2020 Volume 60 Issue 4 Pages 745-755
This study elucidates the solubility product of Ti4C2S2 in steels by means of first-principles calculations and thermodynamic analysis. For this purpose, the Gibbs formation energy of Ti4C2S2 was calculated theoretically by considering the effect of lattice vibration and thermal expansion. In addition, the Gibbs energies of the bcc and fcc phases in the Fe–Ti–C ternary system were also obtained using the cluster expansion and cluster variation method. Although some experimental data were considered as required, those results were evaluated as the calculation of phase diagrams (CALPHAD)-type thermodynamic parameters through fitting to the sublattice model. By using those thermodynamic functions, an approximate expression of the solubility product for Ti4C2S2 was derived. The result agrees with an experimental result measured in a relatively large temperature range. Furthermore, the formation behavior of precipitates in typical interstitial-free steels was discussed, incorporating an earlier thermodynamic analysis on the Fe–Ti–S ternary system. The results show that NiAs-type TiS was the main precipitate at higher temperatures and that Ti4C2S2 was the main precipitate at lower temperatures.
Ti-stabilized interstitial-free (IF) steels have been widely used in automotive industries because of their good formability and drawability. It is well known for these steels that inclusions of S enhance the precipitation of several sulfides and carbosulfide, such as NiAs-type TiS, NaCl-type MnS, and Ti4C2S2; hence, controlling the behavior of these precipitates plays a key role in determining the mechanical properties of the steels.1,2,3,4,5,6,7,8,9,10,11,12) However, because of experimental difficulty in detecting small concentration of C and S in IF steels, the thermodynamic properties of these precipitates have remained unclear. In particular, discrepancies among the reported thermodynamic stabilities for the Ti4C2S23,4,6,7,8,9,10,12) have been found according to a variety of chemical compositions, temperatures, and experimental methods.
Thermodynamic calculations based on the calculation of phase diagrams (CALPHAD) method have been performed to evaluate the thermodynamic properties of materials.11,13) However, the thermodynamic parameters optimized in these calculations were obtained only using the experimental values, such as solubility products and phase equilibria, and the calculation accuracy greatly depends on the reliability of experimental values. Recently, first-principles calculations combined with a statistical thermodynamics technique have been applied to evaluate the thermodynamic stability of sulfides in IF steels at finite temperatures.14) The values calculated by the method are introduced to the CALPHAD approach to estimate the thermodynamic properties in the metastable region. Then, the thermodynamic stability of Ti4C2S2 in a wide temperature range is expected to be evaluated in a more-accurate manner without any limitation for conditions. Based on these backgrounds, the present study attempts to calculate the Gibbs energy of Ti4C2S2 using first-principles calculations and investigates the thermodynamic stability in a wider temperature range. Furthermore, this study investigates the precipitation behavior of several sulfides in IF steel based on the theoretical thermodynamic parameters.
The first-principles calculations based on density functional theory were performed to evaluate thermodynamic functions at 0 K in this work. The details of this method and the expression of the thermodynamic model using in the CALPHAD approach are noted in this section.
2.1. Calculation of Formation Energies for CompoundsThe total energies of the NaCl-type TiC and Ti4C2S2 were calculated using the projector augmented wave method15,16) implemented in the Vienna Ab-Initio Simulation Package (VASP) code.17,18) The generalized gradient approximation proposed by Perdew, Burke, and Ernzerhof19) was applied to treat the exchange and correlation function. The Methfessel–Paxton method of order-one smearing was used, and the width of the smearing was set as 0.2 eV. The plane wave energy cutoff was set as 400 eV, and spin polarization was not included in the calculation. The convergence criteria for energy calculation and structure optimization were set as 10−5 eV/atom and 0.2 eV/nm, respectively. For both structures, gamma-centered k-point meshes were prepared, and the sizes of the k-point meshes were 7 × 7 × 7 and 7 × 7 × 5 for TiC and Ti4C2S2, respectively. The formation energies of the compounds were computed from the difference between the total energy of each component in its stable state, i.e., hcp-Ti, orthorhombic-S, and graphite. Because of a difficulty in considering the van der Waals interaction for C by means of the first-principles calculation, experimental value for energy difference between diamond structure and graphite structure, 2412 J/mol,20) was added to a calculated energy of C of the diamond structure obtained in the present study. Some calculation results for the van der Waals interaction of orthorhombic-S were reported,21) however, discordance not to be overlooked was observed between those calculated values. The largest error was observed in CuS in which the experimental error was reported about 30%. This fact led to the abandonment of this calculation in this study, and we considered that calculation errors up to 15% were inherent in the result of the solubility product Ti4C2S2 containing 25% of S.
2.2. Calculation of Free Energy for fcc and bcc PhasesTo clarify the precipitation behavior in IF steel, thermodynamic parameters for matrix fcc-Fe and bcc-Fe phases are necessary. Therefore, the Gibbs free energy of bcc and fcc phases in the Fe–Ti–C ternary system were calculated using the cluster expansion method (CEM)22) and the cluster variation method (CVM).23,24) The results were fitted to the compound energy formalism expressed by the two-sublattice model25,26) as (Fe,Ti)a(C,Vacancy)c. The values of a and c are the numbers of sites in each sublattice are a=c=1 for the fcc phase and a=1 and c=3 for bcc phase. The composition area of each ordered structure was defined by four end-member structures, i.e., bcc-Fe, bcc-Ti, FeC3, and TiC3 for the bcc phase; and fcc-Fe, fcc-Ti, FeC, and TiC for the fcc phase. Then, the matrix fcc-Fe and the TiC can be described by the same Gibbs free energy surface based on this two-sublattice model. First, for describing the Gibbs energies of these phases, the formation energies of the end-member structures were calculated by the VASP code. Next, there were 890 and 901 ordered structures prepared for the bcc and fcc phases, respectively, to calculate the Gibbs energies of the solid solutions. The convergence criteria for energy calculation and structure optimization were set to be 10−5 eV/atom and 10−4 eV/atom, respectively. In addition, the ferromagnetic state of bcc-Fe was adopted as the reference state of formation energy calculation. Structure volume change was merely allowed in the structure optimization, and the k-point meshes were created with 1000 k-points per reciprocal atom. In the calculation of the bcc-ordered structures, spin polarization was considered, but it was not considered in fcc-ordered structures because the magnetic structure of the fcc-Fe is paramagnetic at or above room temperature.
In the formalism of CEM, the formation energy of a structure is represented as:
| (1) |
| (2) |
| (3) |
To consider the vibrational effect on the free energy of compound phase at finite temperatures, finite displacement method29,30) and quasiharmonic approximation were applied to Ti4C2S2, NaCl-type TiC, NaCl-type FeC, bcc-ordered structure TiC3, and FeC3.
The Helmholtz free energy with lattice vibration contribution F(T) can be described as follows:
| (4) |
| (5) |
Lastly, Gibbs free energy G(T,p) and heat capacity at constant pressure Cp(T) is given by:
| (6) |
| (7) |
To evaluate the value of
| (8) |
The above calculations were performed using phonopy code32,33,34) combined with VASP. Then, 2 × 2 × 2 supercell structures were prepared, and the sizes of the k meshes were 4 × 4 × 4 for NaCl-type and bcc-ordered structures and 4 × 4 × 2 for Ti4C2S2.
2.4. Thermodynamic Modeling 2.4.1. Ti4C2S2 PhaseIn the present study, the Ti4C2S2 phase was treated as being a stoichiometric compound. The molar Gibbs free energy of Ti4C2S2,
| (9) |
| (10) |
The two-sublattice model (Fe,S,Ti)1(C,Vacancy)3 was applied to describe the Gibbs free energy for the bcc phase. The molar Gibbs free energy of the bcc phase for formula unit of the two-sublattice is expressed as follows:
| (11) |
| (12) |
Because of the similarity in crystal structures between the fcc solid solution and the NaCl-type TiC phase, the Gibbs free energy for these phases is described by the two-sublattice model which has a formula of (Fe,S,Ti)1(C,Vacancy)1. Then, the molar Gibbs free energy of the fcc phase for formula unit of the two-sublattice is expressed as follows:
| (13) |
| (14) |
In addition to these models, the contribution to the Gibbs free energy because of magnetic ordering was considered by adding the formalism suggested by Inden36) and Hillert and Jarl37) to the nonmagnetic term of Gibbs free energy. Thermodynamic analysis and phase diagram calculation were performed using the Thermo-Calc software package.38)
The formation energies of TiC and Ti4C2S2 were evaluated using the first-principles calculations. The calculated results and the lattice parameters are listed in Table 1 with the corresponding experimental results. The reference states of Ti, C, and S for these formation energies were set to hcp-Ti, graphite, and orthorhombic-S, respectively. Experimental formation energies of TiC were measured by combustion calorimetry,39) direct synthesis calorimetry,40) Knudsen cell method,41) and solute-solvent drop calorimetry.42) In contrast, the formation energies of Ti4C2S2 have never been evaluated directly in an experiment. Then, the calculated results by first-principles calculation43,44,45) were compared with the present study in Table 1, which shows that the calculated values in this work show reasonable agreement with those by the previous studies.
| Compound | Space group | Calculated lattice parameter, nm | Calculated formation energy, kJ/mol of atom | Experimental39,40,41,42) and calculated43,44,45) results, kJ/mol of atom | Ref. |
|---|---|---|---|---|---|
| TiC | Fm3m | a=0.433 α=90° | −82.4 | −91.7±0.8 | 39) |
| −91.6±9.6 | 40) | ||||
| −95.2 | 41) | ||||
| −92.9±8.9 | 42) | ||||
| Ti4C2S2 | P63/mmc | a=b=0.320 | −122.1 | −117.4 | 43) |
| c=1.126 | −130.5 | 44,45) | |||
| α=β=90°, γ=120° |
Calculations of the heat capacities for TiC and Ti4C2S2 were performed by the finite displacement method, and the results are given in Figs. 1 and 2. Figure 1 shows the calculated isobaric heat capacity of TiC, which is in good agreement with the experimental data46) and evaluated value by Frisk.47) Figure 2 shows the isobaric heat capacity of Ti4C2S2. The solid line denotes the calculated value using optimized parameters in the framework of the CALPHAD method, and the result corresponds well with the results by the finite displacement method.


