Publication Cover
Materials Technology
Advanced Performance Materials
Volume 38, 2023 - Issue 1
416
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Study on superplastic behaviour of cold rolled Inconel 718 alloy with δ phase dispersion distribution

, , , &
Article: 2221548 | Received 30 May 2023, Accepted 31 May 2023, Published online: 10 Jun 2023

ABSTRACT

Cold-rolled Inconel 718 with δ phase dispersion distribution was prepared and exhibited excellent superplasticity at 950°C with a strain rate of 6 × 10−4 ~4 × 10−3 s−1. The elongation to failure of 561% was obtained at a strain rate of 1 × 10−3 s−1. Superplastic deformation accelerates the recrystallization of cold-rolled deformed microstructure, thus improving grain uniformity. The dispersed δ phase promotes the nucleation of recrystallized grains and pin grain boundaries to limit grain size. Diffusion-coordinated grain boundary slip is the main deformation mechanism of the alloy sheet during superplastic stretching. The dense dislocations in the deformed grains act as the channel of atomic diffusion in the early deformation stage. Subsequently, the material conversion between grain boundaries promotes atomic diffusion. The present results have certain reference significance for preparing and forming superplastic Inconel 718 alloy.

Introduction

Superplastic forming technology has the advantages of large deformation, low forming pressure, and precise dimensions. The superplasticity of alloys at high temperatures makes it possible to manufacture complex parts by gas blowing. Inconel 718 is a nickel-based superalloy with high-temperature strength, excellent fatigue resistance, and good corrosion resistance. On the other hand, due to its high deformation resistance and severe work-hardening, the traditional forming process of the alloy has high requirements for moulds and equipment. Therefore, the research on the superplasticity of the alloy is beneficial to reduce production costs and expand its application range. The combination of deformation and heat treatment to improve the microstructure is the main method to enhance the superplasticity of the alloy.

Jafarian et al. [Citation1] found that superplasticity occurs in metals with a grain size of less than 10 µm and good structural stability. Conditions for realizing superplasticity include high deformation temperature (T ≥0.5 Tm) and low strain rate (10−4 ~10−1 s−1). Moreover, Sato and Kuribayashi [Citation2] pointed out that since dynamic recrystallization (DRX) occurs within the temperature range of superplastic deformation, non-annealed alloys can exhibit superplasticity by refining grains through DRX at the initial stage of superplastic deformation. Therefore, it is feasible to achieve superplasticity in some deformed alloys. Yang et al. [Citation3], through analysing the microscopic superplastic deformation mechanism of cold-rolled Inconel 718, proved that atomic diffusion, grain rotation, dislocation creep, and continuous DRX occur during superplastic deformation. Lv et al. [Citation4] carried out a superplastic tensile test on a hot extruded nickel-based superalloy. It was pointed out that the extruded alloy has a high-density dislocation, γ’ phase plays the role of pinning grain boundaries, and grain boundary slip (GBS) controlled by grain boundary diffusion and lattice diffusion is the superplastic deformation mechanism. Yakovtseva et al. [Citation5] reported that DRX is beneficial for grain refinement through PSN and Zener pinning mechanisms, providing a large m value and sufficient GBS for the high strain rate superplastic deformation of aluminium-based alloy. Hidalgo-Manrique [Citation6] obtained equiaxed grains with an average size of less than 500 nm through the multi-pass rolling process and proved that the stored energy and the pinning effect brought by rolling are beneficial to promote recrystallization and refine grains in the preheating stage of superplastic deformation.

Many studies have shown that the δ phase in Inconel 718 alloy achieves grain refinement by both pinning grain boundaries and promoting nucleation. For example, Páramo-Kañetas et al. [Citation7] studied that the δ phase inhibits grain growth and promotes nucleation of recrystallized grains during thermal deformation. Xu et al. [Citation8] investigated the cold rolling effect reduction rate on the δ phase precipitation of Inconel 718 alloy and pointed out that cold rolling increased the nucleation of δ phase and decreased the critical nucleation work, thus promoting the precipitation of δ phase.

