Next Article in Journal
Electric Field-Induced Settling and Flotation of Flocs in Mixed Aqueous Suspensions of Poly(methyl methacrylate) and Aluminosilicate Hollow Particles
Previous Article in Journal
Retention of Pediatric BioFlx Crowns Versus Stainless Steel Crowns Using Different Types of Luting Cements: An In Vitro Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Solution Treatment on Microstructure Evolution and the Mechanical Properties of GH4780 Superalloy

1
Institute for Advanced Studies in Precision Materials, Yantai University, Yantai 264005, China
2
Beijing GAONA Materials and Technology Co., Ltd., Beijing 100081, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(6), 1288; https://doi.org/10.3390/ma18061288
Submission received: 10 February 2025 / Revised: 2 March 2025 / Accepted: 11 March 2025 / Published: 14 March 2025

Abstract

:
This study systematically investigated the microstructural evolution and mechanical properties of GH4780 superalloy under various solution treatment conditions. Experimental results reveal a strong temperature dependence of grain growth kinetics, with the average grain diameter increasing from approximately 20 μm to 194 μm as the solution temperature rises from 1020 °C to 1110 °C. Mechanical testing demonstrates that grain coarsening reduces the yield strength by 19% at room temperature (from 920 MPa to 743 MPa) and by 9.5% at 760 °C (from 707 MPa to 640 MPa), primarily due to decreased grain boundary density and enhanced dislocation mobility. High-temperature deformation mechanisms were characterized, showing that the reduced grain boundary area facilitates dislocation motion while compromising strength. Furthermore, a grain growth kinetic model was developed, providing a quantitative prediction of microstructural evolution. These findings offer significant guidance for improving the high-temperature performance of GH4780 superalloy by optimizing heat treatment processes.

Graphical Abstract

1. Introduction

Ni-based superalloys are extensively utilized in critical aerospace components owing to their high strength and creep rupture lifetime as well as their remarkable resistance to corrosion, oxidation, and fatigue under high-temperature conditions [1,2,3,4]. The hot-section components of modern aero-engines, including turbine blades, nozzle guide vanes, turbine disks, and combustion chambers, are predominantly manufactured from nickel-based superalloys due to their unique combination of high-temperature performance and structural reliability [5,6]. With increasing engine thrust requirements, turbine inlet temperatures continue to rise, demanding corresponding improvements in mechanical properties of superalloys [7]. This technological imperative drives the development of advanced alloy chemistries and processing techniques to enhance high-temperature performance, which is crucial for the evolution of both aviation engines and industrial gas turbines [8]. Within this technological framework, GH4780 superalloy emerges as a prominent nickel-based cast and wrought (C&W) material, demonstrating exceptional mechanical properties and thermal stability at service temperatures of up to 760 °C. These characteristics make GH4780 particularly suitable for demanding high-temperature applications in advanced aerospace and power generation systems.
Heat treatment serves as a critical process for tailoring alloys’ mechanical properties through precise microstructural control, representing the final stage in optimizing material characteristics [9,10]. Proper selection of solution heat treatment temperature enables optimal grain size distribution and precipitation morphology, and therefore the mechanical properties of the alloy can be enhanced [11,12]. Given their fundamental importance, the effects of solution temperature on the microstructure and properties of superalloys have been extensively investigated, with numerous studies focusing on the underlying mechanisms and processing–property relationships. Zhang et al. [13]. investigated the effects of solution treatment temperatures on the microstructure and mechanical properties of Hastelloy X superalloy. The results revealed the temperature-dependent solubility behavior of key precipitated phases, including the γ′ phase and carbides. Optimal solution treatment temperature facilitates the formation of a homogeneous supersaturated solid solution, which is essential for subsequent age hardening. This uniform microstructure enables more consistent and dense precipitation of strengthening phases during aging treatment, ultimately enhancing the mechanical performance of Hastelloy X superalloy.
Li et al. [14] examined the influence of solution treatment on γ′ phase evolution and grain growth kinetics. Their results demonstrated the size-dependent dissolution behavior of γ′ precipitates, with larger particles showing significantly slower dissolution rates. Moreover, γ′ phase dissolution facilitates grain boundary migration, leading to accelerated grain growth with increasing temperature and prolonged holding time. Song et al. [15] studied grain growth behavior in nickel-based superalloys, revealing significant grain boundary-pinning effects from coarse γ′ phases and inert precipitates. Their findings demonstrated distinct grain growth mechanisms under different thermal conditions; at sub-solvus temperatures, grain size is primarily controlled by the size and volume fraction of γ′ precipitates, and at super-solvus temperatures, grain growth is restricted by the thermal stability of MC carbides and yttrium oxides. Complete grain growth observed at 1270 °C was attributed to the coarsening and dissolution of MC carbides and yttrium oxides, which subsequently lost their grain boundary-pinning capability. Zhou et al. [16] systematically investigated the effect of solution temperature on the microstructural evolution and creep behavior of Ni-Co-based superalloys for turbine disk applications. The results showed that microstructural variations following different solution treatments were primarily governed by the distribution of the γ′ phase and the morphological characteristics of grain boundary carbides. As the solution temperature increased, rapid dissolution of primary γ′ phase was observed, accompanied by increased volume fractions of secondary and tertiary γ′ precipitates. This microstructural transformation significantly enhanced the alloy’s mechanical properties, particularly its creep resistance.
As a newly developed Ni-based C&W superalloy, the microstructural evolution mechanisms and property optimization of GH4780 superalloy during heat treatment remain insufficiently understood. The grain size of the GH4780 superalloy can be precisely controlled through the synergistic optimization of temperature–time parameters during solution treatment. This control directly influences critical properties, including strength, plasticity, and creep resistance. This mechanism is fundamentally rooted in balancing the requirements of grain refinement strengthening and high-temperature stability, thereby providing a reliable material foundation for extreme environmental applications, such as aerospace engineering.
In this study, the newly developed GH4780 superalloy by Beijing GAONA Materials and Technology Co., Ltd. (Beijing, China), was thoroughly investigated, focusing on the influence of solution temperature (1020~1110 °C) on its microstructure and mechanical properties. Special emphasis was placed on achieving precise grain size control through the optimization of solution treatment parameters to enhance its mechanical performance. A temperature-dependent grain growth model incorporating MC carbide-pinning effects was uniquely established, representing a critical advancement due to the 1080 °C threshold for rapid coarsening observed in GH4780 superalloy. This behavior fundamentally differs from conventional Ni-Co alloys, which exhibit gradual coarsening above 1050 °C, providing critical theoretical insights for optimizing heat treatment processes. These findings will offer significant contributions to the mission to meet the stringent high-temperature performance requirements of the components of advanced aircraft engines’ hot sections.