The isobaric heat capacity of Ti4C2S2 calculated by the finite displacement method. The solid line represents the result based on the present thermodynamic analysis.
To calculate the solubility products of Ti4C2S2 as well as TiC and TiS in Fe, thermodynamic properties for the fcc and bcc phases in the Fe–Ti–S–C system were evaluated. To accomplish this objective, the Fe–Ti, Ti–C, Fe–C binary systems and the Fe–Ti–C ternary system were thermodynamically reassessed based on the first-principles calculations. Because of the low concentration of C and S in steels, the Fe–C–S and Ti–C–S ternary systems were excluded from the analysis. Thermodynamic parameters for the Fe–S and Ti–S binary systems and the Fe–Ti–S ternary system were obtained from previous studies.14,48)
3.2.1. Fe–Ti, Ti–C, and Fe–C Binary SystemsThe Fe–Ti binary system is composed of the fcc, bcc, hcp, liquid, Fe2Ti, and FeTi phases. Figure 3 shows a comparison of the Gibbs free energies for the fcc and bcc phases calculated by CVM with the values from a previous study.49) Some discrepancies between our CVM calculation and the thermodynamic analysis are found. However, their parameters49) were adopted in the Fe–Ti–S ternary system14) that was accepted in the present work, then in spite of the slight difference between them, the result was used by this analysis. The thermodynamic parameters in the fcc and bcc phases are shown in Table 2, and the calculated Fe–Ti binary phase diagram is presented in Fig. 4.

