Next Article in Journal
A Review on Rotary Generators of Hydrodynamic Cavitation for Wastewater Treatment and Enhancement of Anaerobic Digestion Process
Previous Article in Journal
Leakage Characteristics of Proportional Directional Valve
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Contact Conditions at Wire–Die Interface on Temperature Distribution during Wire Drawing

School of Mechatronics Engineering, Korea University of Technology & Education, Cheonan 31253, Republic of Korea
*
Author to whom correspondence should be addressed.
Processes 2023, 11(2), 513; https://doi.org/10.3390/pr11020513
Submission received: 6 January 2023 / Revised: 3 February 2023 / Accepted: 6 February 2023 / Published: 8 February 2023

Abstract

:
The effects of contact conditions at the wire–die interface on the temperature distribution of the specimen and die are investigated to understand the wire drawing process. Finite element analysis and experiments are performed to analyze the temperature distribution of a drawn wire and die based on different contact conditions using a low-carbon steel wire. The maximum temperature (Tmax) of the die decreases as the contact heat transfer coefficient at the wire–die interface increases, whereas that of the wire increases with the contact heat transfer coefficient. The Tmax of the die and wire decreases with the thermal conductivity of the die. As the thermal conductivity of the die increases, the heat generated by friction is rapidly absorbed into the die, and the Tmax of the die decreases, thus resulting in a decrease in the surface temperature of the wire. The Tmax of both the die and wire linearly increases with the friction factor. In particular, the Tmax of the die more sensitively changes with the friction factor compared with that of the wire. The Tmax of the die linearly increases with the drawing velocity, whereas that of the wire parabolically increases with the drawing velocity. The influence of bearing length on the temperature increase in both the wire and die is insignificant.

1. Introduction

Wire drawing is a simple cold metal forming process involving wire, rod, and bar products, such as cables, electrical wires, springs, musical instruments, tire cords, saw wire, wire rope, and so on [1]. During wire drawing, metals show various deformation behaviors depending on the processing temperature because the underlying mechanism of strengthening and ductility of metals is different based on temperature [2,3,4,5,6,7,8]. In the wire drawing industries, the temperature increase of both the specimen and die is a significant issue because an excessive increase in temperature during the drawing process deteriorates the product quality, drawability, and die life, particularly in pearlitic steels [9,10,11,12,13]. For example, the temperature increase deteriorated the lubrication property during the process [14], leading to the surface delamination or cracks of the specimen, resulting in the reduction of the drawability of the specimen. In addition, as the strength of the specimen increases with the drawing pass owing to the strain-hardening effect of metals, the deformation resistance of the specimen increases, generating more heat [11]. Furthermore, excessive temperature increase followed by fast cooling sometimes leads to the local surface hardening of the specimen due to the generation of martensite structures. Therefore, the drawing velocity or productivity of the wire, rod, and bar products can increase to a certain extent owing to the temperature rise during wire drawing.
During wire drawing, heat is generated and transferred via several mechanisms, as shown in Figure 1a. The temperature of the specimen is increased by the heat generated by plastic deformation and the heat caused by sliding friction at the wire–die interface. By contrast, the heat of the wire is dissipated via convection and radiation. For example, the heat of the wire is lost to the environment, tools, dies, lubricants, and coolants during wire drawing. The high-pressure contact sliding between the wire and die results in intense conduction heat transfer; therefore, the surface region of the specimen experiences a rapid temperature change during the drawing process. The temperature increase of a drawn wire can be determined by the process conditions, such as the drawing velocity (Vd), reduction in area per pass (Rp), and semi-die angle (α), as well as material properties, such as the thermal conductivity, heat capacity (Cp), density (ρ), strain hardening coefficient (K), and strain hardening exponent (n) of a specimen. In addition, the thermal behavior of the specimen depends on the contact conditions of the wire and die, such as contact heat transfer coefficient (hc), friction factor between the specimen and die (m), contact length of the wire and die (Lc), bearing length (Lb), and the ratio of thermal conductivity between die (kd) and wire (kw) as shown in Figure 1b. The influences of process conditions on the temperature increase of materials have been investigated in recent decades [15,16,17,18,19,20,21,22,23,24,25]. Most of the studies reported that an increase in Vd, Rp, Lb, and m increased the temperature of a specimen. In the case of α, an optimum value of α resulted in a minimal temperature increase.
To the best of the author’s knowledge, studies regarding the effects of contact conditions on the temperature increase of a material during wire drawing are insufficient. In particular, studies regarding the effects of contact conditions at the wire–die interface, such as hc, kd/kw, Lc, and Lb on the temperature rise of a specimen are rarely performed. Hence, in this study, the effects of contact conditions on temperature increase are investigated using a low-carbon steel wire to understand the thermal behavior of drawn wire and to improve the wire drawing process from a thermal management standpoint. Finite element analysis (FEA) and experiments are performed to evaluate the temperature increase of deformed specimens and dies based on six contact parameters, i.e., hc, kd, Lc, Lb, Vd, and m.

