Next Article in Journal
On Relationships between Plasma Chemistry and Surface Reaction Kinetics Providing the Etching of Silicon in CF4, CHF3, and C4F8 Gases Mixed with Oxygen
Previous Article in Journal
Thermodynamic Insights into the Oxidation Mechanisms of CrMnFeCoNi High-Entropy Alloy Using In Situ X-ray Diffraction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Contact Pressure on Strain Distribution during Compression-Type Bulk Forming Processes

School of Mechatronics Engineering, Korea University of Technology and Education, Cheonan 31253, Republic of Korea
Materials 2023, 16(14), 5041; https://doi.org/10.3390/ma16145041
Submission received: 6 June 2023 / Revised: 14 July 2023 / Accepted: 16 July 2023 / Published: 17 July 2023

Abstract

:
Inhomogeneity of the material properties of workpieces developed during compression-type bulk forming processes (CBFPs) is an important issue. The effect of contact pressure on the workpiece surface on the strain inhomogeneity in the workpiece was investigated to understand and reduce the formation of strain inhomogeneity during CBFPs. Workpieces fabricated via rod caliber rolling, rod flat rolling, plate flat rolling, and rod compression were analyzed and compared. The extent of strain inhomogeneity in a workpiece differs with the forming process, because the occurrence of macroscopic shear bands (MSBs) is dependent on the workpiece shape and tool design. A flat-rolled rod exhibits the maximum strain inhomogeneity, whereas a flat-rolled plate shows the minimum strain inhomogeneity. The occurrence of MSBs was influenced by the distribution of the normal contact pressure or compression stress. The MSBs were stronger when the contact pressure was higher in the edge region of the surface. For example, the flat-rolled plate exhibited weak MSBs due to the relatively uniform or higher contact pressure on the central region. In contrast, strong MSBs appeared in the flat-rolled rod and compressed rod, because the contact pressure in the edge region of these two processes was high. Thus, the strain inhomogeneity in a workpiece fabricated via CBFPs can be reduced by controlling the contact pressure distribution on the workpiece surface.

1. Introduction

In the metal-forming industry, compression-type bulk forming processes (CBFPs) are used extensively owing to their high formability and productivity. However, the inhomogeneity of the material properties of the product, that is, macroscopic shear bands (MSBs), typically develops in the process [1]. In this study, the MSBs are associated with the regions of the specimen with higher effective strain compared with other regions of the specimen. The MSBs and/or strain inhomogeneity during the CBFPs is primarily caused by the restricted material flow in the workpiece–tool interface because of the frictional stress, as schematically shown in Figure 1, and it depends on both the process conditions and the material properties [2,3,4,5,6,7,8].
Therefore, the strain inhomogeneity in a workpiece has been investigated during CBFPs, such as compression [2,9,10], side pressing [1,3], and rod flat rolling [11,12]. Eom et al. [9] reported that the strain distribution of a specimen was more homogeneous when increasing the strain hardening rate of material during compression based on the comparison of steels with high and low strain hardening rates. They explained this phenomenon in terms of the capability of the strain distribution in metals: a metal with high strain hardening rate can transfer the plastic region far away from the initial deformation regions, i.e., workpiece and tool interface. Kazeminezhad et al. [11] revealed the formation of MSBs during rod flat rolling using the finite and slab element method and validated the results based on the Vickers hardness test and metallographic examination of the cross-section of the flat-rolled copper rod. They showed that the strain inhomogeneity in the flat-rolled rod increased with an increasing friction factor and decreasing reduction ratio. In addition, Hwang [12] reported a very different behavior of hardness, strain, and twinning of flat-rolled rod with a region based on experimental testing and numerical simulation using twinning-induced plasticity (TWIP) steel. He revealed that the center region tended to have the maximum twin density, effective strain, and hardness, whereas the free surface region had the minimum values. Paul et al. [2] reported that the mechanism of MSB formation depended on the crystallographic orientation of material based on the plane strain compressed test using fine-grained aluminum. They showed that the strain-induced rotation of the crystal lattice led to the formation of specific MSBs and they facilitated slip propagation across the grain boundaries. Semiatin et al. [3] revealed that the tendency to form shear bands depends on test temperature and deformation based on an isothermal side pressing test using Ti-6A1-2Sn-4Zr-2Mo0.1Si alloy with a variety of working temperatures and deformation rates.
The issue of strain inhomogeneity of a product becomes significant during the plastic forming process, as the demand for non-heat-treated products is increasing in metal-forming and related industries [13,14]. By eliminating heat treatments of products during the manufacturing process, several advantages can be obtained industrially as follows: production costs can be reduced, distortion of materials occurring during heat treatment can be prevented, production time can be reduced, and environmental pollution can be prevented. The strain inhomogeneity of a workpiece is significant in TWIP steels because they are used as the final products with non-heat treatments after the metal-forming process [15]. Therefore, decreasing the stain inhomogeneity in a workpiece during CBFP is essential to improve the quality and reduce the cost of the final product.
Recently, Hwang [16] reported that the homogeneity of the strain distribution in a workpiece tends to improve with increasing elongation and decreasing pure spreading during CBFP. Based on this, they suggested that low pure spreading and high elongation of a workpiece guarantee a more homogeneous and high-quality product during CBFP. Furthermore, the distribution of the contact pressure (Pc) on the workpiece surface should affect the formation of MSBs as well as the roll force and tool wear [17,18,19,20]. However, to the best of the author’s knowledge, studies regarding the effect of the Pc on the MSBs of a material during CBFP are few. In particular, studies regarding the effect of the Pc on the MSBs of a material during the rolling process are rarely performed. Li et al. [19] reported that the distribution of effective strain in flat-rolled plate was more homogeneous when friction hill type Pc was formed during plate flat rolling compared with the formation of two-peak type Pc. However, they did not reveal the influence of the Pc on strain inhomogeneity or MSB formation during the rolling process. Thus, it can be concluded that there is still a lack of understanding of the behavior of MSBs with the distribution of Pc during CBFP.
Therefore, this study evaluates the effect of the Pc on the strain inhomogeneity in a workpiece to reveal the mechanism of MSB formation and alleviate the strain inhomogeneity in the workpiece during CBFP. For a comprehensive study, four different CBFPs, i.e., rod caliber rolling, rod flat rolling, plate flat rolling, and rod compression, were analyzed and compared via finite element analysis (FEA) and experiments using a TWIP steel. FEA was used to evaluate the complicated distributions of strain, stress, and Pc in the workpiece during the process, and hardness tests were carried out to analyze the strain distribution throughout the surface region of the workpieces.