The Gibbs free energy of the (a) fcc and (b) bcc phases calculated in the Fe–Ti binary system. The reference states are (a) fcc-Fe and fcc-Ti, and (b) bcc-Fe and bcc-Ti, respectively. The black lines denote the results of thermodynamic analysis reported by Kumar et al.49)
| Phase and model | System | Thermodynamic parameters, J/mol | Temperature, K | Ref. |
|---|---|---|---|---|
| bcc (Fe,S,Ti)1(C,Vacancy)3 | Fe–S | 298.15<T<6000 | 48) | |
| Fe–Ti | 298.15<T<6000 | 49) | ||
| 298.15<T<6000 | ||||
| S–Ti | 298.15<T<6000 | 14) | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Fe–C | 298.15<T<6000 | This work | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | 53) | |||
| Ti–C | 298.15<T<6000 | This work | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| fcc (Fe,S,Ti)1(C,Vacancy)1 | Fe–S | 298.15<T<6000 | 48) | |
| Fe–Ti | 298.15<T<6000 | 49) | ||
| Ti–S | 298.15<T<6000 | 14) | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Fe–C | 298.15<T<6000 | This work | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Ti–C | 298.15<T<6000 | 47) | ||
| 298.15<T<6000 | This work | |||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Fe–Ti–C | 298.15<T<6000 | This work | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Ti4C2S2 (Ti)4(C)2(S)2 | Ti–C–S | 298.15<T<6000 | This work |

