Next Article in Journal
Effect of the Annealing Treatment on the Microstructure and Corrosion Resistance of a Cu-Al-Ni Alloy
Previous Article in Journal
The Rollers’ Offset Position Influence on the Counter-Roller Flow-Forming Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mathematical Modeling of Heating and Strain Aging of Steel during High-Speed Wire Drawing

by
Liudmila V. Radionova
1,
Dmitry V. Gromov
1,
Alexandra S. Svistun
1,
Roman A. Lisovskiy
2,
Sergei R. Faizov
2,
Lev A. Glebov
2,
Sergei E. Zaramenskikh
2,
Vitaly A. Bykov
2 and
Ivan N. Erdakov
2,*
1
Department of Metallurgy, Moscow Polytechnic University, Bolshaya Semyonovskaya Street 38, Moscow 107023, Russia
2
Department of Metal Forming, South Ural State University, Lenin Prospect 76, Chelyabinsk 454080, Russia
*
Author to whom correspondence should be addressed.
Metals 2022, 12(9), 1472; https://doi.org/10.3390/met12091472
Submission received: 6 August 2022 / Revised: 29 August 2022 / Accepted: 29 August 2022 / Published: 3 September 2022

Abstract

:
In this article, a mathematical model of the wire’s average temperature change in the process of multiple drawing on high-speed straight-line drawing machines has been developed. The calculation results showed that the average temperature of the wire during a drawing at a speed of up to 45 m/s on straight-line drawing machines could reach 400 °C. Deformation heating of the wire during drawing does not exceed 60 °C, and heating due to sliding friction can reach 300 °C, depending on the friction coefficient, which ranges from 0.05 to 0.15. The average strain rates under the conditions of the modern high-speed drawing process reach 7000 s−1. Over the course of the research, it was found that there are no conditions for the occurrence of dynamic deformation aging due to impurity atoms of carbon, nitrogen and oxygen. At the same time, at the temperature and speed parameters of the high-speed wire drawing, conditions are created for the onset of the dynamic strain aging of steel in the presence of hydrogen atoms. Therefore, during heat treatment and pickling, it is necessary to exclude the hydrogenation of steel. It has been established that in order to exclude static strain aging of steel during drawing, it is necessary to prevent heating the wire above 180–200 °C.

1. Introduction

Wire drawing is a process of plastic deformation of metal, during which a round billet, when pulled through a tool called a die, reduces its cross-section and increases its length. A wire is a long metal product with a diameter of 0.005 to 8 mm of a cylindrical or shaped profile, made of steel or non-ferrous alloys [1].
Wire can be obtained through various metal forming processes, namely rolling [2,3], extrusion [4] and drawing [5,6,7], or an effective combination of them. However, drawing remains the most widespread method of wire production [8,9]. Overall, it seems simple, but this is one of the most difficult types of metal forming. The drawing process is influenced by a large number of various factors, such as drawing speed [10], technological lubrication [11,12,13], sub-lubricating layer [14], deformation degree [15,16], drawing die [17] and drawing machine design, etc.
Consumable tools for wire drawing are dies. Recently, the results of studies have been published to determine the optimal parameters of dies in terms of energy consumption [18,19], wear of the die [20], wire quality [21] and other parameters.
The entire history of the development and improvement of metal drawing is mainly associated with a decrease in the negative effect of friction. Research is underway to reduce the negative effect of friction by improving the lubrication coating [22], developing new types of technological lubricants [23], and optimizing the drawing dies [24]. At the same time, its advantages, such as greater accuracy of geometric shapes and high quality of the wire surface, make it competitive with the emergence of new methods and technologies for the production of wire. In the process of drawing, the wire in the die is heated due to the transformation of the deformation work into heat and the action of contact friction. A significant part of the heat (up to 90%) is accumulated by the wire, while the remaining part causes the die to heat up, especially at the metal–die contact [25].
The need to reduce the cost of wire by increasing productivity, as well as the possibility of using more automation in drawing production, leads to an increase in drawing speed. In turn, an increase in the drawing speed has a significant effect on the technological features of the wire drawing process, primarily on temperature conditions. Modern drawing machines include devices and measures to reduce the heating of the wire during the drawing process [26]. To reduce the heating of the wire during the drawing process, water cooling of the inner part of the pulling drums and the outer surface of the clips of the drawing dies in the soap boxes is used.
Much attention is paid to this issue, among other things, because over the past few decades, the equipment used for wire drawing has changed significantly [27]. Drawing machines of the previous generation—pulley wire drawing machines—made it possible to produce wire at a drawing speed of up to 8 m/s. With the development of production automation, which affected wire production, drawing machines of a new generation appeared capable of developing drawing speeds up to 45 m/s [10,28].
Straight-line drawing machines provide a cooling system for pulling drums, drawing dies, the use of new generation lubricants, etc. However, the question remains, which technological and technical parameters of drawing create the conditions for the onset of strain aging of alloys?
Strain aging consists of blocking free dislocations by impurity atoms and inclusions, which in turn leads to an increase in the strain stress and loss of plastic properties [29,30,31,32]. At the same time, one should not forget that strain aging only occurs if:
(1) If a certain number of "fresh" dislocations is introduced into the metal as a result of deformation (or in another way).
(2) The concentration of impurity atoms that can effectively interact with these dislocations exceeds 10–4 wt %.
A decrease in the mobility of dislocations introduced by deformation due to their blocking by carbon, nitrogen, hydrogen and oxygen atoms is possible only if the dislocation velocity is commensurate with the diffusion rate of impurity atoms determined for a given temperature range and strain rates [7,33].
Previous studies [34,35,36] considered the process of drawing and strain aging during its implementation for the conditions and technologies used earlier. The study [34] evaluated the ductility of carbon steel wire at a strain rate of 10–4 to 10–3 s–1. The paper [35] presents data on the change in the diffusion coefficient depending on the steel temperature. Knowing the value of the diffusion coefficient of hydrogen makes it possible to determine the possibility of blocking dislocations during plastic deformation and premature loss of plastic properties. The study [36] shows the intensity of the temperature increase when the drawing speed changes from 1 to 7 m/s, which corresponds to the drawing machines of the previous generation. The maximum temperature during such drawing does not exceed 70 °C. In studies [37,38], the strain rates were calculated during drawing with a speed of 20 m/s. Using the FEM method, the distribution of strain rates over the wire cross-section was shown with a change in the angle of the die.
The purpose of this study is to determine the conditions for the occurrence of dynamic and static strain aging during steel wire drawing on modern high-speed straight-line machines. The scientific novelty of the study lies in the theoretical analysis of the conditions for dynamic and static strain aging during high-speed wire drawing in monolithic dies.