2. Experimental

2.1. Alloy Preparation and Heat Treatment

In this study, GH4780 superalloy ingots were initially processed in a ZG-0.5LB vacuum induction furnace (Jinzhou Oriental Gold Technology Research Co., Ltd., Jinzhou, China). After homogenization, the ingots underwent thermomechanical processing, commencing with precision forging utilizing a 1-ton electro-hydraulic forging hammer (Jier Machine-Tool Group Co., Ltd., Jinan, China), followed by controlled hot-rolling operations performed on a 280-model horizontal rolling mill (Yantai Haige Machine Tools Co., Ltd., Yantai, China). The actual chemical composition of the alloy was measured and is detailed in Table 1. The uneven ends of the rolled bars were machined using wire electrical discharge machining (Suzhou Sanguang Science & Technology Co., Ltd., Suzhou, China), and then the specimens were cut into specimens with specific dimensions of Ø19 × 70 mm and Ø19 × 20 mm. Subsequently, the samples underwent heat treatment according to the procedure detailed in Table 2. For samples AT1~AT4, air cooling (AC) using a fan was performed after solution and aging heat treatments. Additionally, samples ST1~ST4 were water-quenched (WQ) immediately after solution heat treatment to investigate the microstructural evolution of the alloy.

2.2. Mechanical Tests

The samples AT1~AT4 were subjected to tensile tests at both room temperature and the evaluated temperatures in two sets of parallel experiments. Tensile tests were conducted at room temperature in accordance with the ASTM E8 standard [17] and at high temperature (760 °C) in accordance with the ASTM E21 standard [18], using machined cylindrical threaded dogbone specimens with a gauge length of 65 mm and a gauge diameter of 6 mm, as schematically illustrated in Figure 1. ASTM E8 (room temperature) and ASTM E21 (high temperature) were selected to align with industry standards for aerospace materials, ensuring comparability with existing datasets and certification requirements.

2.3. Microstructural Characterization

The microstructure and phase compositions of the alloy specimens before and after mechanical testing were characterized using an optical microscope (OM, Olympus Corporation, Tokyo, Japan), electron probe micro-analyzer (EPMA, JEOL Ltd., Tokyo, Japan), scanning electron microscope (SEM, JSM-7200F from JEOL Ltd., Tokyo, Japan) equipped with an energy-dispersive spectroscope (EDS in JSM7200F), electron backscatter diffraction (EBSD in JSM7200F), and a transmission electron microscope (TEM). SEM-EDS elemental mapping elucidated carbide distributions, EBSD analysis with kernel average misorientation (KAM) mapping quantified grain boundary characteristics, and TEM imaging revealed stacking faults within M23C6 carbides. This multiscale analytical approach directly linked critical microstructural features (e.g., γ′ precipitates) to mechanical performance. Firstly, metallographic preparation of the samples was performed, which included step-by-step grinding with abrasive paper, followed by mechanical polishing and finally etching in a corrosive solution for approximately 4 min. This corrosive solution contained 1.5 g of CuSO4 (Wanhua Chemical Group Co., Ltd., Yantai, China). 20 mL of C2H6O (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), and 20 mL of HCl (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China). For EBSD analysis, samples were extracted from the longitudinally sectioned fracture surfaces and then subjected to electrolytic polishing in a solution of 80 mL CH3OH (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) and 10 mL H2SO4 (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) using a voltage of 20 V for 5~6 s. For the preparation of SEM samples, an electrochemical etching technique was employed to selectively remove the γ matrix, allowing for a clearer examination of the features of γ′ precipitates. The electrochemical etching was conducted at a voltage of 4.5 V with a duration of 4~6 s in a solution consisting of 15 g Cr2O3 (Shanghai Liantian Material Technology Co., Ltd., Shanghai, China), 10 mL H2SO4, and 150 mL H3PO4 (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China). TEM and STEM analyses were performed using an FEI-Talos F200X microscope (Thermo Fisher Scientific Inc., Hillsboro, OR, USA) equipped with an advanced super-XTM EDS detector. For the preparation of electron transparent disks intended for TEM analysis, a Struers TenuPol-3 twin-jet electro-polisher (Struers A/S, Ballerup, Denmark) was employed. The electrolyte solution consisting of 90% CH3OH and 10% HClO4 (Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) was applied at a temperature of −30 °C with an operating voltage of 20 V.

3. Results and Analysis

3.1. Microstructure of GH4780 Superalloy

Figure 2a–d display EBSD inverse pole figure (IPF) maps of GH4780 superalloys after different solution and aging heat treatments, while the grain size distribution is depicted in Figure 2e–f. It can be observed that the grain size of the alloy remained at approximately 20 µm and did not exhibit significant changes when solution heat treatment was conducted at temperatures of 1020 °C and 1050 °C. When the temperature was increased to 1080 °C and 1110 °C, the average grain size rapidly increased to approximately 187 μm and 194 μm, respectively.
Figure 3a shows the backscattered electron diffraction (BSE) image of the carbides in the AT1 alloy, where coarse carbides are uniformly dispersed throughout the alloy, while fine carbides are located at the grain boundaries. The composition distribution of these carbides was mapped using EPMA, as shown in Figure 3b–l. The coarse carbides were found to be enriched in C, Nb, Ti, and Ta. Their crystal structure was determined to be face-centered cubic (FCC) through TEM selected-area electron diffraction (SAED) analysis, as shown in Figure 3i,j. As a result, these coarse carbides can be identified as MC carbides. The finer carbides at the grain boundaries were found to be rich in C and Cr. These carbides also exhibited an FCC structure but maintained a specific orientation relationship with the γ matrix substrate, Figure 3k,l. The calculated lattice constant of this carbide is 10.66 Å, which is approximately three times the matrix lattice constant of 3.6 Å. Therefore, the fine precipitated phase at the grain boundary can be identified as M23C6 carbides.
Figure 4a–d present the OM images of MC carbides in solution-treated GH4780 superalloy, demonstrating negligible variations in both size distribution and morphological characteristics among the ST1 to ST4 alloy conditions. Figure 4e–h show the TEM high-angle annular dark-field (HAADF) images of M23C6 carbides in the aged specimens of AT1 to AT4, illustrating a pronounced temperature-dependent growth behavior. Specifically, both the size and morphological complexity of M23C6 carbides exhibit substantial enhancement in the AT3 and AT4 alloys as the solution temperature increases. In the GH4780 superalloy, M23C6 carbides nucleate primarily at grain boundaries, and as the solution temperature increases, the grain size enlarges, leading to a reduction in the number and density of the grain boundaries. Consequently, the M23C6 carbides that form become larger and exhibit more complex morphologies due to the decreased frequency of nucleation sites. SEM images of fine spherical secondary γ′ precipitates distributed within the grains are shown in Figure 4i–l. The size distribution, morphological sphericity, and area fraction of the secondary γ′ precipitates in the aged state after different solution heat treatments are essentially uniform across the four superalloys AT1~AT4, with an average particle size and area fraction of approximately 30 nm and 23.8%, respectively. These M23C6 carbides and secondary γ′ phases nucleate and grow during the aging heat treatment at 800 °C for 8 h [19,20].