The calculated Fe–Ti binary phase diagram based on the thermodynamic analysis by Kumar et al.49)
The Ti–C binary system is composed of fcc (which is identical to NaCl-type TiC), bcc, hcp, liquid, and graphite phases. The calculated Gibbs free energies for the fcc and bcc phases in the Ti–C binary system are shown in Fig. 5. The symbols in these figures represent the calculated results using CVM, and the solid lines show the Gibbs energy evaluated by fitting the CVM values using Eqs. (11) and (13). The solid lines in Fig. 5 are shown for comparison with the results from CVM. From this comparison in Fig. 5(a), thermodynamic parameters for fcc phase evaluated in this study agree with the result of CVM. In contrast, we observe a slight difference which originates from influence of short-range ordering in the CVM calculation in Fig. 5(b). The calculated value of the CVM decreases solubility of C in bcc-Ti, resulting in deviation from the experimental values. Therefore, the experimental solubilities of C in bcc-Ti50,51) were considered simultaneously with the evaluated free energy based on the CVM calculation in this thermodynamic analysis, and the free energy in conformity to the experimental values was adopted in this study. The Ti–C binary phase diagram calculated using these parameters is compared with that by the previous study47) in Fig. 6.

The calculated Gibbs free energies of the (a) fcc and (b) bcc phases in the Ti–C binary system. The solid lines denote the results of the thermodynamic analysis in the present study.

The calculated Ti–C binary phase diagram in (a) this study compared with a (b) previous study.47)
In the Fe–C binary system, the fcc, bcc, liquid, and graphite phases constitute the equilibria. The calculated Gibbs free energies for the fcc and bcc phases in the binary system are shown in Fig. 7. The symbols in these figures represent the results of CVM calculation, and the solid lines denote the calculated values in the present study. To check the validity of the optimized thermodynamic parameters, the calculated activity of C was compared with the experimental results. Figure 8 shows the comparison for the activity of C in the (a) fcc and (b) bcc phases with the experimental values.52,53,54) The calculated activities are consistent with the experimental results in both phases, and the calculated Fe–C binary phase diagram is shown in Fig. 9 and compared with the previous result.55)

The Gibbs free energies for the (a) fcc and (b)(c) bcc phases in the Fe–C binary system calculated using CVM. The solid lines denote the results of the thermodynamic analysis in this study.


The calculated Fe–C binary phase diagram (a) in this study compared with a (b) previous study.55)
The ternary fcc phase in the Fe–Ti–C system was thermodynamically analyzed in the present study. Figures 10(a) and 10(b) show the cross-sections of the Gibbs free energy surface along the lines connecting Fe to TiC, and TiC to FeC, respectively. The symbols in this figure represent the results of CVM calculation and the solid lines show the evaluated values by the thermodynamic analysis. The Gibbs energy curves across Fe to TiC are upwardly convex, indicating a two-phase separation between the fcc-Fe and NaCl-type TiC phase. In Fig. 10(a), there exists some discrepancy between the results of thermodynamic analysis and CVM at lower C content, which may be a due to effect of short-range ordering in CVM calculation. When the result of CVM is just used for phase diagram calculation, the solubility of TiC will be evaluated smaller than that obtained in the present analysis. This results in the estrangement from the experimental values. Figure 11 shows a comparison for the activity of C in the Fe–Ti–C ternary system.56) The calculated activities agree well with the experimental results. In Fig. 12, the calculated solubilities of TiC in fcc-Fe are shown with some experimental results.57,58) The good accordance between the results shows that the thermodynamic properties evaluated by this study describe this ternary fcc phase well. In the case of using CVM results faithfully, it is expected that the solubility of TiC in fcc-Fe is calculated smaller because of the instability of CVM results. Then in the present thermodynamic analysis, experimental solubilities of TiC in fcc-Fe57,58) were also considered simultaneously with the evaluated free energy based on the CVM calculation. For the thermodynamic parameters, except for the bcc, fcc, and Ti4C2S2 phases in the Fe–Ti–S–C quaternary system, the results from the previous works on the Fe–S,48) Fe–Ti,49) Ti–C,47) Fe–C,55) Fe–Ti–C,59) Ti–S and Fe–Ti–S14) were adopted. Finally, the thermodynamic parameters for bcc, fcc, and Ti4C2S2 phases are listed in Table 2.

The Gibbs free energy curves calculated in the cross sections of (a) Fe–TiC and (b) TiC–FeC in the fcc phase of the Fe–Ti–C ternary system.