2. Modeling the Temperature of Wire Drawing on Straight-Line Drawing Machines

A significant part of the work spent on the process of forming is converted into thermal energy, which is accumulated by the metal being processed and heating up the tool. The amount of energy that will be converted into heat significantly depends on the process used to obtain the wire and the technological features during its implementation. Contact friction has the greatest influence on the heating of the metal in the forming process. During rolling and roll drawing, rolling friction is realized, due to which the heating of the metal during deformation is not so significant. In monolithic drawing, due to sliding friction, the heating of the wire and die can be significant. During repeated drawing, the temperature of the wire is determined by two opposite processes: heating in the drawing dies and cooling on the pulling drums.
The model of temperature changes that occur during wire drawing consists of two parts: the first is the influx of heat as a result of the work of friction and deformation forces, and the second is the outflow of heat to the die, to the drum of the drawing machine and the air space of the environment. Let us make a balance of inflow and outflow of heat.
The heat influx is the sum of the heat release caused by the work of deformation and friction at the wire–die contact and the heat brought in from heating in the previous drawing die. The outflow of heat is distributed between the drawing die and the metal. The heat balance during drawing will be as follows:
q d e f + q f r + q w / i n = q d i e + q w / o u t
where q d e f is the heat release in the metal caused by the work of deformation, W/m2;
q f r —heat release in metal caused by friction, W/m2;
q w / i n —heat introduced into the deformation zone by the wire from heating in the previous pass, W/m2;
q d i e —heat given off to the die, W/m2;
q w / o u t —heat carried away by the metal, W/m2.
When moving the wire from the i die to i + 1 , it loses heat to the environment by convection and radiation. Heat losses from radiation are insignificant, so we neglect this type of loss [1]. To determine the heat loss from convection, we divide the area from i to i + 1 of the die into three areas (Figure 1).
The heat q w / o u t is transferred to the environment (Sections I and III) and to the wire drawing drum (Section II) (Figure 1).
q w / o u t = q I + q I I + q I I I + q w / i n ,
where q I , q I I , q I I I is the heat outflow in the first, second and third areas, W/m2.
Hence,
q d f + q f r + q w / i n = q d i e + q I + q I I + q I I I + q w / o u t
The total heating of the wire is determined by adding the average temperature rise over the wire cross-section, caused by the heat of friction and the increase in wire temperature due to plastic deformation:
Δ t = Δ t d e f + Δ t f r ,
where Δ t d e f is the increase in wire temperature due to plastic deformation, °C; Δ t f r is the average temperature increase over the wire section, caused by the heat of friction, °C.
The increase in wire temperature due to plastic deformation is defined as
Δ t d e f = q d e f c ρ ,
where c is the heat capacity, J/kg·K; ρ is the metal density, kg/m3.
Heat release in the metal due only to the deformation work, W/m2:
q d e f = σ s ln μ ,  
where σ s is the yield strength alloy of the wire, N/m2; ln μ is the draw ratio.
The average temperature rise over the wire cross-section caused by the heat of friction is defined as
Δ t f r = t max δ d δ 2 0.5 d 2 ,
where t max is the maximum heating temperature of the surface of the wire and die, °C;
δ —thickness of the metal layer heated near the friction surface, m;
d —wire diameter, m.
Then, the maximum heating temperature of the contact surfaces between the wire and the die:
t max = q w / o u t × B ,
where B is the coefficient determined by the formula
B = 4 l π c ρ λ 0 v ,
where λ 0 is the thermal conductivity of the die, W/(m·K);
v —drawing speed, m/s; l —length of the contact zone, m.
The thickness of the metal layer heated near the friction surface
δ = k λ 0 l c ρ v ,
where k is a coefficient taken equal to 1.65.
The heat q f r generated as a result of the work of friction forces on the contact surfaces can be divided into the heat q d i e absorbed by the die and the heat q w / o u t absorbed by the moving wire.
Heat absorbed by the die, including heat inflow and outflow, contact conditions (i.e., coefficient of friction) and the effect of back tension, is defined as
q d i e = q f r B t d i e / o u t R λ 1 ln R o u t R + B ,
where t d i e / o u t is the outside die temperature, °C;
R o u t —outer radius of die, m;
λ 1 —thermal conductivity coefficient of the wire, W/(m·K);
R —arithmetic mean between the input die radius and the output one, m.
R = 0.5 ( R 0 + R 1 )
The heat release in the metal caused by contact friction is defined as
q f r = f σ N v ,
where σ N is the normal pressure of the metal on the die, MPa; f is the friction coefficient.
The normal pressure of the metal on the die is determined by the formula
σ N = σ d r a w R 1 2 sin α ( R 0 2 R 1 2 ) sin ( α + a r c t g ( f ) ) ,
where α is a die half-angle.
Heat absorbed by moving wire
q w / o u t = q f r R λ 1 ln R o u t R + t d i e / o u t R λ 1 ln R o u t R + B
The first area is from the i -th die to the drum. In this area, the wire is straight, and the heat exchange between the wire and the environment can be considered as in the case of a longitudinal flow around the cylinder. According to:
f o r R e < 10 3 t o N u = 0.49 R e 0.5
f o r R e > 10 3 t o N u = 0.245 R e 0.6 ,
