Next Article in Journal
Development and Assessment of a Water-Based Drilling Fluid Tackifier with Salt and High-Temperature Resistance
Next Article in Special Issue
Study of the Hibridation of Ablation Casting and Laser Wire Metal Deposition for Aluminum Alloy 5356
Previous Article in Journal
Use of Hybrid Flame Retardants in Chemically Foamed rPET Blends
Previous Article in Special Issue
Evaluation of Structural Transition Joints Cu-Al-AlMg3 Used in Galvanizer Hangers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Gradient Cooling Behavior on the Microstructure and Mechanical Properties of Al-2at.% Nd Alloy in a Vacuum Environment

1
Key Lab of Electromagnetic Processing of Materials, Ministry of Education, Northeastern University, Shenyang 110819, China
2
Advanced Manufacturing Technology and Engineering Research Centers, Northeastern University, Shenyang 110819, China
3
School of Metallurgy, Northeastern University, Shenyang 110819, China
*
Author to whom correspondence should be addressed.
Crystals 2025, 15(1), 81; https://doi.org/10.3390/cryst15010081
Submission received: 26 December 2024 / Revised: 8 January 2025 / Accepted: 10 January 2025 / Published: 15 January 2025
(This article belongs to the Special Issue Development of Light Alloys and Their Applications)

Abstract

:
Al-2at.% Nd alloy with a gradient cooling rate was prepared using a wedge-shaped mold in a vacuum environment. The relationship between gradient cooling behavior and the microstructure and properties of the Al-2at.% Nd alloy was investigated. The stability of the Al11Nd3 phase and the mechanical properties were confirmed through first-principles calculations. The results indicated that as the cooling rate decreased, the transformation of grain morphology in Al-2at.% Nd occurred as follows: a mixture of columnar grains and equiaxed grains→equiaxed grains. The grain size of the alloy increased. Discontinuous skeletal eutectic phases, α-Al and Al11Nd3, formed within the alloy, resulting in a reduction in the number of phase boundaries and grain boundaries. The hardness of the alloy decreased by 25.53%, and this pattern of change closely aligned with the calculated results.

1. Introduction

Aluminum alloys have garnered significant attention from researchers due to their low density, commendable corrosion resistance, and exceptional formability. These materials are extensively utilized in various sectors, including aerospace, rail transportation, automotive, maritime applications, and national defense technology [1,2,3]. However, with the rapid advancement of technology, the performance requirements for aluminum alloys are becoming increasingly stringent. Consequently, certain aluminum alloys can no longer meet specific application areas’ demands. Therefore, there is a pressing need to develop new aluminum alloys that exhibit superior properties, such as high strength, enhanced corrosion resistance, and elevated temperature resistance, to satisfy the evolving requirements of both production and everyday life [4,5,6,7].
In recent years, the Al-Nd binary alloy system has gained significant attention among researchers as a target material in sputtering coating processes. However, the production and preparation of Al-Nd target materials present considerable challenges. To ensure the quality and performance of the target material, it is essential to use high-purity raw materials and employ precise processing techniques. Alloying represents a critical method for the development of new high-performance aluminum alloys. Research indicates that incorporating rare earth elements into aluminum and aluminum alloys can significantly modify the alloy composition and substantially enhance the microstructure of these material [8,9,10,11,12]. However, there is currently limited literature on Al-Nd target materials. One study indicates that the Al-2at.% Nd alloy is predominantly utilized in advanced thin-film transistor liquid crystal displays (TFT-LCDs) due to its exceptional resistance to hillock and whisker formation, alongside its having a low resistivity (<5 μΩcm) [13]. Therefore, conducting in-depth research on the preparation processes and performance characteristics of Al-Nd target materials is highly significant for advancing the development and progress of related industries. The challenge of homogenizing the microstructure of Al-Nd target materials lies in the formation and distribution of Al-Nd intermetallic compounds. In Al-Nd target materials, Nd primarily exists in the form of the eutectic compound Al11Nd3 [14]. Unlike the research on other Al-Nd intermetallic compounds, little is currently known about the Al11Nd3 phase. Xiao et al. [15] investigated the crystallization process of rapidly solidified Al-Nd-Ni alloys, which include Al11Nd3, Al3Ni, and several unidentified phases. Liu et al. [16] examined the stability and mechanical properties of Al-RE intermetallic compounds (Al2RE, Al3RE, AlRE, Al11RE3, AlRE2, and AlRE3), demonstrating that these intermetallic compounds can act as strengthening phases, enhancing the properties of Al alloys. Therefore, it is essential to investigate the thermodynamic stability and mechanical properties of the Al11Nd3 phase to design and produce high-quality Al-Nd target materials and to advance the development of large-sized liquid crystal display panels.
First-principles calculations, commonly referred to as ab initio calculations, are methodologies grounded in the principles of quantum mechanics that address the Schrödinger equation utilizing only fundamental physical constants, thereby eschewing any reliance on empirical parameters [17,18,19]. Researchers have undertaken comprehensive investigations into the phase structures of binary aluminum alloy systems through these first-principles calculations. By evaluating physical quantities such as electronic structure, energy, and mechanical properties of materials, they can predict the macroscopic properties of these materials. In recent years, significant advancements in material design, synthesis, and simulation calculations have emerged from integrating first-principles calculations based on density functional theory (DFT) with molecular dynamics, establishing it as a crucial foundation and primary methodology in computational materials science. Michael C. Gao [20] conducted a systematic study on the formation and stability of Al-RE intermetallic compounds. Similarly, Michal Jahnátek [21] systematically investigated the plastic deformation behavior of Al3(Sc, Ti, and V) intermetallic compounds in the L12 and D022 structures, providing a comprehensive analysis of various elastic constants.
Adding the element Nd can further improve the thermal stability and enhance the mechanical strength of aluminum alloys. Furthermore, the Al-2at.% Nd alloy exhibits favorable casting properties and is easily processed into various shapes and sizes of components. Therefore, due to considerations of processing efficiency and economic viability, we have chosen to study the Al-2at.% Nd alloy. However, there is still limited research on Al-Nd target materials and the Al11Nd3 phase. The stability, morphological characteristics, and inherent mechanical properties of Al11Nd3 in Al-Nd target alloys are not yet fully understood, and this will be essential in order to comprehend the material’s properties. Therefore, to address this knowledge gap, this study employs wedge-shaped copper molds to cast an Al-2at.% Nd alloy. It examines the influence of cooling rate on the grain morphology, dendrite arm spacing, and eutectic phase morphology of the Al-2at.% Nd alloy. Furthermore, first-principles calculations are utilized to evaluate the alloy’s phase stability and elastic properties, thereby providing a theoretical foundation for the fabrication of Al-Nd alloy targets.