2. Experiment and Deformation Analysis

2.1. Experiments

A TWIP steel ingot weighing 50 kg was cast using vacuum induction. The chemical composition, as determined via spark optical emission spectrometer, was Fe-19.94Mn-0.6C-1.03Al (wt. %). The structure of the cast-ingot with a thickness of 125 mm was homogenized in a box furnace for 12 h at a furnace temperature of 1200 °C to reduce the segregation of chemical composition. The cast-ingot was then rolled to a plate with a thickness of 20 mm using several rolling machines at the final temperature of 950 °C, followed by air cooling at an ambient temperature of 20 °C.
Figure 2a represents the microstructure of the hot-rolled specimen using electron backscatter diffraction (EBSD). The specimen for EBSD observation was ground by mechanical polishing using papers with silicon carbide abrasive and diamond pastes from 6 to 1 μm. Colloidal silica suspensions were then applied to remove the residual stress around the surface of the specimen. EBSD maps of 400 μm × 400 μm were measured using a step size of 0.3 μm with a field emission scanning electron microscope, which is equipped with a TSL data acquisition system. The specimen’s tilting angle was approximately 70° and the acceleration voltage was 20 kV. The acquired EBSD data were analyzed using the commercial orientation imaging microscopy software. Based on the inverse pole figure map obtained via EBSD, recrystallized grains with annealing twins were seen, but deformation twins were not detected, as shown in Figure 2a. The calculated grain size was approximately 28 μm on average. Tensile test specimens (gauge diameter of 5 mm and length of 25 mm) were extracted along the rolling direction from the hot-rolled plate. The specimens were elongated at a slow strain rate of 10−3 s−1 at room temperature (RT, 26 °C). Figure 2b shows the true stress–strain curve of the specimen.
Several rods with a diameter of 13 mm and length of 400 mm were extracted from the hot-rolled plate using a lathe for conducting the rod caliber rolling and rod flat rolling tests. The cross-section of the workpieces was reduced using grooved and flat rolls with a diameter of 400 mm during rod caliber rolling and rod flat rolling, respectively. The rolling velocity was set to 5 revolutions per minute (RPM) to neglect the influence of temperature increase and strain rate. The initial cross-sectional shape of the workpiece, detailed roll design, and reduction in height (Rh) are presented in Figure 3 and Table 1. The Rh of the workpiece in all processes is calculated as follows:
R h = h i h f h i × 100 ( % )
where hi and hf represent the initial and final workpiece heights, respectively. During rod caliber rolling, hf was obtained by calculating the average roll height due to the oval-shaped roll design. A nominal strain (εn) of the deformed workpiece was obtained as follows:
ε n = l n h i h f
For the compression test, the hot-rolled plate was machined into a cylindrical-type compression specimen with a height of 15 mm and diameter of 10 mm. The specimens were compressed using an Instron machine at a rate of 10−3 s−1 at RT.
The Vickers hardness (HV) test was carried out with a load of 1.0 kgf and a dwell time of 15 s for the inhomogeneity evaluation of the mechanical properties of the workpiece.

