Next Article in Journal
Prediction Model of End-Point Phosphorus Content in EAF Steelmaking Based on BP Neural Network with Periodical Data Optimization
Next Article in Special Issue
Effect of Tungsten Addition on Continuous Cooling Transformation and Precipitation Behavior of a High Titanium Microalloyed Steel
Previous Article in Journal
Comparative Study on the Corro-Erosive Properties of Base Cemented Transition Metals TaC and HfC and TaX-HfX-C Coatings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influences of Strain on the Microstructure and Mechanical Properties of High-Carbon Steel

1
The State Key Laboratory of Refractories and Metallurgy, Hubei Collaborative Innovation Center for Advanced Steels, Wuhan University of Science and Technology, Wuhan 430081, China
2
Wuhan Iron and Steel Company Limited, Qingshan District, Wuhan 430080, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(9), 1518; https://doi.org/10.3390/met12091518
Submission received: 3 August 2022 / Revised: 3 September 2022 / Accepted: 11 September 2022 / Published: 14 September 2022

Abstract

:
The effects of strain on the microstructure and mechanical properties of 0.81C-0.22Si-0.31Mn (wt%) high-carbon steel were investigated by thermal simulation, scanning electron microscopy, high-resolution transmission electron microscopy (HRTEM), and an electron backscatter diffractometer (EBSD). It was found that when the steel was deformed at 670 °C (a temperature between A1 and Ar1), a deformation-induced pearlite transformation and cementite spheroidization occurred. The volume fraction of pearlite and the spheroidization ratio of cementite increased with a strain increase from 20% to 75%. The microstructure mainly consisted of pearlite when the deformation strain exceeded 40%. The aspect ratio was at its maximum (5.3) at 40% strain and decreased to 1.4 at 75% strain. In addition, the strength of the steel decreased and the elongation increased rapidly with the increase in strain from 20% to 60% due to the spheroidization of cementite. However, as the strain further increased to 75%, the strength increased slightly due to the refinement of the ferrite matrix. The comprehensive performance of the investigated steel can be improved by applying a strain between A1 and Ar1.

Graphical Abstract

1. Introduction

High-carbon steels have excellent comprehensive performance and are widely used in aerospace, automobile engineering, etc. [1,2]. The microstructure of hot-rolled high-carbon steel is generally dominated by flaky pearlite, which has high strength and low elongation [3]; thus, it is difficult to use hot-rolled high-carbon steels directly in engineering applications. In order to improve the comprehensive performance of high-carbon steels, they are generally treated by spheroidizing annealing after hot rolling [4,5]. Flaky pearlite particles are transformed into micro-ferrite and nano-granular cementite after spheroidizing annealing [6]. However, spheroidizing annealing generally takes a long time, which greatly increases the production time and cost of high-carbon steels [7]. Therefore, how to realize the online spheroidization of high-carbon steels during hot rolling has been one of the main research issues in recent years.
It is reported that deformation-induced pearlite transformation (DIPT) and cementite spheroidization (CP) can cause the transformation and spheroidization of flaky pearlite in a short time during hot rolling [8,9]. Previous studies on the DIPT and spheroidization of high-carbon steels have mainly focused on the deformation rate and temperature [10,11,12,13]. It is known that the deformation rate has a significant effect on pearlite transformation and cementite spheroidization. Normally, with an increase in deformation rate, pearlite transformation and cementite spheroidization accelerate [14,15]. In addition, the microstructure of high-carbon steels depends on the deformation temperature. When the deformation temperature is high, austenite is transformed into a ferrite matrix and lamellar pearlite, whereas at a low deformation temperature, austenite is transformed into a ferrite matrix, pearlite and grain-boundary cementite [16]. A composite structure of the ferrite matrix and finely disseminated cementite can be obtained at a deformation temperature near the pearlite transformation temperature [17]. Moreover, it is reported that with the increase in strain, DIPT and CP accelerate in medium-carbon steels [18,19,20]. However, the effects of strain on the DIPT and CP of high-carbon steels are rarely explored. Therefore, in the present study, the DIPT and cementite spheroidization of 0.81C-0.22Si-0.31Mn (wt%) high-carbon steel were investigated. The effects of strain on the microstructure and mechanical properties were discussed to provide a theoretical and experimental basis for the production of low-cost high-carbon steels.