2. Experimental Methods

2.1. Sample Preparation

Samples were prepared utilizing a commercially available 4N-grade bulk Al-2at.% Nd alloy. Initially, a high-purity graphite crucible was positioned within the coil of a vacuum induction melting furnace, and the interstices were filled with magnesia sand. Subsequently, the magnesia sand was combined with water glass to achieve a paste-like consistency, which was then applied around the crucible and allowed to dry. After thoroughly cleaning the furnace chamber and crucible, we placed the Al-Nd alloy raw material into the crucible. The furnace was subsequently evacuated to establish a vacuum conducive to melting. Upon reaching a temperature of 700 °C, the melts were poured into a wedge-shaped copper mold with an inside diameter of 65 × 50 mm and a height of 120 mm. After cooling for 30 min, we extracted the ingot.
Figure 1a illustrates the internal structure of the wedge-shaped copper mold. The cast ingot is segmented into four distinct regions, designated as A, B, C, and D, arranged from the bottom to the top, as depicted in the diagram.
In order to observe the grain structure of the alloy, anodic oxidation was conducted for approximately 45 s utilizing a 10% HBF4 solution at a voltage of 20 V and a current range of 0.3–0.5 mA. The microstructure of the anodic coating on the cast sample was examined using polarized light microscopy with a Leica DMI5000M optical microscope (Wetzlar, Germany). For deep etching, the 10% NaOH solution was employed at room temperature for a duration of 50 s. In order to study the phase composition of the alloy, rectangular samples measuring 15 mm × 15 mm × 10 mm were cut, polished, and analyzed using X-ray diffraction (XRD, Rigaku, DMAX-2500, Cu-Kα, 18 kW). The morphology and elemental composition of the phases within the cast alloy were characterized using a field emission scanning electron microscope (Tescan, Brno, Czechia, MIRA3) equipped with an X-ray spectrometer. The alloy’s grain size and secondary dendrite arm spacing were statistically analyzed using Image-Pro Plus software (6.0, Media Cybernetics, Rockville, MD, USA). Using wire cutting technology, rectangular samples measuring 20 mm × 15 mm × 10 mm were cut out of areas of ingots A, B, C, and D. The hardness of the alloy at various positions was measured using a microhardness tester (FUTURE TECH, Kawasaki, Japan).