2.2. Deformation Analysis

The plastic deformation of a workpiece is complex during rolling processes, particularly three-dimensional shape rolling such as rod caliber rolling, ring rolling, thread rolling, and rod flat rolling [12,21]. Accordingly, DEFORM 3D software was utilized to analyze the strain, stress, and Pc distributions in the workpiece during plate flat rolling, rod caliber rolling, rod flat rolling, and rod compression because FEA provides reliable output for 3D complex plastic forming processes. The rolls and dies used in the tests were assumed to be rigid bodies. The workpiece was assumed to be a rigid plastic material because the four workpieces were mainly deformed by plastic deformation. In other words, the average nominal strain of the workpieces was approximately 0.42 during deformation as listed in Table 1; therefore, the elastic deformation can be neglected in this analysis owing to the non-consideration of residual stress. In addition, the workpiece was assumed to be an isotropic material because a hot-rolled specimen was used in this study. The influences of the temperature increase and strain rate during the processes were not considered in this study because of the low forming rate, i.e., the slow strain rate. In such a case, among various constitutive models for material properties, the constitutive behavior of a workpiece can be simply modeled via Hollomon’s law, i.e., σ = n [15,22]. K is a constant representing the material strength and n is the strain hardening exponent of material. The necessary K and n values were obtained in the following equation based on the fitting curve in the tensile test, as presented in Figure 2b:
σ = 1970.3ε0.535
The value of R2 in this curve fitting was 0.972, which is relatively high and acceptable for modeling the constitutive behavior of a workpiece to simulate plastic deformation of a material. The deviation of the experimental tensile data and fitted curve was relatively high in the elastic region, as shown in Figure 2b, because Hollomon’s law cannot describe the elastic deformation. However, the workpiece was mainly deformed by plastic deformation in these forming processes, as shown in Table 1, indicating that the elastic deformation was not important compared with the plastic deformation. Furthermore, the fitted curve well described the constitutive behavior of a workpiece from a true strain of 0.05 up to a true strain of 0.42, which is the plastic deformation region of this study.
At the workpiece–tool interface, the shear friction model was applied, and the friction factor was chosen to be 0.3 for all processes [21]. Rods with a diameter of 13 mm were rolled with the 400 mm diameter rolls. The rolling speed was 5 RPM. A quarter of the full geometry was simulated due to the symmetrical process conditions of the four CBFPs used in this study. A hexahedron element was added to all the workpieces, and the number of elements in the workpiece was approximately 16,000. Approximately 250 elements were uniformly put into the cross-section of the workpieces, and 64 elements were inserted along the longitudinal direction of the workpieces, as shown in Figure 4.

3. Validation

To validate the present numerical model, the lateral spreading (Ws) results obtained from the experiments and FEA were compared during rod caliber rolling. The Ws values obtained from the numerical simulation were in good agreement with the measured value, as shown in Figure 5.
In addition, the contact width (Wc) and Ws between the experiment and FEA were compared for the rod flat rolling. The Ws and Wc values from the FE analysis agreed well with the measured values, as shown in Figure 6, despite a small deviation in Ws between the two values. Overall, the results of the comparison show that the proposed FEA model is acceptable for further analysis.