2. Materials and Methods

The investigated steel was melted in a vacuum induction furnace and cast into a 50 kg ingot with a diameter of 150 mm. The ingot was then heated to 1200 °C for two hours in a furnace, and subsequently hot-rolled into a plate of 60 mm thickness and 100 mm width on a laboratory hot-rolling mill. The steel plate was cut into billets of 250 × 100 × 60 mm (length × width × thickness). The billets were heated to 1200 °C for 4 h and then cooled to room temperature so as to make the elements dissolve evenly in the steel. Finally, the billets were machined into thermal simulation samples of 8 mm in diameter and 12 mm in length. The chemical composition of the investigated steel is presented in Table 1.
The influences of deformation strain on the microstructure and mechanical properties of the investigated steel were studied with a THERMECMASTER-Z thermal simulator. Axisymmetric compression deformation was carried out on the specimens. The thermomechanical processing route is schematically presented in Figure 1. The A1 temperature (the pearlite transformation temperature at equilibrium) of the investigated steel was calculated as 720 °C by JmatPro software, and the Ar1 temperature (pearlite transformation temperature during cooling) was measured as 610 °C by thermal simulations at a cooling rate of 20 °C/s after austenitizing. The samples were first heated to 1150 °C at 20 °C/s, then kept at this temperature for 300 s, and finally cooled to 670 °C (a temperature between A1 and Ar1) at 20 °C/s. The samples were then deformed to 0, 20, 40, 60 and 75% strains at a strain rate of 0.01 s–1 and cooled to room temperature at 50 °C/s. The dilation data for these different paths were recorded.
The dog-bone-shaped specimens, with a gauge dimension of 10 mm × 5 mm × 2 mm (length × width × thickness), were machined and tensiled at room temperature with a strain rate of 2 × 10−1 s−1 using an Instron 8801 universal testing machine (Instron Limited, High Wycombe, Britain). The direction of the tensile specimen was perpendicular to the deformation direction. Duplicate tests were carried out to ensure the accuracy of the experiments. The microstructures of different samples were revealed by a field-emission scanning electron microscope (FE-SEM) at an accelerating voltage of 20 kV. Electron backscatter diffraction (EBSD) was employed for crystallographic characterization and grain size measurement. The cross-section of the tensile specimen was the observation surface of the SEM and EBSD microstructure. The EBSD specimens were electro-polished in an electrolyte consisting of 10 vol% of perchloric acid and 90 vol% of glacial acetic acid at a voltage of 40 mV, and a step size of 0.2 µm was applied to obtain high-quality images. The transmission electron microscope (TEM) was employed to observe the fine microstructure and precipitates. Slices for the TEM were cut from the thermal simulation samples, and were subsequently ground to less than 70 µm and punched into 3 mm discs. Thin foils for TEM observation were prepared by twinjet polishing with an electrolyte solution consisting of 6% perchloric acid and 94% ethyl alcohol below −30 °C. The surface of the tensile specimen was the observation surface of the TEM microstructure. In addition, the TEM observation was carried out using a JEM-2100 field-emission high-resolution transmission electron microscope (HRTEM) operated at a voltage of 200 kV.