2.2. First-Principles Calculations

All first-principles calculations conducted in this study, based on Density Functional Theory (DFT), were executed using Vienna ab initio Simulation Package software (VASP 6.3). The Projector Augmented Wave (PAW) [22] method was utilized to represent the atomic pseudopotentials, while the Perdew–Burke–Ernzerhof (PBE) [23] functional was applied to characterize the exchange-correlation energy. Each unit cell was fully optimized until the total energy and force reached the convergence standards of 1.0 × 10−5 eV and 0.01 eV/Å, respectively [24]. Furthermore, all calculations presented in this paper underwent rigorous convergence testing.
The formation enthalpy of a compound serves as a valuable metric for evaluating the stability of alloy compounds. It is typically negative and represents the energy absorbed or released during the formation of a compound from its stable elemental constituents. A smaller formation enthalpy indicates greater stability of the compound. The formula for calculating the formation enthalpy of an alloy is as follows [25]:
Δ H = 1 / ( x + y ) ( E t o t x E b u l k A y E b u l k B )
The AlxNdy alloy phase, E tot , represents the total energy of the system. E b u l k A and E bulk B are the energies of Al atoms and Cu atoms in the crystal structure, respectively. The variables x and y correspond to the quantities of Al and Nd atoms present in the crystal structure of the intermetallic compound, respectively.
Elastic constants are fundamental physical quantities that provide insight into the mechanical stability of materials. These constants characterize various mechanical properties, including strength, toughness, hardness, and brittleness. Within the linear deformation regime of materials, characterized by small strains, the relationship between stress and strain is linear, thereby conforming to Hooke’s Law [26]:
σ i = j = 1 6 C i j ε i j
In this context, σ and ε denote stress and strain, respectively, while Cij represents the elastic stiffness constant. It is important to note that the indices 1 ≤ i ≤ 6 indicate that both strain and stress possess six independent components.

3. Results and Discussion

3.1. The Influence of Gradient Cooling Behavior on the Microstructure of Al-2at.% Nd Alloy

Figure 2a–d illustrate the microstructural grain morphology of samples A, B, C, and D. It is evident that samples A, B, and C comprise a combination of columnar grains and equiaxed grains. In contrast, sample D is exclusively composed of equiaxed grains. As indicated by the arrows in direction 2 of Figure 2a–d, samples A, B, and C display columnar grains adjacent to the mold wall, with the growth direction of these columnar grains oriented perpendicularly to the mold wall. This is because the heat dissipation direction near the mold wall is perpendicular to the mold wall, and the grain growth rate in this direction is higher than that in other directions, resulting in preferential growth. The latent heat released by the preferentially growing dendrites inhibits the growth of dendrites in alternative directions, leading to the solidification of columnar grains in the preferred direction [27]. In contrast, sample D solidifies at the upper section of the mold, where heat dissipation is slower and lacks a specific direction for rapid heat dissipation. Consequently, the grains do not exhibit a preferred growth direction, resulting in the absence of columnar crystals along the mold wall. When the undercooling of the components at the tips of the columnar grain exceeds the critical undercooling necessary for nucleation, equiaxed crystal nuclei will be formed in the melt (as depicted in direction 1 of Figure 2a–d). By studying the grain size of the Al-2at.% Nd alloy under various cooling rates, we can gain insights into the mechanisms of grain growth during solidification. Additionally, by controlling the cooling rate, the grain size of the Al alloy can be adjusted, allowing for the regulation of the alloy’s properties. The grain sizes of samples A, B, C, and D were statistically analyzed utilizing Image-Pro Plus software (6.0, Media Cybernetics, Rockville, MD, USA). For each sample, the grain sizes of 15 different regions were measured, and the average value was calculated. The results presented in Figure 2(a1–d1). Figure 2(a1–d1) illustrate the frequency distribution of grain sizes across the four sample regions: D, C, B, and A, respectively. As the cooling rate diminishes, the grain size progressively increases from 523.28 μm to 743.57 μm.
Figure 3 shows the microstructure of the Al-2at.% Nd alloy at different cooling rates. It is evident that a gradual decrease in the cooling rate results in a notable coarsening of the alloy’s structure. Furthermore, the secondary dendrite arm spacing of the Al-2at.% Nd alloy at different cooling rates was quantified, as depicted in Figure 4. The data indicate that, from region A to region D, the secondary dendrite arm spacing increases from 14.21 μm to 42.72 μm as the cooling rate decreases. This phenomenon occurs because when the alloy cools at a slower rate, the heat loss at the solidification front is slower, leading to a smaller degree of undercooling and a correspondingly reduced nucleation rate. As a result, fewer crystal nuclei are formed during the solidification process, and these nuclei have more time to grow, resulting in larger grains and increased secondary dendrite arm spacing. Conversely, when the cooling rate is rapid, the crystal nuclei do not have sufficient time to grow during solidification, leading to smaller grain sizes and decreased secondary dendrite arm spacing. The reduced secondary dendrite arm spacing results in a greater number of grain boundaries, which can more effectively impede the movement of dislocations and enhance the hardness of the alloy.
To thoroughly investigate the influence of cooling rate on the morphology of the eutectic phase in the Al-2at.% Nd alloy, scanning electron microscopy (SEM) was employed for observation. Based on the Al-Nd binary phase diagram (Figure 5), it is established that the predominant phase in the Al-2at.% Nd alloy is the eutectic Al11Nd3 phase. Using wire cutting technology, a square sample of 15 mm × 15 mm × 10 mm is extracted from the core of the solidified ingot. Point scanning and XRD detection were performed on various morphologies of the second phase to analyze its composition further. By integrating the results from point scanning and face scanning (Figure 6) with the XRD spectrum analysis (Figure 7), it is discernible that the black matrix (point B) corresponds to the α-Al phase. In contrast, the white eutectic phase (points A and C) is attributed to the Al11Nd3 phase.
Figure 8a–d illustrate the microstructure of the Al-2at.% Nd alloy following solidification at various cooling rates. Figure 8e–h correspond to the boxed portions in Figure 8a–d, respectively, illustrating the morphology of eutectic Al11Nd3 after deep etching with NaOH solution. The observations indicate that at higher cooling rates, the morphology of the eutectic Al11Nd3 exhibits a skeletal structure, which is embedded within the aluminum matrix. This phenomenon can be attributed to the high cooling rate and substantial undercooling, which result in an unstable growth interface of the lamellar eutectic, thereby producing a complex skeletal structure of Al11Nd3.