where R e is the Reynolds criterion; N u —Nuselt criterion; L —length of the wire in the first area, m.
Reynolds criterion:
Re = v L η ,
where η is the kinematic viscosity of air, g·m/s.
Considering the correction factor for the angle of attack of the air flow to the wire, the heat transfer coefficient is determined as
α k = 0.4 N u λ a i r d i ,
where λ a i r is the thermal conductivity coefficient of air, N·s.
The wire at the end of the first section is cooled to a temperature
t i I = ( t i t в ) exp ( 2 B i F o ) + t a i r ,
where B i is the Biot criterion; F o is the Fourier criterion; t a i r —ambient air temperature, °С; τ —cooling time, s.
Biot criterion:
B i = α k λ a l l o y × d i 2 ,
where λ a l l o y is the thermal conductivity coefficient of the drawn metal, W/(m·K).
Fourier criterion:
F o = λ a l l o y × d i τ c ρ × ( d i 2 ) 2
The second area is the drum of the drawing machine. The wire on the drum transfers heat to the environment and the drum. We assume that in a stationary process, the wire and the drum have the same temperature; therefore, the calculation of heat losses is carried out as for a vertical cylinder. Considering the correction factor for the angle of attack of the air flow to the drum, the heat transfer coefficient in this section is determined by the formula
α k I I = φ k × 1.04 × λ a i r ( v η D d ) 0.5 P L 1 3 ,
where D d is the diameter of the drum, considering the wire located on it, m; φ k —correction factor for the angle of attack of the air flow 90°; P L —Prandtl coefficient for the cooling air.
Correction factor for 90° airflow angle of attack φ k = 0.55 and Prandtl factor for cooling air P L = 0.703 .
The decrease in the wire temperature in the second area:
t i I I = ( t i t a i r ) exp ( 2 B i F o ) + t a i r
The third area is from the point of descent of the wire from the drum to i + 1 die. The decrease in temperature in the third area is defined as
t i I I I = ( t i t a i r ) exp ( 2 B i F o ) + t a i r
Next, you need to combine the two parts of the heat balance. The wire after each drawing in the die is heated and then cooled, more intensively on the drawing machine drum and less intensively when interacting with the environment in the second and third areas. Therefore, before entering the next die, the heating of the wire will be
t i 0 = t i 1 t i 1 I I I ,
where t i 0 is the temperature of the wire before entering the i -th die, °C.
Based on the described technique, a computer program was developed to simulate the temperature regime of multiple drawing. This computer program allows you to estimate the temperature of the wire along the entire drawing route, as well as determine the change in temperature depending on the technological parameters of drawing.
The average wire temperatures for the passes are given in Table 1. Since a straight-line drawing machine is used in the studies, the drawing speeds along the route increase in each subsequent die, taking into account the draw ratio according to the law of volume constancy. The drawing speed indicates the speed at the exit of the last (seventh) die. The steel (0.2 wt.% С) billet yield strength was σ s = 420 MPa. The calculation results with an error not exceeding 12% are consistent with the results of industrial studies of temperature conditions performed by us earlier [10]. Comparable results were obtained in the study [11].
As follows from the calculation results, when drawing a wire from a steel billet with a yield strength σ s = 420 MPa on straight-line drawing machines with a speed of 45 m/s in the last pass, the average temperature of the wire rises to 381 °C. The average temperature of the wire is greatly influenced by the friction coefficient. An increase in the friction coefficient of 0.05 to 0.15 leads to an increase in temperature by 71 °C. The drawing speed also causes the temperature of the wire to rise. With an increase in the drawing speed of 10 to 45 m/s at a friction coefficient f = 0.05, the average temperature rises from 270 °C to 353 °C.
Wire heating during drawing, according to Equation (4), consists of deformation and heating due to friction.
The effect of deformation parameters on deformation heating is shown in Figure 2.
To analyze the effect of technological parameters (die half-angle and friction coefficient) on the maximum temperature that occurs at the wire–die contact, a calculation was performed, the results of which are shown in Figure 3.
The thickness of the surface layer of the metal heated by friction along the drawing route ( f = 0.15) is given in Table 2.
The calculation results shown in Figure 2 demonstrate that the heating of the wire due to deformation does not exceed 60 °C, but the heating caused by friction at the wire–die contact (Figure 3) with poor preparation of the wire surface and poor-quality technological lubrication can reach 320 °C. It should be noted that with a high friction coefficient, an increase in the die half-angle of 3 to 8° leads to an increase in the maximum temperature by 100 °C.
As the calculation results show, the heating of the wire due to deformation does not exceed 60 °C, but the heating caused by friction can reach 400 °C. The maximum temperatures are localized in the surface layer, but due to the thermal conductivity of the metal, this heat is transferred to the depth of the wire. High-speed drawing processes, even with good heat removal on modern straight-line drawing machines, lead to heat accumulation in the wire and its heating up to 300–400 °C. High-speed drawing should be accompanied by good surface preparation of the workpiece and the use of high-quality lubricants.