4. Results and Discussion

4.1. Strain Distribution in Workpiece

Figure 7 shows a comparison of the cross-section and effective strain (von Mises strain, εeff) of the caliber-rolled rod, flat-rolled rod, flat-rolled plate, and compressed rod. The εeff value varied across the workpiece. Except for the flat-rolled plate, the core region of all specimens showed the peak εeff, and the surface region exhibited the lowest value. The regions along the diagonal direction showed a relatively high εeff, resulting in the occurrence of MSBs. In other words, the regions near the MSBs were harder than the other regions in the workpiece due to the concentration of stress and strain, indicating that the formation of MSBs caused the strain inhomogeneity throughout the region of the workpiece. Eom et al. [9] reported that the peak εeff occurred in the core region of the workpiece during the compression test using steels. In addition, Kazeminezhad et al. [11] showed that the maximum and minimum εeff were developed at the free surface and center regions of the flat-rolled rod based on the copper rod.
The caliber-rolled rod and flat-rolled rod exhibited the minimum εeff in the free surface, whereas the compressed rod demonstrated the minimum value in the contact surface region. Hwang [16] reported that these results are due to the elongation effect during the rolling process. Meanwhile, the distribution of εeff in the flat-rolled plate significantly differed from that in the caliber-rolled rod, flat-rolled rod, and compressed rod. The contact surface region exhibited the peak εeff, and the core region tended to have low strain in the flat-rolled plate.
Figure 8 compares the profiles of εeff along the spreading and compression directions of the four workpieces. The highest εeff was observed in the core region of the caliber-rolled rod, flat-rolled rod, and compressed rod, whereas the peak εeff was observed in the contact surface region of the flat-rolled plate. The deviation in εeff was much higher along the spreading than the compression direction in the flat-rolled rod, caliber-rolled rod, and flat-rolled plate. By contrast, in the compressed workpiece, the difference in εeff was higher along the compression than the spreading direction.
Figure 9 shows a comparison of the deviations in εeff of the four workpieces. The flat-rolled plate exhibited a relatively homogeneous strain distribution throughout its region, and the flat-rolled rod demonstrated the maximum inhomogeneity of εeff.

4.2. Contact Pressure Distribution on Workpiece

The Pc and its distribution affect the workpiece shape, forming energy, and temperature distribution of both the workpiece and tool during the forming process [23,24,25]. Figure 10 shows comparisons of the effective stress (von Mises stress), compression stress (downward stress), and Pc on the surface of the four workpieces. The contours of the effective stress, compression stress, and Pc differed significantly for the four forming processes. The compression stress was strongly associated with the Pc. Figure 11 compares the Pc of the four workpieces using different scales for a more detailed analysis. In the case of the flat-rolled rod, the Pc distribution was high at the entrance zone, edge of the contact region, and exit zone. The Pc on flat-rolled rods was reported previously [26,27,28]. According to previous investigations, the high value of Pc in the entrance zone and edge of the contact zone is related to the restrictions caused by the rigid surrounding material, which still undergoes elastic deformation. Additionally, the high value of Pc in the exit zone is associated with the work hardening rate of the metal during the forming process. The caliber-rolled rod also had Pc similar to that of the flat-rolled rod. However, the peak strain was not observed at the entrance zone owing to the large contact area between the workpiece and roll at the initial rolling time originating from the grooved oval-shaped roll, as shown in Figure 3. In the compression test, the edge regions of the contact region exhibited higher normal pressure. Interestingly, the distribution of Pc on the flat-rolled plate was significantly different from that on the other forming processes. The flat-rolled plate exhibited higher normal pressure in the central region. Hartley et al. [29] reported similar results during plate flat rolling. In should be noted that the distribution of Pc on flat-rolled plates differs with the process conditions, such as the reduction ratio, roll crown, and bending force [30,31]. Figure 12 compares the measured average hardness distribution on the contact surface with the process. The profiles of the measured hardness could reflect the Pc variations during the process. For instance, the hardness of the center region was lower than that of the edge region in the caliber-rolled rod, flat-rolled rod, and compressed rod. In particular, the deviation in hardness on the contact surface was significantly high in the compressed workpiece because the distribution of Pc was quite inhomogeneous along the radial direction of the workpiece.