3.2. The Effect of Gradient Cooling Behavior on the Properties of Al-2at.% Nd Alloy

Figure 9 illustrates the evolution of hardness in Al-2at.% Nd at various cooling rates. Generally, the cooling rate has a positive influence on hardness [29]. The hardness of the Al-2at.% Nd alloy decreases with a decreasing cooling rate and increases by 25.53% with an increasing cooling rate. This is because, as the cooling rate increases, the number of alloy phases and grains also increases, leading to a higher number of phase boundaries and grain boundaries. The greater the resistance encountered during the dislocation process, the higher the hardness of the alloy. Additionally, the sub-eutectic grain refinement of Al-2at.% Nd, the finer eutectic Al11Nd3, and the reduced secondary dendrite arm spacing positively influence hardness.

3.3. Crystal Structure and Phase Stability

Figure 10a shows the crystal cell structure model of the Al matrix, which belongs to the Fm-3m space group and has a face-centered cubic structure, with a = b = c = 4.03893 Å and α = β = γ = 90°. Figure 10b presents the unit cell structure model of the Al11Nd3 phase, which has an orthorhombic structure. The space group of Al11Nd3 is Immm, with lattice constants a = 4.3565 Å, b = 10.0679 Å, c = 12.9989 Å, and α = β = γ = 90°. The lattice constants and formation enthalpy of Al and Al11Nd3 are listed in Table 1. The setting of the lattice constant is consistent with the results obtained from experimental testing. The Monkhorst–Pack method is used to divide the simple Brillouin zone K-point grid of the unit cell model, with the K-point for the Al matrix set to 5 × 5 × 5 and the K-point for the Al11Nd3 phase set to 14 × 6 × 5. The plane wave cutoff energy (Ecut) is set to 400 eV for the structural optimization of the Al matrix and Al11Nd3. From Table 1, it is evident that the enthalpy of formation of Al11Nd3 is negative (−0.42 eV/atom), indicating that the Al11Nd3 phase can exist stably. This also confirms that the addition of the Nd element in the previous experiment can lead to the formation of a stable discontinuous eutectic phase at the grain boundary.

3.4. Mechanical Properties

The elastic constant is a significant physical parameter for investigating the mechanical properties of compounds and comprehending the microscopic mechanical properties of crystal structures. Additionally, to ensure the accuracy of the calculations, we set the ion movement step size to POTIM = 0.015. The face-centered cubic structure of the Al matrix is characterized by three independent elastic constants (C11, C22, and C44). In contrast, the body-centered orthorhombic structure of the Al11Nd3 compound is defined by nine independent elastic constants (C11, C22, C33, C44, C55, C66, C12, C13, and C23). The conditions for mechanical stability are calculated by the following expressions [30]:
Cubic crystal system:
C 11 > 0 C 44 > 0 C 11 > | C 12 | C 11 + 2 C 12 > 0
Orthorhombic crystal system:
C i i > 0 i = 1 , 2 , 3 , 4 , 5 , 6 C 11 + C 22 2 C 12 > 0 C 11 + C 33 2 C 13 > 0 C 22 + C 33 2 C 23 > 0 C 11 + C 22 + C 33 + 2 C 12 + C 13 + C 23 > 0
This study calculates the elastic constants of the Al matrix and the Al11Nd3 phase to investigate the mechanical properties of the Al-Nd binary alloy. According to Table 2 and Formula (4), it can be established that Al11Nd3 satisfies the stability criterion for orthorhombic crystal systems, thereby confirming its stability. The compressibility of the crystal along the x, y, and z axes is directly indicated by the values of the elastic constants C11, C22, and C33; a higher value signifies greater resistance to compression in that particular direction. The crystal’s capacity to withstand pure shear deformation on the (100), (010), and (001) crystal planes is represented by the elastic constants C44, C55, and C66. Table 2 presents the theoretically calculated elastic constants, revealing that C11 > C22 > C33 for Al11Nd3 suggests that the material is more resistant to compression along the x-axis. Additionally, the relationship C44 > C55 > C66 for Al11Nd3 indicates that it exhibits the highest resistance to shear deformation on the (100) crystal plane.
This article also determines the bulk modulus (B), shear modulus (G), Young’s modulus (E), and Poisson’s ratio (σ) of the crystalline compound utilizing the Voigt–Reuss–Hill (VRH) approximation method. The values can be computed using the following formulas [31]:
B H = 1 2 B V + B R
G H = 1 2 G V + G R
E = 9 B H G H 3 B H + G H
σ = 3 B H 2 G H 2 3 B H + G H
In the formula, BH, BV, BR, GH, GV, and GR represent the bulk and shear modulus calculated using the Voigt–Reuss–Hill, Voigt, and Reuss approximation methods [32].
The bulk modulus is a physical property that quantifies the compressive strength of an alloy when subjected to external forces, and it is intrinsically linked to the chemical bonds within the material. The shear modulus represents the capacity of a material to resist shear strain; a lower shear modulus indicates a diminished ability of the material to withstand shear strain. Young’s modulus characterizes the stiffness or resistance to deformation of solid materials. Upon analyzing the data presented in Figure 11 and Table 3, it is observed that the B value of Al11Nd3 is marginally lower than that of the Al matrix, suggesting that Al11Nd3 possesses inferior compressive strength relative to the Al matrix. Conversely, the G and E values of Al11Nd3 exceed those of the Al matrix, indicating that Al11Nd3 can resist deformation and shear strain. Generally, the plasticity and brittleness of materials within a crystalline structure can be inferred from the ratio of the bulk modulus to the shear modulus and Poisson’s ratio [33]. According to the Pugh criterion, the material is considered brittle if the Poisson’s ratio is less than 0.26 or the B/G ratio is less than 1.75. As illustrated in Figure 11b, the B/G ratio of the Al matrix exceeds 1.75, indicating that the Al matrix exhibits favorable ductility. In contrast, the B/G ratio of Al11Nd3 is below 1.75, signifying that the presence of Al11Nd3 contributes to the material’s brittleness. This also demonstrates the formation of Al11Nd3, which makes it difficult for Al-Nd alloys to undergo severe plastic deformation. The hardness changes observed in the previous experiments have been verified.

