Next Article in Journal
Advanced Mass Spectrometric Techniques for the Comprehensive Study of Synthesized Silicon-Based Silyl Organic Compounds: Identifying Fragmentation Pathways and Characterization
Previous Article in Journal
A Review of Numerical Simulation of Laser–Arc Hybrid Welding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simulation of Work Hardening in Machining Inconel 718 with Multiscale Grain Size

1
Hubei Digital Manufacturing Key Laboratory, School of Mechanical and Electronic Engineering, Wuhan University of Technology, Wuhan 430070, China
2
State Key Laboratory of Intelligent Manufacturing Equipment and Technology, Huazhong University of Science and Technology, Wuhan 430074, China
3
Xiamen Golden Egret Special Alloy Corporation Limited, Xiamen 361000, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(9), 3562; https://doi.org/10.3390/ma16093562
Submission received: 21 March 2023 / Revised: 26 April 2023 / Accepted: 4 May 2023 / Published: 6 May 2023

Abstract

:
Machining nickel-based alloys always exhibits significant work-hardening behavior, which may help to improve the part quality by building a hardened layer on the surface, while also causing severe tool wear during machining. Hence, modeling the work-hardening phenomenon plays a critical role in the evaluation of tool wear and part quality. This paper aims to propose a numerical model to estimate the work-hardening layer for a deeper understanding of this behavior, employing both recrystallization-based and dislocation-based models to cover workpieces with multiscale grain sizes. Different user routines are implemented in the finite element method to simulate the increase in hardness in the deformed area due to recrystallization or changes in the dislocation density. The validation of the proposed model is performed with both literature and experimental data for Inconel 718 with small or large grain sizes. It is found that the recrystallization-based model is more suitable for predicting the work-hardening behavior of small-grain-size Inconel 718 and the dislocation-based model is better for that of large-grain-size Inconel 718. Further, as an important type of cutting tool in machining Inconel 718, the chamfered tools with different edge geometries are employed in the simulations of machining-induced work hardening. The results illustrate that the uncut chip thickness and chamfer angle have a significant influence on the work-hardening behavior.

1. Introduction

Inconel 718 is an essential superalloy, commonly utilized in the aerospace industry (e.g., gas turbines), as well as in the shipbuilding, petrochemical, and nuclear industries due to its exceptional corrosion resistance and the ability to maintain its physical properties even under high temperatures [1,2,3]. Nevertheless, the machinability of Inconel 718 is poor due to severe work-hardening behavior during the cutting process [4]. A certain degree of work hardening (DWH) will significantly enhance the strength and wear resistance of the workpiece surface, thereby improving the performance of the workpiece. However, excessive work hardening can lead to drastic changes in the metal’s microstructure, increasing both the cutting force and wear on the tool, which makes continuous cutting challenging. Therefore, studying the phenomenon of work hardening is imperative.
Numerous experimental studies have been conducted to examine the influence of cutting parameters, tool geometries, and cutting conditions on work hardening from a macroscopic point of view. According to research conducted by Hua and Liu [5], a higher cutting speed leads to a higher DWH, with only a slight difference in the depth of the hardening layer. A greater feed rate leads to a higher DWH and a deeper hardening layer. Umbrello [6] discovered that increasing the cutting speed and feed rate resulted in a higher surface hardness and deeper hardening layer on the workpiece, accompanied by a noticeable refinement of the grain structure. Dinesh et al. [7] conducted a study on the effects of cutting speed, feed rate, and cutting depth on the DWH and found that the lowest DWH occurs under conditions of a low cutting speed and feed rate. Pawade [8] utilized Taguchi orthogonal design to thoroughly investigate the influence of the cutting parameter and tool geometry on work-hardening behavior in the cutting process of Inconel 718. The research indicated that both the tool geometry and cutting depth of the tool had the most significant influence on DWH. The depth of the hardening layer was found to be greatest when using a tool with a chamfer angle (i.e., the angle between the chamfer and the tool rake face plane). However, increasing the chamfer angle resulted in a decrease in the depth of the hardening layer.
It is widely recognized that microstructure differences between the machined surface of the workpiece and its subsurface are caused by different machining parameters. Numerical methods are widely employed to simulate this behavior because of their high prediction accuracy and great ability to capture intricate in-process physical variables. Two primary theories of simulation studies on work hardening are recrystallization and dislocation density. For the recrystallization theory, Jafarian et al. [9,10] proposed a model based on the hardness of the material, combined with the model based on the Zener–Hollomon parameter (Z–H model) to study the influence of tool micro-edges including the edge radius and chamfer angle for Inconel 718 microstructure evolution during the cutting process, especially the effect of grain size and hardness. Caruso et al. [11] studied the microstructure evolution of the SAE 8620 cutting process by combining the hardness-based Johnson–Cook (J–C) material constitutive model and the Z–H model. The results indicate that the average prediction error for grain size is approximately 20%, while for surface hardness, it is 4%. Arisoy et al. [12,13] combined the improved temperature-based flow softening material model and the three-dimensional finite element (FE) model and used the recrystallization model based on Johnson–Mehl–Avrami–Kolmogorov (JMAK) to study the microstructure changes of the machined surfaces of Inconel 100 alloy and titanium alloy Ti6Al4V, including the changes in phase distribution, volume fraction, and hardness, during the cutting process. Xu et al. [14] predicted the grain size and work hardening of Ti6Al4V during the cutting process by using the recrystallization model based on JMAK, and studied the grain refinement and work hardening caused by dynamic recrystallization through the microstructure evolution of titanium alloy. When it comes to the dislocation density theory, Ding et al. [15] used the dislocation density-based model (DDB model) to study the cutting speed, feed rate, and tool rake angle on dislocation density, grain size, hardness, and metamorphic layer thickness of Al6061–T6 alloy. A low cutting speed, high feed rate, and negative tool rake angle contributed to a higher dislocation density, thinner deformation layer, and smaller grain size, resulting in improved hardness of the machined surface. Li et al. [16] used the DDB model to study the effect of the cutting speed, feed rate, and radial cutting depth on the grain size evolution of the H13 steel surface during the milling process. With the increase in the cutting speed, feed rate, and radial cutting depth, the grain size of the machined surface changes slightly and gradually returns to its initial size. Liu et al. [17] simulated the cutting process of OFHC copper using the DDB model and concluded from the experiment and simulation results that the fluctuation of a cutting force at high cutting speed was caused by the evolution and distribution of the dislocation density. The dislocation density on the machined surface increased with the increase in cutting speed, while it decreased in the chips.
The different grain sizes of materials under the same processing conditions will show differences in the cutting force, surface hardness, and surface integrity. Komatsu et al. [18,19] researched the effect of stainless-steel grain size on the micro-milling process. The study found that the formation of burrs can be controlled by reducing the grain size. Compared with ordinary grain steel, when cutting ultra-fine-grain steel, the shear force is small, and the surface finish is markedly improved. The findings of Olovsjö et al. [20] demonstrated that large burrs only occurred in large grain size materials of the wrought Alloy 718 during cutting. Meanwhile, the depth of the deformation layer of the large-grain-size material was greater than the small-grain-size material. Yang et al. [21] studied the effect of grain size on the surface characteristics of pure iron during orthogonal cutting and found that the residual stress on the machined surface decreased with an increase in grain size. In addition, the hardness of small-grain-size material is greater than that of large-grain-size material. Wu et al. [22] studied the influence of grain size on cutting force and burr formation in the copper micro-cutting process and found that small grain size is beneficial for reducing cutting force and burrs. In the study of the process of tool wear, Polvorosa et al. [23] found that the small grain size of the material resulted in larger flank wear during the milling of nickel-based superalloys. However, few researchers have studied the influence of grain size on work-hardening behavior in the cutting process.
In this paper, an FE model of orthogonal cutting is established. Moreover, the Z–H model based on recrystallization theory and the DDB model based on dislocation theory are combined to systematically simulate the work-hardening behavior of Inconel 718 with different grain sizes and discuss the applicability of different models. Finally, a suitable model for large-grain-size Inconel 718 is used to study the effects of uncut chip thickness and tool edge geometries (the chamfer angle and chamfer length) on work-hardening indicators (DWH, surface hardness, and hardening layer depth).

