Next Article in Journal
Clean H2 Production by Lignin-Assisted Electrolysis in a Polymer Electrolyte Membrane Flow Reactor
Next Article in Special Issue
Kinetics of Austenite Decomposition in 54SiCr6 Steel during Continuous Slow Cooling Conditions
Previous Article in Journal
Experimental Study on the Influence of Transverse Crack on Chloride Ingress in Concrete Slab Track of High-Speed Railway
Previous Article in Special Issue
Fatigue Properties of Spring Steels after Advanced Processing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Aspects of Austenitization for the Bearing Steel Induction Quenching Design

COMTES FHT a.s., Prumyslova 995, 334 41 Dobrany, Czech Republic
*
Author to whom correspondence should be addressed.
Materials 2023, 16(9), 3523; https://doi.org/10.3390/ma16093523
Submission received: 7 February 2023 / Revised: 27 April 2023 / Accepted: 28 April 2023 / Published: 4 May 2023

Abstract

:
The dissolution of carbides during the heating to the quenching temperature has a significant effect on the martensite oversaturation and the resulting mechanical properties. The kinetics of dissolution can be influenced by various external factors. This work deals with monitoring the carbide dissolution utilizing dilatometer analysis. The austenitization of 100CrMnSi6-4 bearing steel in two initial states was compared—after accelerated spheroidization annealing and conventional soft annealing. The main objective was to determine the amount of undissolved cementite during austenitization in the temperature range where only austenite and cementite are present in the structure. The austenitization temperature determines the degree of cementite dissolution and, consequently, the carbon content in austenite and thus the final properties after quenching. The cementite dissolution was quantified from dilatometric curves and image analysis.

