Metals 13 00371 v3
Metals 13 00371 v3
Article
Numerical Simulation of Low-Pressure Carburizing and Gas
Quenching for Pyrowear 53 Steel
Bartosz Iżowski 1,2,∗ , Artur Wojtyczka 1,2 and Maciej Motyka 2,3
1 Research and Development Center, Pratt & Whitney Rzeszów S.A., 35-078 Rzeszów, Poland
2 Faculty of Mechanical Engineering and Aeronautics, Rzeszow University of Technology,
12 Powstańców Warszawy Ave., 35-959 Rzeszów, Poland
3 Research and Development Laboratory for Aerospace Materials, Rzeszow University of Technology,
4 Żwirki i Wigury Str., 35-959 Rzeszów, Poland
* Correspondence: [email protected]
Abstract: The hardness and phase composition are, among other things, the critical material prop-
erties considered in the quality control of aerospace gears made from Pyrowear 53 steel after high-
pressure gas quenching. The low availability of data on and applications of such demandingstructures
justify investigating the choice of the material and the need to improve its manufacturability. In this
study, computational finite-element analyses of low-pressure carburizing followed by oil and gas
quenching of Pyrowear 53 steel were undertaken, the objective of which was to examine the influence
of the process parameters on the materials’ final phase composition and hardness. The material input
was prepared using JMatPro. The properties computed by the CALPHAD method were calibrated
by the values obtained from physical experiments. The heat transfer coefficient was regarded as
an objective variable to be optimized. A 3D model of the Standard Navy C-ring specimen was
utilized to predict the phase composition after the high-pressure gas quenching of the steel and the
hardness at the final stage. These two parameters are considered good indicators of the actual process
parameters and are used in the industry. The results of the simulation, e.g., optimized heat transfer
coefficients, cooling curves, and hardness and phase composition, are presented and compared with
experimental values. The accuracy of the simulation was validated, and a good correlation of the data
Citation: Iżowski, B.; Wojtyczka, A.; was found, which demonstrates the quality of the input data and setup of the numerical procedure.
Motyka, M. Numerical Simulation of A computational approach to heat treatment processes’ design could contribute to accelerating new
Low-Pressure Carburizing and Gas procedures’ implementation and lowering the development costs.
Quenching for Pyrowear 53 Steel.
Metals 2023, 13, 371. https://doi.org/ Keywords: steel; modeling; simulation; Pyrowear 53; inverse heat transfer; DEFORM3D; low
10.3390/met13020371 pressure carburizing; high-pressure gas quenching; simufact; phase transformations; latent heat;
Academic Editors: Jelena Srnec Koistinen–Marburger
Novak, Francesco De Bona and
Francesco Mocera
properties of Pyrowear 53 are achieved thanks to, among other things, the high content
of silicon, providing resistance to softening during tempering, nickel, responsible for
added toughness and fabricability, molybdenum, to ensure hot hardness, and vanadium,
providing secondary hardening, whichalso increases the material’s susceptibility to grain
refinement [1]. In contrast to typical steel grades for carburizing, Pyrowear 53 contains
molybdenum and copper. The addition of these two elements distinguishes it among
the variety of steel grades. Molybdenum is responsible for increasing the heat and wear
abrasion resistance, and copper improves the lubrication properties and shock load [2].
The carburization process of Pyrowear 53 is recommended at temperatures between 870
and 927 ◦ C, while hardening from 904 to 921 ◦ C. To obtain the maximum hardness and
dimensional stability, subzero treatment is recommended at −73 ◦ C. Double-tempering is
recommended for applications requiring elevated dimensional stability [3,4]. For parts with
tight dimensional tolerances, such as power transmission components, distortions, being
an effect of heat treating, have a substantial economic impact on the following machining
processes, increasing the overall production costs by 20–40% to eliminate dimensional
deviations [5].
The analysis of the metallurgical effects during quenching is of great importance.