2. Experiment and Numerical Simulation

2.1. Experiment

A low-carbon steel was selected as the test material because this study is primarily interested in the temperature distribution of both the specimen and die depending on the contact conditions rather than the mechanical properties or drawability of the wire with drawing conditions. The low-carbon steel wire rod (13 mm diameter) with a chemical composition of Fe–0.1C–0.4Mn–0.1Si (wt.%) was obtained from POSCO, a steel-making company in Pohang, South Korea. This wire rod steel was fabricated by heating a billet at approximately 1150 °C, performing hot shape rolling at approximately 1000 °C, and conducting air cooling at a cooling rate of 3 °C/s. The microstructure and true stress–strain curve of the hot-rolled wire rod are shown in Figure 2b,c, respectively. The microstructure was characterized via scanning electron microscopy using secondary electrons at 15 kV. To perform a tensile test, the hot-rolled wire rod with a diameter of 13 mm was machined into a tensile specimen with a gage length of 25 mm and a diameter of 5 mm; subsequently, it was strained at a strain rate of 10−3 s−1 using Instron equipment at 26 °C.
The wire rod 13 mm in diameter was drawn to a wire measuring 11.63 mm in diameter at a Vd of 0.07 m/s using a single-pass draw bench machine at 26 °C, as shown in Figure 2a. Prior to the drawing, chemical pickling was applied to the wire rod to remove oxidation scales. Subsequently, spray-type molybdenum disulfide (MoS2) solid lubricant was applied to the specimen owing to its low friction coefficient and good anti-seizure ability originating from the easy cleavage and excellent adhesion to the surface of metals. [26,27,28]. The α was 6° and the Rp was approximately 20%. The Rp was calculated as follows:
R p = d 0 2 d f 2 d 0 2 × 100   ( % )
where d0 and df are the initial and final wire diameters, respectively.
The core temperature of the specimen was measured using a K-type thermocouple measuring 1.0 mm in diameter. To prevent temperature disturbances at the specimen surface, the thermocouple was embedded at the bottom of the wire through a hole measuring 1.0 mm in diameter, as shown in Figure 3 [29]. The drawing force was measured using a load cell installed in the draw bench machine.

2.2. Finite Element Analysis

During wire drawing, the thermal behavior exhibited by the wire and die is quite complex, as discussed in the Section 1 (Figure 1). To the best of the author’s knowledge, it is difficult to measure the surface temperature of rounded small specimens using both radiation-type pyrometers and conduction-type thermocouples [30]. In addition, the surface temperature can be significantly different from that of the inner region of the wire, especially in high-speed wire drawing. For this reason, mathematical models were frequently used to predict the thermal behavior of the wire [12,31]. In this study, FEA was performed to analyze the complex temperature distribution of the specimen with several contact conditions during wire drawing. The DEFORM software (version 11.0, Scientific Forming Technologies Corporation, Columbus, OH, USA was used to simulate the drawing process. The wire rod 13 mm in diameter was drawn to a wire 11.63 mm in diameter. The flow stress curve for the numerical simulation was obtained using the results of the tensile stress–strain curve (Figure 2c). The specimen was considered to be isotropic; therefore, the constitutive behavior of the specimen was described using Hollomon’s law, i.e., σ = Kεn. The n and K values of the wire were set as 0.16 and 628 MPa, respectively, by fitting the tensile curve of hot-rolled steel, as shown in Figure 2c. The die was regarded as a rigid body, i.e., the die did not deform during the forming process. Lc is calculated as follows based on Figure 3:
L c = d 0 d f 2 sin α
Friction significantly affects the deformation behavior of a workpiece during plastic forming. In this study, the shear friction model was applied at the die–wire interface owing to the formation of relatively high pressure during wire drawing as follows:
τ = mk
where τ is the shear stress on the contact surface and k is the shear yield stress of the material.
The values of m listed in Table 1 were selected to understand the effect of friction on the temperature increase of the specimen and die during wire drawing.
The temperature increase due to plastic deformation was calculated as follows [32,33]:
Δ T i = Δ u ρ C p = β ρ C p ε 1 ε 2 σ d ε
where Ti, Δu, and β are the temperature rise caused by plastic deformation, the generated heat energy, and the fraction factor from mechanical work to heat energy, respectively. β was selected as 0.9 because only a low amount of mechanical work was stored in the deformed wire as elastic energy [32,33,34]. The thermal properties of the specimen and die, as shown in Figure 3, were assumed to be unaffected by temperature. The following six experimental parameters were selected, and their summary is provided in Table 1:
(i)
hc varied from 1 to 200 kW/m2/°C.
(ii)
kd varied from 12 to 300 W/m/°C.
(iii)
m varied from 0.01 to 0.4.
(iv)
Vd varied from 0.05 to 0.3 m/s.
(v)
Lc varied from 3.17 to 12.61 mm.
(vi)
Lb varied from 1.3 to 7.8 mm.
Table 1. Material properties and process conditions of specimen and die used in FEA. * indicates the standard operating condition.
Table 1. Material properties and process conditions of specimen and die used in FEA. * indicates the standard operating condition.
ParameterWire RodDieWire–Die Interface
Material propertiesFlow stress (MPa)σ = 628ε0.16Rigid body-
Thermal conductivity (k, W/m/°C)60 [35]12, 30, 60 *, 120, 300kd/kw of 0.2, 0.5, 1.0 *, 2.0, 5.0
Specific heat capacity (ρCp, N/mm2/°C)3.6 [32]3.6-
Fraction factor (β)0.9--
Process conditionsInitial wire diameter (do, mm)13.00--
Drawn wire diameter (df, mm)-11.63-
Reduction in area per pass (Rp, %)20--
Drawing velocity (Vd, m/s)--0.05, 0.1 0.15 *, 0.3
Contact conditionsContact length (Lc, mm)-3.17, 4.74, 6.31 *, 9.46, 12.61Lc/do of 0.24, 0.36, 0.49 *, 0.73, 0.97
Bearing length (Lb, mm)-1.3, 3.9 *, 7.8Lb/do of 0.1, 0.3 *, 0.6
Shear friction factor (m)--0.01, 0.1, 0.2 *, 0.4
Contact heat transfer coefficient (hc, kW/m2/°C)--1, 5, 10, 20 *, 40, 80, 200