1. Introduction

Bearing steels are used in the hardened state, i.e., after quenching and tempering. The final properties of the tempered martensite are by a large extent determined by the carbon content in the martensite. Hypereutectoid steels, such as the 100CrMnSi6-4 and similar grades, contain roughly 1 wt.% carbon. This amount is too high to obtain martensite of suitable properties if dissolved in the austenite completely upon austenitization before quenching [1]. Therefore, only partial carbide dissolution is necessary during austenitization. This requires setting the austenitization parameters accordingly [2,3]. Heating in a furnace is usually performed with a sufficient hold at a temperature that can ensure a close-to-equilibrium state. It is possible to use the equilibrium diagram or thermodynamic equilibrium calculation in this case. This approach usually cannot be relied on in the case of induction hardening. It consists of induction heating and subsequent quenching. The material is heated at a high rate (up to 100 °C/s), usually with no hold at the quenching temperature. This treatment does not allow for the equilibrium state to be established. Therefore, austenitization and carbide dissolution kinetics, the heating rate and the type of initial microstructure influence the setting of the quenching temperature required for the desired carbide dissolution stage [4,5].
Comprehensive approaches to the theory of induction heating are available in the literature [6]. The heating itself can be modeled really accurately for the material of known physical properties. Close attention is paid to the induction heating equipment and heating measurement [7,8,9]. However, not so much space is devoted to the setting up of the quenching temperature during induction heating. The quenching temperature directly determines the stage of carbide dissolution and thus the amount of carbon in austenite upon quenching. The setting of this critical parameter is covered in the literature by a rough estimation for given steel ([9]-page 191), with a note that the differences in required austenitization temperature can vary by more than 100 °C due to differences in the heating rate and microstructure. Thus, the quenching temperature setting is left to the trial-and-error method.
Such an approach can be pretty straightforward in the case of the treatment of small and cheap parts, where several or even several dozen test pieces can be easily heated, quenched and measured to see whether they meet the required properties. However, this trial run of the technology can be unfavorable or even unacceptable when the treated parts are large, expensive or simply available in a limited number.
Trial-and-error parameter setting is also problematic due to the fact that the hardness is usually the only required and actually measured material property after the induction quenching. A high hardness is normally required in the case of high-carbon steels, such as the bearing steel 100CrMnSi6-4, 63 HRC (Rockwell hardness—see Table 1) and more. In this hardness interval, there is no clear dependence of hardness on carbon content in parent austenite (i.e., on quenching temperature) as more carbon in austenite does not cause a clear hardness increase, like for lower carbon content [10]. Thus, it is possible to obtain “correct” hardness by the quenching of overheated high-carbon austenite, resulting in coarse high-carbon martensite and a high content of residual austenite. This has an inevitably detrimental effect on the mechanical properties of the final product and can be unnoticed if no metallography analysis of the quenched part is executed.
This article offers an alternative to the trial-and-error method for quenching temperature determination. It describes the use of dilatometry measurement and metallography analysis to determine the quenching temperature based on the demanded hardness. This approach is less demanding on the experimental equipment than, for example, in situ X-ray diffraction analysis or thermomechanical simulation [4,11]. The influence of the carbide size and heating rate on final hardness was examined for widely used bearing steel 100CrMnSi6-4.
The carbide size and shape gain importance when non-equilibrium states of carbide dissolution are considered [12,13,14]. Finer carbides dissolve more rapidly upon austenitization and are also able to reach finer and harder martensite, as documented in previous work [14,15,16]. Bearing steels are usually soft annealed to form the microstructure of spheroidized carbides in the ferrite matrix with good machinability and formability. The soft annealing (SA) [17], which can last up to tens of hours [18], ensures carbide spheroidization in a ferrite matrix [19]. Much finer spheroidized carbides can be produced by accelerated carbide spheroidization and refinement (ASR) [20]. The microstructures obtained by SA and ASR were used as initial structures for this experimental induction quenching. They are morphologically identical (spherical carbides in a ferrite matrix) and just differ in carbides size.
The carbide dissolution was monitored by sample dilatation until full austenitization at various heating rates. Then, the quenching temperature was set for each heating rate according to the arbitrarily chosen degree of carbide dissolution determined from the dilatometer curve. The undissolved carbide content was determined by image analysis of the microstructure. It was proved that the dilatometer record can be used for setting the quenching temperature to achieve a certain degree of carbide dissolution. Moreover, the quantification of undissolved carbide content enables the setting of the quenching temperature to the desired amount of carbon dissolved into the austenite and thus the desired hardness of the material after quenching.
The carbon content in austenite was also measured indirectly by the measurement of the martensite start (Ms) temperature. Generally, when more carbon is in the austenite the Ms is decreasing [21].
Dilatometer and metallography analysis can be used for setting the parameters of induction heating in cases when an extensive run of test trials is not possible due to costs or due to a lack of spare pieces for the test trials.

2. Materials and Methods

2.1. Experimental Material

The experimental material was the 100CrMnSi6-4 bearing steel grade with the chemical composition given in Table 2. The material was supplied in the form of hot-rolled 21 mm diameter bars. The as-received microstructure consisted of lamellar pearlite with a small amount of secondary cementite. The hardness of the as-received material was 383 HV10 [22].

2.2. Spheroidization Annealing

The as-received state of the lamellar pearlite was further processed by two annealing methods to obtain a structure with different sizes of globular carbides. The samples were 16 mm in diameter and 100 mm in length. The finer globular cementite was prepared by accelerated spheroidization and refinement (ASR) annealing using induction heating in a medium-frequency source (fmax = 17 kHz) with a maximum power of 25 kW. The ASR process consisted of induction heating of 15 °C/s to 800 °C with a 15 s residence time and air cooling to 680 °C. This cycle was repeated twice more, followed by the cooling of the material in air to ambient temperature [22].
The resulting microstructure was a fine spheroidized cementite (Figure 1). Coarser globular cementite (Figure 2) was prepared by conventional soft annealing (SA) in an atmospheric furnace. The conventional soft annealing regime consisted of a 15 hr temperature hold at 810 °C and controlled cooling in the furnace at 13 °C/hr. These two spheroidized microstructures were used as initial states for the experiment.