Due to the complexity of the interrelation between the phenomena and their transient
characteristics, it is challenging to capture the sequence and magnitude of these effects
experimentally. Numerical simulations of heat treatment processes have been developed
and used in the industry to analyze transient processes such as quenching thanks to con-
stantly growing computational capabilities and commercial software availability. There are
multiple systems based on the finite-element method, specifically dedicated to heat treat-
ment processes’ design and optimization; among them, DANTE [6], DEFORM2D/3D [7,8],
and Simufact [9] are either fully dedicated or provide specific modules to solve the thermo-
mechano-metallurgical problems often coupled with diffusion models for carburization or
nitriding analyses. For the computational approach to be efficient, this requires detailed
input data, which, in the case of steels’ heat treatment analysis, may be difficult to find in the
literature or to determine experimentally, as the key to numerical simulation is predicting
that the final property of the material is an effect of the stress evolution caused by the
phase transformation, as well as the difference between the properties of the coexisting
phases. Therefore, the precise definition of the input data for specific phases present in the
system is crucial. To define the temperature-dependent thermophysical and mechanical
properties of each phase or the TTT/CCT diagrams, software such as JMatPro [10–12] or
Thermo-Calc [13] based on the CALPHAD method can be used.
Witness specimens are commonly used to eliminate the need to cut up production
parts, leading to increased costs for development efforts. Standard Navy C-ring-type
specimens are known to reflect the dimensional change of the actual components under
investigation. Its asymmetric geometry prevents uniform cooling and delays the begin-
ning of phase transformation in thicker sections, causing distortion. The relatively simple
geometry of the specimens makes it easy to use as, usually, three characteristics are mea-
sured: the outer diameter, inner diameter, and gap. Numerous research works utilizing the
computational approach and using the C-ring-type specimens as the baseline have been
conducted to predict the distortions that are an effect of the heat treatment, confirming the
accuracy of simulations’ capabilities and the geometry of the specimen [5,14–16].
This paper describes the simulation procedure to predict the influence of the carbon
content and quenching process parameters on the phase composition and hardness dis-
tribution after heat treatment. Because of the relatively new process of high-pressure gas
quenching of Pyrowear 53 for applications in demanding structures, such as aerospace
gears, and the limited amount of literature references related to its properties, the objective
of this work was to research them. To achieve this, numerical simulations were used to
predict the material’s behavior under specific process conditions. Quenching by high-
pressure gas quenching (HPGQ) and OH70 quenching oil was employed to validate the
computational results.
Metals 2023, 13, 371 3 of 17
Figure 1. The geometry of the C-ring specimen used in the research. Dimensions are in mm.
3500
HTC OH70
3000
2500
HTC (W/m2 °C)
2000
1500
1000
500
0
0 200 400 600 800
Temperature (°C)
Figure 2. Heat transfer Heat transfer coefficients of OH70 oil’s dependence on temperature (adapted
with permission from Ref. [21], Copyright 2019, Stal, Metale & Nowe Technologie).
Hardness was measured using the NEXUS 4303 tester and Vickers method under a
500 g load in the transverse and longitudinal cross-sections of the specimens, as shown in
Figure 4, performed according to aerospace standards to confirm the effective case depth.
Metals 2023, 13, 371 5 of 17
(a) (b)
Figure 4. Hardness check locations in C-ring cross-sections: transverse (a) and longitudinal (b).
Item C Si Ni Cu Mn Cr Mo V S P Al
Result 0.138 0.945 1.873 1.776 0.319 1.060 3.320 0.092 0.005 0.007 0.002
SD * 0.004 0.008 0.011 0.011 0.001 0.041 0.019 - - - -
* Standard deviation.
In order to obtain an accurate description of the materials’ properties for each phase
present in the material during the whole thermal cycle, which may be further used, the JMat-
Pro simulation was utilized. The CAPLHAD results were then validated and adjusted
by the experimental results. The main adjustments of the inputs made by the authors
consisted of defining the carbon concentration’s influence on the martensite properties,
such as hardness, as shown in Figure 5.
800
700
Hardness [HV0 5]
600
500
400
300
0 0.2 0.4 0.6 0.8 1 1.2
Carbon concentration [wt %]
Figure 5. Dependency of the martensite hardness on carbon concentration for Pyrowear 53 steel.