3. Results

3.1. Microstructure

The microstructures at different strains are exhibited in Figure 2.
It is observed that when the steel was directly cooled to room temperature at a cooling rate of 50 °C/s without deformation, the microstructure was composed of martensite (Figure 2a). Under deformation with 20% strain, partial DIPT occurred, and the microstructure mainly consisted of pearlite (red arrows) and martensite (yellow arrows) (Figure 2b). Martensite was formed during the cooling process after deformation, and lamellar pearlite appeared during DIPT. When the deformation strain was 40%, a large amount of DIPT appeared during the deformation of the austenite, and lamellar pearlite (red arrows) together with some spherical ones (white arrows) were formed (Figure 2c). When the strain increased to 60%, spherical pearlite (white arrows) with a small amount of short lamellar cementite (red arrows) appeared (Figure 2d). When the deformation strain further increased to 75%, the microstructure was composed of uniform and spherical pearlite grains (Figure 2e).
The TEM microstructures at different strains are presented in Figure 3. It is observed that at 20% strains, the microstructure was dominated by flaky pearlite containing small cementite particles. The vertical distance between two flaky pearlites is the lamellar spacing of pearlite. The Image-Pro software was used to measure the pearlite flake spacing in Figure 3a, and the average lamellar spacing of pearlite was 163 nm. When the strain increased to 40%, pearlite particles deformed and a small proportion of the cementite bent (the white-dotted line in Figure 3b) or fractured (the red-dotted line in Figure 3b). At 60% strain, cementite was distributed as particles or short rods (the red circles in Figure 3c), and when the deformation strain increased to 75%, carbon atoms re-precipitated in the supersaturated ferrite matrix in the form of cementite (Figure 3d). Moreover, two types of cementite particles (fused and spheroidized) were detected in the microstructure. Large spheroidized particles (type 1 in Figure 3d) were mainly distributed on the ferrite matrix grain boundaries, and their average diameter was 129–427 nm, while smaller cementite particles with an average diameter of 87–155 nm (type 2 in Figure 3d) were distributed in the ferrite matrix.
The transformation fraction of pearlite and the aspect ratio of cementite at different strains are presented in Figure 4 and Figure 5, respectively. Image-Pro software was used to calculate the phase proportion of the different samples. It is clear from Figure 4 that the fraction of pearlite transformation increased significantly with the increase in strain from 0 to 40%. The microstructure mainly consisted of pearlite when the deformation strain exceeded 40% with the proportion of pearlite more than 97%. Figure 5 reveals that the aspect ratio of cementite decreased rapidly from 40% to 60% strain and then slowly decreased from 60% to 75% strain. The aspect ratio was the maximum (5.3) at 40% strain and decreased to 1.4 at 75% strain. The smaller the aspect ratio, the higher the spheroidization ratio of cementite. The particle size distribution of cementite at 75% strain is displayed in Figure 6. The cementite particles smaller than 400 nm were predominant and accounted for 85.71% of size distribution.
Figure 7 presents the calibration diagram of the bright-field image, dark-field image, and diffraction pattern of the precipitated particles at 75% strain. The precipitated particles were calibrated by connecting adjacent diffraction spots through the central spot of the diffraction pattern to form a parallelogram with its diagonal [21]. The two sides and diagonal of the diffraction pattern were denoted as R1, R2, and R3, respectively (vectors correspond to the diagonal), and the corresponding crystal plane spacings were denoted as d1, d2, and d3, respectively (crystal plane spacing corresponding to the diagonal). The angles between the diagonal and two sides were defined as <R1, R3> and <R2, R3>, respectively. The crystal plane indices were detected as (31 1 ¯ ), (102), and (411); hence, the precipitated phase could be confirmed as Fe3C.

3.2. Grain Size and Orientation

An EBSD was performed at the cross-section of the tensile specimens to quantify their mean crystallographic unit size. The approximate surface area of the microstructure was 30,000 µm2. Figure 8 displays the variation in crystallographic orientation and grain size distributions at different deformation strains. It is observed that the structure of the (101) crystallographic orientation increased significantly with the increase in deformation strain. The fractions of (101) crystallographic orientation at 20, 40, 60 and 75% strains were 11.8, 22.3, 29.3 and 40.9%, respectively. It is seen that the crystallographic orientation of the ferrite matrix was not random at high strain. When analyzing the grain size map, the critical orientation angle was set to be 10°, and the pixel pairs exceeding the critical orientation angle were identified as the grain boundaries. Hence, the areas of each grain can be measured, and then transferred to a circle of equal area. The grain size of each grain is the diameter of the equal area circle. The grain size of the larger strain samples (such as 60 and 75% strain) should be replaced by the width of the grain. It is obvious that the average grain size of the ferrite matrix decreased gradually with the increase in deformation strain (Figure 8a’–d’). The average grain sizes of the ferrite matrix were 10.43 and 3.55 µm at 20% and 75% strains, respectively, indicating that the grains were refined at a larger strain.
Figure 9 displays the ferrite matrix grain size distributions at different strains. When the deformation strain increased from 20% to 75%, the grain size distribution curve moved to the upper left corner, indicating that the ferrite matrix was refined gradually with the increase in deformation strain.
Figure 10 presents the grain boundary distribution maps at different strains, where large-angle grain boundaries (≥15°), small-angle grain boundaries (5–15°), and sub-angle grain boundaries (2–5°) are indicated by blue, green and red lines, respectively. The lengths of the large-angle grain boundaries at 20, 40, 60 and 75% strains were 1.20, 1.61, 2.14 and 2.55 cm, respectively, further proving that the ferrite matrix was refined gradually with the increase in strain.