2.3. Dilatometric Experiments for the Monitoring of Carbides Dissolution

The cylinders with a diameter of 4 mm and a length of 10 mm were extracted from 100CrMnSi6-4 steel samples in two initial states: after soft annealing (SA) and ASR treatment. The specimens were treated in the quenching dilatometer L78RITA (Rapid Inductive Thermal Analysis, producer Linseis, Selb, Germany). The dilatometer allows the specimen temperature to be controlled precisely and with a very short response time during heating and cooling at a rate of up to 200 °C/s. The specimens were heated by the induction coil and cooled by flowing helium. They were heated by the linear regime with a constant heating rate (labelled as “lin”) as well as by the regime simulating heating by an industrial inductor with a fixed power output (labelled as “sim”). Constant heating output results roughly in linear heating when the thermal capacity is constant; however, the heating stops when a transformation with a latent heat proceeds (such as the austenitization). This creates a delay in heating. The simulation regime was obtained from the induction heating of a 16 mm diameter, 100 mm long rod with a welded thermocouple. It was performed in a medium-frequency converter (fmax = 17 kHz) with a maximum power of 25 kW. According to measured data, heating was then simulated in a quenching dilatometer (Figure 3). This simulation replicates the real processing process, thus minimizing the influence of factors commonly used in linear dilatometer heating simulations.
The samples were heated by heating rates 6, 13 and 22 °C/s in the austenitic state. An explanation of the marking of the samples is given in two examples below:
ASR lin-6: initial state after accelerated ASR processing, linearly heated to the austenitizing temperature at 6 °C/s.
SA sim-13: initial state after conventional soft annealing, simulated heating to the austenitizing temperature at approx. 13 °C/s according to the heating of the bar in medium-frequency induction equipment.

2.4. Image Analysis

The amounts of carbides in the structures were determined by image analysis. The dilatometric specimens were cut longitudinally and the section was mechanically ground and polished with the last mechanical polishing step being 1 µm diamond suspension. The final polishing step was chemo-mechanical polishing by OP-S colloidal silica suspension (provided by Struers). Its average grain size is 50 nm and it causes a slight etching of the steel surface during polishing. OP-S polishing was carried out for 3 min.
The specimens were observed in a scanning electron microscope (SEM) with an acceleration voltage of 5 kV. Ten images were captured from random locations on the sample. The carbide content was evaluated by the point method [23]. Ten randomly chosen image fields were used for each sample, and a grid of 19 × 14 points was evaluated for each field.

3. Results and Discussion

The first group of dilatometric specimens was heated up to 1300 °C to determine the austenite homogenization temperature interval. The dilatometer curves were recorded for the purpose of austenite formation and homogenization tracking (Figure 4). AC1 and ACm transformation temperatures were determined from the dilatometer curves (Table 3). The specimens in the second group were heated at the same heating rates. The heating was interrupted by sudden quenching from various temperatures to capture carbide dissolution into austenite in a particular stage. These specimens were used for metallographic quantitative analysis of carbides content. The quenching temperatures TQ were determined from the first specimen group.

3.1. Determination of Transformation Temperatures

Heating to 1300 °C was performed to obtain dilatometer curves through the whole austenitization interval up to the state of homogeneous austenite. The heating rates were 6, 13 and 22 °C/s and they were monitored mainly over temperature Ac1. In addition to linear heating at a certain rate, which is commonly used for heating during dilatometric measurements, a simulation of real induction heating was performed (Figure 3).
The total specimen elongation at heating is the sum of the thermal expansion of all phases, decrease in length due to austenite forming and increase in length due to carbide dissolution. Temperature intervals for phase transformation and carbide dissolution certainly overlap. The dilatometric effect of carbide dissolution prevails over the effect of austenite formation above temperature T1 shown in Figure 4. The transformation temperatures Ac1 and Ms were determined by the lever rule as 1% elongation deviation from the linear trend before transformation to the linear trend after transformation. The temperature ACm was estimated as the point where the elongation deviation is only 1% from the trend for homogeneous austenite compared to its maximum value at point T1 shown in Figure 4. An example of the use of the lever rule for determining Ms temperature is given in Figure 5.