2. Overview of the Work

This paper focuses on the simulations of the work-hardening phenomenon when cutting Inconel 718 with multiscale grain size, and the overview is shown in Figure 1. First, multiple cutting simulations are performed based on DEFORM. Then, two different models (Z–H and DDB models) are employed to predict the hardness in the subsurface. The model constants will be calibrated until the maximum prediction error is obtained. For Inconel 718 with a large grain size, orthogonal cutting experiments are conducted, and the hardness in the subsurface area is measured by nanoindentation. Further, the microstructure images are taken for model validation. For Inconel 718 with a small grain size, literature data from reference [6] are used to compare the performance of the two models. Then, the predicted results given by the Z–H and DDB models are compared, and their performances are discussed. Finally, the effect of the uncut thickness and tool edge geometries (the chamfer angle and chamfer length) on the work-hardening behavior of Inconel 718 with a large grain size is further investigated using the DDB model.

3. Simulation and Experimentation

3.1. FE Models of the Orthogonal Process

DEFORMTM v11.0 developed by Scientific Forming Technologies Corporation (SFTC, Columbus, OH, USA) was employed to simulate the orthogonal cutting process of Inconel 718. The plane strain coupling thermos-mechanical analysis was carried out under the orthogonal assumption. During the simulation, the updated Lagrange code and remeshing technique were adopted to realize the thermo-steady-state and mechanical-steady-state conditions. Very fine elements were defined near the cutting region (1 μm) to obtain more precise results and better chip geometry shape.
The mesh and boundary conditions of the FE model are shown in Figure 2. The tool was deprived of all displacements, and the workpiece could move only in the X direction. The bottom and left of the workpiece and the top and right of the tool were set to room temperature (Troom). The top and right of the workpiece and the bottom and left of the tool (marked in red) were allowed to exchange heat with the environment. During the cutting process, the heat generated in the cutting zone was transferred to the workpiece, tool, chips, and environment [24]. A very high heat transfer coefficient of 105 kW/(m2K) [25] was selected between the chip, tool, and workpiece so that the temperature field could reach a stable state in a short time. The hybrid friction model regarding the chip–tool and tool–workpiece interface was considered. Specifically, the shear friction factor was employed in the sticking contact area, and the coulomb friction factor was employed in the sliding area. In this model, the shear friction factor and coulomb friction factor were set to 1 and 0.3, respectively [25].
In this paper, Cockroft and Latham’s damage criterion [26] was considered to simulate the formation of serrated chips during orthogonal cutting. The criterion is expressed as:
0 ε f σ 1 d ε = D
where εf is the effective strain, σ1 is the maximum principal stress, and D is the critical damage value. When the integral of the largest tensile principal stress of any element reaches the D value, the fracture will occur, and chip segmentation will start. In this model, D = 120.
The flow stress behavior was defined using the J–C material constitutive model, which describes the plastic deformation of materials at different ranges of strain, strain rate, and temperature. The model is illustrated as follows:
σ = A + B ε n 1 + C ln ε ˙ ε ˙ 0 1 T T r o o m T m e l t T r o o m m
where σ is the flow stress, ε is the plastic strain, ε ˙ is the strain rate, ε ˙ 0 is the reference strain rate, T (°C) is the workpiece temperature, Troom is the room temperature (20 °C), and Tmelt is the melting temperature of the material (1300 °C for Inconel 718). A, B, C, m, and n are J–C material constitutive model parameters, where A (MPa) is the yield strength, B (MPa) is the hardening modulus, C is the strain rate sensitivity coefficient, m is the thermal softening coefficient, and n is the hardening coefficient. The J–C material constitutive model relevant parameters are listed in Table 1.