4.3. Effect of Contact Pressure on Strain Distribution

The authors believe that the strain inhomogeneity or occurrence of MSBs is influenced by the distribution of Pc. Figure 13 shows a comparison of the effective stress and compression stress on the four workpieces. The pattern of the effective stress and compression stress of the four workpieces was similar. However, the occurrence of MSBs in the cross-sections of the four workpieces was related to the compression stress on the surface of the workpieces rather than the effective stress on the surface. Accordingly, the Pc and εeff values were compared, as shown in Figure 14, because the distribution of Pc was strongly related to the compression stress (Figure 10). The distribution of εeff or the strain inhomogeneity depended on the distribution of Pc. The MSBs were stronger as the Pc was higher in the edge region of the workpiece because the sound metal flow in the width direction of the workpiece was disturbed as the Pc increased in the edge region. For instance, the flat-rolled rod and compressed rod had strong MSBs due to the higher Pc value in the edge region. Conversely, the flat-rolled plate exhibited weak MSBs owing to the higher Pc value in the central region. The MSBs in the flat-rolled rod were stronger than those in the caliber-rolled rod owing to the high Pc in the edge region of the flat-rolled rod.
In order to confirm the above suggestions, the distribution of the compression stress during the forming steps was examined, as shown in Figure 15. Evidently, the inhomogeneity of the compression stress on the workpiece surface caused product inhomogeneity during the forming steps. The formation of the strain inhomogeneity is schematically shown in Figure 16 from the perspective of the Pc. The results show that stress concentration occurred in the workpiece at the point where the Pc was relatively high. In order words, the occurrence of MSBs in the workpieces has a strong correlation with the distribution of the compression stress or Pc on the surface of the workpiece during CBFP.
By understanding the mechanism of MSB generation during CBFP, we can derive conceptual solutions to improve the strain homogeneity of the material properties in the products by controlling the Pc. The Pc on the workpiece surface should be homogeneous during the forming process by controlling the initial shape of the workpiece, reduction ratio, lubricants, tool design, and temperature of both the workpiece and tool [19,30,31,32]. Cavaliere et al. [30] reported that the distribution of Pc on the flat-rolled plate varied with the friction coefficient using several numerical simulations. They showed that the homogeneity of Pc on the flat-rolled plate increased with a decreasing friction coefficient. Their results also indicated that the distribution of Pc can be controlled using roll design, i.e., roll crown. In this study, the caliber-rolled rod had a smaller peak Pc compared with the flat-rolled rod, which originated from the large contact area between the workpiece and roll due to the grooved roll shape or caliber roll, indicating that roll design can control the distribution of Pc on the workpiece during the rolling process. Li et al. [19] revealed that the distribution of Pc changed from two-peak type to friction hill type with an increasing reduction ratio during the plate flat rolling. And they insisted that the εeff in the specimen with the friction hill type Pc was more uniform compared with the specimen with two-peak type Pc. This result showed that the distribution of Pc affected the strain inhomogeneity of material, and it depended on the reduction ratio during the rolling process. Hwang [12] showed that the strain inhomogeneity gradually increased and then decreased with the reduction ratio during rod flat rolling, which means that the maximum strain inhomogeneity occurred at the specific reduction ratio. The author believes that this phenomenon is related to the distribution of Pc on the flat-rolled rod with the reduction ratio. In addition, Li et al. [19] showed that the distribution of Pc depended on the strain hardening rate of material based on the comparison simulation of aluminum and copper. The result indicated that the distributions of Pc and εeff were more uniform with an increasing strain hardening rate of material. It is known that the strain hardening rate of material affects the strain distribution of the workpiece during the metal forming process. Temperature has an indirect effect on the distribution of Pc during the forming process because the friction coefficient at the material and die interface, and the strain hardening rate and strain rate sensitivity of material, changed with temperature.
Alternatively, the deviation of Pc on the surface of the workpiece should be reduced by decreasing Pc in the edge region or increasing Pc in the central region on the surface of the workpiece. For example, in the case of the flat-rolled plate, the Pc in the central region of the surface was relatively high, thereby enhancing the homogeneity of the mechanical properties of the workpiece, as shown in Figure 9.

5. Conclusions

Based on the comparative study of the distributions of the strain, stress, and Pc in TWIP steel during rod caliber rolling, rod flat rolling, plate flat rolling, and rod compression, the following conclusions were derived:
  • Strain inhomogeneity or MSBs appeared in the workpiece during the CBFPs. Except for the flat-rolled plate, the core region of all specimens showed peak εeff, and the free surface region exhibited the lowest value. The regions along the diagonal direction showed a relatively high εeff, resulting in the occurrence of MSBs. However, the level of strain inhomogeneity in the workpiece was different from the forming process, because the formation of MSBs was dependent on the initial workpiece shape and tool design. The flat-rolled rod exhibited the maximum strain inhomogeneity, whereas the flat-rolled plate showed the minimum strain inhomogeneity.
  • The occurrence of MSBs is related to the distribution of Pc or compression stress. The MSBs were stronger when the Pc was higher in the edge region of the workpiece because the sound metal flow in the width direction of the workpiece was disturbed as the Pc increased in the edge region. For example, the flat-rolled plate had weaker MSBs due to the relatively uniform Pc or higher Pc value on the central region. By contrast, strong MSBs appeared in the flat-rolled rod and compressed rod because the Pc value in the edge region of these two processes was high.
  • The strain homogeneity in a workpiece fabricated via CBFP can be improved by tailoring the distribution of Pc on the surface of the workpiece. To improve the strain homogeneity of the workpiece, the Pc on the workpiece surface should be homogeneous by controlling the initial shape of the workpiece, reduction ratio, lubricants, tool design, and strain hardening rate of material.