The model of carbon diffusivity Equation (1) described in [22] and supplemented
with an alloy correction factor Equation (2) [23] that incorporates the influence of alloying
elements on the diffusivity of carbon was used in this research. Thermal data such as
Version February 6, 2023 submitted to Metals 6 of 2
Metals 2023, 13, 371 6 of 17
heat capacity also required adjustments; therefore, experimental data were used [24]. A 146
thermal conductivity
simulation that utilizedand heat capacity
diffusion alsoasrequired
coefficient adjustments;
a function therefore, experimental
of carbon concentration, tempera- 147
ture and alloying elements concentration was performed. The results are presented in thisof
data were used [24]. A simulation that utilized the diffusion coefficient as a function 148
carbon concentration, temperature, and alloying elements’ concentration was performed.
paper. 149
The results are presented in this paper.
−(37000 − 6600%C )
D = 0.47 · exp(−1.6%C ) · exp −(37, 000 − 6600%C ) · q (1)
D = 0.47 · exp(−1.6%C ) · exp R·T ·q (1)
R·T
where R and T are Boltzmann constant and the carburizing temperature, respectively. 150
where R(2)
The equation and T are the
describes Boltzmann
the constant
alloy correction and the carburizing temperature, respec-
factor. 151
tively. Equation (2) describes the alloy correction factor.
Heat Treatment
The Continuous Cooling Transformation (CCT) diagram shown in Figure 6 was de- 153
The experimentally
termined continuous cooling transformation
[24]. According to(CCT) diagram shown
it, martensitic in Figure 6starts
transformation was determined
at 438 °C. 154
experimentally [24]. According to it, martensitic transformation starts at 438 ◦ C. Other data may
Other data may be found in the literature where values of 437°C[25], 510°C[4], or 460°C[26] 155
be mentioned.
are found in the Because
literatureofwhere
the untypical 437 ◦ C [25],
values of chemical 510 ◦ C [4], or
composition of 460
◦ C [26] are mentioned.
the Pyrowear 53, which 156
Because
may exceedof the untypical
JMatPro chemical
limits, it wascomposition
found thatofcompared
Pyrowear 53, which mayresults,
to empirical exceed data
the JMatPro
lacks 157
limits, it was
precision. found that,
Therefore compared
attempts to thetaken
have been empirical results,
to adjust andthe data lack
validate precision. model
a numerical Therefore,
of 158
attempts
the material.were made to adjust and validate a numerical model of the material. 159
1000
Ac3=918 ◦ C
Ferrite + Carbides
Temperature (°C)
600
Ms = 438°C Bainite
400
Martensite
200
Figure6.6.CCT
Figure CCTdiagram
diagramfor
forPyrowear
Pyrowear5353.
Theadjustment
The adjustmentofof the
the martensite
martensite start temperature
temperature(M (Mss))consisted
consistedofofaamodification
modificationof 160
ofthe model
the modelpresented
presentedin [27] by the
in [27] wasauthor.
made byThethe
results
author. of the modification
Results were compared
of modification were 161
to the experimental
confronted resultsresults
by experimental and showed
and showa good
a good fit,fit,asasillustrated
illustratedinin Figure
Figure 7. ItItmust
must 162
material’showever,
hardness; hardness;
inhowever, in this simulation,
this simulation, martensite ismartensite
the only phase is theformed
only phase
fromformed from
austenite’s 164
austenite’s decomposition; its hardness depends on the carbon content. Both the carbon
Metals 2023, 13, 371 7 of 17
concentration and hardness shown in Figure 5 are empirical data, and their origin, whether
from structural martensite or carbides, are negligible from the numerical point of view in
this case.
500
Original model
Martensite start temperature [◦ C] Modified model
400 Exp. MS of core
Exp. MS of case
300
200
100
0
0 0.2 0.4 0.6 0.8 1 1.2
Carbon concentration [wt %]
Figure 7. Comparison of martensite start temperature models. Original vs. modified by the author.
(a) (b)
Figure 8. Computational domain (a) and FE model with control points P1 and P2 and symmetry
planes (b).
transfer during quenching. Because no precise data are available, the heat transfer co-
efficients for the high-pressure gas quenching were optimized, using the DEFORM3D
Inverse Heat Transfer module, to obtain a realistic input for the quenching simulation. The
cooling curves captured from the experiment were used as the optimization objectives,
Broyden–Fletcher–Goldfarb–Shannon as the optimization algorithm, and B-spline as the
interpolation method [8].