3. Model Validation

Prior to analyzing the temperature distribution of the specimen and die via FEA with the contact conditions, the reliability of the FEA model was verified by comparing the numerically simulated and experimentally measured drawing forces and core temperatures of the low-carbon steel. In this case, m was set as 0.1765 based on a previous study [36]. Owing to the limited ability of the draw bench machine used in this study, the drawing velocity was set to 0.07 m/s. The other operating conditions were identical to standard conditions, as listed in Table 1.
Figure 4 shows a comparison of the drawing forces and core temperatures of the specimen between the experiments and FEA. The simulated drawing force agreed well with the measured value. The prediction error was 1.4%, as listed in Table 2. In terms of temperature, the temperature numerically predicted was slightly higher than that experimentally obtained. The prediction error was 4.0%, which is associated with the friction factor assumed in the FEA. Overall, based on the prediction error for both the drawing force and core temperature, the results of the FEA model are acceptable for further analysis.

4. Results and Discussion

4.1. Effect of Contact Heat Transfer Coefficient

During the bulk forming process, hc is affected by several process parameters, i.e., forming speed, reduction ratio, lubricant, surface roughness of workpiece and tool, tool shape, and specimen temperature. Therefore, researchers have used different values for these parameters to simulate the bulk forming process: the most typically used range for hc was 5 to 80 kW/m2/°C [36,37,38,39,40,41,42]. For example, Moon et al. [36] determined hc as 10 kW/m2/°C for the wire drawing of plain carbon steel. Notably, obtaining the optimum heat transfer coefficient during wire drawing is difficult due to the complexity of deriving hc and the limitations of the experiments conducted in this study. Hence, the author assumed an appropriate hc based on the literature review, and then the thermal behaviors of the wire and die were qualitatively compared via FEA.
Figure 5a compares the temperature distribution of the drawn wire calculated via FEA using different values of hc. The temperatures of both the wire and die varied with hc. The surface region of the wire exhibited the highest temperature, and the center region of the wire had the lowest temperature during wire drawing. Figure 5b shows the temperature profiles along the radial direction of the drawn wire at the die exit. The wire temperature increased with hc, particularly at the surface region of the wire. To provide a general overview, Figure 5c shows a summary of the maximum temperature (Tmax) of the wire and die during the process against hc. The Tmax of the die decreased with hc, whereas that of the wire increased with hc. The heat caused by friction readily transferred from the die to the wire as hc increased. From the standpoint of die wear, these results suggest that die wear can be reduced by increasing hc during wire drawing.

4.2. Effect of Thermal Conductivity of Die

Thermal conductivity is the material’s intrinsic ability to conduct heat. Thermal conduction occurs through molecular agitation, not the bulk movement of the solid. Heat moves along a temperature gradient from a high-temperature region to a low-temperature region until thermal equilibrium is reached. Therefore, the heat transfer rate depends on the magnitude of the temperature gradient and the thermal conductivity of the material. For example, Hwang [43] reported that the wire temperature differs with the thermal conductivity of the specimen during wire drawing. The thermal behavior of the specimen and die can differ with the thermal conductivity of the die during wire drawing. Figure 6a shows a comparison of the temperature contours of the drawn wires for different values of kd. In this study, kd was normalized by kw and compared by considering the field applicability. Both the specimen and die temperatures decreased as kd/kw increased, which is confirmed by the temperature profiles along the radial direction of the wire at the die exit, as shown in Figure 6b. Figure 6c shows the variation in the Tmax of the wire and die during the process with kd/kw. The Tmax of the die and wire decreased with increasing kd/kw. As kd increased, the heat generated by friction was rapidly absorbed into the die and the Tmax of the die decreased. Accordingly, the surface temperature of the wire in the contact region decreased with kd. In summary, the increase in kd decreased the die wear as well as the Tmax of the wire. Meanwhile, the authors believe that it is necessary to study the performance of various drawing dies [44] in view of kd.

4.3. Effect of Friction

Friction is the non-conservative force resisting the relative motion of the two solid surfaces or fluid layers. When contact surfaces move relative to each other, the frictional force between the two surfaces converts kinetic energy into thermal energy, leading to the temperature increase of the two specimens, particularly the wire and die during wire drawing. Figure 7a shows a comparison of the temperature contours of the wire and die with the different m, and Figure 7b shows the temperature profiles along the radial direction of the wire at the die exit. Although m value varies depending on the contact location between the wire and die due to the different contact pressure with position within the deformation zone [45], a constant value of m was assumed in this study regardless of wire position. The temperature gradient along the radial direction of the wire increased with m. The heat caused by friction at the wire–die interface increased with m. Therefore, the temperature in the surface region of the specimen significantly increased with m, which is consistent with the results of previous studies [46,47]. Figure 7c compares the variation in the Tmax of wire and die during the process against m. The Tmax of both the die and wire linearly increased with m. In particular, the Tmax of the die more sensitively changed with m compared with that of the wire.
Meanwhile, it should be noted that the friction coefficient at the wire–die interface and wear of wire typically increased with increasing wire temperature during wire drawing because high temperature softened wire materials and sometimes promoted the oxidation of the metal. The softened materials can be easily sheared and removed under frictional stress, resulting in high wear depth [48]. Therefore, to overcome this drawback, the research on the streamlined die [49,50], hydrodynamic lubrication [51,52], die coating [53], and ultrasonically oscillating dies [54,55,56] have been conducted in the wire drawing research fields. In addition, the frictional stress at the wire–die interface affected the material properties at the surface of the wire during wire drawing [57].