3.2. Tensile Properties of GH4780 Superalloy

The tensile properties of GH4780 superalloys, including yield strength (YS), ultimate tensile strength (UTS), and elongation, are systematically presented in Figure 5. Comparative analysis reveals that the AT1 and AT2 alloys exhibit nearly identical mechanical performance, with both demonstrating superior YS and UTS values compared to the AT3 and AT4 alloys across all testing temperatures. Notably, the AT2 alloy achieves optimal strength characteristics, with peak values of 920 MPa YS and 1290 MPa UTS at room temperature. When subjected to elevated temperature conditions, it exhibits only a 23.2% reduction in YS (from 920 MPa to 707 MPa) and retains 37.2% of its UST (from 1290 MPa to 810 MPa), demonstrating exceptional thermal stability. In terms of ductility, the AT4 alloy demonstrates exceptional elongation properties, particularly at high temperature, where it reaches a maximum elongation of 17%, significantly exceeding that of other tested alloys.
Figure 6 illustrates the fracture characteristics of the GH4780 superalloys after room-temperature tensile testing. Figure 6a,b depict the macroscopic fracture morphology of the AT1 and AT2 alloys, exhibiting a mixed fracture mode consisting of transcrystalline rupture and ductile fracture features. Figure 6c,d present the macroscopic fracture morphology of the AT3 and AT4 alloys, revealing a combination of intergranular fracture and ductile fracture patterns. At the microscopic level, Figure 6e,f,i,j show the detailed fracture surfaces of the AT1 and AT2 alloys, characterized by transcrystalline fracture mechanisms and well-defined dimple morphologies, indicative of their ductile behavior. In contrast, Figure 6g,h,k,l display the microscopic fracture morphologies of the AT3 and AT4 alloys, dominated by intergranular fracture features accompanied by dimple structures.
The distinct fracture behaviors observed in the AT1/AT2 and AT3/AT4 alloys highlight the significant influence of solution temperature on the microstructural characteristics and mechanical performance of GH4780 superalloys. The observed variation in fracture modes among the alloys subjected to different solution temperatures can be attributed to microstructural evolution, particularly grain growth, which reduces the density of grain boundaries. This microstructural change promotes dislocation accumulation at the remaining grain boundaries, leading to localized stress concentration. Under tensile loading, crack initiation preferentially occurs in these stress-concentrated regions, followed by propagation along the grain boundaries, ultimately resulting in intergranular fracture.
Under high-temperature tensile testing, the AT1~AT4 alloys predominantly exhibit intergranular fracture as the primary failure mode. As shown in Figure 7, the macroscopic fracture surfaces display clear intergranular fracture characteristics, while microscopic analysis reveals a distinct rock candy-like morphology along the grain boundaries, as illustrated in Figure 7e–h. This fracture behavior is consistent with the well-documented phenomenon in superalloys wherein elevated temperatures significantly reduce grain boundary strength, promoting crack propagation along these weakened boundaries during deformation [21].
The enhanced elongation observed in AT3 and AT4 alloys, as shown in Figure 5b, can be attributed to microstructural modifications resulting from increased solution temperatures, particularly grain growth, which reduces the density of grain boundaries. However, this microstructural evolution also facilitates the precipitation of M23C6 carbides at grain boundaries, which usually has a detrimental effect on mechanical properties. As demonstrated in Figure 8a, the aggregation and coarsening of M23C6 carbides create localized stress concentrations, serving as preferential sites for crack initiation during high-temperature deformation. This mechanism significantly reduces the plastic deformation capability of the matrix, thereby reducing the yield strength of the alloy. Furthermore, Figure 8b–d reveal the formation of extensive stacking faults within grain boundary M23C6 carbides during deformation. This structural instability leads to carbide softening at elevated temperatures, promoting crack nucleation within the carbides themselves. The combined effects of carbide softening and stress concentration at grain boundaries severely compromise boundary integrity, ultimately leading to a substantial reduction in the overall strength of the superalloy.