ζ n +1 − ζ n ∆ζ
L = ∆E = ∆E · (4)
t n +1 − t n ∆t
where ∆E is the enthalpy of austenite transformation to a child phase (see Figure 9) and
ζ n+1 and ζ n are the volume fractions of the specific phase that transformed at tn and tn+1
simulation time steps, respectively [29]. The initial condition at time step t = 0 is defined
as
T |t=0 = T0 ( x, y, z) (5)
and the boundary condition for temperature field calculation during quenching can be
expressed as
∂T
−λ = hc ( Ts − Te ) + R · e Ts4 − Te4 (6)
∂n
where Te and Ts are the temperatures of the environment and the steel surface. Because the
product of the Boltzmann constant R and the emissivity e is considered as the radiation
heat transfer coefficient,
R · e = hr (7)
Equation (6) can be written as
∂T
−λ = hc ( Ts − Te ) + hr Ts4 − Te4 = h( Ts − Te ) (8)
∂n
where hc and hr are the convection and radiation coefficients, respectively, and where h
is the total heat transfer coefficient [30,31], which combines the effect of convective and
radiative heat transfer, as both of these phenomena take place simultaneously with a varied
intensity along the quenching time. The determination of the total heat transfer coefficient
was the main goal of the inverse heat transfer analysis.
Metals 2023, 13, 371 9 of 17
700
%C=0.13
600 %C=0.98
500
Enthalpy (J/g)
400
300
200
100
0
−100
0 200 400 600 800 1000
Temperature (°C)
∂C
= ∇( D (∇C )) (9)
∂t
where D is the diffusion coefficient, t is time, and C is the carbon concentration at a
specific position. This type of equation requires initial and boundary conditions’ definition.
The initial condition for t = 0 is given as
C ( x, y, z) = C0 (10)
where C0 is the initial carbon content in the steel before carburizing. In this work, carbides’
precipitation was neglected, and the assumption of no soot formation on the steel surface
was made; therefore, the equation for the Neumann-type boundary condition expressing
the mass balance at the surface may be simplified [32,33] as
where Ce is the carbon concentration in equilibrium with the atmosphere and Cs is the
carbon content in the gas–solid interface that reacts with the hot surface of the steel [33].
The surface carbon concentration Ce is required to calculate the mass transfer coefficient
β. The value of this coefficient was determined from the experiment, where the carbon
saturation was constant in time and equal to 60 min. Then, for processes realized at a
constant temperature, T = 921 ◦ C in this research, the value of the mass transfer coefficient
β can be calculated from Equation (12).
∆m
A
β= (12)
t(Ce − CS )
where ∆m is the mass of carbon that the specimen could absorb with no soot formation on
the surface and A is the specimen’s surface that underwent carburizing. For cyclic processes
such as low-pressure carburizing, the determinative parameters are the number of cycles
and the total time of the saturation kinetic model given as a function of temperature and
Metals 2023, 13, 371 10 of 17
time factors may be used [33–35]. The approach described above allows determining
the carbon concentration in equilibrium with the atmosphere Ce and the mass transfer
coefficient β as 1.47 (wt.%) and 9.87 × 10−3 ( mm
s ), respectively.
where ζ ( F,P,B) stands for the volume fraction of ferrite, pearlite, and bainite, respectively,
and b and n are material kinetic parameters defined as:
ln(1−ζ )
ln ln(1−ζ 1 )
n= 2 (14)
ln tt12
ln(1 − ζ 1 )
b=− (15)
t1n
The isothermal times of a certain temperature, t1 and t2 , and the corresponding
volume phase fractions ζ 1 and ζ 2 are defined by means of TTT diagram analysis [29].
Volume fractions of 1% and 99% of the specific phase are usually used as the beginning and
the end of certain phase transformations.
The martensitic transformation is a diffusionless transformation that occurs upon
quick quenching of the austenite phase, and it is characterized by the martensite start
and martensite finish temperatures. As the martensitic transformation is athermal, that
is not being controlled by the thermal history of the material, the volume fraction of the
transformed phase is calculated based on an equation incorporating the degree of under-
cooling of the material. For transformation’s kinetics description, a Koistinen–Marburger
model [38] was used.
ζ M = ζ A 1 − e−K M ( MS −T (t)) (16)
where
MS − M f
KM = (17)
ln 0.1
While both the martensite start and martensite finish temperatures depend on the
alloying elements’ concentration, the K M parameter is considered [39] as
where T (t) is the temperature at a specific time step and ζ M and ζ A represent the volume
phase fractions of martensite and austenite, respectively.