4. Conclusions

This article employs a wedge-shaped copper mold to fabricate Al-2at.% Nd alloy ingots and analyses the alloy’s microstructure and phase structure in conjunction with first-principles calculations. The findings lead to the following conclusions:
  • Under varying cooling rates, the microstructure of the Al-2at.% Nd alloy displays distinct morphologies. The large grain morphology near the mold wall shows columnar crystals. Conversely, at the center of the ingot, a reduction in the cooling rate leads to a transition in grain morphology from a mixed structure of columnar and equiaxed grains to a predominantly equiaxed grain structure.
  • As the cooling rate diminishes, there is a corresponding increase in grain size, which ranges from 523.28 μm to 743.57 μm. Additionally, the spacing between secondary dendrite arms expands from 14.21 μm to 42.72 μm.
  • At higher cooling rates, the morphology of the Al11Nd3 phase is characterized as skeletal and embedded within the aluminum matrix. The eutectic phase of the alloy comprises α-Al and Al11Nd3. The hardness of the Al-2at.% Nd alloy decreases with a decreasing cooling rate and increases by 25.53% with an increasing cooling rate.
  • According to first-principles calculations, the Al-Nd binary intermetallic compound’s formation enthalpy is negative, indicating that Al11Nd3 is thermodynamically stable. The Poisson’s ratio of Al11Nd3 is measured at 0.23, and the B/G value is calculated to be 1.53, suggesting that the compound exhibits intrinsic brittleness in its ground state.