3.3. Mechanical Properties and Fracture Surface

Figure 11 and Table 2 present the tensile test results of the investigated steel at different strains. The yield and tensile strengths first decreased and then increased with the increase in strain, whereas the total elongation first increased and then decreased. At 20% strain, the yield (1297 ± 13 MPa) and tensile strengths (1443 ± 11 MPa) were found to be the largest. The maximum total elongation (15.3 ± 0.5%) was obtained at 60% strain. This indicates that an improved drawability can be achieved. This is also evidence of an enhanced formability after deformation.
After testing, the representative fracture surfaces are examined and illustrated in Figure 12, which exhibits the fracture morphologies of the tensile specimens at the macroscopic and microscopic levels with different strains. It is noted that the fractographs at 20% and 40% strains (Figure 12a’,b’) predominantly comprise cleavage sheaths (shown by the yellow arrows) along with a few dimples (shown by the red arrow). There are some dimples, but the presence of a higher amount of cleavage sheaths clearly indicates that the specimen predominantly exhibited a brittle fracture, while the fractographs at 60% and 75% strain reveal the formation of plentiful dimples (shown by the red arrows in Figure 12c’,d’). It is clear that the tensile fractures were dominated by ductile fractures at 60% and 75% strain. When the deformation strain increased from 20% to 75%, dimples in the fractured specimens became smaller and more uniform with the increase in strain.

4. Discussion

The microstructure of supercooled austenite mainly experienced pearlite dynamic transformation and pearlite dynamic spheroidization during the deformation process (Figure 2, Figure 3 and Figure 4). The expansion curves during the cooling process after deformation at 20% to 75% strains are presented in Figure 13. The straight line during cooling shows that no pearlite transformation occurred during the cooling process, indicating all the pearlite was formed during the deformation process. The deformation temperature was 670 °C, which was in the range of A1 and Ar1. Pearlite transformation could not occur without deformation in this temperature range [22]. During the deformation process, stress was applied; thus, internal energy was generated in the deformed steel. A small proportion of this energy was transferred to the environment in the form of heat energy, and the remaining energy existed in dislocations and defects in the form of deformation energy [23]. Deformation generated high internal energy in the steel, making the material unstable. In order to restore the equilibrium state, the steel tended to spontaneously shift to a lower energy state; therefore, the deformation energy increased the driving force for phase transformation and caused pearlite transformation [24]. In addition, deformation increased the dislocation density, the number of nucleation sites and the nucleation rate, and consequently accelerated the pearlite transformation. Therefore, when the deformation strain increased to 20%, DIPT occurred during the deformation process.
The internal stress in the microstructure was nonuniform when the stress was small in the initial stage of deformation; hence, the stress concentration occurred near the original austenite grain boundaries, and subsequently [25] the stored deformation energy at these grain boundaries increased [26]. Therefore, when the deformation strain was small, the phase transformation of pearlite preferentially occurred at the grain boundaries. As strains increased and deformation gradually extended to the inside of the austenite, the driving force for phase transformation was increased [27]. Subsequently, the density of internal defects in the austenite increased; thus, the pearlite volume fraction increased with strain (Figure 2 and Figure 4).
With the increase in strain, pearlite particles became spheroidized (Figure 2 and Figure 5). Two types of granular cementite (type 1 and type 2) were formed during spheroidization (Figure 3d). As the strain increased, due to the difference in the mechanical properties of the ferrite matrix (soft phase) and the cementite (hard phase) in pearlite, the deformation of these two phases did not occur in a coordinated state; thus, the cementite lamellar fractured and bent. The curvature radius in the sag areas formed after the fracture of cementite lamellar was significantly smaller than that in the flat areas, causing higher carbon concentration near the smaller curvature radius and forming a carbon concentration gradient at the phase interface [28]. Carbon atoms diffused from the high-carbon zone to the low-carbon zone at the phase interface, promoting the dissolution and spheroidization of cementite lamellar and forming coarse cementite particles on the ferrite matrix boundary (type 1) (Figure 3d).
When the formation energy of the carbon atoms was lower than the binding energy of carbon atoms with dislocations, the carbon atoms dissolved in the dislocations; thus, the dissolution of cementite occurred, and the carbon atoms that had combined with the dislocations dissolved in the ferrite matrix to form a Cottrell gas mass. As the deformation strain increased, dislocations continued to move, and the ferrite matrix recovered and recrystallized [29]. As the dislocation density decreased, the carbon atoms that had dissolved in the dislocations precipitated in the form of fine cementite and were distributed in the ferrite matrix (type 2) (Figure 3d).
The spheroidization of cementite was a thermal activation process [30]. The internal energy and the dislocation density both increased with strain. The generation of a large number of dislocations promoted the diffusion of carbon atoms and accelerated the dissolution and reprecipitation of carbon atoms; hence, when the deformation strain increased from 40% to 60%, the spheroidization rate of cementite rapidly increased. With the further increase in strain, the increase in internal energy and the dislocation density tended to be saturated; thereby, the spheroidization rate of cementite increased slowly as the strain increased from 60% to 75% (Figure 4).
The ferrite matrix grain size is fine (see Figure 9) due to the uniform dispersion of the fine cementite particles in the microstructure. These cementite particles pin the grain boundaries (i.e., Zener drag effect [31]) and drastically reduce the mobility of the grain boundaries, thus inhibiting the recrystallization process [32]. Hence, it is obvious that the ferrite matrix grain size decreases with the increase in the spheroidization ratio of cementite. The results of the EBSD (Figure 8, Figure 9 and Figure 10) also indicated that the ferrite matrix was refined gradually as the deformation strain increased from 20% to 75%. The refinement of the ferrite matrix was beneficial for improving the comprehensive performance of the investigated steel.
It is noticeable from Figure 11 that the tensile and yield strengths of the steel decreased rapidly, and the total elongation increased as the deformation strain increased from 20% to 40%. When the strain was 20%, partial DIPT occurred in the steel, and the microstructure mainly consisted of lamellar pearlite and martensite. When the strain reached 40%, more DIPT occurred, and the microstructure consisted of only lamellar pearlite (Figure 2). In comparison to martensite, lamellar pearlite had lower strength and better plasticity [33]; therefore, the strength decreased rapidly, and the total elongation increased. Furthermore, with the increase in strain from 40% to 60%, cementite particles became spheroidized. The larger the deformation strain, the higher the proportion of cementite spheroidization. On the one hand, in comparison to lamellar pearlite, granular cementite has lower strength and better plasticity [17]; hence, the strength continued to decrease, and the total elongation increased. On the other hand, the refinement of the ferrite matrix was beneficial for increasing the strength and elongation of the investigated steel. Therefore, when the strain increased from 40% to 60%, the strength decreased slightly, while the elongation increased significantly. When the strain reached 60%, the volume fraction of pearlite and the spheroidization rate of cementite tended to be saturated (Figure 5). As the deformation strain further increased to 75%, it had a higher spheroidization rate and a finer ferrite matrix. On the one hand, in comparison to martensite and lamellar pearlite, the higher spheroidization rate had lower strength and better plasticity. On the other hand, the finer ferrite matrix was beneficial for increasing the strengths and elongations of the investigated steel. Therefore, the comprehensive properties of the steels were affected by the strain amount. When the deformation reached 75%, the better comprehensive properties were obtained.
In addition to strength and ductility, toughness is another important index of steel performance. The toughness is normally estimated by Charpy impact tests, but the impact tests could not be performed due to the thinness of the samples. The area encompassed by the stress–strain curve was used to present the toughness [34,35]. The toughness value at 75% strain was higher than others, as shown in Figure 11. This is because the spheroidized structure suppresses the stress concentration and as a result delays the onset of fracture, thus increasing both ductility and toughness. It is expected that toughness increases with the increase in strain due to enhanced spheroidization. It is also the reason why the fracture of the tensile specimen changes from a brittle fracture to a ductile fracture with the increase in deformation strain.