3.2. FE Models of Microstructure Changes

3.2.1. The Z–H Model

The FE simulation software DEFORMTM v11.0 can simulate the macroscopic deformation and microstructure evolution of the metal-forming process and provide the secondary development subroutine interface. A user subroutine for the Z–H model was implemented in this software to predict microstructure changes such as recrystallization grain size and hardness. In this paper, the Zener–Hollomon equation (Equations (3) and (4) is employed to predict recrystallization [28], and the Hall–Petch equation (Equation (5)) is employed to calculate the hardness change [29]. A more detailed description of these equations is as follows:
Z = ε ˙ exp Q R T
d = b Z m
H V = C 0 + C 1 d 0.5
where Z is the Z–H parameter, ε ˙ is the strain rate, Q is the apparent activation energy (368.5 kJ/mol) [25], R is the universal gas constant (8.3145 J/K·mol) [25], T is the temperature (K), d is the recrystallized grain size, and HV is the value of the Vickers hardness. In addition, b, m, C0, and C1 are material constants, which will be given in Section 3.2.3. When the strain generated on the material surface during cutting exceeds the critical strain, the recrystallization process begins. The critical strain εcr was first proposed by Sui et al. [30] (for the hot continuous rolling process of Inconel 718 alloy), and then Jafarian et al. [25] calibrated the formula for the cutting process by adding a constant “c”. The value of critical strain εcr is related to strain rate ε ˙ and temperature T. The final critical strain εcr is represented by the following formula:
ε c r = 0.00234 ε ˙ 0.1293 exp 5729.863 T + 273 / c
where εcr is the critical strain, ε ˙ is the strain rate, T is the temperature (°C), and c is the calibration constant.
The implemented strategy for the prediction of hardness variation during the orthogonal cutting process of Inconel 718 is shown in Figure 3. In each time step of the simulation, the critical strain εcr of all elements is calculated and compared with the corresponding strain obtained from the FE simulation. When the strain exceeds the critical strain εcr, recrystallization occurs, and the hardness changes. As a result, the hardness value is updated. Otherwise, the hardness will not change.

3.2.2. The DDB Model

A user subroutine for the DDB model is implemented in DEFORMTM v11.0 software to predict microstructure evolution. Estrin et al. [31] first introduced the model to study the grain refinement of copper during isotropic corner extrusion, and then Ding et al. [32,33] applied the model to predict the evolution of the material dislocation density, chip formation, and grain refinement in the cutting process for the titanium alloy. In this model, the dislocations are generated by the plastic deformation of the workpiece, and there are two types of dislocation densities, namely cell interior dislocation density (ρc) and cell wall dislocation density (ρw), which can be further subdivided into geometrically necessary dislocation density (ρwg) and statistical dislocation density (ρws). The relevant dislocation density evolution rate equations are as follows:
ρ ˙ c = α * 1 3 b ρ w γ ˙ w r β * 6 b d ( 1 f ) 1 / 3 γ ˙ c r k 0 γ ˙ c r γ ˙ 0 r 1 / n ρ c γ ˙ c r
ρ ˙ w s = β * 3 ( 1 f ) f b ρ w γ ˙ c r + ( 1 ξ ) β * 6 ( 1 f ) 2 / 3 b f d γ ˙ c r k 0 γ ˙ w r γ ˙ 0 r 1 / n ρ w s γ ˙ w r
ρ ˙ w g = ξ β * 6 ( 1 f ) 2 / 3 b f d γ ˙ c r
ρ ˙ w = ρ ˙ w s + ρ ˙ w g = β * 3 ( 1 f ) f b ρ w γ ˙ c r + β * 6 ( 1 f ) 2 / 3 b f d γ ˙ c r k 0 γ ˙ w r γ ˙ 0 r 1 / n ρ w γ ˙ w r
The three terms on the right side of Equations (7) and (10) represent the contributions of different mechanisms to the dislocation density. The first one is the dislocation growth caused by the Feank–Read dislocation source, the second one is the dislocation transformation from interior dislocations to cell wall dislocations due to plastic deformation during the processing, and the third one is the dislocation annihilation caused by thermal dynamic reversion at high temperatures. Parameter ξ is the proportion of geometrically necessary dislocation density in the cell wall dislocation density, and 𝑏 is the Burgers vector, γ ˙ 0 r is the reference shear strain rate. α*, β*, and k0 are parameters to control the evolution rate of dislocation density, n is the parameter of temperature sensitivity, f is the volume fraction of the cell wall dislocation density, and the parameters of n, f, and d are reported in the following equations:
k 0 = 0.016 T + 0.4
n = B / T
f = f + ( f 0 f ) e γ r / γ ˜ r
where B is a constant, T is the temperature (K), γ r is the shear strain, γ ˜ r is the reference shear strain, and f0 and f are the proportions of the initial and final dislocation density in the total dislocation density, respectively.
Parameters γ ˙ w r and γ ˙ c r are the shear strain rates of the cell wall and cell interior, respectively. In order to satisfy the strain compatibility, the following equations are used to describe the shear strain rates:
γ ˙ w r = γ ˙ c r = γ ˙ r
γ ˙ r = M ε ˙
where M is the Taylor factor and ε ˙ is the equivalent strain rate.
The final grain size d is related to the total dislocation density ρtot.
ρ t o t = f ρ w + ( 1 f ) ρ c
d = K / ρ t o t
where K is the material coefficient. The strengthening of material hardness depends on the dislocation density due to severe plastic deformation (SPD). The relationship between the hardness (ΔHVSPD) of the material and the dislocation can be added to the current model, and the equation of hardness (ΔHVSPD) strengthening [34] is as follows:
Δ H V S P D = k h M a G b ρ t o t
where kh and a are constants. M is the Taylor factor, G is the shear modulus, and b is the Burgers vector. In this paper, a constant µ is defined as:
μ = k h M a G b
The flow chart of the implemented strategy for the prediction of hardness variation during the orthogonal cutting process of Inconel 718 is shown in Figure 4. At each step of the simulation, the current time is compared to the predetermined total simulation time. If the current time exceeds the set total simulation time, the subroutine will stop executing the calculation and deliver the result; otherwise, the subroutine will continue to execute.