3.2. Carbide Dissolution Quantification

The specimens for the metallographic determination of the absolute amount of carbides were heated in the same way as samples for determining the austenite homogenization interval, but they were quenched by a cooling rate of 200 °C/s after reaching a temperature at which a significant occurrence of carbides in the structure simultaneously with full austenitization was ensured. This point was found at three-quarters of the band between the linearized trends of T1 and Ti according to the heating rate (Figure 4). It was chosen arbitrarily, the only necessity being the full austenitization of the ferritic matrix. Thus, all specimens from the second group were quenched from different temperatures TQ but at the same point regarding the general shape of the dilatometer curve.
The following Table 3 shows transformation temperatures that were evaluated from the individual dilatometric curves (for examples, see Figure 4 and Figure 5), the hardness and the amount of undissolved carbides.
Metallograph sections from the quenched specimens were prepared, as well as sections from the initial states, ASR- and SA-treated steel. The volume content of cementite was measured for these specimens by image analysis. SEM images showed a significant dissolution of the finer carbides after heating to the TQ temperature (see Figure 6 and Figure 7).

3.3. Carbide Content Quantification in Dilatometer Curve

Image analysis gave a value of carbide content for one particular temperature during the specimen heating. This enabled the relation of the dilatation of the sample to the amount of undissolved carbides, as is shown in Figure 8 and Figure 9. The representation of dilatometer curves as a direct measurement of carbide content is valid only with an assumption that no process other than carbide dissolution determines specimen dilation. This can be assumed only above temperature T1, where the specimen’s length monotonously decreased, as expected during carbide dissolution. Other phenomena obviously contributed to the specimen dilation at lower temperatures—mainly the phase transformation of the matrix.

3.4. Carbide Dissolution and Final Hardness

The carbon is distributed between the austenite matrix and undissolved carbides during the austenitization. The ratio of the initial volume of carbides and volume of undissolved carbides represents the distribution of carbon between the austenite and the carbides. Thus, the carbon content in austenite Cγ (in wt.%) was computed from the undissolved cementite volume as follows:
C γ = 0.94   1 C E M Q C E M i n i t
where 0.94 is the total amount of carbon in steel in wt.% (see Table 2), CEMQ is the volume of carbides in the quenched sample and CEMinit is the volume of carbides in an initial microstructure. The dependence of specimen hardness and Cγ can be plotted for all quenched specimens (see Figure 10). All specimens quenched from the TQ temperature were grouped together in two distinctive groups for ASR and SA initial state. This points to two things. First, the stage of carbide dissolution (and the resulting value of Cγ) was quite precisely defined by setting TQ according to the dilatometer curve. Second, the initial state of the material strongly influences the carbide dissolution rate. Based on these results, two additional specimens were heated by the linear regime at the rate of 22 °C/s and quenched from temperatures 30 °C and 60 °C below TQ. They reached both lower hardness and lower carbide dissolution, as can be expected. These specimens showed the hardness–Cγ relationship, which follows the famous graph published by Krauss [10].
To set the quenching temperature for desired hardness, one has to follow these steps:
  • Obtain Cγ from Figure 10 for desired hardness;
  • Calculate the amount of undissolved carbides CEMQ from Equation (1);
  • Determine the quenching temperature from the dilatometer curve recorded at the corresponding heating rate of the appropriate initial microstructure (e.g., Figure 9).
A similar plot was drawn for hardness–Ms temperature dependence (Figure 11). Ms temperature can be used similarly for carbide dissolution quantification. However, it depends on the Cγ and also on the prior austenite grain size [1]. Thus, metallography analysis is still necessary for the precise assessment of Cγ based on the Ms temperature. A rough estimate of Cγ can still be made, especially as just a quick check of whether the steel was overheated and whether too much carbon dissolved into austenite.
The spread of points representing specimens quenched from the TQ temperature is noticeably larger in the value of Cγ than in the measured hardness. This can be attributed to the hardness measurement being more precise than quantitative image analysis of undissolved carbides. The main obstacle in the image analysis was presumably the low absolute content of undissolved carbides, leading to the large scatter of results.
The type of heating regime—linear or simulated industrial heating—did not affect the results in terms of carbide dissolution and hardness. This confirms that linear heating regimes can be used for predictions of the carbide dissolution and final hardness of the real industrial heating process. The real heating process with a fixed power output has a significantly higher heating rate when the material is paramagnetic (below the Ac1 temperature) and also a much more pronounced delay in heating at the Ac1 temperature. These differences do not change the resulting microstructure and properties as long as the heating rate in the austenitic state is the same as for the linear regime.