5. Conclusions

The effects of strain on the microstructure and mechanical properties of high-carbon steel were investigated. The main conclusions can be drawn as follows.
(1) When the deformation temperature was between A1 and Ar1, deformation-induced pearlite transformation took place as the deformation strain increased to 20%. The volume fraction of pearlite increased with strain. When the deformation strain exceeded 40%, the microstructure mainly consisted of pearlite.
(2) When the strain was less than 40%, the microstructure of the investigated steel was mainly composed of lamellar pearlite. With the increase in strain, the spheroidization of cementite occurred, the aspect ratio of cementite decreased rapidly from 40% to 60% strain, and then slowly decreased from 60% to 75% strain. The ferrite matrix was refined gradually and the structure of the (101) crystallographic orientation increased significantly as the deformation strain increased from 20% to 75%.
(3) The yield and tensile strengths of the steel decreased, and the total elongation increased as the deformation strain increased from 20% to 60%. As the strain further increased to 75%, the strength increased slightly. The strain can effectively improve the comprehensive properties of high-carbon steels.

Author Contributions

X.G. conceived and designed the experiments; Z.C. conducted experiments and analyzed the data; Y.L. conducted experiments; S.L. conducted experiments and analyzed the data; S.B. conducted experiments; G.X. conceived the experiments. All authors participated in the discussion of experimental results. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support from the National Nature Science Foundation of China (No. 51874216), and the Project of Wuhan Iron and Steel Company Limited (No. 2021-4201-12-000240).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, nor interpretation of data; in the writing of the manuscript; nor in the decision to publish the results.