3.2.3. Models Calibration

To ensure accurate predictions, a calibration phase is required to define material constants using the experimental results. Therefore, conducting a sensitivity analysis to understand the influence of individual parameters on the hardness is crucial.
For the Z–H model, FE simulation analysis reveals that the surface hardness and DWH vary significantly when parameter b is changed; however, the change in the work-hardening layer depth is not significant. Parameter c primarily influences the work-hardening layer depth. To begin the calibration, the same value utilized by Jafarian [25] is chosen as the starting value for the sensitivity analysis since these values are good predictors of Inconel 718 surface hardness. Parameters m, C0, and C1 are set at 0.01, 378, and 298.4, respectively, and b is a function of uncut chip thickness, as follows:
b = 600 h 2 139.8 h + 10.2
Subsequently, calibration is conducted to acquire the suitable value for parameter c. For different initial grain sizes of Inconel 718, parameter c is set to different values: 100 and 1000 for the small grain size (18 μm) and large grain size (115 μm), respectively.
For the test condition of h = 0.05 mm, the calibration results of hardness variation near the machined surface can be seen in Figure 5. In the machined surface and subsurface, the hardness has a significant change. The maximum hardness is in the range of 10~20 μm from the machined surface, which is not only affected by the mechanical load but also by the thermal load to a substantial extent. Along the depth direction from the machined surface, the influence of thermal load gradually decreases, but the hardness in the range of 50 μm from the machined surface is also significantly larger than the bulk hardness. The hardness variation curve at the Q point of the machined surface along the depth direction decreases rapidly from the machined surface to 80 μm and begins to converge at approximately 115 μm, which is in good agreement with the experimentally measured hardening layer depth of 104.5 μm [6,25].
A comparison between the predicted and experimental results for the three test conditions (h = 0.05, 0.075, and 0.1 mm) is shown in Figure 6. Very good agreement is observed between the predicted and experimental results for the hardness of the machined surfaces for all test conditions. In terms of the depth of the hardening layer, the average error is 5.17%, and the simulation result is generally good, although under some test conditions (1 out of 3 experiments), the prediction result is higher than the corresponding experimental result (13.6% error). The overall average error (Figure 7) of the machined surface hardness and depth of the hardening layer fluctuates between 2.05% and 7.9%.
The DDB model is adapted to the cutting process of the Inconel 718 by modifying the important parameters of the model.
The microstructure pictures (Figure 8) of the cross-section view of the workpiece are obtained via Scanning Electron Microscopy (SEM) (Zeiss, Germany), using test 7 (cutting speed vc = 50 m/min, uncut chip thickness h = 0.08 mm, chamfer length L = 100 μm, chamfer angle θ = 15°) as an example. The results indicate that slip lines can be observed in the subsurface area within the range of 20 µm under the machined surface, which means an increase in the dislocation density.
Under test 7 conditions, the predicted dislocation density, grain size, and surface hardness are shown in Figure 9. The prediction results show that there is an obvious deformation layer along the depth direction from the machined surface, which is consistent with the SEM pictures (Figure 8). The dislocation density and hardness are the highest at the machined surface and gradually decrease along the depth direction of the machined surface until they are equal to the bulk material. The grain size variation trend is opposite to the hardness variation.
In the paper by Hu et al. [35], the DDB model parameters for the shot-peened process of Inconel 718 are given. On this basis, some of the parameters are modified for the cutting process of Inconel 718 in this paper.
The specific parameters for the model are listed in Table 2. Specifically, the parameter K and the previously defined constant µ have different values for different grain sizes of Inconel 718, and for different initial grain sizes of Inconel 718, parameters K and µ are set to different values. For small-grain-size Inconel 718, the parameters K and µ are set to 119 and 0.0153, respectively. For large-grain-size Inconel 718, parameters K and µ are set to 769 and 0.03, respectively.

3.3. Experimentation

The accuracy of the outcomes obtained by the FE method heavily depends on the accuracy of the input values after calibration. Thus, to obtain an accurate prediction of work-hardening phenomena, a suitable experimental scheme needs to be devised.
Experimental tests for the large-grain-size Inconel 718 under dry cutting conditions on a CAK5058nzj CNC lathe were performed as shown in Figure 10. The uncoated carbide triangle blades with a chamfer angle are employed to provide a 0° rake angle and a 6° clearance angle in the orthogonal cutting experiment. The blade is mounted on a Sandvik shank. The workpiece dimensions are 50 mm × 25 mm × 3 mm (length × width × height) (Figure 10c) and the material is the large-grain-size Inconel 718 with an average Vickers hardness of 345 and an average grain size of 115 μm. The related parameters of the cutting experiment are listed in Table 3 test 1~9. After processing, the workpiece is cut into 10 mm × 10 mm × 3 mm (length × width × height) small samples. The sample surface is made smooth by grinding and polishing operations and then preserved by pressing the samples into the toner using a hot press inlay machine. The NanoTest Vantage4 is employed for nanoindentation measurements with a Berkovich indenter, applying a fixed maximum load of 50 mN in this study. We measured 75 points in total along the depth direction from the machined surface, with a spacing of 20 μm between every individual indentation, and 5 indentations are set for a specific depth value, as shown in Figure 10a,b.
For small-grain-size Inconel 718, testing conditions are conducted at the same cutting speeds (vc = 70 m/min) but different uncut chip thicknesses (h = 0.05, 0.075, and 0.1 mm). These are literature data from D. Umbrello [6]. The workpiece material has an average hardness value of 429 HV and an average grain size of 18 μm.