4. Conclusions

  • Dilatometer analysis can be utilized as a tool for setting parameters of induction heating. It was proved that the amount of dissolved carbides in hypereutectoid steel 100CrMnSi6-4 is represented in the dilatometric curve. The dissolution stage can be attributed to a certain point of the curve based on its general shape. Thus, it is possible to determine the quenching temperature on the actual carbon content in the austenite, regardless of the heating rate.
  • The initial structure proved to have a strong impact on the carbide dissolution rate and resulting hardness. The microstructure of the austenitized steel has to be taken into account, rather than only the nominal chemical composition.
  • It is possible to relate the dilatometer curve to the resulting hardness by hardness measurement and the quantification of the undissolved carbide content for specimens with interrupted heating. This can be used for the setting of an appropriate quenching temperature at the given heating rate to reach the required hardness.
  • Linear heating rates normally used in dilatometer measurements can be used as a simulation of real heating regimes by industrial induction devices without any significant result distortion.

Author Contributions

D.N. designed the experiment and processed the samples by heat treatment for dilatometric analysis. J.D. performed metallographic analysis and hardness measurements. P.M. measured and evaluated the dilatometric analysis. J.K. evaluated the graphs from the dilatometric analysis and performed the graphical editing of the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Education, Youth and Sports of Czech Republic, grant number CZ02.1.01/0.0/0.0/16_019/0000836.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used in this article are part of the ongoing experimental program and can not be shared.