3. Conditions for the Development of Dynamic and Static Strain Aging

To calculate the average strain rate ε ˙ during drawing, the authors in the study [10] propose to use the formula
ε ˙ = 6 tan α v ln μ d ( μ 3 1 ) .
The results of calculating the average strain rate depending on the technological parameters of the drawing process are shown in Figure 4.
As can be seen in Figure 4a, an increase in the drawing speed of 5–45 m/s with a wire diameter of 3 mm and a drawing ratio of 1.35 at different die half-angles leads to an increase in the average strain rate of 276 to 6616 s−1. Only by changing the die half-angle to 3–8° at a drawing speed of 45 m/s does the strain rate increase by 2482 to 6616 s−1, i.e., more than 2.5 times. Less but still significantly, the value of the elongation coefficient also affects the average strain rate (Figure 4b). A drop in the average strain rate by 2916 to 2277 s−1 is observed with an increase in the draw ratio of 1.1 to 1.5 at a drawing speed of 45 m/s. The diameter of the wire being processed has a significant effect on the average strain rate (Figure 4c, so a decrease in the diameter of 6 to 2 mm, other things being equal, leads to an increase in the average speed of 1656 to 4968 s−1, which is a threefold increase. For comparison, Figure 4d shows the technological conditions of the drawing process, under which the average strain rates are in the range of 20–180 s−1. Such significant changes in the strain rate affect the ductility of the alloy and its machinability. It is necessary to determine whether there are prerequisites for the occurrence of dynamic deformation processes of aging since it is the strain rates that play a decisive role in this. The conditions under which dynamic interaction between dislocations and impurity atoms is ensured are determined precisely by the critical strain rate, and as we can see, the spread of these values during drawing is possible from 20 up to 12,000 s−1. The diffusion coefficient is a characteristic that is very sensitive to temperature.
The critical strain rate at which a dynamic interaction between dislocations and impurity atoms is observed is defined as
ε ˙ c r = 2 × 10 2 × D ρ d
where ε ˙ c r is the rate of plastic deformation, s−1; D is the diffusion coefficient of dissolved atoms, m2/s; ρ d is the dislocation density, m−2.
The dependence of the diffusion coefficient on temperature obeys the Arrhenius law [39,40]:
D = D 0 exp ( Q k T )
where Q is the activation energy, eV; k is the Boltzmann constant; D 0 is the pre-exponential factor.
This dependence of the diffusion coefficient on temperature has been experimentally confirmed for many systems [39]. The parameters for determining impurity diffusion in α-Fe are presented in Table 3 [40].
For metal temperatures achieved during wire drawing, the change in diffusion coefficients is shown in Figure 5.
The density of dislocations during plastic deformation, according to [41,42], is in the range of 1012–1015 m−2 (Figure 6). The maximum density of dislocations during plastic deformation, according to the Bochvar-Oding curve, does not exceed 1015–1016 m−2. A further increase in the dislocation density leads to the formation of cracks.
Substituting the values of the diffusion coefficients for the elements C, N, H and O (Figure 5) and the density of dislocations ρ d = 1013 m−2 in Equation (26), we obtain the dependence of the critical strain rate (Figure 7) for the temperature range calculated for the drawing process under study (Table 1). As we see in Figure 7, the calculated critical strain rate for impurity hydrogen atoms (H) varies in the range of 2000 to 4200 s−1, which corresponds to the strain rate in the drawing process (Figure 4). At the same time, the critical strain rates for C, N, and O impurity atoms are in the range of 0 to 0.07 s. Such strain rates during drawing have not been found. Therefore, hydrogen atoms can be the cause of dynamic strain aging since the strain rates during drawing are commensurate with the critical strain rate for hydrogen atoms. This condition is met at a dislocation density ρ d = 1013 m−2, which corresponds to a degree of deformation of 7% (Figure 6).