4. Results and Discussions

4.1. Two Different Models Predict Results

The Z–H model and DDB model are employed to predict the work-hardening behavior of Inconel 718, considering the small grain size (18 μm) and large grain size (115 μm).
For small-grain-size Inconel 718, FE simulations are conducted under two models for three experimental conditions (h = 0.05, 0.075, and 0.1 mm), respectively, and the fitting results of the simulated result and experimental data are shown in Figure 11.
The simulation results of both the Z–H model and DDB model can fit the experimental data of small-grain-size Inconel 718 (Figure 12), which indicates that the microstructure models based on recrystallization and dislocation can be used to explain the work-hardening phenomenon of small-grain-size Inconel 718. However, the Z–H model, based on recrystallization, is more suitable for small-grain-size Inconel 718.
For the large-grain-size Inconel 718, FE simulations were performed under four test conditions (tests 1, 2, 4, and 7) in Table 3 using the Z–H model and DDB model, respectively. The fitting of simulation results and experimental data for each of the four test conditions is shown in Figure 12.
The fitting of the predicted results and experimental data of the two models under the conditions of tests 1, 2, 4, and 7 is shown in Figure 12. It is apparent that the DDB model better fits the predicted results and experimental values. As can be seen under the optical microscope in Figure 13, grain refinement does not occur on either the machined surface or the subsurface, which indicates that recrystallization does not occur on the machined surface of the large-grain-size Inconel 718. The SEM pictures (Figure 8) show severe plastic deformation and obvious slip lines, which are reflections of dislocation densities. This further supports the conclusion that the DDB model, based on dislocation theory, is more suitable than the Z–H model, based on recrystallization theory, to explain the work-hardening phenomenon of large-grain-size Inconel 718.
After determining that the DDB model can better predict the large-grain-size Inconel 718, the effects of the uncut chip thickness, chamfer length, and chamfer angle on work-hardening behavior are studied by using the DDB model under different test conditions (test 1~9).

4.2. Analysis of Work-Hardening Behavior

Under different cutting parameters, tool chamfer angles, and chamfer lengths, the variation trends of work-hardening indices (DWH, surface hardness, and hardening layer depth) are shown in Figure 14.
The work-hardening characteristics of Inconel 718 are a limiting factor with poor machinability. As a result, the DWH needs to be controlled in the cutting of Inconel 718. Hence, in this study, the DWH calculated by Equation (21) illustrates the severity of the work-hardening behavior of the machined material. Equation (21) is described as follows [8,36]:
DWH ( % ) = H V s H V 0 H V 0 100 %
where HVs and HV0 represent the hardness of surface and bulk material, respectively.
For the DWH, the maximum value is obtained when the large uncut chip thickness (0.08 mm) and small chamfer angle (15°) are used, and the minimum value is acquired when the small uncut chip thickness (0.04 mm) and small chamfer angle (15°) are used. The DWH is positively correlated with the chamfer angle and uncut chip thickness. When the large chamfer angle (25°) is used, the DWH is negatively correlated with the chamfer length. When the small chamfer angle (15°) is used, the DWH increases at first and then decreases with the increase in the chamfer length. The result of variance analysis shows that the uncut chip thickness and chamfer angle have a significant influence on the DWH, with the uncut chip thickness exerting the most significant influence, followed by the chamfer angle.
For the surface hardness, its changing trend is consistent with the DWH. The most significant impact on the surface hardness is the uncut chip thickness, followed by the chamfer angle. A larger uncut chip thickness leads to a greater cutting force and more severe plastic deformation on the machined surface, resulting in a larger plowing area under the cutting edge and aggravating the work hardening.
For the hardening layer depth, the maximum value is obtained when a small uncut chip thickness (0.04 mm) and large chamfer angle (25°) are used, and the minimum value is acquired when a large uncut chip thickness (0.08 mm) and small chamfer angle (15°) are used. The hardening layer depth is negatively correlated with the uncut chip thickness and positively correlated with the chamfer angle. With a small uncut chip thickness (0.04 mm), the hardening layer depth is negatively correlated with the chamfer length. With a large uncut chip thickness (0.08 mm), the hardening layer depth is positively correlated with the chamfer length. The result of variance analysis shows that the uncut chip thickness has a significant influence on the depth of the hardening layer.