4.4. Effect of Drawing Velocity

During wire drawing, the wire temperature as well as material properties are significantly affected by Vd [16,18,21,31,58,59]. Figure 8a shows the temperature contours of the specimen and die for the different values of Vd, and Figure 8b shows a comparison of the temperature profiles along the radial direction of the wire at the die exit. As expected, the temperatures of both the wire and die increased with Vd, which is consistent with the previous results [59]. The temperature gradient along the radial direction of the specimen increased with Vd due to the effect of frictional heating. As Vd increased, the temperature gradient inside the wire became stronger because the heat generated by friction at the wire–die interface did not have enough time to be transferred into the wire interior or atmosphere. Meanwhile, the Tmax of the die linearly increased with Vd, whereas that of the wire parabolically increased with Vd, as shown in Figure 8c. Notably, the temperatures of wire and die can decrease with increasing Vd in view of friction because it is known that the friction coefficient is reduced with increasing Vd in copper wire drawing [60].

4.5. Effect of Contact Length

The plastic deformation of the specimen takes place within Lc, that is, the deformation zone, and this length was determined by Rp and α during the typical conical die drawing process [61]. Figure 9a compares the temperature contours of the drawn wire and die based on different values of Lc. Both the specimen and die temperatures increased with Lc, as confirmed by the temperature profiles along the radial direction of the wire at the die exit (Figure 9b). Figure 9c shows a comparison of the variation in the Tmax of wire and die during the process with Lc. The Tmax of the die and wire increased with Lc owing to the longer wire and die contact time, indicating that the temperatures of wire and die increased with decreasing α in this range of contact conditions during wire drawing.

4.6. Effect of Bearing Length

Lb is an important process parameter in the wire drawing process because it determines the final shape and residual stress of the drawn wire [62]. Figure 10a shows the temperature contours of the wire and die for different values of Lb, and Figure 10b shows the temperature profiles along the radial direction of the wire at the die exit. In addition, Figure 10c compares the Tmax of the die and wire against Lb/do. The influence of Lb on the temperature distributions of both the wire and die was insignificant.

4.7. Effect of Contact Conditions on Temperature Increase

Figure 11 summarizes the effect of the contact conditions on the temperature increase of the wire and die during wire drawing based on the numerical simulations. The effects of the contact conditions on the temperature distributions of both the wire and die were associated with the heat generated by friction at the wire–die interface. Based on the classical theory of wire drawing [11,63], the total temperature increase of a wire (ΔTt) comprises temperature rise from ideal plastic deformation (ΔTi)—as expressed in Equation (4)—temperature increase caused by frictional work (ΔTf), and temperature increase caused by redundant work (ΔTr), as follows:
ΔTt = ΔTi + ΔTf + ΔTr
where ΔTi and ΔTr depend on Rp, α, n, and K values instead of the contact conditions. The surface temperature of the wire was slightly higher than the central temperature owing to the higher ΔTr in the surface region originating from the higher effective strain during wire drawing [64,65].
By contrast, ΔTf is significantly affected by the contact conditions during wire drawing. Therefore, the heat generated by friction at the wire–die interface should be prioritized to tailor the temperature distributions of the wire and die. Practically, the die temperature should be decreased during the process to reduce die wear. The die temperature decreased as hc and kd increased and Vd, m, and Lc decreased. Therefore, hc needs to be increased and m should be decreased by selecting the appropriate lubricants and lubrication conditions. In addition, a die material with a high kd should be used. In the case of wire, hc, Vd, m, and Lc should be decreased to prevent an increase in temperature of the wire because the temperature rise of the wire deteriorated the performance of lubricants [14], increased the wear of the wire [48], and induced the wire breaks during drawing due to the flow localization originating from the dynamic strain aging effect in plain carbon steels [66]. Notably, the effect of hc on the temperature increase in the wire and die was different. As hc increased, the wire temperature increased but the die temperature decreased.
Based on the above results, the engineers in the wire drawing mill should derive optimal process conditions for the drawing process according to the mill situation. Meanwhile, it is necessary to consider that process variables move together. For example, as Vd increases, m decreases [60,67], and the relationship between Vd and m was different with lubricant type and die material [68]. In addition, the Vd and α simultaneously affected the temperature rise of the wire [69], which means that the optimum wire drawing condition varies with Vd depending on the α. The effect of these complex process conditions was not considered in this study.

5. Conclusions

Based on a parametric study of the effects of the contact conditions at the wire–die interface on the temperature distributions of the specimen and die during wire drawing, the following conclusions were obtained:
  • The Tmax of the die decreased with increasing contact heat transfer coefficient, whereas that of the wire increased with contact heat transfer coefficient. The heat generated by friction at the wire–die interface was readily transferred from the die to the wire with an increasing contact heat transfer coefficient.
  • The Tmax of the die and wire decreased with increasing thermal conductivity of the die. As the thermal conductivity of the die increased, the heat generated by friction was rapidly absorbed into the die, thus causing the Tmax of the die to decrease. Accordingly, the surface temperature of the wire at the contact region decreased with the thermal conductivity of the die.
  • The Tmax of both the die and wire linearly increased with the friction factor. In particular, the Tmax of the die more sensitively changed with the friction factor compared with that of the wire.
  • The temperature gradient along the radial direction of the wire increased with the drawing velocity because the heat generated by friction at the wire–die interface did not have enough time to be transferred into the wire interior or atmosphere. Meanwhile, the Tmax of the die linearly increased with drawing velocity, whereas that of the wire parabolically increased with drawing velocity.
  • The Tmax of the die and wire increased with the contact length of the wire and die owing to the longer wire and die contact time. By contrast, the effect of the bearing length on the temperature increase of both the wire and die was insignificant.