As shown in Figure 6, the dislocation density in the drawing process takes values of 1012–1015. According to Equation (26), changing the dislocation density to 1012 without changing other parameters will allow us to obtain dependences similar to those in Figure 7, with the only difference being that the change in the critical strain rate must be reduced by a factor of 10. Therefore, the critical strain rate for hydrogen atoms will be in the range of 200 to 420 s−1. Increasing the dislocation density up to 1014 will lead to an increase in the critical strain rate for hydrogen atoms up to 200,000–420,000 s−1. If strain rates of about 200 s−1 can be obtained by drawing with a low drawing speed and small draw ratios (Figure 4d), then strain rates of 200,000 s−1 during drawing are unattainable.
The critical strain rates for impurity carbon, nitrogen, and oxygen atoms do not exceed 7 s−1 at a dislocation density of 1015.
Consequently, the calculations showed that the conditions for the occurrence of dynamic strain aging could only be created by hydrogen atoms. At the same time, we know that this requires a concentration of hydrogen atoms greater than 10−4%. The above analysis was conducted for ferrite, i.e., low-carbon steel.
Figure 8 shows the change in the diffusion coefficient depending on the temperature of steel with different structures [35].
The diffusion coefficient for carbon steel varies slightly with temperature and is about 105 m2/s. Substituting this value into Equation (26) shows that, depending on the dislocation density, the critical strain rate varies from 200,000 s−1 and above. There are no such strain rates during the drawing process, so we can talk about the aversion of carbon steel to dynamic strain aging. The diffusion coefficient of pearlite is close to that of ferrite. Therefore, when calculating the critical strain rate, we obtain values of 100 to 200,000 s−1. These values lie in the region of strain rates during drawing. Consequently, pearlite, just like ferrite, in the presence of impurity hydrogen atoms, is prone to dynamic strain aging.
The diffusion coefficient of austenitic steel is very sensitive to temperature. In the temperature range of 20–150 °C, the diffusion coefficient varies from 10™12 to 10™7 m2/s. Above a temperature of 150 °C, the diffusion coefficient remains constant. At such values, the critical strain rate varies from 0.2 to 200,000 s−1, depending on the dislocation density and steel temperature. Therefore, austenitic steel is also prone to dynamic strain aging in the presence of hydrogen atoms.
In the process of preparing the workpiece for drawing (heat treatment and etching) and the application of galvanic coatings, steel hydrogenation can occur. Therefore, great attention should be paid to measures that reduce or eliminate steel hydrogenation.
The start time of static strain aging, depending on the steel temperature, varies from several hours to fractions of a second (Table 4) [10,14].
From the analysis of Table 1, Table 2 and Table 4, it follows that the heating of the wire during drawing on high-speed straight-line machines provides the conditions for the passage of static strain aging of steel.
To eliminate the conditions for the occurrence of static strain aging of steel, it is necessary to limit the heating temperature of the wire to 180–200 °C.
To do this, it is necessary to pay great attention to the quality of the wire surface preparation before drawing, the properties of technological lubrication, and the cooling of the die and the pulling drum.
It is especially important to avoid winding the finished wire in a warm state onto a spool. A tightly wound wire at a temperature of 100–140 °C will stay on the spool for a long time, and this will allow the processes of static strain aging to take place, which will lead to a loss or decrease in the plastic properties of the wire.

4. Conclusions

The conducted studies have shown:
  • A mathematical model has been developed for calculating the wire average temperature during drawing on straight-line drawing machines;
  • The calculation results showed that the average temperature of the wire during drawing at a speed of up to 45 m/s on straight-line drawing machines can reach 400 °C;
  • Deformation heating of the wire during drawing does not exceed 60 °C, and heating due to sliding friction can reach 300 °C;
  • Average strain rates on modern high-speed drawing machines reach 7000 s−1;
  • In modern temperature and speed regimes of the drawing process, conditions are created for the occurrence of dynamic deformation aging of steel in the presence of hydrogen atoms;
  • During heat treatment and pickling, it is necessary to exclude the hydrogenation of steel;
  • To exclude static deformation aging of steel during drawing, it is necessary to prevent heating of the wire above 180–200 °C.