Comparison of the calculated activities of C in austenitic Fe–Ti alloys with the experimental values.56)

In this section, an approximate expression for the solubility product of Ti4C2S2 is derived. When the Ti4C2S2 phase is treated as a stoichiometric compound, the equilibrium between Ti4C2S2 with fcc-Fe can be expressed by the following equation:60)
| (15) |
| (16) |
| (17) |
| (18) |
| (19) |
By applying the thermodynamic parameters listed in Table 2 and neglecting the composition dependency for interaction parameter L, Eq. (20) can be obtained:
| (20) |
| (21) |
Therefore, the equation for the solubility product can be described as Eq. (22):
| (22) |
The solubility product of Ti4C2S2 calculated by using Eq. (22) is listed in Table 3. Figure 13 shows the calculated solubility product of Ti4C2S2 in the fcc phase, which is compared to some experimental values.4,6,9,10) The filled squares are the solubility product listed in Table 3. In the solubility lines determined by experiments, the portion given by the bold line shows the measured experimental region. In general, the solubility product changes linearly with the reciprocal temperature, at least in a small temperature range. That is one of the reasons why the temperature dependency of the solubility product significantly differs outside the region of each measurement. Regarding the observation by Mitsui et al.,10) a wider temperature range was covered compared with that observed in other measurements.4,6,9) The solubility product obtained in the present study coincides well with their results.
| Temperature, T / K | Solubility product log[mass%Ti][mass%S]0.5[mass%C]0.5 | |
|---|---|---|
| 1250 | 8 | −5.62 |
| 1428.5714 | 7 | −4.43 |
| 1666.6667 | 6 | −3.22 |
| 2000 | 5 | −2.01 |

Figure 14 presents the calculated amount of precipitates in Fe–Ti–S–C quaternary system and Fig. 14(a) shows the results of Fe–0.0010%C–0.02%Ti–0.025%S alloy varying with temperature. The main precipitate in austenite is NiAs-type TiS in a higher temperature range, while the Ti4C2S2 phase is formed with descending temperature. For the thermodynamic properties of the NiAs-type TiS, estimated values by Hirata et al.14) were accepted in the present study as listed in Table 4.

Calculated amount of precipitates in (a) Fe–0.0010%C–0.02%Ti–0.025%S alloy varying with temperature, and effect of alloying contents of (b) S and (c) C on the formation behavior of the precipitates.
| Phase and model | System | Thermodynamic parameters, J/mol | Temperature, K | Ref. |
|---|---|---|---|---|
| NiAs–TiS (Fe,Ti,Vacancy)1(S,Vacancy)1 | Fe–S | 298.15<T<6000 | 48) | |
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Ti–S | 298.15<T<6000 | 14) | ||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| 298.15<T<6000 | ||||
| Fe–Ti–S | 298.15<T<6000 | 14) | ||
| 298.15<T<6000 |
Figures 14(b) and 14(c) show the results for the effect of S and C contents in the alloy on the formation behavior of the precipitates. The increase of the S content leads to an increase of the amount of TiS; however, the formation temperature range of Ti4C2S2 and its total amount of precipitation little change, as shown in Fig. 14(b). In contrast, as seen in Fig. 14(c), the increase of the C content in the alloy scarcely influences the formation behavior of TiS; however, it yields an increase in the amount of Ti4C2S2. This fact confirms that the formation of Ti4C2S2 is mainly influenced by the C levels in IF steels.
This work evaluates the solubility and formation behavior of Ti4C2S2 in steel, which have not been clarified because of experimental difficulty, by a combination of the first-principles calculations and the thermodynamic analysis. Based on this result, the formation temperatures for the Ti4C2S2 and other sulfides in the IF steel was calculated, and the formation behavior was discussed. As a result, the following conclusions were obtained.
(1) The free energies of formation for the Ti4C2S2 and NaCl-type TiC were evaluated by first-principles calculations and vibrational analysis. The free energies of bcc and fcc phases in the Fe–Ti–C ternary system were also calculated by the cluster expansion and CVM to clarify the formation behavior of these precipitates in IF steels. Thermodynamic parameters for these phases were evaluated by fitting those results as well as the thermodynamic analysis in which some experimental data were considered to the sublattice model.
(2) By using the evaluated thermodynamic parameters, an approximate expression of the solubility product for Ti4C2S2 was derived. The result agrees well with the experimental result measured in a relatively large temperature range.
(3) The formation behavior of these precipitates in some typical IF steels was calculated by considering the result of the thermodynamic analysis on the Fe–Ti–S ternary system by a previous work. NiAs-type TiS was the main precipitate at higher temperatures, while the Ti4C2S2 phase was formed with descending temperature. The results show that these precipitates are influenced by the contents of alloying elements, and the amount of precipitation of Ti4C2S2 increases as C concentration increases in the steels.
This study was supported by JSPS KAKENHI (grant number, 16H02387). The authors also gratefully acknowledge the useful discussions with Dr. Masanori Enoki at Tohoku University.