5. Conclusions

In this paper, the Z–H model and DDB model are used to predict the work hardening behavior of Inconel 718 with large or small grain sizes, respectively, during the orthogonal cutting process, and the applicability of these two models to workpieces of different grain sizes is discussed. Based on the DDB model, the effects of the uncut chip thickness, chamfer angle, and chamfer length on the work-hardening behavior of large-grain-size Inconel 718 are also discussed. The main contributions can be drawn as follows:
(1)
For the small-grain-size Inconel 718, the fitting effect of the Z–H model (i.e., recrystallization-based model) is better than the DDB model. For the large-grain-size Inconel 718, the predicted results of the DDB model are in better agreement with the experimental results.
(2)
The SEM pictures of the experiment show that there are slip lines in the surface area, but no obvious grain refinement is found. This indicates that the recrystallization-based model of work hardening is not suitable for the workpiece with a large-grain-size Inconel 718.
(3)
The DWH and the surface hardness are significantly affected by the uncut chip thickness and chamfer angle, among which uncut chip thickness is the most significant factor. With the increase in both the uncut chip thickness and chamfer angle, the surface hardness and DWH increase.
(4)
For the depth of the hardening layer, the uncut chip thickness has a significant influence. The depth of the hardening layer will decrease with increasing uncut chip thickness. In general, the uncut chip thickness has the most significant influence on the hardening behavior, followed by the chamfer angle and chamfer length.
In this study, the uncut chip thickness is relatively low (0.05, 0.075, and 0.1 mm). More work should be performed in the future under the conditions of a large uncut chip thickness where the work-hardening phenomenon is more significant.

Author Contributions

Conceptualization, J.W.; methodology, J.W.; formal analysis L.Z.; investigation, J.W. and C.F.; data curation, Z.W. and C.F.; writing—original draft preparation, C.F. and Z.W.; writing—review and editing J.W. and K.Z.; visualization, Z.W.; supervision L.Z.; project administration, K.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The State Key Laboratory of Intelligent Manufacturing Equipment and Technology (IMETKF2023001) and the Natural Science Foundation of Hubei Province (2021CFB304).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

All authors have reviewed this paper and agreed to submit it.

Data Availability Statement

Not applicable.