The overall hardness of the material is calculated based on the rule of phases mixture,
which defines the total material’s hardness as a weighted sum of the volume fraction of
where ζ i is the volume fraction of the i-th phase and HVi is its hardness for i = M, B, P, F,
representing martensite, bainite, pearlite, and ferrite, respectively.
Multiple models are available in the literature describing models for hardness calcu-
lations, among which the Maynier model Equation (20) is considered reliable, although
limited to carbon concentration up to 0.5%wt [40]. An additional Equation (21) for marten-
site hardness containing over 0.5% of carbon was proposed by Leslie [41]:
• For %C < 0.5,
Both cases were used for the computations of the carburized case in steel, as the appro-
priate carbon concentration point has a trespassed point of 0.5% of carbon concentration.
of around 1.1 mkW2 ·◦ C were determined for the specimen’s flat faces (Zone 3). In contrast,
the lowest heat transfer of a maximum of 0.87 mkW 2 ·◦ C was found for the inner diameter
(Zone 2).
The results showed the heat transfer coefficients as a function of temperature at the
computational boundary, so influenced by the thermal field only with no accounting of
time. The dashed lines in Figure 10 represent the results of the simulation, which did
not take the latent heat of martensitic transformation into account; therefore, a dimple
in the curves, reaching the minimum at 0.5 mkW 2 ·◦ C , is visible. This is a numerical artifact
caused by the apparent heat capacity of the material. Further analyses of the heat transfer
allowed us to optimize the apparent heat capacity and compute the value of the latent heat
of the martensitic transformation by integrating the function described by the heat capacity
peak’s data points. Integration with the trapezoid method and 100 steps were utilized,
and eventually, the latent heat of 244 gJ was determined and used in further simulations.
Metals 2023, 13, 371 12 of 17
Figure 11 shows the optimization results for the heat capacity and the cooling, where the
determined value of martensitic transformation latent heat was used.
1.2
0.8
0.6
0.4
Zone 1
0.2 Zone 2
Zone 3
0
0 200 400 600 800 1000
Temperature (°C)
Figure 10. Optimized values of the heat transfer coefficients for specific zones.
40 1000
Baseline EXP P1
Peak SIM P1
800
Heat capacity ( mmN2 ·K )
30
Temperature (°C)
600
20
400
10
200
0 0
0 200 400 600 800 1000 0 50 100 150 200
Temperature (°C) Time (s)
(a) (b)
Figure 11. The latent heat of martensitic transformation’s optimization results. Optimized apparent
heat capacity (a) and simulated cooling curve with optimized latent heat (b).
A more detailed analysis, which may include a time factor on the change of the
quenching gas temperature and, therefore, impacting the heat transfer between the gas and
the specimen’s surface, would require a CFD approach and computations of the transient
thermal field in the whole volume of the vacuum chamber, which was not the goal of the
current research.
1.2
GD-OES
0.8
0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Depth (mm)
(a) (b)
Figure 12. Simulated carbon concentration. Visualization of the carburized case (a) and comparison
of the experimental and predicted concentration profiles (b) .
1000 1000
EXP P1 HPGQ SIM P2 HPGQ
SIM P1 HPGQ
800 800
Temperature (°C)
Temperature (°C)
600 600
400 400
200 200
0 0
0 50 100 150 200 250 0 25 50 75 100 125 150
Time (s) Time (s)
(a) (b)
Figure 13. Cooling curves for the core material at P1 (a) and the carburized case at P2 (b).
Metals 2023, 13, 371 14 of 17
The computed cooling rates presented similarity to those obtained from the experiment.
The choice of a cooling rate of 11 ◦ C/s was found for a temperature T = 715 ◦ C, which
corresponds to a time point of t = 26 s of the quenching process. The decrease of the
cooling rate value corresponding to a temperature of 422 ◦ C was related to the temperature
dropduring the exothermic martensitic transformation, as may be observed in Figure 14.
12 12
EXP P1 EXP P1
SIM P1 10 SIM P1
10
8 8
6 6
4 4
2 2
0 0
0 50 100 150 200 250 300 350 400 1000 800 600 400 200 0
Time (s) Temperature (°C)
(a) (b)
Figure 14. Cooling rates as a function of the process time (a) and the temperature (b).