4. Discussion

4.1. Effect of Solution Treatment on the Microstructural Evolution of GH4780 Superalloy

The above experimental results demonstrate a strong correlation between solution temperature and grain size evolution in GH4780 superalloy, which directly governs its mechanical performance. As shown in Figure 2, the GH4780 superalloy undergoes a distinct grain growth transition within the critical temperature range of 1050~1080 °C. Notably, when the solution temperature exceeds 1080 °C, a marked acceleration in grain growth kinetics is observed during thermal processing. Therefore, this paper intends to establish a comprehensive kinetic model to describe the grain growth behavior of the alloy during solution treatment, focusing on the temperature range of 1080~1110 °C. The Sellars model, originally developed for predicting austenite grain growth kinetics, was adapted and implemented to quantitatively characterize the grain size evolution of GH4780 superalloy under various solution treatment conditions [22]:
D n = D 0 + A t e x p ( Q R T )
In Equation (1), the parameters are defined as follows. D denotes the final average grain diameter (μm), D0 represents the initial grain size prior to solution treatment (μm), t corresponds to the isothermal holding time (s), T indicates the solution temperature (K), Q stands for the activation energy of grain boundary migration (J/mol), and R is the universal gas constant (8.314 J·mol−1·K−1). Experimental observations reveal that grain growth behavior within the temperature range of 1050~1110 °C demonstrates size-independent characteristics, allowing for the elimination of the initial grain size term D0. Consequently, the Sellars model can be simplified to the following form, as shown in Equation (2):
D n = A t e x p ( Q R T )
Taking the logarithm on both sides of the simplified Sellars model yields Equation (3):
ln D = 1 n l n A + 1 n l n t Q n R 1 T
Based on the experimental data, the relationship curves between ln D − ln t and between ln D − 1000/T were acquired, as depicted in Figure 9a,b, respectively. By linear fitting, the values of n = 0.99, Q = 256,978.26 J·mol−1, and A = 4.03 × 108 were calculated. Substituting these parameters into Equation (2), the grain growth equation for GH4780 superalloy under different solution temperatures and holding times can be obtained, as shown in Equation (4):
D 0.99 = 4.03 × 10 8 · t · e x p ( 256,978.26 8.314 T )
The good agreement presented in Figure 9 clearly demonstrates that this equation is capable of accurately depicting the trend of variation in grain size in the GH4780 superalloy throughout the solution treatment process.
Moreover, the microstructures of the superalloys at different solution temperatures were analyzed to explore the impact of microstructure on grain growth. Grain growth represents a microstructural evolution process propelled by grain boundary migration, which commences with a decrease in surface energy. As the solution temperature increases, the thermal energy of atoms increases, causing an elevation in their vibrational amplitude. This increased thermal energy promotes thermally activated atomic motion and boosts atomic diffusion. As a result, the diffusion of atoms in the vicinity of grain boundaries is enhanced, making it more probable that these atoms transfer into adjacent grains. This transfer leads to the gradual coalescence and growth of the grains [23,24]. Furthermore, at high temperatures, atoms at grain boundaries possess higher energies, enabling them to more readily surmount energy barriers and migrate along the grain boundaries. This migration of grain boundaries causes the interface between adjacent grains to gradually disappear, thereby resulting in an increase in grain size [25].
Furthermore, the second-phase particles present in the superalloy can act as obstacles to the migration of grain boundaries, impeding the movement of grain boundaries and thus preventing the growth grains [26,27]. As is evident in Figure 10a,c for the ST1 and ST2 alloys, the grain boundaries are pinned by MC carbides, which in turn inhibits grain growth. When the solution temperature is increased, grain boundaries begin to migrate, as observed in Figure 10b,d for the ST3 and ST4 alloys. In these alloys, the grain boundaries are no longer under the pinning influence of MC carbides, thereby enabling the process of grain growth to take place. A schematic diagram that vividly illustrates the migration of grain boundaries with the increase in solution temperature is presented in Figure 10e.

4.2. Effect of Grain Size on Tensile Properties of GH4780 Superalloy at Room Temperature

Grain size significantly influences the tensile properties of the GH4780 superalloy. The relationship between the grain size and the yield strength of the alloy follows the Hall–Petch relationship, as described by Equation (5) [28]:
σ S = σ 0 + k d 1 2
In this equation, σs stands for the yield strength of the alloy, k is the proportional coefficient known as the Hall–Petch constant, d denotes the average grain size, and σ0 represents the intrinsic lattice friction stress of the alloy. The experimental results demonstrate a clear correlation between solution temperature and grain size evolution, where elevated temperatures promote grain coarsening, consequently reducing the yield strength of the alloy, as quantitatively illustrated in Figure 5a. At room temperature, grain boundaries exhibit superior mechanical strength compared to intragranular regions, thereby playing a crucial role in the strengthening mechanisms of GH4780 superalloy. A fine grain size gives rise to a higher density of grain boundaries that effectively impedes the motion of dislocations during the deformation process, resulting in enhanced yield strength and tensile strength [29,30,31]. This mechanism is further supported by EBSD IPF and KAM measurements, as presented in Figure 11, which reveal a high dislocation density at the grain boundary in the AT3 alloy. However, it should be noted that while grain refinement strengthens the alloy, it simultaneously increases the defect density. Conversely, grain coarsening reduces the grain boundary area, thereby decreasing defect concentration and improving plastic deformation capability [32]. Therefore, among the studied alloys, the AT4 alloy demonstrates superior plasticity.

4.3. Effect of Microstructure on High-Temperature Tensile Properties of GH4780 Superalloy

The tensile properties of the GH4780 superalloy at high temperatures are inherently associated with its resistance to crack initiation and propagation. During the deformation process, the density of geometrically necessary dislocations in the fine-grained regions is significantly higher than that in the coarse-grained regions. At high temperatures, grain boundaries are in a more viscous state, making them more susceptible to sliding and migration under the influence of external forces [33,34]. This thermally activated process diminishes the strengthening effect typically associated with grain refinement, suggesting that controlled grain coarsening may effectively mitigate grain boundary sliding and potentially enhance high-temperature mechanical performance.
Experimental data demonstrate distinct reductions in yield strength at high temperatures, of approximately 24% for the AT1 and AT2 alloys and 17% for the AT3 and AT4 alloys, when compared to the values at room temperature. This differential response can be attributed to the more uniform deformation behavior of coarse-grained structures at high temperatures, where homogeneous dislocation motion reduces localized stress concentrations and minimizes damage accumulation. This enhanced high-temperature performance is further supported by the EBSD KAM analysis in Figure 12, which compares the dislocation distributions in the AT1 and AT3 alloys after high-temperature tensile testing. The KAM maps clearly indicate higher dislocation densities in fine-grained regions, consistent with the observed mechanical behavior.
Under high-temperature service conditions, superalloys exhibit distinct deformation mechanisms that significantly differ from room-temperature behavior, making the investigation of these mechanisms essential for performance optimization of GH4780 superalloy. Microtwins (MTs) are one of the primary deformation mechanisms during high-temperature treatment, and their formation is closely related to the crystal structure of the superalloy. In the FCC structure, the layered atomic arrangement enables relative slip between atomic layers through twin boundaries, resulting in the formation of MTs. During alloy plastic deformation, dislocation motion and interactions serve as the main driving forces for MT formation. When external stress is applied, dislocations slip and climb within the crystal. Under specific conditions, dislocation accumulation can induce atomic layer slip, forming micro-twin boundaries and creating MTs. However, MT propagation may cause stress concentration during deformation, leading to crack initiation and strength degradation. Excessive MT formation can increase alloy brittleness, as micro-twin boundaries may provide paths for crack propagation, ultimately reducing the performance of the superalloy [35,36,37]. As shown in Figure 13, TEM bright-field (BF) images reveal the synergistic effect of grain boundaries in hindering MT propagation. During deformation, MT expansion is effectively constrained by the high density of grain boundaries in the fine-grained AT1 and AT2 alloys, thereby enhancing their strength.