Acknowledgments

The authors appreciate the financial support from the State Key Laboratory of Intelligent Manufacturing Equipment and Technology (IMETKF2023001) and the Natural Science Foundation of Hubei Province (2021CFB304).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liang, X.; Liu, Z.; Wang, B. State-of-the-art of surface integrity induced by tool wear effects in machining process of titanium and nickel alloys: A review. Measurement 2019, 132, 150–181. [Google Scholar] [CrossRef]
  2. Wang, B.; Liu, Z. Influences of tool structure, tool material and tool wear on machined surface integrity during turning and milling of titanium and nickel alloys: A review. Int. J. Adv. Manuf. Technol. 2018, 98, 1925–1975. [Google Scholar] [CrossRef]
  3. Yin, Q.; Liu, Z.; Wang, B.; Song, Q.; Cai, Y. Recent progress of machinability and surface integrity for mechanical machining Inconel 718: A review. Int. J. Adv. Manuf. Technol. 2020, 109, 215–245. [Google Scholar] [CrossRef]
  4. Ulutan, D.; Ozel, T. Machining induced surface integrity in titanium and nickel alloys: A review. Int. J. Mach. Tools Manuf. 2011, 51, 250–280. [Google Scholar] [CrossRef]
  5. Hua, Y.; Liu, Z. Effects of cutting parameters and tool nose radius on surface roughness and work hardening during dry turning Inconel 718. Int. J. Adv. Manuf. Technol. 2018, 96, 2421–2430. [Google Scholar] [CrossRef]
  6. Umbrello, D. Investigation of surface integrity in dry machining of Inconel 718. Int. J. Adv. Manuf. Technol. 2013, 69, 2183–2190. [Google Scholar] [CrossRef]
  7. Thakur, D.G.; Ramamoorthy, B.; Vijayaraghavan, L. Effect of cutting parameters on the degree of work hardening and tool life during high-speed machining of Inconel 718. Int. J. Adv. Manuf. Technol. 2011, 59, 483–489. [Google Scholar] [CrossRef]
  8. Pawade, R.S.; Joshi, S.S.; Brahmankar, P.K. Effect of machining parameters and cutting edge geometry on surface integrity of high-speed turned Inconel 718. Int. J. Mach. Tools Manuf. 2008, 48, 15–28. [Google Scholar] [CrossRef]
  9. Jafarian, F.; Umbrello, D.; Jabbaripour, B. Identification of new material model for machining simulation of Inconel 718 alloy and the effect of tool edge geometry on microstructure changes. Simul. Model. Pract. Theory 2016, 66, 273–284. [Google Scholar] [CrossRef]
  10. Jafarian, F.; Masoudi, S.; Umbrello, D.; Filice, L. New strategies for improvement of numerical model accuracy in machining of nickel-based alloy. Simul. Model. Pract. Theory 2019, 94, 134–148. [Google Scholar] [CrossRef]
  11. Caruso, S.; Rotella, G.; Del Prete, A.; Umbrello, D. Finite element modeling of microstructural changes in hard machining of SAE 8620. Appl. Sci. 2019, 10, 121. [Google Scholar] [CrossRef]
  12. Arısoy, Y.M.; Guo, C.; Kaftanoğlu, B.; Özel, T. Investigations on microstructural changes in machining of Inconel 100 alloy using face turning experiments and 3D finite element simulations. Int. J. Mech. Sci. 2016, 107, 80–92. [Google Scholar] [CrossRef]
  13. Arısoy, Y.M.; Özel, T. Prediction of machining induced microstructure in Ti–6Al–4V alloy using 3-D FE-based simulations: Effects of tool micro-geometry, coating and cutting conditions. J. Mater. Process. Technol. 2015, 220, 1–26. [Google Scholar] [CrossRef]
  14. Xu, X.; Zhang, J.; Outeiro, J.; Xu, B.; Zhao, W. Multiscale simulation of grain refinement induced by dynamic recrystallization of Ti6Al4V alloy during high speed machining. J. Mater. Process. Technol. 2020, 286, 116834. [Google Scholar] [CrossRef]
  15. Ding, L.; Zhang, X.; Liu, C.R. Finite element modeling on dislocation density and grain size evolution in machined surface. In Proceedings of the ASME 2013 International Manufacturing Science and Engineering Conference, MSEC2013-1130, Madison, WI, USA, 10–14 June 2013; p. 10. [Google Scholar]
  16. Li, B.; Zhang, S.; Hu, R.; Zhang, X. Dislocation density and grain size evolution in hard machining of H13 steel: Numerical and experimental investigation. J. Mater. 2020, 9, 4241–4254. [Google Scholar] [CrossRef]
  17. Liu, H.; Zhang, J.; Xu, X.; Qi, Y.; Liu, Z.; Zhao, W. Effects of dislocation density evolution on mechanical behavior of OFHC Copper during high-speed machining. Materials 2019, 12, 2348. [Google Scholar] [CrossRef]
  18. Komatsu, T.; Matsumura, T.; Torizuka, S. Effect of grain size in stainless steel on cutting performance in micro-scale cutting. Int. J. Automot. 2011, 5, 334–341. [Google Scholar] [CrossRef]
  19. Komatsu, T.; Yoshino, T.; Matsumura, T.; Torizuka, S. Effect of crystal grain size in stainless steel on cutting process in micromilling. Procedia CIRP 2012, 1, 150–155. [Google Scholar] [CrossRef]
  20. Olovsjö, S.; Wretland, A.; Sjöberg, G. The effect of grain size and hardness of wrought Alloy 718 on the wear of cemented carbide tools. Wear 2010, 268, 1045–1052. [Google Scholar] [CrossRef]
  21. Yang, X.; Zhang, B.; Bai, Q.; Kang, R.; Tang, J. Effect of grain size on subsurface characterization of pure iron subjected to orthogonal cutting. Int. J. Adv. Manuf. Technol. 2022, 120, 5793–5806. [Google Scholar] [CrossRef]
  22. Wu, X.; Li, L.; He, N.; Zhao, M.; Zhan, Z. Investigation on the influence of material microstructure on cutting force and bur formation in the micro cutting of copper. Int. J. Adv. Manuf. Technol. 2015, 79, 321–327. [Google Scholar] [CrossRef]
  23. Polvorosa, R.; Suárez, A.; de Lacalle, L.L.; Cerrillo, I.; Wretland, A.; Veiga, F. Tool wear on nickel alloys with different coolant pressures: Comparison of Alloy 718 and Waspaloy. J. Manuf. Process. 2017, 26, 44–56. [Google Scholar] [CrossRef]
  24. Weng, J.; Saelzer, J.; Berger, S.; Zhuang, K.; Bagherzadeh, A.; Budak, E.; Biermann, D. Analytical and experimental investigations of rake face temperature considering temperature-dependent thermal properties. J. Mater. Process. Technol. 2023, 314, 117905. [Google Scholar] [CrossRef]
  25. Jafarian, F.; Imaz Ciaran, M.; Umbrello, D.; Arrazola, P.J.; Filice, L.; Amirabadi, H. Finite element simulation of machining Inconel 718 alloy including microstructure changes. Int. J. Mech. Sci. 2014, 88, 110–121. [Google Scholar] [CrossRef]
  26. Arrazola, P.J.; Özel, T.; Umbrello, D.; Davies, M.; Jawahir, I.S. Recent advances in modelling of metal machining processes. CIRP Ann. Manuf. Technol. 2013, 62, 695–718. [Google Scholar] [CrossRef]
  27. De Mange, J.J.; Prakash, V.; Pereira, J.M. Effects of material microstructure on blunt projectile penetration of a nickel-based super alloy. Int. J. Impact Eng. 2009, 36, 1027–1043. [Google Scholar] [CrossRef]