In this paper, a rolled Inconel 718 alloy with a dispersed distribution of δ was prepared by two-stage rolling and intermediate annealing. The superplasticity of the alloy prepared by the above process was characterized. The microstructure evolution and related mechanism during deformation were studied in detail. In addition, we focus on the role of the δ phase in the superplastic deformation.

Experimental procedures

Preparation process

The initial microstructure with an average grain size of 5.1 μm is indicated in . The alloy was solution-treated at 1050°C for 1 h and cut into 5 mm thickness plates. Subsequently, the samples were cold-rolled (CR) to a thickness of 1.6 mm with a reduction of 68%. The samples were heated in a tube furnace for 5 h at 890°C to precipitate δ phase. The microstructure of plates after δ phase precipitation treatment (CD plates) is shown in . Then, the CD plates were CR with a reduction of 16% to obtain δ phase dispersed cold-rolled sheets (CDC sheets, ). The volume fraction of the δ phase was calculated as 10.72%. The detailed preparation method and subsequent superplastic tensile test are indicated in .

Figure 1. (a) Optical microscopy (OM) image of Inconel 718. Electron microprobe (SEM) images of (b) 68% CR + 890°C × 5h and (c) 68% CR+890°C×5h +16% CR.

Figure 1. (a) Optical microscopy (OM) image of Inconel 718. Electron microprobe (SEM) images of (b) 68% CR + 890°C × 5h and (c) 68% CR+890°C×5h +16% CR.

Figure 2. Flow chart of the preparation process and superplastic tensile test.

Figure 2. Flow chart of the preparation process and superplastic tensile test.

Superplastic tensile test

The CDC sheets were heated to 950°C and maintained for 5 min. Then, the alloy was stretched at a rate of 6 × 10−4 to 4 × 10−3 s−1 using an AG-X plus 100 N. To understand the microstructural evolution, a few samples were stretched to an elongation of 50%, 150%, and 350% at a strain rate of 1 × 10−3 s−1 and 950°C. Fractured and interrupted specimens from the above stretching process were cooled to room temperature by water quenching.

Microstructure detection

The specimens for OM and SEM observation were ground and polished, followed by chemical etching using a solution of 20 mL HCL +20 mL C2H5OH +5 g CuSO4. The content of the δ phase in 4000 magnification images was counted by Image-Pro Plus. Electron backscatter diffraction (EBSD) measurements were executed using a NordlysMax3 equipped with HKL Channel 5 software. The samples for EBSD investigation were electrolytically polished with a solution of 90% CH3CH2OH +10% HClO4 at 30 V for 15 s. Transmission electron microscope (TEM) was employed to examine the morphology of the δ phase and microstructure on a JEOL JEM-F200 at 200 kV.

Results

Characterization of superplastic mechanical properties

It is shown in that the stress increases at first and then decreases with the increase of strain. The maximum flow stress is below 140 MPa, which proves low flow stress in superplastic forming. shows that the elongation exceeds 430% at all strain rates. It proves that the rolled sheets with a dispersed distribution of the δ phase exhibit excellent superplasticity. At a strain of 0.3, the strain-rate sensitivity index (m value) is 0.368, calculated from the true stress and true strain, as shown in . This indicates GBS is the main deformation mechanism. As depicted in , the true stress-strain curves with strains of 0.41, 0.92, and 1.88 coincide, which proves the reliability of the intermediate data.

Figure 3. (a) True stress-strain curves at 950°C and different strain rates. (b) Relationship between the elongation and strain rate (c) Plot of logarithmic stress- logarithmic strain rate. (d) True stress-strain curves at different strains at a strain rate of 10−3 s−1.

Figure 3. (a) True stress-strain curves at 950°C and different strain rates. (b) Relationship between the elongation and strain rate (c) Plot of logarithmic stress- logarithmic strain rate. (d) True stress-strain curves at different strains at a strain rate of 10−3 s−1.

Evolution of microstructure

shows the evolution of the δ phase under different conditions. shows that δ phase content of the tensile specimen at strain e = 0.41 shows a marked drop in comparison with initial CDC plates (). It illustrates that during superplastic tensile, the δ phase is significantly dissolved at a higher temperature. With the increase of strain, as shown in , the coarsening of the δ phase indicates that the deformation promotes the accumulation of the fine δ phase. In , during the static annealing, the δ phase is dissolved without obvious growth.