900
Exp LPC+HPGQ
800 Exp LPC+Oil
SIM LPC+HPGQ
Hardness (HV0.5)
700 SIM LPC+Oil
600
500
400
ECD
300
0 0.4 0.8 1.2 1.6 2
Depth (mm)
Figure 15. Results of hardness profile simulation comparison with the experimentally measured after
the HPGQ and oil quenching processes.
Austenite
0.5
Martensite
0.25
M Vol.Frac = 0.927
0
0 75 150 225 300 375 450 525 600 675
Time [s]
Figure 16. Computed phase volume fraction changes during quenching for a control point P1.
4. Conclusions
This paper presented the results of heat-treated specimens made from Pyrowear
53 low-alloy steel. The simulation of the quenching of previously carburized and non-
carburized specimens was analyzed. The analysis of the C-ring, which went through LPC,
followed by different quenching techniques, also concluded that high-pressure gas quench-
ing provides results comparable to standard oil quenching. The numerical simulation of
the heat treatment processes could accelerate new procedures’ implementation and lower
development costs. The simulation of the heat-treatment process using Simufact Forming
supplied with the input of JMatPro material’s properties, although requiring validations
and some adjustments, provided a good correlation with the experimental measurements
for all analyzed variables. A computational approach as presented in this paper’s procedure
for phase composition and hardness analysis may be successfully applied to the analy-
sis of production parts made from different alloys to optimize the process by providing
metallurgical analysis with a limited testing campaign.
Author Contributions: Conceptualization, B.I. and M.M.; methodology, B.I. and A.W.; software, B.I.;
validation, B.I. and M.M.; formal analysis, M.M.; investigation, B.I. and A.W.; resources, B.I. and
Metals 2023, 13, 371 16 of 17
A.W.; data curation, B.I.; writing—original draft preparation, B.I.; writing—review and editing, B.I.;
visualization, B.I.; supervision, M.M.; project administration, B.I.; funding acquisition, M.M. All
authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Data Availability Statement: Not applicable.
Acknowledgments: For calculations and simulations in this work, some of the results gathered within
the framework of the development project in the TECHMATSTRATEG2/406725/1/NCBR/2020
program of the National Centre for Research and Development were used. In particular, the CCT
diagram (Figure 6) was prepared by W. Zalecki from the Łukasiewicz Research Network - Upper
Silesian Institute of Technology. The authors would also like to express their gratitude to FIN Sp. z o.o.
for the preparation of the C-ring specimens that were used in the experiments described in this article.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. National Academy of Sciences. Materials for Helicopter Gears, Final Report No. NMAB-351; Technical Report; Committee on
Helicopter Transmission Gear Materials: Washington, DC, USA, 1979.
2. Korecki, M.; Wołowiec-Korecka, E.; Bazel, M.; Sut, M.; Kreuzaler, T. Thomas Detlef. Outstanding Hardening of Pyrowear® Alloy
53 with Low Pressure Carburizing. In Proceedings of the 70º Congresso Anual da ABM, Rio de Janeiro, Brazil, 17–21 August 2005.
[CrossRef]
3. AMS6308 rev. F-Steel, Bars and Forgings 0.90Si-1.0Cr-2.0Ni-3.2Mo-2.0Cu-0.10V (0.07–0.13C) Vacuum Arc or Electroslag Remelted; SAE
International: Warrendale, PA, USA, 2018
4. Technology, C. CarTech® Pyrowear® Alloy 53—Technical Datasheet.
5. Easton, D.; Perez, M.; Huang, J.; Rahimi, S. Effects of Forming Route and Heat Treatment on the Distortion Behaviour of
Case-Hardened Martensitic Steel type S156. Heat Treat 2017, 2017,
6. Dowling, W.E. Development of a Carburizing and Quenching Simulation Tool: Program Overview. In Proceedings of the 2nd
International Conference on Quenching and Control of Distortion, Materials Park, OH, USA, 9–13 September 1996; pp. 349–355.
7. Arimoto, K. Development of heat treatment simulation system Deform-HT. In Proceedings of the 18th Heat Treating Conference,
ASM International, Materials Park, OH, USA, 12–15 October 1998.