  28. Yanagimoto, J.; Karhausen, K.; Brand, A.J.; Kopp, R. Incremental formulation for the prediction of flow stress and microstructural change in hot forming. J. Manuf. Sci. Eng. 1998, 120, 316–322. [Google Scholar] [CrossRef]
  29. Hughes, G.D.; Smith, S.D.; Pande, C.S.; Johnson, H.R.; Armstrong, R.W. Hall–Petch strenghening for the microhardness of twelve nanometer grain diameter electrodeposited nickel. Scr. Mater. 1986, 20, 93–97. [Google Scholar]
  30. Sui, F.-L.; Zuo, Y.; Liu, X.-H.; Chen, L.-Q. Microstructure analysis on IN 718 alloy round rod by FEM in the hot continuous rolling process. Appl. Math. Model. 2013, 37, 8776–8784. [Google Scholar] [CrossRef]
  31. Estrin, Y.; Toth, L.S.; Molinari, A.; Bréchet, Y. A dislocation-based model for all hardening stages in large strain deformation. Acta Mater. 1998, 46, 5509–5522. [Google Scholar] [CrossRef]
  32. Ding, H.; Shin, Y.C. Dislocation density-based grain refinement modeling of orthogonal cutting of commercially pure titanium. In Proceedings of the ASME 2011 International Manufacturing Science and Engineering Conference, MSEC2011-50220, Corvallis, OR, USA, 13–17 June 2011; p. 10. [Google Scholar]
  33. Ding, H.; Shin, Y.C. Dislocation density-based grain refinement modeling of orthogonal cutting of Titanium. J. Manuf. Sci. Eng. 2014, 136, 041003. [Google Scholar] [CrossRef]
  34. Ding, H.; Shin, Y.C. Multi-physics modeling and simulations of surface microstructure alteration in hard turning. J. Mater. Process. Technol. 2013, 213, 877–886. [Google Scholar] [CrossRef]
  35. Hu, D.; Gao, Y.; Meng, F.; Song, J.; Wang, R. Experimental and numerical analysis of microstructures and stress states of shot-peened GH4169 superalloys. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2018, 49, 1397–1409. [Google Scholar] [CrossRef]
  36. Liu, C.R.; Barash, M.M. The mechanical state of the sublayer of a surface generated by chip-removaI process. J. Ind. Eng. Int. 1976, 1, 1192–1201. [Google Scholar] [CrossRef]
Figure 1. Overview of this study.
Figure 1. Overview of this study.
Materials 16 03562 g001
Figure 2. Mesh and boundary conditions for the FE model.
Figure 2. Mesh and boundary conditions for the FE model.
Materials 16 03562 g002
Figure 3. Flow chart of Z–H model subroutine.
Figure 3. Flow chart of Z–H model subroutine.
Materials 16 03562 g003
Figure 4. Flow chart of DDB model subroutine.
Figure 4. Flow chart of DDB model subroutine.
Materials 16 03562 g004
Figure 5. Simulated result of Z–H model under test condition h = 0.05 mm.
Figure 5. Simulated result of Z–H model under test condition h = 0.05 mm.
Materials 16 03562 g005
Figure 6. Comparison between experimental and predicted results of hardness variation depth (a) and surface hardness (b).
Figure 6. Comparison between experimental and predicted results of hardness variation depth (a) and surface hardness (b).
Materials 16 03562 g006
Figure 7. Overall average error for machined surface hardness and hardening layer depth.
Figure 7. Overall average error for machined surface hardness and hardening layer depth.
Materials 16 03562 g007
Figure 8. SEM of surface and subsurface layer (a) and subsurface deformation after machining at vc = 50 m/min, h = 0.08 mm, L = 100 μm, and θ = 15° (b).
Figure 8. SEM of surface and subsurface layer (a) and subsurface deformation after machining at vc = 50 m/min, h = 0.08 mm, L = 100 μm, and θ = 15° (b).
Materials 16 03562 g008
Figure 9. Simulated results of DDB model calibrated under test 7 condition. Dislocation density (a), grain size (b), and surface hardness (c).
Figure 9. Simulated results of DDB model calibrated under test 7 condition. Dislocation density (a), grain size (b), and surface hardness (c).
Materials 16 03562 g009
Figure 10. Settings of cutting experiments and hardness measurements. The nanoindentation measurement (a,b) and the workpiece dimensions (c).
Figure 10. Settings of cutting experiments and hardness measurements. The nanoindentation measurement (a,b) and the workpiece dimensions (c).
Materials 16 03562 g010
Figure 11. Comparison of the performance of DDB and Z–H models for Inconel 718 with small grain size based on literature data [6]. h = 0.05 (a), h = 0.075 (b), and h = 0.1 mm (c).
Figure 11. Comparison of the performance of DDB and Z–H models for Inconel 718 with small grain size based on literature data [6]. h = 0.05 (a), h = 0.075 (b), and h = 0.1 mm (c).
Materials 16 03562 g011
Figure 12. Comparison of the performance of DDB and Z–H models for Inconel 718 with large grain size based on experiments, see Table 3. Tests 1 (a), 2 (b), 4 (c), and 7 (d).
Figure 12. Comparison of the performance of DDB and Z–H models for Inconel 718 with large grain size based on experiments, see Table 3. Tests 1 (a), 2 (b), 4 (c), and 7 (d).
Materials 16 03562 g012
Figure 13. Microscope photos of machined surface and subsurface.
Figure 13. Microscope photos of machined surface and subsurface.
Materials 16 03562 g013
Figure 14. Variation trends of DWH (a), surface hardness (b), and hardness layer depth (c).
Figure 14. Variation trends of DWH (a), surface hardness (b), and hardness layer depth (c).
Materials 16 03562 g014
Table 1. J–C material constitutive model parameters of Inconel 718 [27].
Table 1. J–C material constitutive model parameters of Inconel 718 [27].
ParameterA (MPa)B (MPa)Cnm
Value12908950.0160.5261.55
Table 2. DDB model parameters of Inconel 718 [35].
Table 2. DDB model parameters of Inconel 718 [35].
Parameterα*β*Bb (mm) ƒ 0 ƒ
Value0.080.0414,900 2.5 × 1070.250.06
ParameterKM γ ˙ 0 r γ ˜ r ρ w t = 0 (mm−2) ρ c t = 0 (mm−2)
Value119 or 7693.061.0 × 1053.22.5 × 1075.0 × 107
Table 3. Cutting parameters for Inconel 718 machining process.
Table 3. Cutting parameters for Inconel 718 machining process.
TestCutting Speed:
vc (m/min)
Uncut Chip Thickness:
h (mm)
Chamfer Angle:
θ (°)
Chamfer Length:
L (μm)
1500.0425100
2500.0425150
3500.0425200
4500.0415100
5500.0415150
6500.0415200
7500.0815100
8500.0815150
9500.0815200
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

Zhuang, K.; Wang, Z.; Zou, L.; Fu, C.; Weng, J. Simulation of Work Hardening in Machining Inconel 718 with Multiscale Grain Size. Materials 2023, 16, 3562. https://doi.org/10.3390/ma16093562

AMA Style

Zhuang K, Wang Z, Zou L, Fu C, Weng J. Simulation of Work Hardening in Machining Inconel 718 with Multiscale Grain Size. Materials. 2023; 16(9):3562. https://doi.org/10.3390/ma16093562

Chicago/Turabian Style

Zhuang, Kejia, Zhuo Wang, Linli Zou, Changni Fu, and Jian Weng. 2023. "Simulation of Work Hardening in Machining Inconel 718 with Multiscale Grain Size" Materials 16, no. 9: 3562. https://doi.org/10.3390/ma16093562

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