Author Contributions

J.-K.H. designed, wrote, and reviewed the paper and Y.-C.C. performed the data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Education and Research promotion program of KOREATECH in 2022” and “Regional Innovation Strategy (RIS)” through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (MOE)(2021RIS-004).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Cpspecific heat (J/kg/K)
dfdrawn wire diameter (mm)
doinitial wire diameter (mm)
hccontact heat transfer coefficient (kW/m2/°C)
Kstrain hardening coefficient (MPa)
kshear yield stress of the material (MPa)
kdthermal conductivity of die (W/m/°C)
kwthermal conductivity of wire (W/m/°C)
Lbbearing length (mm)
Lccontact length (mm)
mshear friction factor
nstrain hardening exponent
Rpreduction in area per pass (%)
Ttemperature (°C)
ΔTitemperature rise from ideal plastic deformation (°C)
ΔTftemperature increase caused by frictional work (°C)
ΔTrtemperature increase caused by redundant work (°C)
ΔTttotal temperature increase of wire (°C)
Vddrawing velocity (m/s)
αsemi-die angle (degree)
βfraction factor between the mechanical work and heat energy
ρdensity (kg/m3)
τshear stress on the contact surface (MPa)

References

  1. Newbury, B.D.; Notis, M.R. The history and evolution of wiredrawing techniques. JOM 2004, 56, 33–37. [Google Scholar]
  2. Strzępek, P.; Mamala, A.; Zasadzińska, M.; Franczak, K.; Jurkiewicz, B. Research on the drawing process of Cu and CuZn wires obtained in the cryogenic conditions. Cryogenics 2019, 100, 11–17. [Google Scholar] [CrossRef]
  3. Kauffmann, A.; Freudenberger, J.; Klauß, H.; Schillinger, W.; Subramanya Sarma, V.; Schultz, L. Efficiency of the refinement by deformation twinning in wire drawn single phase copper alloys. Mater. Sci. Eng. A 2015, 624, 71–78. [Google Scholar]
  4. Hwang, J.K. Effect of drawing speed on microstructure distribution and drawability in twinning-induced plasticity steel during wire drawing. J. Iron Steel Res. Int. 2020, 27, 577–587. [Google Scholar]
  5. Hillery, M.T.; McCabe, V.J. Wire drawing at elevated temperatures using different die materials and lubricants. J. Mater. Process. Technol. 1995, 55, 53–57. [Google Scholar]
  6. Wei, D.; Min, X.; Hu, X.; Xie, Z.; Fang, F. Microstructure and mechanical properties of cold drawn pearlitic steel wires: Effects of drawing-induced heating. Mater. Sci. Eng. A 2020, 184, 139341. [Google Scholar] [CrossRef]
  7. Liu, J.P.; Chen, J.X.; Liu, T.W.; Li, C.; Chen, Y.; Dai, L.H. Superior strength-ductility CoCrNi medium-entropy alloy wire. Scr. Mater. 2020, 181, 19–24. [Google Scholar]
  8. Kubota, H.; Akimoto, Y.; Saito, K.; Sakurazawa, W.; Yoshida, K. Residual Stress Control in Drawn Bar and Wire by Heating-Cooling-Drawing Process. ISIJ Int. 2021, 61, 2792–2797. [Google Scholar] [CrossRef]
  9. Lee, S.K.; Lee, S.B.; Kim, B.M. Process design of multi-stage wet wire drawing for improving the drawing speed for 0.72 wt% C steel wire. J. Mater. Process. Technol. 2010, 210, 776–783. [Google Scholar] [CrossRef]
  10. Lee, S.K.; Kim, D.W.; Jeong, M.S.; Kim, B.M. Evolution of axial surface residual stress in 0.82 carbon steel wire during multi pass drawing process considering heat generation. Mater. Des. 2012, 34, 363–371. [Google Scholar] [CrossRef]
  11. Lee, S.K.; Ko, D.C.; Kim, B.M. Pass schedule of wire drawing process to prevent delamination for high strength steel cord wire. Mater. Des. 2009, 30, 2919–2927. [Google Scholar]
  12. Jo, H.H.; Lee, S.K.; Kim, M.A.; Kim, B.M. Pass schedule design system in the dry wire-drawing process of high carbon steel. Proc. IMechE Part B J. Eng. Manuf. 2002, 216, 365–373. [Google Scholar]
  13. Muskalski, Z. Selected problems from the high-carbon steel wire drawing theory and technology. Arch. Metall. Mater. 2014, 59, 527–535. [Google Scholar] [CrossRef]
  14. Suliga, M. Analysis of the heating of steel wires during high speed multipass drawing process. Arch. Metall. Mater. 2014, 59, 1475–1480. [Google Scholar]
  15. Hwang, J.K. Fracture behavior of twinning-induced plasticity steel during wire drawing. J. Mater. Res. Technol. 2020, 9, 4527–4537. [Google Scholar]
  16. Haddi, A.; Imad, A.; Vega, G. Analysis of temperature and speed effects on the drawing stress for improving the wire drawing process. Mater. Des. 2011, 32, 4310–4315. [Google Scholar] [CrossRef]
  17. Cetinarslan, C.S.; Guzey, A. Tensile properties of cold-drawn low-carbon steel wires under different process parameters. Mater. Technol. 2013, 47, 245–252. [Google Scholar]
  18. Vega, G.; Haddi, A.; Imad, A. Temperature effects on wire-drawing process: Experimental investigation. Int. J. Mater. Form. 2009, 2, 229–232. [Google Scholar] [CrossRef]