Figure 4. SEM image of the deformation zone of the alloy sheet with a strain of (a) 0.41, (b) 0.92, and (c)1.88 at 950°C and a strain rate of 1 × 10−3 s−1. (d) SEM image of the gripping zone.

Figure 4. SEM image of the deformation zone of the alloy sheet with a strain of (a) 0.41, (b) 0.92, and (c)1.88 at 950°C and a strain rate of 1 × 10−3 s−1. (d) SEM image of the gripping zone.

shows the volume fraction and average size of the δ phase for different treatment conditions. It can be quantified from that comparing the CDC sheets, the volume fraction of the δ phase within the tensile specimens all decreases significantly due to the high temperature. In the deformation area, the content of the δ phase grew slightly with the increase of deformation, while δ phase content for the static annealing decreases more apparently. shows that the size of the δ phase gradually increases with the increase of deformation, which maximumly reached 0.54 μm at a true strain of 1.88. However, under the static annealing condition, the size of the δ phase is still 0.28 μm, which means that deformation promotes the aggregation of the δ phase while also inhibiting its dissolution.

Figure 5. (a) Volume fraction and (b) average size of the δ phase at different deformation conditions.

Figure 5. (a) Volume fraction and (b) average size of the δ phase at different deformation conditions.
shows the microstructure of the CRC sheets under different deformation conditions. shows that most of the grains in the CRC sheet have recrystallized after 5 min of preheating. High-density dislocations generated during the two stages of cold rolling provide a large amount of stored energy, which incr6a, bs the number and rate of recrystallization nucleation. The red box shows that a few coarse deformed grains are still unannealed. Comparing , it is found that the deformed grain undergoes DRX during superplastic deformation and transforms into an annealed structure without lattice distortion. shows the gradual growth of grains during the superplastic stretching. For the specimen with a strain of e = 1.88, the size of the grain in the deformation zone is larger than that in the static annealing zone. This suggests that deformation promotes grain growth.

Figure 6. Kernel average misorientation (KAM) graphs the deformation zone with strains of (a) 0, (b) 0.41, (c) 0.92, and (d) 1.88 under 950°C at a strain rate of 1×10−3 s−1. (e) KAM image of the gripping zone (e = 1.88).

Figure 6. Kernel average misorientation (KAM) graphs the deformation zone with strains of (a) 0, (b) 0.41, (c) 0.92, and (d) 1.88 under 950°C at a strain rate of 1×10−3 s−1. (e) KAM image of the gripping zone (e = 1.88).

shows the distribution of grain boundaries and the average grain size under different deformation conditions. shows that the average misorientation angle increases and then tends to be stable with an increased strain rate. DRX occurs at the early stage of deformation, resulting in the transformation of low-angle grain boundaries (LAGBs) to high-angle grain boundaries (HAGBs). shows that the grain size increases gradually with increasing deformation and reaches a maximum of 2.56 μm at strain e = 1.88. However, in the region of static annealing, the grains grew slightly. The results suggest that the deformation promotes grain enlargement.

Figure 7. (a) the average misorientation angle and (b) the average size of grain in different conditions.

Figure 7. (a) the average misorientation angle and (b) the average size of grain in different conditions.

shows that after 5 min of preheating, a small amount of deformed structure and the granular δ phase precipitated during the CDC process are present in the alloy and that some of the grains have undergone DRX. For the specimen with a strain of 0.41, it is observed in that the dense δ phase particles are distributed around the grain boundaries of the recrystallized grains, thus acting as pinning of the grain boundaries and limiting grain growth. Moreover, the matrix has been completely transformed into a recrystallized structure without lattice distortion.

Figure 8. TEM bright-field images with strains of (a) 0, and (b) (c) 0.41 of the deformation zone.

Figure 8. TEM bright-field images with strains of (a) 0, and (b) (c) 0.41 of the deformation zone.

Discussion

The effect of the δ phase

The shear bands and dislocations formed in the alloy during the first cold rolling process provide a position for the nucleation of the particles, thus promoting δ phase precipitation [Citation9]. The acicular δ phase in the CD plate () is transformed from the pre-precipitated γ″ phase in the matrix at 890°C. In the second cold rolling process with a reduced rate of 16%, the fully precipitated δ phase particles are crushed and uniformly distributed in the matrix [Citation10] ().