5. Conclusions

The present study systematically investigated the effects of solution temperature on the microstructural evolution and mechanical properties of GH4780 superalloy. Through comprehensive microstructural characterization and mechanical testing, the following conclusions could be drawn.
(1) In GH4780 superalloy, grain size evolution strongly depends on solution treatment temperatures. MC carbides are crucial for pinning grain boundaries, effectively restraining grain growth below 1080 °C. However, above 1080 °C, thermally activated grain boundary migration significantly accelerates grain growth kinetics. Furthermore, a grain growth kinetic model has been established within the temperature range of 1080~1110 °C.
(2) The mechanical properties of GH4780 superalloy demonstrate a significant grain size dependence. The AT2 alloy, with an average grain size of 20 µm, exhibits superior room-temperature yield strength of 920 MPa and tensile strength of 1290 MPa. This strength advantage is maintained at elevated temperatures, demonstrating yield strength of 707 MPa and tensile strength of 810 MPa at 760 °C. In contrast, the coarse-grained AT4 alloy (average grain size ≈ 194 µm) shows enhanced ductility, achieving maximum elongation of 17% at 760 °C.
(3) High-temperature tensile loading induces grain boundary weakening in GH4780 superalloy, primarily causing intergranular fracture. This process is accelerated by M23C6 carbides at grain boundaries, where stress concentrations initiate cracks that propagate along boundary networks. During deformation, dislocation interactions generate microtwins (MTs) that reduce strength when propagating. However, grain boundaries effectively hinder MT expansion, enhancing alloy strength through a pinning mechanism, thereby strengthening the alloy.
(4) The grain growth kinetics model and mechanical property data established in this study provide a comprehensive quantitative framework for optimizing the heat treatment parameters of GH4780 superalloys in advanced aero-engine and industrial gas turbine applications. By accurately predicting a critical threshold temperature of 1080 °C, the model enables precise control of alloy grain size, ensuring optimal mechanical properties and long-term operational safety under extreme service conditions.

Author Contributions

Conceptualization, X.-F.X. and L.-L.Z.; methodology, T.-H.F.; software, Y.L.; validation, X.-F.X. and L.-L.Z.; formal analysis, J.-L.Q. and S.-M.L.; investigation, T.-H.F.; resources, X.-F.X. and L.-L.Z.; data curation, T.-H.F.; writing—original draft preparation, T.-H.F.; writing—review and editing, X.-F.X. and L.-L.Z.; visualization, J.-J.R.; supervision, J.-H.D.; funding acquisition, X.-F.X. and L.-L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Nos. 52471147, 52074092), the Shandong Natural Science Foundation of China (No. ZR2021QE102), the Taishan Scholars Program of Shandong Province (No. tsqn201909081), the Science Foundation Program for Distinguished Young Scholars (Overseas) of Shandong Province (2022HWYQ-084), and the Talent Training Program for Shandong Province Higher Educational Youth Innovative Teams (2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