Author Contributions

Conceptualization, L.V.R. and I.N.E.; methodology, A.S.S. and D.V.G.; software, R.A.L.; validation, L.A.G. and V.A.B.; formal analysis, S.R.F.; investigation, L.V.R.; data curation, S.R.F.; writing—original draft preparation, L.V.R.; writing—review and editing, I.N.E.; visualization, S.E.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wright, R.N. Wire Technology, 2nd ed.; Butterworth-Heinemann: Oxford, UK, 2016; 311p. [Google Scholar] [CrossRef]
  2. Hwang, J.-K. Hardening and Softening Behavior of Caliber-Rolled Wire. Materials 2022, 15, 2939. [Google Scholar] [CrossRef] [PubMed]
  3. Hwang, J.-K.; Kim, S.-J.; Kim, K.-J. Influence of Roll Diameter on Material Deformation and Properties during Wire Flat Rolling. Appl. Sci. 2021, 11, 8381. [Google Scholar] [CrossRef]
  4. Faizov, S.; Sarafanov, A.; Erdakov, I.; Gromov, D.; Svistun, A.; Glebov, L.; Bykov, V.; Bryk, A.; Radionova, L. On the Direct Extrusion of Solder Wire from 52In-48Sn Alloy. Machines 2021, 9, 93. [Google Scholar] [CrossRef]
  5. 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] [CrossRef]
  6. Rodríguez-Alabanda, Ó.; Romero, P.E.; Molero, E.; Guerrero-Vaca, G. Analysis, Validation and Optimization of the Multi-Stage Sequential Wiredrawing Process of EN AW-1370 Aluminium. Metals 2019, 9, 1021. [Google Scholar] [CrossRef]
  7. Haddi, A.; Imad, A.; Vega, G. Analysis of temperature and speed effects on the drawing stress for improving the wire drawing process. Mater. Design 2011, 32, 4310–4315. [Google Scholar] [CrossRef]
  8. Oliveira Anício Costa, I.M.; Batková, M.; Batko, I.; Benabou, A.; Mesplont, C.; Vogt, J.-B. The Influence of Microstructure on the Electromagnetic Behavior of Carbon Steel Wires. Crystals 2022, 12, 576. [Google Scholar] [CrossRef]
  9. 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]
  10. Radionova, L.V.; Shirokov, V.V.; Faizov, S.R.; Zhludov, M.A. Studies of Influence of Process Parameters on the Strain Rate at High-Speed Wire Drawing. MSF 2019, 946, 832–838. [Google Scholar] [CrossRef]
  11. Suliga, M.; Wartacz, R.; Michalczyk, J. The influence of the angle of the working part of the die on the high speed drawing processof low carbon steel wires. Metall. Mater. Trans. A 2017, 62, 483–487. [Google Scholar] [CrossRef]
  12. 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] [CrossRef]
  13. Tzou, G.-Y.; Chai, U.-C.; Hsu, C.-M.; Hsu, H.-Y. FEM simulation analysis of wire rod drawing process using the rotating die under Coulomb friction. In MATEC Web of Conferences; EDP Sciences: Les Ulis, France, 2017; Volume 123, p. 00033. [Google Scholar] [CrossRef]
  14. Radionova, L.V.; Kharitonov, V.A.; Zyuzin, V.I.; Rol’shchikov, L.D. New technological lubricants for steel wire drawing. Steel Trans. 2001, 12, 49–50. [Google Scholar]
  15. Hwang, J.-K.; Yi, I.-C.; Son, I.-H.; Yoo, J.-Y.; Kim, B.; Zargaran, A.; Kim, N.J. Microstructural evolution and deformation behavior of twinning-induced plasticity (TWIP) steel during wire drawing. Mater. Sci. Eng. A 2015, 644, 41–52. [Google Scholar] [CrossRef]
  16. Radionova, L.V.; Lisovskiy, R.A.; Svistun, A.S.; Erdakov, I.N. Change in Mechanical Properties During Drawing of Wire from Ni 99.6. In Materials Science Forum; Trans Tech Publications, Ltd.: Wollerau, Switzerland, 2022. [Google Scholar] [CrossRef]
  17. Massé, T.; Fourment, L.; Montmitonnet, P.; Bobadilla, C.; Foissey, S. The optimal die semi-angle concept in wire drawing, examined using automatic optimization techniques. Int. J. Mater. Form. 2013, 6, 377–389. [Google Scholar] [CrossRef]
  18. Majzoobi, G.H.; Saniee, F.F.; Aghili, A. An investigation into the effect of redundant shear deformation in bar drawing. J. Mater. Processing Technol. 2008, 201, 133–137. [Google Scholar] [CrossRef]
  19. Radionova, L.V.; Lisovsky, R.A.; Lezin, V.D. Theory of Energy Conservation as the Basis for the Design of Wire Drawing. In Proceedings of the 6th International Conference on Industrial Engineering (ICIE 2020). ICIE 2021. Lecture Notes in Mechanical Engineering; Radionov, A.A., Gasiyarov, V.R., Eds.; Springer: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  20. Kumar, R.; Singh, S.; Aggarwal, V.; Singh, S.; Pimenov, D.Y.; Giasin, K.; Nadolny, K. Hand and Abrasive Flow Polished Tungsten Carbide Die: Optimization of Surface Roughness, Polishing Time and Comparative Analysis in Wire Drawing. Materials 2022, 15, 1287. [Google Scholar] [CrossRef]