In the early stage of stretching, most of the alloy structure is transformed into recrystallized grains (). Compared to the preheat structure of Inconel 718 sheets obtained by direct large deformation cold rolling [Citation3], the preheat structure obtained in the process designed in this paper has a significantly higher proportion of recrystallized grains. It is suggested that the homogeneous and fine δ phase which precipitated in the previous process promotes recrystallization nucleation during the preheating [Citation11]. During a brief period of warming and preheating, most of the deformed structures are transformed into equiaxed grains (). This state of structure facilitates the GBS during subsequent deformation [Citation12].

Grain growth during superplastic deformation includes both static and dynamic growth. As a result of the homogeneous and fine δ phase pins the grain boundaries and limits the large-scale migration of the grain boundaries (), the grain size does not change significantly at the initial stage of deformation (). In the later stages of superplastic deformation, a significant increase in grain size occurs (), because of the decrease in the pinning effect due to the increase in size and decrease in the amount of the δ phase () [Citation13]. During superplastic tensile processes, the δ phase coarsens as a result of interfacial plasticity-induced processes, which rapidly reduces the cohesiveness of the second phase and the substrate [Citation14].

Mechanism of superplastic deformation

In the early stage of deformation, most of the grains have completed DRX, and relatively small equiaxed crystals are formed (). The redistribution of the δ phase reflects grain rotation and GBS. Kim [Citation15] pointed out that when the grain size is small, the main mechanism of superplastic deformation shifts to a GBS mechanism regulated by diffusion. The m value greater than 0.3 obtained in this study also reflects the occurrence of GBS. It is difficult for GBS to carry out all deformation alone, and a coordination mechanism is needed [Citation12]. It is observed that there are still a few deformed grains that have not been recrystallized (). A large number of dislocations exist in the deformed grains (), which provides a channel for atomic diffusion [Citation16]. The diffusion of atoms in grains serves as a way of superplastic deformation. In addition, LAGBs are consumed by recrystallized grains as deformed substructures (), DRX occurs in the alloy [Citation17], while the nature of DRX is a thermal diffusion process [Citation11]. Li et al. [Citation18] pointed out that relatively large size grains have reduced grain boundary density and the contribution of GBS to the deformation mechanism is not significant. As a result, the coarse deformed structure with irregular grain boundary shape achieves superplasticity of the alloy by atomic diffusion.

In the late stage of superplastic stretching, all the grains have completed recrystallization during the deformation process, which provides convenient conditions for GBS. Moreover, the obvious grain growth phenomenon () proves that the exchange of substances between grain boundaries is more frequent. The additional channel of grain boundary diffusion accelerates the transport of material between two adjacent grains, thus promoting grain growth [Citation19]. The average grain size in the deformation zone is significantly larger than that in the static annealing zone (). This indicates that the superplastic deformation accelerates the grain boundary migration [Citation20] and the material exchange between grain boundaries [Citation21]. Therefore, the late deformation mechanism of superplastic deformation is the GBS regulated by grain boundary diffusion.

Conclusion

In this paper, a rolled Inconel 718 alloy plate with a dispersed distribution of δ was prepared and the superplasticity of the plate is characterized. Several important conclusions are as follows.

  1. The superplastic elongation of the CDC plate exceeds 400% as strain rates of 6 × 10−4 to 4 × 10−3 s−1 at 950°C. The maximum elongation is 561% at a strain rate of 1 × 10−3 s−1 and the m value is about 0.367.

  2. At the early stage of deformation, the alloy grains do not grow significantly due to the pinning effect of the δ phase. At the later stage, the coarsening of the δ phase is caused by grain rotation during deformation induction and leads to grain growth, which was the main reason for the fracture of the plate.

  3. The deformed grains at the early stage realize the superplasticity through atomic diffusion, and the fine equiaxed grains realize the superplasticity through GBS. During the superplastic tensile process, DRX occurs in the deformed grains, and the deformation mechanism at a later stage transforms into GBS with a grain boundary diffusion coordination.

Acknowledgments

The authors would like to thank the State Key Laboratory of Rolling and Automation (RAL) for their assistance and support.