Authors Tian-Hao Feng, Xing-Fei Xie, Yang Liu, Jing-Long Qu, Shao-Min Lyu, Jin-Hui Du were employed by Beijing GAONA Materials and Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Bryndza, G.; Tchuindjang, J.T.; Chen, F.; Habraken, A.M.; Sepúlveda, H.; Tuninetti, V.; Mertens, A.; Duchêne, L. Review of the Microstructural Impact on Creep Mechanisms and Performance for Laser Powder Bed Fusion Inconel 718. Materials 2025, 18, 276. [Google Scholar] [CrossRef] [PubMed]
  2. He, D.; Chen, X.; Lin, Y.C.; Yan, X.; Xie, H. Influences of stress triaxiality and forming parameters on microstructural evolution and fracture mechanisms in a Ni–Cr–Mo-based superalloy. J. Mater. Res. Technol. 2024, 32, 378–394. [Google Scholar] [CrossRef]
  3. Hardy, M.C.; Detrois, M.; Mcdevitt, E.T.; Argyrakis, C.; Saraf, V.; Jablonski, P.D.; Hawk, J.A.; Buckingham, R.C.; Kitaguchi, H.S.; Tin, S. Solving Recent Challenges for Wrought Ni-Base Superalloys. Metall. Mater. Trans. A 2020, 51, 2626–2650. [Google Scholar] [CrossRef]
  4. Zhu, Q.; Cheng, L.; Wang, C.; Chen, G.; Qin, H.; Zhang, P. Effect of δ phase on size effect in microtensile deformation of a nickel-based superalloy. Mater. Sci. Eng. A 2019, 766, 138405. [Google Scholar] [CrossRef]
  5. Liu, Z.; Liu, W.; Zhang, H.; Ruan, J.; Huang, H.; Zhou, X.; Meng, F.; Zhang, S.; Jiang, L. Dramatically improving thermoplasticity of FGH4096 superalloy by a novel sub-solvus temperature holding followed by extremely slow cooling. J. Mater. Res. Technol. 2023, 24, 1973–1990. [Google Scholar] [CrossRef]
  6. Lv, Y.; Ouyang, X.; Tian, Y.; Yu, H.; Wei, G. Insight into the hot corrosion behavior of GH4251 high-temperature superalloy fabricated by selective laser melting. J. Alloys Compd. 2024, 1002, 175166. [Google Scholar] [CrossRef]
  7. Rajabinezhad, M.; Bahrami, A.; Mousavinia, M.; Seyedi, S.J.; Taheri, P. Corrosion-fatigue failure of gas-turbine blades in an oil and gas production plant. Materials 2020, 13, 900. [Google Scholar] [CrossRef]
  8. Cheng, X.; Xu, J.; Yang, Z.; Guo, Q.; Li, C.; Zhou, J.; Chen, S.-L.; Liu, S.; Ma, Z. A novel nickel-based superalloy with excellent high temperature performance designed for laser additive manufacturing. Mater. Sci. Eng. A 2024, 911, 146926. [Google Scholar] [CrossRef]
  9. Xi, X.; Lin, D.; Song, X.; Luo, X.; Ma, R.; Shi, Z.; Bian, H.; Fu, W.; Dong, Z.B.; Tan, C. Strength-plasticity transition mechanism after the solution treatment of GH3230 superalloy fabricated via laser powder bed fusion. Mater. Sci. Eng. A 2023, 876, 145124. [Google Scholar] [CrossRef]
  10. Tian, Z.-L.; Jiang, S.-B.; Chen, Z.-Z.; Bao, H.-S.; Liu, Z.-D. Microstructural evolution and mechanical properties of a new Ni-based heat-resistant alloy during aging at 750 °C. J. Iron Steel Res. Int. 2017, 24, 513–519. [Google Scholar] [CrossRef]
  11. Wang, Y.; Li, N.; Jia, C.; Liu, W.; Wu, Y.; Gao, C.; Xiong, H. Effects of solution treatment on the microstructural evolution and tensile properties of GH 3230 superalloy manufactured by laser melting deposition. Mater. Charact. 2022, 194, 112339. [Google Scholar] [CrossRef]
  12. Zhang, X.; Han, H.; Zhou, Y.; Chen, J.; Feng, S.; Tang, P.; Xiao, D.; Fu, J.; Zhang, J. Effect of solution treatment on the microstructure and elevated temperature tensile properties of forged Rene 41 superalloy. Materials 2024, 17, 6150. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, W.; Zheng, Y.; Liu, F.; Wang, D.; Liu, F.; Huang, C.; Li, Q.; Lin, X.; Huang, W. Effect of solution temperature on the microstructure and mechanical properties of Hastelloy X superalloy fabricated by laser directed energy deposition. Mater. Sci. Eng. A 2021, 820, 141537. [Google Scholar] [CrossRef]
  14. Li, Z.; Wang, D.; Wang, L.; Jiang, C.; Zhou, L.; Zhou, Z.; Wang, H.; Zhai, Y. The behaviour and kinematic model for γ′ evolution and grain growth during solution treatment of a powdered metallurgical Ni-based superalloy. Mater. Charact. 2023, 199, 112820. [Google Scholar] [CrossRef]
  15. Song, K.; Aindow, M. Grain growth and particle pinning in a model Ni-based superalloy. Mater. Sci. Eng. A 2008, 479, 365–372. [Google Scholar] [CrossRef]
  16. Zhou, Z.; Zhang, R.; Cui, C.; Zhou, Y.; Sun, X. Influence of solution temperature on microstructure and creep behaviors of a Ni–Co base superalloy for turbine disc. Mater. Sci. Eng. A 2022, 853, 143741. [Google Scholar] [CrossRef]
  17. ASTM E8; Standard Test Methods for Tension Testing of Metallic Materials. ASTM International: West Conshohocken, PA, USA, 2022.
  18. ASTM E21; Standard Test Methods for Elevated Temperature Tension Tests of Metallic Materials. ASTM International: West Conshohocken, PA, USA, 2020.
  19. Velasco-Castro, M.; León-Cázares, F.D.; Galindo-Nava, E.I. A comprehensive review of microstructural heterogeneities in the laser powder bed fusion of nickel-base superalloys with high γ′ content. Mater. Des. 2024, 247, 113416. [Google Scholar] [CrossRef]
  20. Zhao, S.; Xie, X.; Smith, G.D.; Patel, S.J. Gamma prime coarsening and age-hardening behaviors in a new nickel base superalloy. Mater. Lett. 2004, 58, 1784–1787. [Google Scholar] [CrossRef]
  21. Masoumi, F.; Jahazi, M.; Shahriari, D.; Cormier, J. Coarsening and dissolution of γ′ precipitates during solution treatment of AD730™ Ni-based superalloy: Mechanisms and kinetics models. J. Alloys Compd. 2016, 658, 981–995. [Google Scholar] [CrossRef]
  22. Sellars, C.; Whiteman, J. Recrystallization and grain growth in hot rolling. Met. Sci. 1979, 13, 187–194. [Google Scholar] [CrossRef]
  23. Qiao, S.; Wang, Y.; Lv, L.; Huang, Z.; Tan, G. Normal and abnormal grain growth in a FGH96 superalloy during thermomechanical treatment. J. Mater. Res. Technol. 2021, 15, 7033–7049. [Google Scholar] [CrossRef]
  24. Rohrer, G.S. “Introduction to Grains, Phases, and Interfaces—An Interpretation of Microstructure,” Trans. AIME, 1948, vol. 175, pp. 15–51, by C.S. Smith. Metall. Mater. Trans. B 2010, 41, 457–494. [Google Scholar] [CrossRef]
  25. Zhang, B.; Ning, Y.; Wang, Z.; Liu, Q. PPB structure elimination, DRX nucleation mechanisms and grain growth behavior of the 3rd-generation PM superalloy for manufacturing aviation components. Chin. J. Aeronaut. 2024, 37, 325–344. [Google Scholar] [CrossRef]
  26. Qiao, S.; Liu, S.; Li, N.; Gao, X.; Wen, X.; Wang, F.; Yuan, C. Abnormal grain growth in the Ni-based wrought superalloy GH4698 bar during heat treatment. J. Mater. Res. Technol. 2024, 30, 6563–6575. [Google Scholar] [CrossRef]
  27. Lan, J.; Huang, H.; Mao, H.; Hua, L. Phase transformation and grain growth behaviors of superalloy IN718 during heat treatment. Mater. Today Commun. 2020, 24, 101347. [Google Scholar] [CrossRef]
  28. Reed, R. The Superalloys: Fundamentals and Applications; Cambridge University Press: Cambridge, UK, 2006; pp. 1–372. [Google Scholar]
  29. Hong, S.I.; Laird, C. Mechanisms of slip mode modification in FCC solid solutions. Acta Metall. Et Mater. 1990, 38, 1581–1594. [Google Scholar] [CrossRef]
  30. Alexandre, F.; Deyber, S.; Pineau, A. Modelling the optimum grain size on the low cycle fatigue life of a Ni based superalloy in the presence of two possible crack initiation sites. Scr. Mater. 2004, 50, 25–30. [Google Scholar] [CrossRef]
  31. Joseph, C.; Persson, C.; Colliander, M.H. Influence of heat treatment on the microstructure and tensile properties of Ni-base superalloy Haynes 282. Mater. Sci. Eng. A 2017, 679, 520–530. [Google Scholar] [CrossRef]
  32. Alabort, E.; Barba, D.; Sulzer, S.; Lißner, M.; Petrinic, N.; Reed, R.C. Grain boundary properties of a nickel-based superalloy: Characterisation and modelling. Acta Mater. 2018, 151, 377–394. [Google Scholar] [CrossRef]
  33. Zhang, S.; Lin, X.; Wang, L.; Yu, X.; Yang, H.; Lei, L.; Huang, W. Influence of grain inhomogeneity and precipitates on the stress rupture properties of Inconel 718 superalloy fabricated by selective laser melting. Mater. Sci. Eng. A 2021, 803, 140702. [Google Scholar] [CrossRef]
  34. Martelli, P.A.; Sabirov, I.; Monclus, M.A.; Bassini, E.; Marchese, G.; Ugues, D. The effect of temperature and strain rate on the grain boundary sliding in a CM247 LC Ni-based superalloy processed with laser based powder bed fusion. J. Mater. Res. Technol. 2024, 28, 2466–2477. [Google Scholar] [CrossRef]
  35. Dong, J.; Yu, H.; Cao, R.; Liu, Z.; Wang, X.; Wang, J.; Bi, Z. The effect of pre-strain on textures, nano-twins and mechanical properties of a novel Nickel-based superalloy GH4251 with low stacking fault energy. Mater. Sci. Eng. A 2024, 915, 147271. [Google Scholar] [CrossRef]