  19. Suliga, M.; Kruzel, R.; Garstka, T.; Gazdowicz, J. The influence of drawing speed on structure changes in high carbon steel wires. Metalurgija 2015, 54, 161–164. [Google Scholar]
  20. Pilarczyk, J.W.; Markowski, J.; Dyja, H.; Golis, B. FEM modeling of drawing of wires for prestressed concrete. Wire J. Int. 2004, 37, 118c123. [Google Scholar]
  21. Nemec, I.; Golis, B.; Pilarczyk, J.W.; Budzik, R.; Waszkielewicz, W. Effect of high-speed drawing on properties of high-carbon steel wires. Wire J. Int. 2007, 40, 63–68. [Google Scholar]
  22. Rittmann, K. Temperaturerhöhungen der Ziehholoberfläche beim Kaltziehen von Stabstahl mit großen Ziehgeschwindigkeiten. Arch. Für Das Eisenhüttenwesen 1971, 42, 265–271. [Google Scholar] [CrossRef]
  23. Zhang, G.L.; Wang, Z.W.; Zhang, S.H.; Cheng, M.; Song, H.W. A fast optimization approach for multipass wire drawing processes based on the analytical model. Proc. IMechE Part B J. Eng. Manuf. 2013, 227, 1023–1031. [Google Scholar]
  24. Strzępek, P.; Mamala, A.; Zasadzińska, M.; Kiesiewicz, G.; Knych, T.A. Shape analysis of the elastic deformation region throughout the axi-symmetric wire drawing process of ETP grade copper. Materials 2021, 14, 4713. [Google Scholar]
  25. Suliga, M.; Jabłońska, M.; Hawryluk, M. The effect of the length of the drawing die sizing portion on the energy and force parameters of the medium-carbon steel wire drawing process. Arch. Metall. Mater. 2019, 64, 1353–1359. [Google Scholar]
  26. Farr, J.P.G. Molybdenum disulphide in lubrication. A review. Wear 1975, 35, 1–22. [Google Scholar]
  27. Vadiraj, A.; Kamaraj, M.; Sreenivasan, V.S. Effect of solid lubricants on friction and wear behaviour of alloyed gray cast iron. Sadhana 2012, 37, 569–577. [Google Scholar] [CrossRef] [Green Version]
  28. Winer, W.O. Molybdenum disulfide as a lubricant: A review of the fundamental knowledge. Wear 1967, 10, 422–452. [Google Scholar]
  29. Hwang, J.K. Thermal Behavior of a Rod during Hot Shape Rolling and Its Comparison with a Plate during Flat Rolling. Processes 2020, 8, 327. [Google Scholar]
  30. Hwang, J.K. Effect of Air Temperature on the Thermal Behavior and Mechanical Properties of Wire Rod Steel during Stelmor Cooling. ISIJ Int. 2022, 62, 2343–2354. [Google Scholar]
  31. Kemp, I.P.; Pollard, G.; Bramley, A.N. Temperature distributions in the high speed drawing of high strength steel wire. Int. J. Mech. Sci. 1985, 27, 803–811. [Google Scholar] [CrossRef]
  32. Curtze, S.; Kuokkala, V.T. Dependence of tensile deformation behavior of TWIP steels on stacking fault energy, temperature and strain rate. Acta Mater. 2010, 58, 5129–5141. [Google Scholar]
  33. Yang, H.K.; Zhang, Z.J.; Dong, F.Y.; Duan, Q.Q.; Zhang, Z.F. Strain rate effects on tensile deformation behaviors for Fe-22Mn-0.6C-(1.5Al) twinning-induced plasticity steel. Mater. Sci. Eng. A 2014, 607, 551–558. [Google Scholar]
  34. Lee, H.W.; Kwon, H.C.; Im, Y.T.; Hodgson, P.D.; Zahiri, S.H. Local austenite grain size distribution in hot bar rolling of AISI 4135 steel. ISIJ Int. 2005, 45, 706–712. [Google Scholar] [CrossRef]
  35. Neu, R.W. Performance and characterization of TWIP steels for automotive applications. Mater. Perform. Charact. 2013, 2, 244–284. [Google Scholar]
  36. Moon, C.; Kim, N. Analysis of wire-drawing process with friction and thermal conditions obtained by inverse engineering. J. Mech. Sci. Technol. 2012, 26, 2903–2911. [Google Scholar] [CrossRef]
  37. Na, D.H.; Lee, Y. A study to predict the creation of surface defects on material and suppress them in caliber rolling process. Int. J. Precis. Eng. Manuf. 2013, 14, 1727–1734. [Google Scholar] [CrossRef]
  38. Nalawade, R.S.; Puranik, A.J.; Balachandran, G.; Mahadik, K.N.; Balasubramanian, V. Simulation of hot rolling deformation at intermediate passed and its industrial validity. Int. J. Mech. Sci. 2013, 77, 8–16. [Google Scholar]
  39. Bagheripoor, M.; Bisadi, H. Effects of rolling parameters on temperature distribution in the hot rolling of alumimum strips. Appl. Therm. Eng. 2011, 31, 1556–1565. [Google Scholar] [CrossRef]
  40. Wu, J.; Liang, X.; Tang, A.; Pan, F. Interfacial heat transfer under mixed lubrication condition for metal rolling: A theoretical calculation study. Appl. Therm. Eng. 2016, 106, 1002–1009. [Google Scholar] [CrossRef]
  41. Serajzadeh, S.; Karimi, T.A.; Nejati, M.; Izadi, J.; Fattahi, M. An investigation on strain inhomogeneity in hot strip rolling process. J. Mater. Process. Technol. 2002, 128, 88–99. [Google Scholar]
  42. Malinowski, Z.; Lenard, J.G.; Davies, M.E. A study of the heat-transfer coefficient as a function of temperature and pressure. J. Mater. Process. Technol. 1994, 41, 125–142. [Google Scholar] [CrossRef]