Author Contributions

Conceptualization, X.W. and S.S.; methodology, X.W.; validation, X.W.; formal analysis, X.W.; investigation, X.Z.; resources, X.W. and S.S.; data curation, X.Z.; writing—original draft preparation, X.Z.; writing—review and editing, W.W. and X.W.; visualization, X.W.; supervision, X.W.; project administration, X.W. and S.S.; and funding acquisition, X.W. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program (2024YFB3714001, 2023YFB3710401, 2022YFB3504401), the National Natural Science Foundation (52271094), the Qingyuan City Science and Technology Plan Project (2023YFJH003), Central Basic Research Funds (N2309003), the Bingtuan Science and Technology Program (2023AA003), and the Major Science and Technology Projects in Xinjiang Uygur Autonomous Region (2023A01001).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Some or all data that support the findings are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Chaudhary, B.; Jain, N.K.; Murugesan, J. Development of friction stir powder deposition process for repairing of aerospace-grade aluminum alloys. CIRP J. Manuf. Sci. Technol. 2022, 38, 252–267. [Google Scholar] [CrossRef]
  2. Xu, Z.; Liu, C.; Zhang, B.; Huang, H.; Cheng, W. Effects of base metal state on the microstructure and mechanical properties of Al–Mg–Si alloy friction stir-welded joints. J. Manuf. Process. 2020, 56, 248–257. [Google Scholar] [CrossRef]
  3. Mo, Y.F.; Liu, C.Y.; Teng, G.B.; Jiang, H.J.; Chen, Y.; Yang, Z.X.; Han, S.C. Fabrication of 7075-0.25Sc-0.15Zr Alloy with Excellent Damping and Mechanical Properties by FSP and T6 Treatment. J. Mater. Eng. Perform. 2018, 27, 4162–4167. [Google Scholar] [CrossRef]
  4. Berry, L.; Wheatley, G.; Ma, W.; Nejad, R.M.; Berto, F. The influence of milling induced residual stress on fatigue life of aluminum alloys. Forces Mech. 2022, 7, 100096. [Google Scholar] [CrossRef]
  5. Chen, B.; Zhang, C.; Jin, Y. First-principles calculation of interface binding strength and fracture performance of β’/Al interface in Al–Mg–Si–Cu alloy. J. Alloys Compd. 2020, 830, 154515. [Google Scholar] [CrossRef]
  6. Yin, Q.; Chen, G.; Shu, X.; Zhang, B.; Li, C.; Dong, Z.; Cao, J.; An, R.; Huang, Y. Analysis of interaction between dislocation and interface of aluminum matrix/second phase from electronic behavior. J. Mater. Sci. Technol. 2023, 136, 78–90. [Google Scholar] [CrossRef]
  7. Huang, J.; Li, M.; Liu, Y.; Chen, J.; Lai, Z.; Hu, J.; Zhou, F.; Zhu, J. A first-principles study on the doping stability and micromechanical properties of alloying atoms in aluminum matrix. Vacuum 2023, 207, 111596. [Google Scholar] [CrossRef]
  8. Lu, Z.; Li, X.; Zhang, L. Thermodynamic Description of Al-Si-Mg-Ce Quaternary System in Al-Rich Corner and Its Experimental Validation. J. Phase Equilibria Diffus. 2018, 39, 57–67. [Google Scholar] [CrossRef]
  9. He, Y.; Liu, J.; Qiu, S.; Deng, Z.; Zhang, J.; Shen, Y. Microstructure evolution and mechanical properties of Al-La alloys with varying La contents. Mater. Sci. Eng. A 2017, 701, 134–142. [Google Scholar] [CrossRef]
  10. Huang, X.; Yan, H. Effect of trace La addition on the microstructure and mechanical property of as-cast ADC12 Al-Alloy. J. Wuhan Univ. Technol. Mater. 2013, 28, 202–205. [Google Scholar] [CrossRef]
  11. Pan, F.; Yang, M.; Chen, X. A Review on Casting Magnesium Alloys: Modification of Commercial Alloys and Development of New Alloys. J. Mater. Sci. Technol. 2016, 32, 1211–1221. [Google Scholar] [CrossRef]
  12. Ding, W.; Zhao, X.; Chen, T.; Zhang, H.; Liu, X.; Cheng, Y.; Lei, D. Effect of rare earth Y and Al–Ti–B master alloy on the microstructure and mechanical properties of 6063 aluminum alloy. J. Alloys Compd. 2020, 830, 154685. [Google Scholar] [CrossRef]
  13. Sarkar, J.; Saimoto, S.; Mathew, B.; Gilman, P. Microstructure, texture and tensile properties of aluminum2at.% neodymium alloy as used in flat panel displays. J. Alloys Compd. 2009, 479, 719–725. [Google Scholar] [CrossRef]
  14. Zhu, S.; Gibson, M.; Nie, J.; Easton, M.; Abbott, T. Microstructural analysis of the creep resistance of die-cast Mg-4Al-2RE alloy. Scr. Mater. 2008, 58, 477–480. [Google Scholar] [CrossRef]
  15. Xiao, Y.D.; Li, W.X.; Ma, Z.Q. Crystallization process in rapidly solidified Al-Nd-Ni amorphous alloy prepared by melt spinning. Trans. Nonferrous Met. Soc. China 2004, 14, 665–669. [Google Scholar]
  16. Liu, T.; Ma, T.; Li, Y.; Ren, Y.; Liu, W. Stabilities, mechanical and thermodynamic properties of Al–RE intermetallics: A first-principles study. J. Rare Earths 2022, 40, 345–352. [Google Scholar] [CrossRef]
  17. Liu, S.; Esteban-Manzanares, G.; LLorca, J. First Principles Prediction of the Al-Li Phase Diagram. Metall. Mater. Trans. A 2021, 52, 4675–4690. [Google Scholar] [CrossRef]
  18. Guo, Y.; Wang, W.; Huang, H.; Zhao, H.; Jing, Y.; Yi, G.; Luo, L.; Liu, Y. Effect of doping Zn atom on the structural stability, mechanical and thermodynamic properties of AlLi phase in Mg–Li alloys from first-principles calculations. Philos. Mag. 2020, 100, 1849–1867. [Google Scholar] [CrossRef]
  19. Gao, M.; Deng, Y.; Wen, D.; Zhao, H. A first-principles study on the site occupancy behavior of transition metals in L12-Al3Li phase. Mater. Res. Express 2020, 7, 046519. [Google Scholar] [CrossRef]
  20. Gao, M.C.; Rollett, A.D.; Widom, M. Lattice stability of aluminum-rare earth binary systems: A first-principles approach. Phys. Rev. B 2007, 75, 174120. [Google Scholar] [CrossRef]
  21. Jahnátek, M.; Krajcí, M.; Hafner, J. Response of trialuminides to [110] uniaxial loading: An ab initio study for Al3(Sc, Ti, V). Phys. Rev. B 2007, 76, 014110. [Google Scholar] [CrossRef]
  22. Joubert, D.; Kresse, G. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar]
  23. Burke, K.; Ernzerhof, M.; Perdew, J.P. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar]
  24. Xue, B.; Xiao, W.; Li, X.; Gao, G.; Li, X.; Zhang, Y.; Wang, L.; Xiong, B. Comprehensive investigation on the structural, electronic and mechanical properties of T-Mg32(Al, Zn)49 phases in Al-Mg-Zn alloys. J. Mater. Sci. Technol. 2024, 173, 237–246. [Google Scholar] [CrossRef]