Funding

This study was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT, South Korea) (No. 2021R1A2C1011700).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Semiatin, S.L.; Jonas, J.J. Formability and Workability of Metals: Plastic Instability and Flow Localization; American Society for Metals USA: Russell Township, OH, USA, 1984. [Google Scholar]
  2. Paul, H.; Driver, J.H.; Tarasek, A.; Wajda, W.; Miszczyk, M.M. Mechanism of macroscopic shear band formation in plane strain compressed fine-grained aluminium. Mater. Sci. Eng. A 2015, 642, 167–180. [Google Scholar] [CrossRef]
  3. Semiatin, S.L.; Lahoti, G.D. The occurrence of shear bands in isothermal, hot forging. Metall. Mater. Trans. A 1982, 13, 275–288. [Google Scholar] [CrossRef]
  4. Lambiase, F.; Ilio, A.D. Deformation inhomogeneity in roll drawing process. J. Manuf. Process. 2012, 14, 208–215. [Google Scholar] [CrossRef]
  5. Jia, N.; Eisenlohr, P.; Roters, F.; Raabe, D.; Zhao, X. Orientation dependence of shear banding in face-centered-cubic single crystals. Acta Mater. 2012, 60, 3415–3434. [Google Scholar] [CrossRef]
  6. Cheng, L.; Xue, X.; Tang, B.; Liu, D.; Li, J.; Kou, H.; Li, J. Deformation behavior of hot-rolled IN718 superalloy under plane strain compression at elevated temperature. Mater. Sci. Eng. A 2014, 606, 24–30. [Google Scholar] [CrossRef]
  7. Tang, B.; Xiang, L.; Cheng, L.; Liu, D.; Kou, H.; Li, J. The formation and evolution of shear bands in plane strain compressed nickel-base superalloy. Metals 2018, 8, 141. [Google Scholar] [CrossRef] [Green Version]
  8. Lee, D.J.; Yoon, E.Y.; Park, L.J.; Kim, H.S. The dead metal zone in high-pressure torsion. Scr. Mater. 2012, 67, 384–387. [Google Scholar] [CrossRef]
  9. Eom, J.G.; Son, Y.H.; Jeong, S.W.; Ahn, S.T.; Jang, S.M.; Yoon, D.J.; Joun, M.S. Effect of strain hardening capability on plastic deformation behaviors of material during metal forming. Mater. Des. 2014, 54, 1010–1018. [Google Scholar] [CrossRef]
  10. Hor, A.; Morel, F.; Lebrun, J.L.; Germain, G. An experimental investigation of the behaviour of steels over large temperature and strain rate ranges. Int. J. Mech. Sci. 2013, 67, 108–122. [Google Scholar] [CrossRef] [Green Version]
  11. Kazeminezhad, M.; Karimi Taheri, A. Deformation inhomogeneity in flattened copper wire. Mater. Des. 2007, 28, 2047–2053. [Google Scholar] [CrossRef]
  12. Hwang, J.K. Deformation behaviors of flat rolled wire in twinning-induced plasticity steel. Met. Mater. Int. 2020, 26, 603–616. [Google Scholar] [CrossRef]
  13. Hwang, J.K. Enhanced homogeneity of a flat-rolled wire in twinning-induced plasticity steel using the pass schedule design. ISIJ Int. 2020, 60, 2493–2502. [Google Scholar] [CrossRef]
  14. Hwang, J.K. Effects of caliber rolling on microstructure and mechanical properties in twinning-induced plasticity (TWIP) steel. Mater. Sci. Eng. A 2018, 711, 156–164. [Google Scholar] [CrossRef]
  15. Kim, S.J.; Lee, T.; Hwang, J.K. High-strength bolt manufactured by an extrusion-based forming process using twinning-induced plasticity steel. J. Manuf. Process. 2020, 59, 33–42. [Google Scholar] [CrossRef]
  16. Hwang, J.K. Inhomogeneity of microstructure and mechanical properties in compression-type bulk formed specimens. Proc. IMechE Part B J. Eng. Manuf. 2023, 237, 1485–1496. [Google Scholar] [CrossRef]
  17. Orowan, E. The calculation of roll pressure in hot and cold flat rolling. Proc. Inst. Mech. Eng. 1943, 150, 140–167. [Google Scholar] [CrossRef]
  18. Carlsson, B. The contact pressure distribution in flat rolling of wire. J. Mater. Process. Technol. 1998, 73, 1–6. [Google Scholar] [CrossRef]
  19. Li, G.J.; Kobayashi, S. Rigid-plastic finite-element analysis of plane strain rolling. J. Eng. Ind. 1982, 104, 55–63. [Google Scholar] [CrossRef]
  20. Rudnytskyj, A.; Vorlaufer, G.; Leimhofer, J.; Jech, M.; Gachot, C. Estimating the real contact area in lubricated hot rolling of aluminium. Tribol. Int. 2023, 180, 108283. [Google Scholar] [CrossRef]