Acknowledgments

The result was supported by ERDF Research of advanced steels with unique properties, No. CZ02.1.01/0.0/0.0/16_019/0000836.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bhadeshia, H.K.D.H. Steels for Bearings. Prog. Mater. Sci. 2012, 57, 268–435. [Google Scholar] [CrossRef]
  2. Lee, S.J.; Clarke, K.D. A Quantitative Investigation of Cementite Dissolution Kinetics for Continuous Heating of Hypereutectoid Steel. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2015, 46, 3917–3923. [Google Scholar] [CrossRef]
  3. Chae, J.Y.; Jang, J.H.; Zhang, G.; Kim, K.H.; Lee, J.S.; Bhadeshia, H.K.D.H.; Suh, D.W. Dilatometric Analysis of Cementite Dissolution in Hypereutectoid Steels Containing Cr. Scr. Mater. 2011, 65, 245–248. [Google Scholar] [CrossRef]
  4. Eggbauer, A.; Lukas, M.; Ressel, G.; Prevedel, P.; Mendez-Martin, F.; Keckes, J.; Stark, A.; Ebner, R. In Situ Analysis of the Effect of High Heating Rates and Initial Microstructure on the Formation and Homogeneity of Austenite. J. Mater. Sci. 2019, 54, 9197–9212. [Google Scholar] [CrossRef]
  5. Wu, Y.X.; Wang, L.Y.; Sun, W.W.; Styles, M.J.; Studer, A.J.; Bréchet, Y.; Arlazarov, A.; Hutchinson, C.R. Austenite Formation Kinetics from Multicomponent Cementite-Ferrite Aggregates. Acta Mater. 2020, 196, 470–487. [Google Scholar] [CrossRef]
  6. Sidebotham, G. Heat Transfer Modeling: An Inductive Approach; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar]
  7. Rapoport, E.; Pleshivtseva, Y. Optimal Control of Induction Heating Processes; Mechanical Engineering; Taylor & Francis: Boca Raton, FL, USA, 2006. [Google Scholar]
  8. Rudnev, V.; Loveless, D.; Cook, R.L. Handbook of Induction Heating; Manufacturing Engineering; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2017. [Google Scholar]
  9. Haimbaugh, R.E. Practical Induction Heat Treating, 2nd ed.; ASM International: Almere, The Netherlands, 2015. [Google Scholar]
  10. Krauss, G. Martensite in Steel: Strength and Structure. Mater. Sci. Eng. A 1999, 273–275, 40–57. [Google Scholar] [CrossRef]
  11. Khajuria, A.; Kumar, R.; Bedi, R. Characterizing Creep Behaviour of Modified 9Cr1Mo Steel by Using Small Punch Impression Technique for Thermal Powerplants Effects of Heat Treatments on Grain Size and Hardness of P91 and Boron Added P91 Steels View Project FE Modelling & Analysis of Impression Creep in P91 Steels View Project. J. Mech. Mech. Eng. 2018, 4, 47–61. [Google Scholar]
  12. Podgornik, B.; Leskovšek, V.; Godec, M.; Senčič, B. Microstructure Refinement and Its Effect on Properties of Spring Steel. Mater. Sci. Eng. A 2014, 599, 81–86. [Google Scholar] [CrossRef]
  13. Dlouhy, J.; Hauserova, D.; Novy, Z. Carbide Morphology and Ferrite Grain Size after Accelerated Carbide Spheroidisation and Refinement (ASR) of C45 Steel. Mater. Tehnol. 2015, 49, 625–628. [Google Scholar] [CrossRef]
  14. Dlouhý, J.; Hauserová, D.; Kövér, M. Induction Quenching of Steel 100CrMnSi6-4: Coarse vs. Fine Structure in View of Transformation Kinetics. In Proceedings of the Materials Science and Technology Conference and Exhibition 2015, MS and T 2015, Columbus, OH, USA, 4–8 October 2015; Volume 2, pp. 961–968. [Google Scholar]
  15. Vidhyasagar, M.; Balachandran, G. Spheroidization of 100Cr6 Bearing Steel by Warm Forging. Trans. Indian Inst. Met. 2021, 74, 767–774. [Google Scholar] [CrossRef]
  16. Dlouhy, J.; Motycka, P. Austenitization of Lamellar Pearlite in 100CrMnSi6-4 Bearing Steel. In Proceedings of the METAL 2016—25th Anniversary International Conference on Metallurgy and Materials, Conference Proceedings, Brno, Czech Republic, 25–27 May 2016; pp. 635–640. [Google Scholar]