  25. Lin, L.; Wang, F.; Yang, L.; Chen, L.; Liu, Z.; Wang, Y. Microstructure investigation and first-principle analysis of die-cast AZ91 alloy with calcium addition. Mater. Sci. Eng. A 2011, 528, 5283–5288. [Google Scholar] [CrossRef]
  26. Fung, Y.C.; Drucker, D.C. Foundation of Solid Mechanics. J. Appl. Mech. 1966, 33, 238. [Google Scholar] [CrossRef]
  27. Yao, X.; Wang, H.; He, B.; Zhou, X. Modeling of Columnar-to-Equiaxed Transition in Solidified Al-Si Alloys. Mater. Sci. Forum 2005, 475–479, 3141–3144. [Google Scholar] [CrossRef]
  28. Okamoto, H. Al-Nd (Aluminum-Neodymium). J. Phase Equilib. 2000, 21, 206. [Google Scholar] [CrossRef]
  29. Yang, Q.; Shi, W.; Wang, M.; Jia, L.; Wang, W.; Zhang, H. Influence of cooling rate on the microstructure and mechanical properties of Al–Cu–Li–Mg–Zn alloy. J. Mater. Res. Technol. 2023, 25, 3151–3166. [Google Scholar] [CrossRef]
  30. Ahmad, S.; Ahmad, R.; Jalali-Asadabadi, S.; Ali, Z.; Ahmad, I. First principle studies of structural, magnetic and elastic properties of orthorhombic rare-earth diaurides intermetallics RAu2 (R=La, Ce, Pr and Eu). Mater. Chem. Phys. 2018, 212, 44–50. [Google Scholar] [CrossRef]
  31. Boucetta, S. Structural and elastic properties of Mg3CuH0.6 ternary hydride by ab initio study. J. Magnes. Alloys 2018, 6, 90–94. [Google Scholar] [CrossRef]
  32. Liu, Y.; Hu, W.-C.; Li, D.-J.; Li, K.; Jin, H.-L.; Xu, Y.-X.; Xu, C.-S.; Zeng, X.-Q. Mechanical, electronic and thermodynamic properties of C14-type AMg2 (A=Ca, Sr and Ba) compounds from first principles calculations. Comput. Mater. Sci. 2015, 97, 75–85. [Google Scholar] [CrossRef]
  33. Abraham, J.A.; Pagare, G.; Chouhan, S.S.; Sanyal, S.P. Theoretical calculations of structural, electronic, optical, elastic, and thermal properties of YX3 (X = In, Sn, Tl, and Pb)compounds based on density functional theory. J. Mater. Sci. 2015, 50, 542–554. [Google Scholar] [CrossRef]
Figure 1. (a) Wedge-shaped copper mold and (b) schematic diagram of the sample cutting.
Figure 1. (a) Wedge-shaped copper mold and (b) schematic diagram of the sample cutting.
Crystals 15 00081 g001
Figure 2. Grain morphology of Al-2at.% Nd at different cooling rates: (a) sample A; (b) sample B; (c) sample C; and (d) sample D. Relative frequency of grain size in Al-2at.% Nd alloy: (a1) sample D; (b1) sample C; (c1) sample B; and (d1) sample A.
Figure 2. Grain morphology of Al-2at.% Nd at different cooling rates: (a) sample A; (b) sample B; (c) sample C; and (d) sample D. Relative frequency of grain size in Al-2at.% Nd alloy: (a1) sample D; (b1) sample C; (c1) sample B; and (d1) sample A.
Crystals 15 00081 g002
Figure 3. Microstructure of Al-2at.% Nd alloy at different cooling rates: (a) sample A; (b) sample B; (c) sample C; and (d) sample D.
Figure 3. Microstructure of Al-2at.% Nd alloy at different cooling rates: (a) sample A; (b) sample B; (c) sample C; and (d) sample D.
Crystals 15 00081 g003
Figure 4. Variation in dendrite arm spacing in Al-2at.% Nd alloy at different cooling rates.
Figure 4. Variation in dendrite arm spacing in Al-2at.% Nd alloy at different cooling rates.
Crystals 15 00081 g004
Figure 5. Phase diagram of Al-Nd system [28].
Figure 5. Phase diagram of Al-Nd system [28].
Crystals 15 00081 g005
Figure 6. Microstructure of as-cast Al-2Nd alloy. (a) Backscattered electron (BSE) image. (b,c) Distributions of Al and Nd for the region indicated in the dashed box in a. EDS point scan results: (d) point A; (e) point B; and (f) point C.
Figure 6. Microstructure of as-cast Al-2Nd alloy. (a) Backscattered electron (BSE) image. (b,c) Distributions of Al and Nd for the region indicated in the dashed box in a. EDS point scan results: (d) point A; (e) point B; and (f) point C.
Crystals 15 00081 g006
Figure 7. XRD spectrum of Al-2at.% Nd alloy.
Figure 7. XRD spectrum of Al-2at.% Nd alloy.
Crystals 15 00081 g007
Figure 8. Morphology of eutectic Al11Nd3 at different cooling rates: (a) sample A; (b) sample B; (c) sample C; and (d) sample D. Morphology of eutectic Al11Nd3 after deep corrosion: (e) sample A; (f) sample B; (g) sample C; and (h) sample D. The red boxes in images (ad) correspond to (eh) respectively.
Figure 8. Morphology of eutectic Al11Nd3 at different cooling rates: (a) sample A; (b) sample B; (c) sample C; and (d) sample D. Morphology of eutectic Al11Nd3 after deep corrosion: (e) sample A; (f) sample B; (g) sample C; and (h) sample D. The red boxes in images (ad) correspond to (eh) respectively.
Crystals 15 00081 g008
Figure 9. Hardness of Al-2at.% Nd alloy at different cooling rates.
Figure 9. Hardness of Al-2at.% Nd alloy at different cooling rates.
Crystals 15 00081 g009
Figure 10. Schematic diagram of crystal structure: (a) Al and (b) Al11Nd3.
Figure 10. Schematic diagram of crystal structure: (a) Al and (b) Al11Nd3.
Crystals 15 00081 g010
Figure 11. Elastic moduli of Al and Al11Nd3 (a) B, G, and E and (b) B/G and σ.
Figure 11. Elastic moduli of Al and Al11Nd3 (a) B, G, and E and (b) B/G and σ.
Crystals 15 00081 g011
Table 1. Lattice constants (a, b, and c/Å), angles (α, β, and γ/°), and formation enthalpy ΔH (eV/atom) of Al and Al11Nd3.
Table 1. Lattice constants (a, b, and c/Å), angles (α, β, and γ/°), and formation enthalpy ΔH (eV/atom) of Al and Al11Nd3.
CompoundsabcαβγΔH
Al11Nd34.439.9912.94909090−0.42
Al4.044.044.04909090
Table 2. Elastic constants Cij (GPa) of Al and Al11Nd3 in the ground state.
Table 2. Elastic constants Cij (GPa) of Al and Al11Nd3 in the ground state.
CompoundsC11C12C13C22C23C33C44C55C66
Al11462 32
Al11Nd3130.843.243.3126.751.8125.955.450.347.9
Table 3. Volume modulus, shear modulus, Young’s modulus, Poisson’s ratio, and B/G of Al and Al11Nd3.
Table 3. Volume modulus, shear modulus, Young’s modulus, Poisson’s ratio, and B/G of Al and Al11Nd3.
CompoundsB (Gpa)G (Gpa)E (Gpa)σB/G
Al79.229.371.10.342.70
Al11Nd370.246.0113.20.231.53
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