  21. Inoue, T.; Yin, F.; Kimura, Y. Strain distribution and microstructural evolution in multi-pass warm caliber rolling. Mater. Sci. Eng. A 2007, 466, 114–122. [Google Scholar] [CrossRef]
  22. Ghosh, A.K. The influence of strain hardening and strain-rate sensitivity on sheet metal forming. J. Eng. Mater. Technol. 1977, 99, 264–274. [Google Scholar] [CrossRef]
  23. Wusatowski, Z. Fundamentals of Rolling; Pergamon Press: London, UK; New York, NY, USA, 1969. [Google Scholar]
  24. Spuzic, S.; Strafford, K.N.; Subramanian, C.; Savage, G. Wear of hot rolling mill rolls: An overview. Wear 1994, 176, 261–271. [Google Scholar] [CrossRef]
  25. Duan, X.; Sheppard, T. Three dimensional thermal mechanical coupled simulation during hot rolling of aluminium alloy 3003. Int. J. Mech. Sci. 2002, 44, 2155–2172. [Google Scholar] [CrossRef]
  26. Kazeminezhad, M.; Karimi Taheri, A. Calculation of the rolling pressure distribution and force in wire flat rolling process. J. Mater. Process. Technol. 2006, 171, 253–258. [Google Scholar] [CrossRef]
  27. Vallellano, C.; Cabanillas, P.A.; Garcia-Lomas, F.J. Analysis of deformations and stresses in flat rolling of wire. J. Mater. Process. Technol. 2008, 195, 63–71. [Google Scholar] [CrossRef]
  28. Hwang, J.K. Strain hardening rate dependency of deformation shape, strain distribution, and contact pressure during wire flat rolling. J. Iron Steel Res. Int. 2023, 30, 475–486. [Google Scholar] [CrossRef]
  29. Hartley, P.; Sturgess, C.E.N.; Liu, C.; Rowe, G.W. Experimental and theoretical studies of workpiece deformation, stress, and strain during flat rolling. Int. Mater. Rev. 1989, 34, 19–34. [Google Scholar] [CrossRef]
  30. Cavaliere, M.A.; Goldschmit, M.B.; Dvorkin, E.N. Finite element analysis of steel rolling processes. Comput. Struct. 2001, 79, 2075–2089. [Google Scholar] [CrossRef]
  31. Liu, X.; Shi, X.; Li, S.; Xu, J.; Wang, G. FEM analysis of rolling pressure along strip width in cold rolling process. J. Iron Steel Res. Int. 2007, 14, 22–26. [Google Scholar] [CrossRef]
  32. Tieu, A.K.; Liu, Y.J. Friction variation in the cold-rolling process. Tribol. Int. 2004, 37, 177–183. [Google Scholar] [CrossRef]
Figure 1. (a,b) Schematics of the effect of friction at the workpiece–tool interface and (c) distribution of effective strain during uniaxial compression.
Figure 1. (a,b) Schematics of the effect of friction at the workpiece–tool interface and (c) distribution of effective strain during uniaxial compression.
Materials 16 05041 g001
Figure 2. (a) Microstructure and (b) true tensile curve of hot-rolled TWIP steel.
Figure 2. (a) Microstructure and (b) true tensile curve of hot-rolled TWIP steel.
Materials 16 05041 g002
Figure 3. Workpiece and tool geometries for (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Figure 3. Workpiece and tool geometries for (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Materials 16 05041 g003
Figure 4. Mesh system for (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Figure 4. Mesh system for (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Materials 16 05041 g004
Figure 5. (a) Measured and (b) simulated cross-section of caliber-rolled rod and (c) comparison of lateral spreading of caliber-rolled rod obtained from experiment and FEA.
Figure 5. (a) Measured and (b) simulated cross-section of caliber-rolled rod and (c) comparison of lateral spreading of caliber-rolled rod obtained from experiment and FEA.
Materials 16 05041 g005
Figure 6. (a) Measured and (b) simulated cross-section of flat-rolled rod; (c) comparison of contact width and lateral spreading of flat-rolled rod obtained from experiment and FEA.
Figure 6. (a) Measured and (b) simulated cross-section of flat-rolled rod; (c) comparison of contact width and lateral spreading of flat-rolled rod obtained from experiment and FEA.
Materials 16 05041 g006
Figure 7. Comparison of shape and effective strain of (a) caliber-rolled rod, (b) flat-rolled rod, (c) flat-rolled plate, and (d) compressed rod.