  21. Milenin, A.; Wróbel, M.; Kustra, P.; Němeček, J. Experimental and Numerical Study of Surface Roughness of Thin Brass Wire Processed by Different Dieless Drawing Processes. Materials 2022, 15, 35. [Google Scholar] [CrossRef]
  22. Liu, S.; Shan, X.; Cao, H.; Xie, T. Finite Element Analysis on Ultrasonic Drawing Process of Fine Titanium Wire. Metals 2020, 10, 575. [Google Scholar] [CrossRef]
  23. Reshmin, Y.A.; Frolov, V.I.; Stolyarov, A.Y.; Zhilkina, V.F. Use of an Emulsion Based on the Lubricant Sinapol in Draw Benches to Make Galvanized Wire. Metallurgist 2004, 48, 45–47. [Google Scholar] [CrossRef]
  24. Sas-Boca, I.M.; Tintelecan, M.; Pop, M.; Iluţiu-Varvara, D.-A.; Mihu, A.M. The Wire Drawing Process Simulation and the Optimization of Geometry Dies. Procedia Eng. 2017, 181, 187–192. [Google Scholar] [CrossRef]
  25. Taheri, A.K.; 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] [CrossRef]
  26. Olowu, T.O.; Ariyo, F.K.; Onibonoje, M.O. Performance Analysis of Direct Torque Controlled, Inverter-fed PMSM Drive for Wire Drawing Machines. Int. J. Eng. Trends Technol. (IJETT) 2018, 55, 1–8. [Google Scholar] [CrossRef]
  27. Kharitonov, V.A.; Usanov, M.Y. Choice of a method of carbon wire drawing. Ferr. Metall. Bull. Sci. Tech. Econ. Inf. 2021, 77, 1177–1185. (In Russian). [Google Scholar] [CrossRef]
  28. Available online: https://www.mflgroup.com/EN/industry/kgt-series (accessed on 27 July 2022).
  29. Hwang, J.-K. Fracture behavior of twinning-induced plasticity steel during wire drawing. J. Mater. Res. Technol. 2020, 9, 4527–4537. [Google Scholar] [CrossRef]
  30. Yamaguchi, I.; Yonemura, M. Recovery and Recrystallization Behaviors of Ni–30 Mass Pct Fe Alloy During Uniaxial Cold and Hot Compression. Metall. Mater. Trans. A 2021, 52, 3517–3529. [Google Scholar] [CrossRef]
  31. Shakhova, I.; Belyakov, A.; Yanushkevich, Z.; Tsuzaki, K.; Kaibyshev, R. On Strengthening of Austenitic Stainless Steel by Large Strain Cold Working. ISIJ Int. 2016, 56, 1289–1296. [Google Scholar] [CrossRef]
  32. Jiang, S.Y.; Wang, R.H. Manipulating nanostructure to simultaneously improve the electrical conductivity and strength in microalloyed Al-Zr conductors. Sci. Rep. 2018, 8, 6202. [Google Scholar] [CrossRef] [Green Version]
  33. Song, Y.; Garcia-Gonzalez, D.; Rusinek, A. Constitutive Models for Dynamic Strain Aging in Metals: Strain Rate and Temperature Dependences on the Flow Stress. Materials 2020, 13, 1794. [Google Scholar] [CrossRef]
  34. Fetisov, V.Р. Evaluation of plasticity during deformation of carbon steel. Litiyo Metall. (Foundry Prod. Metall.) 2019, 3, 85–88. (In Russian). [Google Scholar] [CrossRef]
  35. Saborío-González, M.; Rojas-Hernández, I. Revisión: Fragilización por hidrógeno de metales y aleaciones en motores de combustión. Rev. Tecnol. Marcha 2018, 31, 3–13. [Google Scholar] [CrossRef]
  36. Vega, G.; Haddi, A.; Imad, A. Temperature effects on wire-drawing process: Experimental investigation. Int. J. Mater. Form. 2009, 2, 229. [Google Scholar] [CrossRef]
  37. 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] [CrossRef]
  38. Radionova, L.V.; Lisovskiy, R.A.; Svistun, A.S.; Gromov, D.V.; Erdakov, I.N. FEM Simulation Analysis of Wire Drawing Process at Different Angles Dies on Straight-Line Drawing Machines. In Proceedings of the 8th International Conference on Industrial Engineering. ICIE 2022. Lecture Notes in Mechanical Engineering; Radionov, A.A., Gasiyarov, V.R., Eds.; Springer: Cham, Switzerland, 2023. [Google Scholar] [CrossRef]
  39. Mehrer, H.; Eggersmann, M.; Gude, A.; Salamon, M.; Sepiol, B. Diffusion in intermetallic phases of the Fe–Al and Fe–Si systems. Mater. Sci. Eng. A 1997, 239–240, 889–898. [Google Scholar] [CrossRef]
  40. Heumann, T. Diffusion in Metallen: Grundlagen, Theorie, Vorgänge in Reinmetallen und Legierungen; Springer: Berlin/Heidelberg, Germany, 1992. [Google Scholar] [CrossRef]