  36. Cai, Y.Q.; Tan, Y.B.; Wang, L.X.; Shi, W.; Ji, X.M.; Xiang, S. Multiple strengthening mechanisms induced by nanotwins and stacking faults in CoNiCr-superalloy MP159. Mater. Sci. Eng. A 2022, 853, 143793. [Google Scholar] [CrossRef]
  37. Smith, T.M.; Good, B.S.; Gabb, T.P.; Esser, B.D.; Egan, A.J.; Evans, L.J.; McComb, D.W.; Mills, M.J. Effect of stacking fault segregation and local phase transformations on creep strength in Ni-base superalloys. Acta Mater. 2019, 172, 55–65. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the GH4780 specimens for tensile testing.
Figure 1. Schematic diagram of the GH4780 specimens for tensile testing.
Materials 18 01288 g001
Figure 2. EBSD IPF maps and grain size distribution plots of GH4780 superalloys after different heat treatments: (a,e) AT1; (b,f) AT2; (c,g) AT3; (d,h) AT4.
Figure 2. EBSD IPF maps and grain size distribution plots of GH4780 superalloys after different heat treatments: (a,e) AT1; (b,f) AT2; (c,g) AT3; (d,h) AT4.
Materials 18 01288 g002
Figure 3. Composition and crystal structure analyses for the MC and M23C6 carbides in the AT1 alloy: (a) BSE image highlighting the location and morphology of the carbides; (bh) elemental distributions in the carbides; (i) TEM bright-field (BF) image and (j) obtained SAED patterns of the MC carbide from the region A in (i); (k) TEM BF image and (l) obtained SAED patterns of the M23C6 carbide from the region B in (k).
Figure 3. Composition and crystal structure analyses for the MC and M23C6 carbides in the AT1 alloy: (a) BSE image highlighting the location and morphology of the carbides; (bh) elemental distributions in the carbides; (i) TEM bright-field (BF) image and (j) obtained SAED patterns of the MC carbide from the region A in (i); (k) TEM BF image and (l) obtained SAED patterns of the M23C6 carbide from the region B in (k).
Materials 18 01288 g003
Figure 4. The microstructure of the GH4780 superalloy after various heat treatments, illustrating the precipitation of carbides and secondary γ′ phases: (ad) OM images showing the MC carbides in the ST1~ST4 alloys; (eh) TEM HAADF images showing the M23C6 carbides in the AT1~AT4 alloys; (il) SEM images showing the fine secondary γ′ phase in the AT1~AT4 alloys.
Figure 4. The microstructure of the GH4780 superalloy after various heat treatments, illustrating the precipitation of carbides and secondary γ′ phases: (ad) OM images showing the MC carbides in the ST1~ST4 alloys; (eh) TEM HAADF images showing the M23C6 carbides in the AT1~AT4 alloys; (il) SEM images showing the fine secondary γ′ phase in the AT1~AT4 alloys.
Materials 18 01288 g004
Figure 5. Tensile properties of the GH4780 superalloys tested at (a) room temperature and (b) a high temperature of 760 °C.
Figure 5. Tensile properties of the GH4780 superalloys tested at (a) room temperature and (b) a high temperature of 760 °C.
Materials 18 01288 g005
Figure 6. Fracture morphology of the GH4780 superalloys after tensile testing at room temperature: (ad) macroscopic fracture surfaces of the AT1~AT4 alloys; (el) high-magnification SEM images of the microscopic fracture morphology in the AT1~AT4 alloys, revealing detailed features of dimple structures and fracture mechanisms.
Figure 6. Fracture morphology of the GH4780 superalloys after tensile testing at room temperature: (ad) macroscopic fracture surfaces of the AT1~AT4 alloys; (el) high-magnification SEM images of the microscopic fracture morphology in the AT1~AT4 alloys, revealing detailed features of dimple structures and fracture mechanisms.
Materials 18 01288 g006
Figure 7. Fracture morphology of the GH4780 superalloy after tensile testing at a high temperature of 760 °C: (ad) macroscopic fracture surfaces of the AT1~AT4 alloys; (eh) high-magnification SEM images of the microscopic fracture morphology in the AT1~AT4 alloys.
Figure 7. Fracture morphology of the GH4780 superalloy after tensile testing at a high temperature of 760 °C: (ad) macroscopic fracture surfaces of the AT1~AT4 alloys; (eh) high-magnification SEM images of the microscopic fracture morphology in the AT1~AT4 alloys.
Materials 18 01288 g007
Figure 8. TEM analysis of crack initiation mechanisms associated with M23C6 carbides: (a) crack propagation behavior adjacent to M23C6 carbides at grain boundaries; (b) formation of stacking faults (SFs) within M23C6 carbides; (c) magnified view of region A from (b), demonstrating crack nucleation and growth within M23C6 carbides; (d) corresponding Fourier transform pattern of (c), showing distinct diffraction spots that confirm the crystallographic features of the cracked carbide structure.
Figure 8. TEM analysis of crack initiation mechanisms associated with M23C6 carbides: (a) crack propagation behavior adjacent to M23C6 carbides at grain boundaries; (b) formation of stacking faults (SFs) within M23C6 carbides; (c) magnified view of region A from (b), demonstrating crack nucleation and growth within M23C6 carbides; (d) corresponding Fourier transform pattern of (c), showing distinct diffraction spots that confirm the crystallographic features of the cracked carbide structure.
Materials 18 01288 g008
Figure 9. Quantitative analysis of grain growth kinetics using the Sellars model: (a) ln D − ln t curves for the index n; (b) ln D − 1000/T curves for the activation energy Q.
Figure 9. Quantitative analysis of grain growth kinetics using the Sellars model: (a) ln D − ln t curves for the index n; (b) ln D − 1000/T curves for the activation energy Q.
Materials 18 01288 g009
Figure 10. Schematic illustration of grain boundary migration mechanisms: (a) pinning effect of MC carbides on grain boundary motion in ST1 alloy; (b) unrestricted grain boundary migration in ST3 alloy; (c) and (d) magnified views of regions A and B in (a) and (b), respectively; (e) proposed mechanism of temperature-dependent grain boundary migration.
Figure 10. Schematic illustration of grain boundary migration mechanisms: (a) pinning effect of MC carbides on grain boundary motion in ST1 alloy; (b) unrestricted grain boundary migration in ST3 alloy; (c) and (d) magnified views of regions A and B in (a) and (b), respectively; (e) proposed mechanism of temperature-dependent grain boundary migration.
Materials 18 01288 g010
Figure 11. EBSD images showing the grains and dislocation densities in the AT3 alloy after room-temperature tensile testing: (a) IPF map; (b) KAM map.
Figure 11. EBSD images showing the grains and dislocation densities in the AT3 alloy after room-temperature tensile testing: (a) IPF map; (b) KAM map.
Materials 18 01288 g011
Figure 12. EBSD images showing the grains and dislocation densities in the AT1 and AT3 alloys after high-temperature tensile testing: IPF maps for (a) the AT1 alloy and (c) the AT3 alloy; KAM maps for (b) the AT1 alloy and (d) the AT3 alloy.
Figure 12. EBSD images showing the grains and dislocation densities in the AT1 and AT3 alloys after high-temperature tensile testing: IPF maps for (a) the AT1 alloy and (c) the AT3 alloy; KAM maps for (b) the AT1 alloy and (d) the AT3 alloy.
Materials 18 01288 g012
Figure 13. Microstructural characterization of MT interactions with grain boundaries in the AT3 alloy: (a) TEM BF image demonstrating MT inhibition at grain boundaries; (b) SAED pattern of MTs; (c) high-resolution TEM (HRTEM) image revealing MT morphology and atomic structure.
Figure 13. Microstructural characterization of MT interactions with grain boundaries in the AT3 alloy: (a) TEM BF image demonstrating MT inhibition at grain boundaries; (b) SAED pattern of MTs; (c) high-resolution TEM (HRTEM) image revealing MT morphology and atomic structure.
Materials 18 01288 g013
Table 1. Measured composition of the GH4780 superalloy (wt.%).
Table 1. Measured composition of the GH4780 superalloy (wt.%).
CoCrWNbTaAl + TiC + B + ZrSiNi
17.9921.271.720.630.933.000.0730.053Bal.
Table 2. Heat treatment procedures explored for the GH4780 superalloy.
Table 2. Heat treatment procedures explored for the GH4780 superalloy.
No.Solution TreatmentAging Treatment
AT11020 °C/1 h (AC)800 °C/8 h (AC)
AT21050 °C/1 h (AC)800 °C/8 h (AC)
AT31080 °C/1 h (AC)800 °C/8 h (AC)
AT41110 °C/1 h (AC)800 °C/8 h (AC)
ST11020 °C/1 h (WQ)-
ST21050 °C/1 h (WQ)-
ST31080 °C/1 h (WQ)-
ST41110 °C/1 h (WQ)-
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