  17. Nam, W.J.; Bae, C.M. Coarsening Behavior of Cementite Particles at a Subcritical Temperature in a Medium Carbon Steel. Scr. Mater. 1999, 41, 313–318. [Google Scholar] [CrossRef]
  18. Luzginova, N.V.; Zhao, L.; Sietsma, J. The Cementite Spheroidization Process in High-Carbon Steels with Different Chromium Contents. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2008, 39, 513–521. [Google Scholar] [CrossRef]
  19. Tian, Y.L.; Kraft, R.W. Mechanisms of Pearlite Spheroidization. Metall. Trans. 1987, 18, 1403–1414. [Google Scholar] [CrossRef]
  20. Kučerová, L.; Jirková, H.; Mašek, B. Various Approaches to Accelerated Carbide Spheroidization of 54SiCr Steel. Key Eng. Mater. 2015, 647, 3–8. [Google Scholar] [CrossRef]
  21. Zhang, G.H.; Chae, J.Y.; Kim, K.H.; Suh, D.W. Effects of Mn, Si and Cr Addition on the Dissolution and Coarsening of Pearlitic Cementite during Intercritical Austenitization in Fe-1mass%C Alloy. Mater. Charact. 2013, 81, 56–67. [Google Scholar] [CrossRef]
  22. Hauserova, D.; Dlouhy, J.; Novy, Z. Microstructure and Properties of Hardened 100CrMnSi6-4 Bearing Steel after Accelerated Carbide Spheroidization and Long-Duration Annealing. ASTM Spec. Technol. Publ. 2015, STP 1580, 389–409. [Google Scholar] [CrossRef]
  23. Goutam, D. Image Analysis in Quantitative Metallography. In Materials Characterization Techniques-Principles and Applications; Goswanri, N.G., Sridha, G., Chowdhunv, G., Eds.; NML: Jamshedpur, India, 1999; pp. 135–150. [Google Scholar]
Figure 1. Microstructure after ASR (accelerated spheroidization and refinement) process.
Figure 1. Microstructure after ASR (accelerated spheroidization and refinement) process.
Materials 16 03523 g001
Figure 2. Microstructure after conventional soft annealing (SA).
Figure 2. Microstructure after conventional soft annealing (SA).
Materials 16 03523 g002
Figure 3. Dilatometric curves from linear heating (ASR lin-13) and the simulation of real induction heating in a medium-frequency converter (ASR sim-13).
Figure 3. Dilatometric curves from linear heating (ASR lin-13) and the simulation of real induction heating in a medium-frequency converter (ASR sim-13).
Materials 16 03523 g003
Figure 4. The elongation of the sample ASR-lin-6 and its temperature derivative. A1 is the starting point of transformation, T1 is the point at which the elongation derivative equals to the derivative for homogeneous austenite, Ti is the point of the maximum elongation derivative. TQ is the quenching temperature for the metallographic sample for the evaluation of the absolute amount of carbides.
Figure 4. The elongation of the sample ASR-lin-6 and its temperature derivative. A1 is the starting point of transformation, T1 is the point at which the elongation derivative equals to the derivative for homogeneous austenite, Ti is the point of the maximum elongation derivative. TQ is the quenching temperature for the metallographic sample for the evaluation of the absolute amount of carbides.
Materials 16 03523 g004
Figure 5. The illustration of the use of lever rule to determine Ms temperature.
Figure 5. The illustration of the use of lever rule to determine Ms temperature.
Materials 16 03523 g005
Figure 6. Specimen SA—initial state. Globular carbides in the ferrite matrix. Polished specimen.
Figure 6. Specimen SA—initial state. Globular carbides in the ferrite matrix. Polished specimen.
Materials 16 03523 g006
Figure 7. Specimen SA-sim-13, quenched from TQ temperature. Undissolved carbides in a martensitic matrix. Polished specimen.
Figure 7. Specimen SA-sim-13, quenched from TQ temperature. Undissolved carbides in a martensitic matrix. Polished specimen.
Materials 16 03523 g007
Figure 8. Dilatometric curve of ASR-lin-6 (green line); amount of undissolved carbides (brown line). TQ for this sample was 923 °C, undissolved carbide content at TQ was 2.5 vol.% (see Table 3).