References

  1. Xu, P.W.; Liang, Y.; Li, J.; Meng, C. Further improvement in ductility induced by the refined hierarchical structures of pearlite. Mater. Sci. Eng. A 2019, 745, 176–184. [Google Scholar] [CrossRef]
  2. Sun, J.J.; Lian, F.L.; Liu, H.J.; Jiang, T.; Guo, S.W.; Du, L.X.; Liu, Y.N. Microstructure of warm rolling and pearlitic transformation of ultrafine-grained GCr15 steel. Mater. Charact. 2014, 95, 291–298. [Google Scholar] [CrossRef]
  3. Wang, M.M.; Zhang, F.C.; Yang, Z.N. Effects of high-temperature deformation and cooling process on the microstructure and mechanical properties of an ultrahigh-strength pearlite steel. Mater. Des. 2017, 114, 102–110. [Google Scholar] [CrossRef]
  4. Song, R.; Ponge, D.; Raabe, D.; Speer, J.G.; Matlock, D.K. Overview of processing, microstructure and mechanical properties of ultrafine grained bcc steels. Mater. Sci. Eng. A 2006, 441, 1–17. [Google Scholar] [CrossRef]
  5. Tikhonova, M.; Belyakov, A.; Kaibyshev, R. Strain-induced grain evolution in an austenitic stainless steel under warm multiple forging. Mater. Sci. Eng. A 2013, 564, 413–422. [Google Scholar] [CrossRef]
  6. Doherty, R.D.; Szpunar, J.A. Kinetics of sub-grain coalescence-A reconsideration of the theory. Acta Mater. 1984, 32, 1789–1798. [Google Scholar] [CrossRef]
  7. Yang, Z.; Yang, R. Formation of ultra-fine grain structure of plain low carbon steel through deformation induced ferrite transformation. ISIJ Int. 2003, 43, 761–766. [Google Scholar] [CrossRef]
  8. Wu, T.; Wang, M.Z.; Gao, Y.W.; Li, X.P.; Zhao, Y.C.; Zou, Q. Effects of plastic warm deformation on cementite spheroidization of a eutectoid steel. J. Iron Steel Res. Int. 2012, 19, 60–66. [Google Scholar] [CrossRef]
  9. Gavriljuk, V.G. Decomposition of cementite in pearlitic steel due to plastic deformation. Mater. Sci. Eng. A 2003, 345, 81–89. [Google Scholar] [CrossRef]
  10. Mishra, K.; Khiratkar, V.N.; Singh, A. Change of deformation mechanism through nano-structuring of pearlite: An in-situ study. Mater. Charact. 2020, 167, 110487. [Google Scholar] [CrossRef]
  11. Zhang, S.L.; Sun, X.J.; Dong, H. Effect of deformation on the evolution of spheroidization for the ultra high carbon steel. Mater. Sci. Eng. A 2006, 432, 324–332. [Google Scholar] [CrossRef]
  12. Zhang, X.; Godfrey, A.; Huang, X.; Hansen, N.; Liu, Q. Microstructure and strengthening mechanisms in cold-drawn pearlitic steel wire. Acta Mater. 2011, 59, 422–3430. [Google Scholar] [CrossRef]
  13. Kapp, M.W.; Hohenwarter, A.; Wurster, S.; Yang, B.; Pippan, R. Anisotropic deformation characteristics of an ultrafine and nanolamellar pearlitic steel. Acta Mater. 2016, 106, 239–248. [Google Scholar] [CrossRef]
  14. Bae, C.M.; Lee, C.S.; Nam, W.J. Effect of carbon content on mechanical properties of fully pearlitic steels. Mater. Sci. Technol. 2002, 18, 1317–1321. [Google Scholar] [CrossRef]
  15. Offerman, S.E.; Van Wilderen, L.J.G.W.; Van Dijk, N.H.; Sietsma, J.; Rekveldt, M.T.; Van der Zwaag, S. In-situ study of pearlite nucleation and growth during isothermal austenite decomposition in nearly eutectoid steel. Acta Mater. 2003, 51, 3927. [Google Scholar] [CrossRef]
  16. Mishra, K.; Singh, A. Effect of interlamellar spacing on fracture toughness of nano-structured pearlite. Mater. Sci. Eng. A 2017, 706, 22–26. [Google Scholar] [CrossRef]
  17. Rastegari, H.; Kermanpur, A.; Najafifizadeh, A. Effect of initial microstructure on the work hardening behavior of plain eutectoid steel. Mater. Sci. Eng. A 2015, 632, 103–109. [Google Scholar] [CrossRef]
  18. Tsuzaki, K.; Sato, E.; Furimoto, S. Formation of an (α+θ) microduplex structure without thermomechanical processing in superplastic ultrahigh carbon steels. Scripta Mater. 1999, 40, 675–681. [Google Scholar] [CrossRef]