  41. Odnobokova, M.; Belyakov, A.; Kaibyshev, R. Development of Nanocrystal-line 304L Stainless Steel by Large Strain Cold Working. Metals 2015, 5, 656–668. [Google Scholar] [CrossRef]
  42. Ghosh, C.; Shome, M. Dynamic strain aging during wire drawing and its effect on electrochemical behavior. Ironmak. Steelmak. 2017, 44, 789–795. [Google Scholar] [CrossRef]
Figure 1. Wire drawing scheme on straight-line machines: 1—drawing drum; 2—die.
Figure 1. Wire drawing scheme on straight-line machines: 1—drawing drum; 2—die.
Metals 12 01472 g001
Figure 2. Deformation heating of the steel wire.
Figure 2. Deformation heating of the steel wire.
Metals 12 01472 g002
Figure 3. Maximum heating of the steel wire due to contact friction in the die.
Figure 3. Maximum heating of the steel wire due to contact friction in the die.
Metals 12 01472 g003
Figure 4. The strain rate’s dependence on drawing parameters: (а) α = 3°, d = 3 mm; (b) μ = 1.35, d = 3 mm; (c) α = 3°, μ = 1.35; (d) μ = 1.1, v = 1 m/s.
Figure 4. The strain rate’s dependence on drawing parameters: (а) α = 3°, d = 3 mm; (b) μ = 1.35, d = 3 mm; (c) α = 3°, μ = 1.35; (d) μ = 1.1, v = 1 m/s.
Metals 12 01472 g004
Figure 5. Dependence of diffusion coefficients for impurity atoms in α-Fe on temperature.
Figure 5. Dependence of diffusion coefficients for impurity atoms in α-Fe on temperature.
Metals 12 01472 g005
Figure 6. Change in the dislocation densities in ferritic steel depending on the deformation degree.
Figure 6. Change in the dislocation densities in ferritic steel depending on the deformation degree.
Metals 12 01472 g006
Figure 7. Dependence of the critical strain rate for impurity carbon (С), nitrogen (N), hydrogen (H) and oxygen (O) atoms on temperature at a dislocation density ρ d = 1013 m−2.
Figure 7. Dependence of the critical strain rate for impurity carbon (С), nitrogen (N), hydrogen (H) and oxygen (O) atoms on temperature at a dislocation density ρ d = 1013 m−2.
Metals 12 01472 g007
Figure 8. Diffusion coefficient of hydrogen (H) for different types of steel depending on temperature.
Figure 8. Diffusion coefficient of hydrogen (H) for different types of steel depending on temperature.
Metals 12 01472 g008
Table 1. Calculation results average temperature of steel (0.2 wt.% C) wire along the drawing route.
Table 1. Calculation results average temperature of steel (0.2 wt.% C) wire along the drawing route.
Pass NumberDrawing Route d , mm Friction   Coefficient   f
0.050.15
Drawing   Speed   v 7 ,   m / s
102045102045
Billet3.50------
13.00666361817469
22.60117117117142136132
32.25156164170185187189
41.95184200215216227237
51.70211235259244264284
61.50230262296265293323
71.30270310353305340381
Table 2. Heating of the steel wire surface by friction during drawing.
Table 2. Heating of the steel wire surface by friction during drawing.
Drawing Route3.002.602.251.951.701.501.30
Maximum   heating   of   the   steel   wire   surface   Δ t f r , °С256286305329359397402
Thickness   of   the   surface   layer   of   metal   heated   by   friction   δ , mm0.820.660.530.430.360.280.23
Table 3. Parameters for determining impurity diffusion in α-Fe.
Table 3. Parameters for determining impurity diffusion in α-Fe.
Impurity AtomThe Pre-Exponential Factor D 0 ,   m 2 / s Activation Energy
Q , eV
C3.96 × 10−70.83
N3.00 × 10−70.80
H4.20 × 10−80.04
O3.78 × 10−70.96
Table 4. Start time of static strain aging.
Table 4. Start time of static strain aging.
Steel Temperature during Drawing, °С100140180220260300
Start Time, s700032002001.50.20.004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/met12091472

AMA Style

Radionova LV, Gromov DV, Svistun AS, Lisovskiy RA, Faizov SR, Glebov LA, Zaramenskikh SE, Bykov VA, Erdakov IN. Mathematical Modeling of Heating and Strain Aging of Steel during High-Speed Wire Drawing. Metals. 2022; 12(9):1472. https://doi.org/10.3390/met12091472

Chicago/Turabian Style

Radionova, Liudmila V., Dmitry V. Gromov, Alexandra S. Svistun, Roman A. Lisovskiy, Sergei R. Faizov, Lev A. Glebov, Sergei E. Zaramenskikh, Vitaly A. Bykov, and Ivan N. Erdakov. 2022. "Mathematical Modeling of Heating and Strain Aging of Steel during High-Speed Wire Drawing" Metals 12, no. 9: 1472. https://doi.org/10.3390/met12091472

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