Figure 7. Comparison of shape and effective strain of (a) caliber-rolled rod, (b) flat-rolled rod, (c) flat-rolled plate, and (d) compressed rod.
Materials 16 05041 g007
Figure 8. Profiles of effective strain along (a) spreading and (b) compression directions of four workpieces.
Figure 8. Profiles of effective strain along (a) spreading and (b) compression directions of four workpieces.
Materials 16 05041 g008
Figure 9. Comparison of average, maximum, and minimum effective strains of four workpieces. Nominal strain was calculated using Equation (2).
Figure 9. Comparison of average, maximum, and minimum effective strains of four workpieces. Nominal strain was calculated using Equation (2).
Materials 16 05041 g009
Figure 10. Distributions of effective stress, compression stress, and normal contact pressure on the surface of four workpieces.
Figure 10. Distributions of effective stress, compression stress, and normal contact pressure on the surface of four workpieces.
Materials 16 05041 g010
Figure 11. Distribution of normal contact pressure on the surface in a workpiece during (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Figure 11. Distribution of normal contact pressure on the surface in a workpiece during (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Materials 16 05041 g011
Figure 12. Comparison of the measured average hardness on the contact surface of workpieces along the width direction for different forming processes.
Figure 12. Comparison of the measured average hardness on the contact surface of workpieces along the width direction for different forming processes.
Materials 16 05041 g012
Figure 13. Comparison of effective stress and compression stress of four workpieces.
Figure 13. Comparison of effective stress and compression stress of four workpieces.
Materials 16 05041 g013
Figure 14. Comparison of contact pressure and effective strain of four workpieces.
Figure 14. Comparison of contact pressure and effective strain of four workpieces.
Materials 16 05041 g014
Figure 15. Compression stress variations in workpiece during (a) rod flat rolling and (b) rod compression with forming steps.
Figure 15. Compression stress variations in workpiece during (a) rod flat rolling and (b) rod compression with forming steps.
Materials 16 05041 g015
Figure 16. Schematics of effect of contact pressure on the surface on strain inhomogeneity in the workpiece during (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Figure 16. Schematics of effect of contact pressure on the surface on strain inhomogeneity in the workpiece during (a) rod caliber rolling, (b) rod flat rolling, (c) plate flat rolling, and (d) rod compression.
Materials 16 05041 g016
Table 1. Process conditions for the CBFPs used in the present research.
Table 1. Process conditions for the CBFPs used in the present research.
CBFPHeight Reduction
(Rh, %)
Nominal Strain
(εn)
Evaluation Method
Rod caliber rolling35.20.43Experiment and FEA
Rod flat rolling34.60.42Experiment and FEA
Plate flat rolling35.00.43FEA
Rod compression34.50.42Experiment and FEA
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

Hwang, J.-K. Effect of Contact Pressure on Strain Distribution during Compression-Type Bulk Forming Processes. Materials 2023, 16, 5041. https://doi.org/10.3390/ma16145041

AMA Style

Hwang J-K. Effect of Contact Pressure on Strain Distribution during Compression-Type Bulk Forming Processes. Materials. 2023; 16(14):5041. https://doi.org/10.3390/ma16145041

Chicago/Turabian Style

Hwang, Joong-Ki. 2023. "Effect of Contact Pressure on Strain Distribution during Compression-Type Bulk Forming Processes" Materials 16, no. 14: 5041. https://doi.org/10.3390/ma16145041

APA Style

Hwang, J.-K. (2023). Effect of Contact Pressure on Strain Distribution during Compression-Type Bulk Forming Processes. Materials, 16(14), 5041. https://doi.org/10.3390/ma16145041

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