8. Scientific Forming Technologies Corporation. DEFORM V13.0.1 Manual; Scientific Forming Technologies Corporation: Columbus,
OH, USA, 2022.
9. Farivar, H.; Rothenbucher, G.; Prahl, U.; Bernhardt, R. ICME-Based Process and Alloy Design for Vacuum Carburized Steel Components
with High Potential of Reduced Distortion, Proceedings of the 4th World Congress on Integrated Computational Materials Engineering
(ICME 2017); Springer: Berlin/Heidelberg, Germany, 2017; pp. 133–144.
10. Schillé, J.P.; Guo, Z.; Saunders, N.; Miodownik, A.P. Modeling phase transformations and material properties critical to processing
simulation of steels. Mater. Manuf. Process. 2011, 26, 137–143. [CrossRef]
11. Guo, Z.; Saunders, N.; Schillé, J.; Miodownik, A. Material properties for process simulation. Mater. Sci. Eng. A 2009, 499, 7–13.
[CrossRef]
12. Guo, Z.L.; Turner, R.; Da Silva, A.D.; Sauders, N.; Schroeder, F.; Cetlin, P.R.; Schillé, J.P. Introduction of materials modeling into
processing simulation. Trans Tech. Publ. 2013, 762, 266–276.
13. Jo, A.; Thomas, H.; Lars, H.; Pingfang, S.; Bo, S. Thermo-Calc & DICTRA, computational tools for materials science. Calphad 2002,
26, 273–312.
14. da Silva, A.; Pedrosa, T.; Gonzalez-Mendez, J.; Jiang, X.; Cetlin, P.; Altan, T. Distortion in quenching an AISI 4140 C-ring –
Predictions and experiments. Mater. Des. 2012, 42, 55–61. [CrossRef]
15. Nunes, M.M.; da Silva, E.M.; Renzetti, R.A.; Brito, T.G. Analysis of Quenching Parameters in AISI 4340 Steel by Using Design of
Experiments. Mater. Res. 2019, 22, 1–7. [CrossRef]
16. Sims, J.; Li, Z.; Ferguson, B.L. Causes of Distortion during High Pressure Gas Quenching Process of Steel Parts. In Proceedings of
the Heat Treat 2019: 30th ASM Heat Treating Society Conference, Detroit, MI, USA, 15–17 October 2019; pp. 228–236. [CrossRef]
17. Manivannan, M.; Northwood, D.; Stoilov, V. Use of Navy C-rings to study and predict distortion in heat treated components:
Experimental measurements and computer modeling. Int. Heat Treat. Surf. Eng. 2014, 8, 168–175. [CrossRef]
18. Boyle, E.; Bowers, R.; Northwood, D.O. The use of navy C-ring specimens to investigate the effects of initial microstructure and heat
treatment on the residual stress, retained austenite, and distortion of carburized automotive steels. SAE Trans. 2007, 116, 253–261.
19. Yu, H.; Yang, M.; Sisson, R.D. Application of C-Ring Specimen for Controlling the Distortion of Parts During Quenching. Met.
Sci. Heat Treat. 2021, 63, 220–228. [CrossRef]
20. Heuer, V. Gas Quenching. In Steel Heat Treating Fundamentals and Processes; ASM Handbook; ASM International: Almere, The
Netherlands, 2013; Volume 4A.
21. Adrian, H.; Karbowniczek, M.; Dobosz vel Sypulski, A.; Kowalski, J.; Kozdroń, S. Badania wpływu temperatury na zdolność
chłodzac ˛ a˛ wybranych olejów hartowniczych. Stal. Met. Nowe Technol. 2019, 9, 75–80.
Metals 2023, 13, 371 17 of 17
22. Tibbets, G.G. Diffusivity of carbon in iron and steels at high temperatures. J. Appl. Phys. 1980, 51, 4813–4816. [CrossRef]
23. Neumann, F.; Person, B. A contribution to the metallurgy of gas carburization. The effects of alloying elements on the relationship
between the carbon potential of the gas phase and that of the workpiece. Haerterei-Tech. Mitteilungen 1968, 23, 296–310.