Figure 8. Dilatometric curve of ASR-lin-6 (green line); amount of undissolved carbides (brown line). TQ for this sample was 923 °C, undissolved carbide content at TQ was 2.5 vol.% (see Table 3).
Materials 16 03523 g008
Figure 9. Dilatometric curve of SA-lin-6 (green line); amount of undissolved carbides (brown line). TQ for this sample was 954 °C; undissolved carbide content at TQ was 4.1 vol.% (see Table 3).
Figure 9. Dilatometric curve of SA-lin-6 (green line); amount of undissolved carbides (brown line). TQ for this sample was 954 °C; undissolved carbide content at TQ was 4.1 vol.% (see Table 3).
Materials 16 03523 g009
Figure 10. The plot of “Hardness-Cγ” relationship. Encircled groups of points represent TQ specimens, heated by both linear and simulated regimes. The points out of the circles represent specimens quenched at temperatures TQ—30°C and TQ—60 °C.
Figure 10. The plot of “Hardness-Cγ” relationship. Encircled groups of points represent TQ specimens, heated by both linear and simulated regimes. The points out of the circles represent specimens quenched at temperatures TQ—30°C and TQ—60 °C.
Materials 16 03523 g010
Figure 11. The plot of “Hardness-Ms” relationship.
Figure 11. The plot of “Hardness-Ms” relationship.
Materials 16 03523 g011
Table 1. List of symbols and abbreviations.
Table 1. List of symbols and abbreviations.
AC1The start temperature of austenite formation during heating
ACmThe temperature of the last cementite dissolution during heating
ASRAccelerated carbide spheroidization and refinement
CEMinitThe volume of carbides in an initial microstructure
CEMQThe volume of carbides in the quenched sample
CγCarbon content in austenite
HRCRockwell hardness, C-scale
HVVickers hardness
MsMartensite start temperature during quenching
SASoft annealing
SEMScanning electron microscope
T1The point at which the elongation derivative equals to the derivative for homogeneous austenite
TiThe point of maximum elongation derivative
TQThe quenching temperature
Table 2. Chemical composition of 100CrMnSi6-4 bearing steel, wt.%.
Table 2. Chemical composition of 100CrMnSi6-4 bearing steel, wt.%.
CSiMnPSCrNiAlCu
0.940.651.160.0140.0121.540.030.030.02
Table 3. Results of dilatometric measurement, image analysis and hardness measurement.
Table 3. Results of dilatometric measurement, image analysis and hardness measurement.
Heating to 1300 °CHeating Interrupted by Quenching at TQ
Heating Rate (°C/s)Transformation Temperatures
AC1 (°C) ACm (°C)
Quenching Temperature
TQ (°C)
Martensite Start Temperature
Ms (°C)
Hardness HV10Undissolved Carbides (vol.%)
ASR-lin-6676110769232178952.5
ASR-sim-6674810529232149012.9
ASR-lin-131376111109431948922.2
ASR-sim-131375311019431928882.3
ASR-lin-222276411529671928792.3
ASR-sim-222276811279672179032.1
SA-lin-6677311519542528304.1
SA-sim-6676210849542898223.9
SA-lin-131377611679732898213.7
SA-sim-131377511249732758024.3
SA-lin-222278511749972488164.0
SA-sim-222278311439972738214.1
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

Nachazelova, D.; Dlouhy, J.; Motycka, P.; Kotous, J. Aspects of Austenitization for the Bearing Steel Induction Quenching Design. Materials 2023, 16, 3523. https://doi.org/10.3390/ma16093523

AMA Style

Nachazelova D, Dlouhy J, Motycka P, Kotous J. Aspects of Austenitization for the Bearing Steel Induction Quenching Design. Materials. 2023; 16(9):3523. https://doi.org/10.3390/ma16093523

Chicago/Turabian Style

Nachazelova, Daniela, Jaromir Dlouhy, Petr Motycka, and Jakub Kotous. 2023. "Aspects of Austenitization for the Bearing Steel Induction Quenching Design" Materials 16, no. 9: 3523. https://doi.org/10.3390/ma16093523

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