  43. Hwang, J.K. Comparison of Temperature Distribution between TWIP and Plain Carbon Steels during Wire Drawing. Materials 2022, 15, 8696. [Google Scholar] [CrossRef]
  44. Chen, C.; Sun, M.; Wang, B.; Zhou, J.; Jiang, Z. Recent Advances on Drawing Technology of Ultra-Fine Steel Tire Cord and Steel Saw Wire. Metals 2021, 10, 1590. [Google Scholar] [CrossRef]
  45. Lazzarotto, L.; Dubar, L.; Dubois, A.; Ravassard, P.; Oudin, J. Identification of Coulomb's friction coefficient in real contact conditions applied to a wire drawing process. Wear 1997, 211, 54–63. [Google Scholar]
  46. Haddi, A.; Imad, A.; Vega, G. The influence of the drawing parameters and temperature rise on the prediction of chevron crack formation in wire drawing. Int. J. Fract. 2012, 176, 171–180. [Google Scholar]
  47. Radionova, L.V.; Gromov, D.V.; Svistun, A.S.; Lisovskiy, R.A.; Faizov, S.R.; Glebov, L.A.; Zaramenskikh, S.E.; Bykov, V.A.; Erdakov, I.N. Mathematical modeling of heating and strain aging of steel during high-speed wire drawing. Metals 2022, 12, 1472. [Google Scholar] [CrossRef]
  48. Huang, M.; Fu, Y.; Qiao, X.; Chen, P. Investigation into Friction and Wear Characteristics of 316L Stainless-Steel Wire at High Temperature. Materials 2023, 16, 213. [Google Scholar] [CrossRef] [PubMed]
  49. Liu, X.Y.; Zhang, S.H. The design of a drawing die based on the logistic function for the energy analysis of drawing force. Appl. Math. Model. 2022, 109, 833–847. [Google Scholar]
  50. Lo, S.W.; Lu, Y.H. Wire drawing dies with prescribed variations of strain rate. J. Mater. Process. Technol. 2002, 123, 212–218. [Google Scholar]
  51. Lowrie, J.; Ngaile, G. Analytical modeling of hydrodynamic lubrication in a multiple-reduction drawing die. J. Manuf. Process 2017, 27, 291–303. [Google Scholar]
  52. Suliga, M.; Wartacz, R.; Hawryluk, M.; Kostrzewa, J. Effect of Drawing in Conventional and Hydrodynamic Dies on Structure and Corrosion Resistance of Hot-Dip Galvanized Zinc Coatings on Medium-Carbon Steel Wire. Materials 2022, 15, 6728. [Google Scholar]
  53. Nilsson, M.; Olsson, M. Tribological testing of some potential PVD and CVD coatings for steel wire drawing dies. Wear 2011, 273, 55–59. [Google Scholar] [CrossRef]
  54. Siegert, K.; Möck, A. Wire drawing with ultrasonically oscillating dies. J. Mater. Process. Technol. 1996, 60, 657–660. [Google Scholar]
  55. Pohlman, R.; Lehfeldt, E. Influence of ultrasonic vibration on metallic friction. Ultrasonics 1966, 4, 178–185. [Google Scholar]
  56. Liu, S.; Xie, T.; Han, J.; Shan, X. Stress superposition effect in ultrasonic drawing of titanium wires: An experimental study. Ultrasonics 2022, 125, 106775. [Google Scholar] [CrossRef]
  57. Stolyarov, A.; Polyakova, M.; Atangulova, G.; Alexandrov, S. Effect of die angle and frictional conditions on fine grain layer generation in multipass drawing of high carbon steel wire. Metals 2020, 10, 1462. [Google Scholar]
  58. Suliga, M.; Wartacz, R.; Michalczyk, J. High speed multi-stage drawing process of hot-dip galvanised steel wires. Int. J. Adv. Manuf. Technol. 2022, 120, 7639–7655. [Google Scholar]
  59. Wei, D.; Wang, L.; Hu, X.; Mao, X.; Xie, Z.; Fang, F. Effect of drawing strain rate on microstructure and mechanical properties of cold-drawn pearlitic steel wires. J. Mater. Sci. 2022, 57, 8924–8939. [Google Scholar] [CrossRef]
  60. Santana Martinez, G.A.; Qian, W.L.; Kabayama, L.K.; Prisco, U. Effect of process parameters in copper-wire drawing. Metals 2020, 10, 105. [Google Scholar] [CrossRef]
  61. Alexandrov, S.; Hwang, Y.M.; Tsui, H.S.R. Determining the Drawing Force in a Wire Drawing Process Considering an Arbitrary Hardening Law. Processes 2022, 10, 1336. [Google Scholar]
  62. Överstam, H. The influence of bearing geometry on the residual stress state in cold drawn wire, analysed by the FEM. J. Mater. Process. Technol. 2006, 171, 446–450. [Google Scholar] [CrossRef]
  63. Luis, C.J.; Leon, J.; Luri, R. Comparison between finite element method and analytical methods for studying wire drawing processes. J. Mater. Process. Technol. 2005, 164–165, 1218–1225. [Google Scholar] [CrossRef]
  64. Luksza, J.; Majta, J.; Burdek, M.; Ruminski, M. Modeling and measurements of mechanical behaviour in multi-pass drawing process. J. Mater. Process. Technol. 1998, 80–81, 398–405. [Google Scholar] [CrossRef]
  65. Chin, R.K.; Stelf, P.S. A computational study of strain inhomogeneity in wire drawing. Int. J. Mach. Tools Manuf. 1995, 35, 1087–1098. [Google Scholar]
  66. Karimi Taheri, A.; Maccagno, T.M.; Jonas, J.J. Dynamic Strain Aging and the Wire Drawing of Low Carbon Steel Rods. ISIJ Int. 1995, 35, 1532–1540. [Google Scholar]