  19. Miyamoto, G.; Karube, Y.; Furuhara, T. Formation of grain boundary ferrite in eutectoid and hypereutectoid pearlitic steels. Acta Mater. 2016, 103, 370–381. [Google Scholar] [CrossRef]
  20. Storojeva, L.; Ponge, D.; Kaspar, R.; Raabe, D. Development of microstructure and texture of medium carbon steel during heavy warm deformation. Acta Mater. 2004, 52, 2209–2220. [Google Scholar] [CrossRef] [Green Version]
  21. Zhou, X.L.; Zhou, R.; Zhang, P.X.; Jiang, Y.H.; Li, Z.; Zhao, H. Magnetic field induced pearlite transformation in Fe-0.47C-2.31Si-3.15Mn steel above the eutectoid temperature. Mater. Sci. Eng. A 2009, 525, 42–47. [Google Scholar] [CrossRef]
  22. Mishra, K.; Khiratkar, V.N.; Singh, A. Improvement of sub-critical fatigue crack growth life by nano-structuring of pearlite. Int. J. Fatigue 2019, 122, 84–92. [Google Scholar] [CrossRef]
  23. Khiratkar, V.N.; Mishra, K.; Srinivasulu, P.; Singh, A. Effect of inter-lamellar spacing and test temperature on the charpy impact energy of extremely fine pearlite. Mater. Sci. Eng. A 2019, 754, 622–627. [Google Scholar] [CrossRef]
  24. Zhang, Y.D.; Esling, C.; Lecomte, J.S.; He, C.S.; Zhao, X.; Zuo, L. Grain boundary characteristics and texture formation in a medium carbon steel during its austenitic decomposition in a high magnetic field. Acta Mater. 2005, 53, 5213–5221. [Google Scholar] [CrossRef]
  25. Elwazri, A.M.; Wanjara, P.; Yue, S. Empirical modelling of the isothermal transformation of pearlite in hypereutectoid steel. Mater. Sci. Technol. 2006, 22, 542–546. [Google Scholar] [CrossRef]
  26. Takahama, Y.; Santofimia, M.J.; Mecozzi, M.G.; Zhao, L.; Sietsma, J. Phase field simulation of the carbon redistribution during the quenching and partitioning process in a low-carbon steel. Acta Mater. 2012, 60, 2916–2926. [Google Scholar] [CrossRef]
  27. Saeidi, N.; Ashrafifizadeh, F.; Niroumand, B.; Barlat, F. EBSD study of micromechanisms involved in high deformation ability of DP steels. Mater. Des. 2015, 87, 130–137. [Google Scholar] [CrossRef]
  28. Porter, D.A.; Easterling, K.E.; Smith, G.D.W. Dynamic studies of the tensile deformation and fracture of pearlite. Acta Metall. 1978, 26, 1405–1422. [Google Scholar] [CrossRef]
  29. Zhou, S.; Zuo, Y.; Li, Z.; Wang, X.; Yong, Q. Microstructural analysis on cleavage fracture in pearlitic steels. Mater. Charact. 2016, 119, 110–113. [Google Scholar] [CrossRef]
  30. Moon, J.; Jeong, H.; Lee, J.; Lee, C. Particle coarsening kinetics considering critical particle size in the presence of multiple particles in the heat-affected zone of a weld. Mater. Sci. Eng. A 2008, 483, 633–636. [Google Scholar] [CrossRef]
  31. Maazi, N.; Rouag, N. Consideration of Zener drag effect by introducing a limiting radius for neighbourhood in grain growth simulation. J. Cryst. Growth 2002, 243, 361–369. [Google Scholar] [CrossRef]
  32. Momeni, A.; Dehghani, K. Microstructural evolution and flow analysis during hot working of a Fe-Ni-Cr superaustenitic stainless steel. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2011, 42, 1925–1932. [Google Scholar] [CrossRef]
  33. Thompson, S.W.; Howell, P.R. On the early stages of pearl1tic formation inhypoeutectoid steels. Scr. Metall. 1988, 22, 1775–1778. [Google Scholar] [CrossRef]
  34. Zhao, J.; Wang, T.S.; Lv, B.; Zhang, F.C. Microstructures and mechanical properties of a modified high-C-Cr bearing steel with nano-scaled bainite. Mater. Sci. Eng. A 2015, 628, 327–331. [Google Scholar] [CrossRef]
  35. Prasad, C.; Bhuyan, P.; Kaithwas, C.; Saha, R.; Mandal, S. Microstructure engineering by dispersing nano-spheroid cementite in ultrafine-grained ferrite and its implications on strength-ductility relationship in high carbon steel. Mater. Des. 2018, 139, 324–335. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of thermo−mechanical processing.
Figure 1. Schematic illustration of thermo−mechanical processing.