Wang, X.; Zhang, X.; Wu, W.; Sun, S. The Effect of Gradient Cooling Behavior on the Microstructure and Mechanical Properties of Al-2at.% Nd Alloy in a Vacuum Environment. Crystals 2025, 15, 81. https://doi.org/10.3390/cryst15010081

AMA Style

Wang X, Zhang X, Wu W, Sun S. The Effect of Gradient Cooling Behavior on the Microstructure and Mechanical Properties of Al-2at.% Nd Alloy in a Vacuum Environment. Crystals. 2025; 15(1):81. https://doi.org/10.3390/cryst15010081

Chicago/Turabian Style

Wang, Xiangjie, Xinyu Zhang, Wenjie Wu, and Shuchen Sun. 2025. "The Effect of Gradient Cooling Behavior on the Microstructure and Mechanical Properties of Al-2at.% Nd Alloy in a Vacuum Environment" Crystals 15, no. 1: 81. https://doi.org/10.3390/cryst15010081

APA Style

Wang, X., Zhang, X., Wu, W., & Sun, S. (2025). The Effect of Gradient Cooling Behavior on the Microstructure and Mechanical Properties of Al-2at.% Nd Alloy in a Vacuum Environment. Crystals, 15(1), 81. https://doi.org/10.3390/cryst15010081

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