6. Comparison between the predictive value of the constitutive model fitted by a power-exponential function and the experimental stress-strain curves at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, respectively.
Figure 6. Comparison between the predictive value of the constitutive model fitted by a power-exponential function and the experimental stress-strain curves at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, respectively.
Metals 12 01787 g006
Figure 7. Comparison between the predictive value of the constitutive model fitted by a fourth-order polynomial and the experimental stress-strain curves at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, respectively.
Figure 7. Comparison between the predictive value of the constitutive model fitted by a fourth-order polynomial and the experimental stress-strain curves at temperature of (a) 320 °C, (b) 380 °C, (c) 440 °C and (d) 500 °C, respectively.
Metals 12 01787 g007
Figure 8. The correlation between the predicted flow stress value of the model fitted by a fourth-order polynomial and the experimental value.
Figure 8. The correlation between the predicted flow stress value of the model fitted by a fourth-order polynomial and the experimental value.
Metals 12 01787 g008
Figure 9. The XRD patterns the deformed alloys: (a) original sample, compressed sample: (b) 320 °C/0.001 s−1; (c) 380 °C/0.001 s−1; (d) 440 °C/0.001 s−1; (e) 500 °C/0.001 s−1.
Figure 9. The XRD patterns the deformed alloys: (a) original sample, compressed sample: (b) 320 °C/0.001 s−1; (c) 380 °C/0.001 s−1; (d) 440 °C/0.001 s−1; (e) 500 °C/0.001 s−1.
Metals 12 01787 g009
Figure 10. (a) Schematic precipitate microstructure of the compressed samples; (b) strong interaction between dislocation and T1 phase in sample at 320 °C/0.001 s−1; (c) fracture and dissolution of T1 phase in sample at 380 °C/0.001 s−1; (d) the BF TEM image of representative precipitates in compressed sample at 320 °C/0.001 s−1.
Figure 10. (a) Schematic precipitate microstructure of the compressed samples; (b) strong interaction between dislocation and T1 phase in sample at 320 °C/0.001 s−1; (c) fracture and dissolution of T1 phase in sample at 380 °C/0.001 s−1; (d) the BF TEM image of representative precipitates in compressed sample at 320 °C/0.001 s−1.
Metals 12 01787 g010
Figure 11. EBSD IPF images of compressed samples: (a) 320 °C/1 s−1; (b) 320 °C/0.001 s−1; (c) 440 °C/1 s−1; (d) 440 °C/0.001 s−1.
Figure 11. EBSD IPF images of compressed samples: (a) 320 °C/1 s−1; (b) 320 °C/0.001 s−1; (c) 440 °C/1 s−1; (d) 440 °C/0.001 s−1.
Metals 12 01787 g011
Figure 12. Misorientation angle distribution histograms of the samples: (a) 320 °C/1 s−1; (b) 320 °C/0.001 s−1; (c) 440 °C/1 s−1; (d) 440 °C/0.001 s−1.
Figure 12. Misorientation angle distribution histograms of the samples: (a) 320 °C/1 s−1; (b) 320 °C/0.001 s−1; (c) 440 °C/1 s−1; (d) 440 °C/0.001 s−1.
Metals 12 01787 g012
Table 1. The modified Johnson–Cook constitutive model parameters.
Table 1. The modified Johnson–Cook constitutive model parameters.
ModelA1A2A3A4A5A6A7A8A9A10
Polynomial of second order74.1956−30.202317.79640.156564−0.012550.0021−0.000158905---
Polynomial of third order73.8173−23.176−5.0285519.07500.156445−0.012550.0021−0.000158905--
Polynomial of fourth order72.239517.4164−235.842466.424−276.5190.156543−0.012550.0021−0.000158905-
Polynomial of fifth order69.6888103.5759−993.8623003.928−3835.081764.7050.1563425−0.012550.0021−0.000158846
Power-exponential function122.498−62.66140.091060.156746−0.012550.0021−0.000158905---
Table 2. The values of R and AARE of constitutive models.
Table 2. The values of R and AARE of constitutive models.
ModelsPolynomial of 2nd OrderPolynomial of 3rd OrderPolynomial of 4th OrderPolynomial of 5th OrderPower-Exponential Function
AARE (%)10.67229.69648.78128.723810.6481
R0.96390.98420.98430.98470.9845
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, D.; Zhang, J.; Liu, T.; Li, W.; Wan, Z.; Han, T.; Che, C.; Cheng, L. A Modified Johnson–Cook Model and Microstructure Evolution of As-Extruded AA 2055 Alloy during Isothermal Compression. Metals 2022, 12, 1787. https://doi.org/10.3390/met12111787

AMA Style

Jiang D, Zhang J, Liu T, Li W, Wan Z, Han T, Che C, Cheng L. A Modified Johnson–Cook Model and Microstructure Evolution of As-Extruded AA 2055 Alloy during Isothermal Compression. Metals. 2022; 12(11):1787. https://doi.org/10.3390/met12111787

Chicago/Turabian Style

Jiang, Dongmei, Jian Zhang, Tiejun Liu, Wei Li, Zhong Wan, Tingzhuang Han, Chaojie Che, and Liren Cheng. 2022. "A Modified Johnson–Cook Model and Microstructure Evolution of As-Extruded AA 2055 Alloy during Isothermal Compression" Metals 12, no. 11: 1787. https://doi.org/10.3390/met12111787

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