Metals 12 01518 g001
Figure 2. SEM microstructures at different deformation strains: (a) 0%, (b) 20%, (c) 40%, (d) 60%, (e) 75%.
Figure 2. SEM microstructures at different deformation strains: (a) 0%, (b) 20%, (c) 40%, (d) 60%, (e) 75%.
Metals 12 01518 g002
Figure 3. TEM microstructures at different deformation strains: (a) 20%, (b) 40%, (c) 60%, (d) 75%.
Figure 3. TEM microstructures at different deformation strains: (a) 20%, (b) 40%, (c) 60%, (d) 75%.
Metals 12 01518 g003
Figure 4. Pearlite fractions at different deformation strains.
Figure 4. Pearlite fractions at different deformation strains.
Metals 12 01518 g004
Figure 5. Aspect ratios of cementite at different deformation strains.
Figure 5. Aspect ratios of cementite at different deformation strains.
Metals 12 01518 g005
Figure 6. Distribution of cementite particles at 75% strain.
Figure 6. Distribution of cementite particles at 75% strain.
Metals 12 01518 g006
Figure 7. HRTEM images of precipitates at 75% strain: (a) bright field, (b) dark field, (c) diffraction pattern and index.
Figure 7. HRTEM images of precipitates at 75% strain: (a) bright field, (b) dark field, (c) diffraction pattern and index.
Metals 12 01518 g007
Figure 8. Variation of crystallographic orientation and grain size distributions at different deformation strains: (a,a′) 20%, (b,b′) 40%, (c,c′) 60%, (d,d’) 75%.
Figure 8. Variation of crystallographic orientation and grain size distributions at different deformation strains: (a,a′) 20%, (b,b′) 40%, (c,c′) 60%, (d,d’) 75%.
Metals 12 01518 g008
Figure 9. Grain size distributions at different deformation strains.
Figure 9. Grain size distributions at different deformation strains.
Metals 12 01518 g009
Figure 10. Grain boundary maps at different deformation strains: (a) 20%, (b) 40%, (c) 60%, (d) 75%.
Figure 10. Grain boundary maps at different deformation strains: (a) 20%, (b) 40%, (c) 60%, (d) 75%.
Metals 12 01518 g010
Figure 11. Stress−strain curves at different strains.
Figure 11. Stress−strain curves at different strains.
Metals 12 01518 g011
Figure 12. Fracture surfaces of tensile samples at (ad) macroscopic and (a′d′) microscopic levels with different strains: (a) 20%, (b) 40%, (c) 60%, (d) 75%.
Figure 12. Fracture surfaces of tensile samples at (ad) macroscopic and (a′d′) microscopic levels with different strains: (a) 20%, (b) 40%, (c) 60%, (d) 75%.
Metals 12 01518 g012
Figure 13. Expansion−temperature curves during the cooling process at different deformation strains.
Figure 13. Expansion−temperature curves during the cooling process at different deformation strains.
Metals 12 01518 g013
Table 1. Chemical composition of the investigated steel (wt%).
Table 1. Chemical composition of the investigated steel (wt%).
CSiMnPSNFe
0.810.220.310.0110.0030.004Bal.
Table 2. Mechanical properties at different strains.
Table 2. Mechanical properties at different strains.
Strain (%)20406075
Yield stress (MPa)1297 ± 13734 ± 8683 ± 15728 ± 6
Tensile stress (MPa)1443 ± 11898 ± 13802 ± 7883 ± 9
Elongation (%)3.1 ± 0.34.8 ± 0.215.3 ± 0.514.5 ± 0.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cai, Z.; Gan, X.; Li, Y.; Liu, S.; Bao, S.; Xu, G. Influences of Strain on the Microstructure and Mechanical Properties of High-Carbon Steel. Metals 2022, 12, 1518. https://doi.org/10.3390/met12091518

AMA Style

Cai Z, Gan X, Li Y, Liu S, Bao S, Xu G. Influences of Strain on the Microstructure and Mechanical Properties of High-Carbon Steel. Metals. 2022; 12(9):1518. https://doi.org/10.3390/met12091518

Chicago/Turabian Style

Cai, Zhen, Xiaolong Gan, Yanqi Li, Sheng Liu, Siqian Bao, and Guang Xu. 2022. "Influences of Strain on the Microstructure and Mechanical Properties of High-Carbon Steel" Metals 12, no. 9: 1518. https://doi.org/10.3390/met12091518

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