  67. Kabayama, L.K.; Taguchi, S.P.; Martínez, G.A.S. The influence of die geometry on stress distribution by experimental and FEM simulation on electrolytic copper wiredrawing. Mat. Res. 2009, 12, 281–285. [Google Scholar] [CrossRef]
  68. Martínez, G.A.S.; Rodriguez-Alabanda, O.; Prisco, U.; Tintelecan, M.; Kabayama, L.K. The influences of the variable speed and internal die geometry on the performance of two commercial soluble oils in the drawing process of pure copper fine wire. Int. J. Adv. Manuf. Technol. 2022, 118, 3749–3760. [Google Scholar]
  69. Prisco, U.; Martinez, G.A.S.; Kabayama, L.K. Effect of die pressure on the lubricating regimes achieved in wire drawing. Prod. Eng. 2020, 14, 667–676. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration showing (a) heat transfer mechanisms in wire drawing process and (b) general parameters affecting temperature distribution of specimen and die during wire drawing.
Figure 1. Schematic illustration showing (a) heat transfer mechanisms in wire drawing process and (b) general parameters affecting temperature distribution of specimen and die during wire drawing.
Processes 11 00513 g001
Figure 2. (a) Photograph of obtained hot-rolled wire rod steel, (b) its microstructure, and (c) true stress–strain curve in tensile test.
Figure 2. (a) Photograph of obtained hot-rolled wire rod steel, (b) its microstructure, and (c) true stress–strain curve in tensile test.
Processes 11 00513 g002
Figure 3. Schematic illustration of wire drawing process performed in this study.
Figure 3. Schematic illustration of wire drawing process performed in this study.
Processes 11 00513 g003
Figure 4. Comparison of core temperature and drawing force between experiment and FEA using low-carbon steel wire rod.
Figure 4. Comparison of core temperature and drawing force between experiment and FEA using low-carbon steel wire rod.
Processes 11 00513 g004
Figure 5. Comparison of temperature (a) contours and (b) profiles along the radial direction of wire at die exit; (c) Tmax of wire and die vs. contact heat transfer coefficient.
Figure 5. Comparison of temperature (a) contours and (b) profiles along the radial direction of wire at die exit; (c) Tmax of wire and die vs. contact heat transfer coefficient.
Processes 11 00513 g005
Figure 6. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. thermal conductivity of die.
Figure 6. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. thermal conductivity of die.
Processes 11 00513 g006aProcesses 11 00513 g006b
Figure 7. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. shear fraction factor.
Figure 7. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. shear fraction factor.
Processes 11 00513 g007
Figure 8. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. drawing velocity.
Figure 8. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. drawing velocity.
Processes 11 00513 g008
Figure 9. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. contact length.
Figure 9. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. contact length.
Processes 11 00513 g009
Figure 10. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. bearing length.
Figure 10. Comparison of temperature (a) contours and (b) profiles along radial direction of wire at die exit; (c) Tmax of wire and die vs. bearing length.
Processes 11 00513 g010aProcesses 11 00513 g010b
Figure 11. Schematic illustration showing effects of contact conditions on temperature increase in (a) specimen and (b) die during wire drawing.
Figure 11. Schematic illustration showing effects of contact conditions on temperature increase in (a) specimen and (b) die during wire drawing.
Processes 11 00513 g011
Table 2. Comparison of equilibrium drawing force and core temperature between experiment and FEA.
Table 2. Comparison of equilibrium drawing force and core temperature between experiment and FEA.
ParameterExperimentFEAError (%)
Equilibrium drawing force (kN)22.5 ± 1.8 22.8 ± 1.41.4
Equilibrium core temperature (°C)72.2 ± 3.775.1 ± 0.34.0
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.; Chang, Y.-C. Effects of Contact Conditions at Wire–Die Interface on Temperature Distribution during Wire Drawing. Processes 2023, 11, 513. https://doi.org/10.3390/pr11020513

AMA Style

Hwang J-K, Chang Y-C. Effects of Contact Conditions at Wire–Die Interface on Temperature Distribution during Wire Drawing. Processes. 2023; 11(2):513. https://doi.org/10.3390/pr11020513

Chicago/Turabian Style

Hwang, Joong-Ki, and Young-Chul Chang. 2023. "Effects of Contact Conditions at Wire–Die Interface on Temperature Distribution during Wire Drawing" Processes 11, no. 2: 513. https://doi.org/10.3390/pr11020513

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