24. Research project “Development of high pressure gas quenching technology for the satellite gears of the FDGS engine’s epicyclic
gearbox, made of Pyrowear 53 steel and operating under long-term and cyclically fluctuating loads” funded by National Center
for Research and Development (Narodowe Centrum Badan i Rozwoju) under the TECHMATSTRATEG2/406725/1/NCBR/2020
program. Unpublished Research Data.
25. Freborg, A.M. Investigating and Understanding the Role of Transformation Induced Residual Stress to Increase Fatigue Life of
High Strength Steel Used in Transmission Gears. Ph.D. Thesis, University of Akron, Akron, OH, USA, 2013.
26. Horstemeyer, M.F. Integrated Computational Materials Engineering (ICME) for Metals: Concepts and Case Studies; John Wiley & Sons:
Hoboken, NJ, USA, 2018.
27. van Bohemen, S.; Sietsma, J. Effect of composition on kinetics of athermal martensite formation in plain carbon steels. Mater. Sci.
Technol. 2009, 25, 1009–1012. [CrossRef]
28. Şimşir, C.; Gür, C.H. Simulation of quenching. In Handbook of Thermal Process Modeling of Steels; Routledge: Abingdon, UK, 2009;
pp. 341–425.
29. Li, J.; Feng, Y.; Zhang, H.; Min, N.; Wu, X. Thermomechanical Analysis of Deep Cryogenic Treatment of Navy C-Ring Specimen.
J. Mater. Eng. Perform. 2014, 23, 4237–4250. [CrossRef]
30. Li, H.; Zhao, G.; Huang, C.; Niu, S. Technological parameters evaluation of gas quenching based on the finite element method.
Comput. Mater. Sci. 2007, 40, 282–291. [CrossRef]
31. Sugianto, A.; Narazaki, M.; Kogawara, M.; Shirayori, A.; Kim, S.Y.; Kubota, S. Numerical simulation and experimental
verification of carburizing-quenching process of SCr420H steel helical gear. J. Mater. Process. Technol. 2009, 209, 3597–3609. .:
10.1016/j.jmatprotec.2008.08.017. [CrossRef]
32. Khan, D.; Gautham, B. Integrated modeling of carburizing-quenching-tempering of steel gears for an ICME framework. Integr.
Mater. Manuf. Innov. 2018, 7, 28–41. [CrossRef]
33. Zajusz, M.; Tkacz-Śmiech, K.; Dychtoń, K.; Danielewski, M. Pulse carburization of steel–model of the process. Trans. Tech. Publ.
2014, 354, 145–152. [CrossRef]
34. Wierzba, B.; Romanowska, J.; Kubiak, K.; Sieniawski, J. The cyclic carburization process by bi-velocity method. High Temp. Mater.
Process. 2015, 34, 373–379. [CrossRef]
35. Smirnov, A.; Ryzhova, M.Y.; Semenov, M.Y. Choice of boundary condition for solving the diffusion problem in simulation of the
process of vacuum carburizing. Met. Sci. Heat Treat. 2017, 59, 237–242. [CrossRef]
36. Avrami, M. Kinetics of phase change. I General theory. J. Chem. Phys. 1939, 7, 1103–1112. [CrossRef]
37. William, J.; Mehl, R. Reaction kinetics in processes of nucleation and growth. Trans. Metall. Soc. AIME 1939, 135, 416–442.
38. Dp, K.; Re, M. A general equation prescribing the extent of the austenite-martensite transformation in pure iron-carbon alloys
and plain carbon steels. Acta Metall. 1959, 7, 59–60.
39. Simufact Engineering GmbH. Simufact Forming Users Manual; Simufact Engineering: Hamburg, Germany, 2022.
40. Maynier, P.; Dollet, J.; Bastien, P. Hardenability concepts with applications to steels. In Proceedings of the International Conference
on Artificial Intelligence in Medicine, New York, NY, USA, 1978; Volume 518.
41. Leslie, W. Interstitial atoms in alpha iron. Leslie WC—The Physical Metallurgy of Steels; McGraw-Hill: New York, NY, USA, 1982;
pp. 85–91.
42. Wu, C.; Xu, W.; Wan, S.; Luo, C.; Lin, Z.; Jiang, X. Determination of Heat Transfer Coefficient by Inverse Analyzing for Selective
Laser Melting (SLM) of AlSi10Mg. Crystals 2022, 12, 1309. [CrossRef]
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.