Feng, T.-H.; Xie, X.-F.; Liu, Y.; Qu, J.-L.; Lyu, S.-M.; Du, J.-H.; Ruan, J.-J.; Zhu, L.-L. Effects of Solution Treatment on Microstructure Evolution and the Mechanical Properties of GH4780 Superalloy. Materials 2025, 18, 1288. https://doi.org/10.3390/ma18061288

AMA Style

Feng T-H, Xie X-F, Liu Y, Qu J-L, Lyu S-M, Du J-H, Ruan J-J, Zhu L-L. Effects of Solution Treatment on Microstructure Evolution and the Mechanical Properties of GH4780 Superalloy. Materials. 2025; 18(6):1288. https://doi.org/10.3390/ma18061288

Chicago/Turabian Style

Feng, Tian-Hao, Xing-Fei Xie, Yang Liu, Jing-Long Qu, Shao-Min Lyu, Jin-Hui Du, Jing-Jing Ruan, and Li-Long Zhu. 2025. "Effects of Solution Treatment on Microstructure Evolution and the Mechanical Properties of GH4780 Superalloy" Materials 18, no. 6: 1288. https://doi.org/10.3390/ma18061288

APA Style

Feng, T.-H., Xie, X.-F., Liu, Y., Qu, J.-L., Lyu, S.-M., Du, J.-H., Ruan, J.-J., & Zhu, L.-L. (2025). Effects of Solution Treatment on Microstructure Evolution and the Mechanical Properties of GH4780 Superalloy. Materials, 18(6), 1288. https://doi.org/10.3390/ma18061288

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