Disclosure statement

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

  • Jafariana HR, Mousavi Anijdan SH, Miyamotoc H. Observation of low-temperature superplasticity in an ultrafine grained AA6063 alloy. Mater Sci Eng A. 2020;795:140015.
  • Sato E, Kuribayashi K. Superplasticity and deformation induced grain growth. ISIJ Int. 1993;33(8):825–7.
  • Yang X, Chen SN, Wang BX, et al. Superplastic deformation behavior of cold-rolled Inconel 718 alloy at high strain rates. J Mater Process Technol. 2022;308:117696.
  • Lv SM, Jia CL, He XB, et al. Superplastic deformation and dynamic recrystallization of a novel disc superalloy GH4151. Materials. 2019;12(22):3667.
  • Yakovtseva OA, Sitkina MN, Kotov AD, et al. Experimental study of the superplastic deformation mechanisms of high-strength aluminum-based alloy. Mater Sci Eng A. 2020;788:139639.
  • Hidalgo-Manrique P, Orozco-Caballero A, Cepeda-Jiménez CM, et al. Influence of the accumulative roll bonding process severity on the microstructure and superplastic behaviour of 7075 AL alloy. J Mater Sci Technol. 2016;32(8):774–782.
  • Páramo-Kañetas PJ, Ozturk U, Calvo J, et al. Analysis of strain-induced precipitates by delta-processing in Inconel 718 superalloy. Mater Charact. 2021;173:110926.
  • Xue H, Zhao JQ, Liu YK, et al. δ-phase precipitation regularity of cold-rolled fine-grained GH4169 alloy plate and its effect on mechanical properties. Trans Nonferrous Met Soc China. 2020;30(12):3287–3295.
  • Singh RP, Hyzak JM, Howson TE, et al. Recrystallization behavior of cold rolled alloy 718. Superalloys. 1991;718:205–215.
  • Páramo-Kañetas P, Özturk U, Calvo J, et al. High-temperature deformation of delta-processed Inconel 718. J Mater Process Technol. 2018;255:204–211.
  • Chen MS, Zou ZH, Lin YC, et al. Microstructural evolution and grain refinement mechanisms of a Ni-based superalloy during a two-stage annealing treatment. Mater Charact. 2019;151:445–456.
  • Masuda H, Sato E. Diffusional and dislocation accommodation mechanisms in superplastic materials. Acta Mater. 2020;197:235–252.
  • Chen MS, Zou ZH, Lin YC, et al. Effects of annealing parameters on microstructural evolution of a typical nickel-based superalloy during annealing treatment. Mater Charact. 2018;141:212–222.
  • Jouiad M, Marin E, Devarapalli RS, et al. Microstructure and mechanical properties evolutions of alloy 718 during isothermal and thermal cycling over-aging. Mater Des. 2016;102:284–296.
  • Kim WJ. Explanation for deviations from the hall–petch relation based on the creep behavior of an ultrafine-grained Mg–Li alloy with low diffusivity. Scripta Mater. 2009;61(6):652–655.
  • Sergueeva AV, Mara NA, Mukherjee AK. Grain boundary sliding in nanomaterials at elevated temperatures. J Mater Sci. 2007;42(5):1433–1438.
  • Huang LJ, Feng Q, Hua PT, et al. Discontinuous dynamic recrystallization of inconel 718 superalloy during the superplastic deformation. Metall Mater Trans A. 2015;46(9):4276–4285.
  • Li MJ, Pan QL, Shi YJ, et al. High strain rate superplasticity in an Al–Mg–Sc–Zr alloy processed via simple rolling. Mater Sci Eng A. 2017;687:298–305.
  • Kim BN, Hiraga K, Sakka Y, et al. A grain-boundary diffusion model of dynamic grain growth during superplastic deformation. Acta Mater. 1999;47(12):3433–3439.
  • Wilkinson DS, Caceres CH. On the mechanism of strain-enhanced grain growth during superplastic deformation. Acta Metall. 1984;32(9):1335–1345.
  • Valiev RZ, Kaibyshev OA, Khannanov SK. Grain boundaries during superplastic deformation. Phys Status Solidi A. 1979;52(2):447–453.