ISIJ International
Online ISSN : 1347-5460
Print ISSN : 0915-1559
ISSN-L : 0915-1559
Fundamentals of High Temperature Processes
Coupled Experimental Study and Thermodynamic Modeling of the Al2O3–Ti2O3–TiO2 System
Sourav Kumar PandaIn-Ho Jung
Author information
JOURNAL OPEN ACCESS FULL-TEXT HTML

2020 Volume 60 Issue 1 Pages 31-41

Details
Abstract

A complete critical evaluation and re-optimization of phase diagrams and thermodynamic properties of the Al2O3–Ti2O3–TiO2 system at 1 atm pressure has been performed. Equilibration and quenching experiment in the Al2O3–TiO2 system in air was also performed to constrain the solubility limit of Al2O3 in TiO2 rutile solution at high temperatures. The molten oxide phase was described by the Modified Quasichemical Model considering the short-range ordering in molten oxide. While Al2TiO5 and Ti3O5 were treated as separate stoichiometric phases in the previous optimization, they were described in this study, using the Compound Energy Formalism, as part of pseudobrookite solid solution with a miscibility gap based on new experimental data. Corundum and rutile solutions were also described based on their crystal structures. New high temperature phase, Al6Ti2O13, was also considered for the first time. A set of optimized model parameters of all phases was obtained, which reproduces all available and reliable literature data within experimental error limits from 25°C to above the liquidus temperatures under oxygen partial pressures from metallic saturation to 1 atm. The newly optimized database was applied to calculate the inclusion diagram and reoxidation of Al-killed and Ti bearing steels.

1. Introduction

Titanium addition to steel creates various forms of non-metallic inclusions such as oxide, nitride, carbide, and sulphide, depending on the steel composition. The effect of these non-metallic inclusions to microstructure and mechanical properties of solid steel is quite important to understand to produce steel with desired properties.1) Moreover, TiOx inclusions easily get agglomerate and forms clusters by decreasing the final steel properties and causes a severe submerged entry nozzle clogging problem during continuous casting of Ti-bearing steel. To understand this phenomenon thoroughly, accurate knowledge on thermodynamics and phase equilibria in the Al–Ti–O and Fe–Al–Ti–O system is crucial. The present author (Jung et al.2)) already performed the comprehensive critical evaluation and optimization of the Al–Ti–O (Al2O3–Ti2O3–TiO2) system and presented the inclusion stability diagram of the Fe–Al–Ti–O system for better understanding the inclusion chemistry in Ti-bearing and Al-killed steel. In the previous study, Al2TiO5 and Ti3O5 were treated as separate stoichiometric phases. However, it is found out that both solid phases belong to pseudobrookite solid solution, which will be discussed later.

In the present study, all the experimental thermodynamic properties and phase diagram data for the Al2O3–TiO2–Ti2O3 system at 1 atm total pressure with oxygen partial pressures from metallic saturation to 1 atm were critically evaluated and re-optimized to obtain the models with a set of model parameters which will be able to reproduce all reliable experimental data in the system. In order to assist the optimization, equilibration and quenching experiment in the Al2O3–TiO2 system in air was also performed at high temperatures. In the end, a revised inclusion diagram of the Fe–Al–Ti–O system was calculated based on the revised thermodynamic database. All the thermodynamic calculations in the present study were performed using the FactSage thermochemical software.3)

2. Equilibrium Phases and Thermodynamic Models

The list below shows all the solutions and stoichiometric compounds that are observed in the Al2O3–Ti2O3–TiO2 system:

(i) Liquid phase (Liq): AlO1.5–TiO1.5–TiO2 in the molten state

(ii) Pseudobrookite solid solution (Psb): (Al3+, Ti3+, Ti4+)4c[Al3+, Ti3+, Ti4+]28f O5

(iii) Ilmenite solid solution (Ilm): (Al3+, Ti3+)A [Al3+, Ti3+, Ti4+]B O3

(iv) Corundum (Cor) solid solution: (Al, Ti)2O3

(v) Rutile solid solution (Rut): TiO2–TiO1.5–AlO1.5

(vi) Stoichiometric compounds: Several Magnéli phases (TinO2n−1, n ≥ 4) and Al6Ti2O13

Cations shown within a set of brackets occupy the same sublattice. The abbreviation of solution phase name was used throughout this study. The metallic phases and gas phases are taken from FACT pure substance database.3) The optimized model parameters for each phase are listed in Table 1.

Table 1. Optimized model parameters in the Al2O3–TiO2–Ti2O3 system (J/mol and J/mol·K).
Liquid phasea: AlO1.5–TiO1.5–TiO2
q TiO 1.5 ,    TiO 2 00 =-11,715.20
q AlO 1.5 ,    TiO 2 01 =13,388.80
q TiO 1.5 ,    TiO 2 20 =16,736.00
q AlO 1.5 ,    TiO 1.5 00 =16,736.00
q TiO 1.5 ,    TiO 2 01 =25,104.00
q AlO 1.5 ,    TiO 1.5 01 =16,736.00
q AlO 1.5 ,    TiO 2 00 =5,230.00
q AlO 1.5 ,    TiO 1.5 (TiO 2 ) 001 =20,920.00
Coordination numbers (CN) of Al3+, Ti3+ and Ti4+ was set to be ZAl3+ = 2.0662, ZTi3+ = 2.0662 and ZTi4+ = 2.7548, respectively.
Pseudobrookite solutionb, Al2TiO5–Ti3O5: (Al 3+ ,    Ti 3+ ,    Ti 4+ ) 4c (Al 3+ ,    Ti 3+ ,    Ti 4+ ) 2 8f O 5
GAT = 2 G°(Al2TiO5)α – 4 RT ln (0.5) – GAA
GTA = G°(Al2TiO5)α + 20, 920.00 – 4.60 T
ΔAA:TT = (GAA + GTT) – (GTA + GAT) = 0
ΔTAG = (GTA + GAG) – (GAA + GTG) = 0
ΔGTA = (GTT + GGA) – (GGT + GTA) = 154, 808.00
Ilmenite solutionc, Ti2O3-rich: (Al 3+ ,    Ti 3+ ) A (Al 3+ ,    Ti 3+ ,    Ti 4+ ) 2 B O 5
GAA = G°(Al2O3)β + 1000
GAG = 0.5 GAA + 0.5 GGG + 41, 840
ΔGA:AG = (GGA + GAG) – (GGG + GAA) = 89, 956
ΔAT:GA = (GAT + GGA) – (GAA + GGT) = 0
Corundum: (Al, Ti)2O3
G°(Al2O3) = G°(Al2O3)βG°(Ti2O3) = G°(Ti2O3)δ + 1000 q Al 2 O 3 ,    Ti 2 O 3 00 =46,024.00
Rutile: TiO2 – TiO1.5 – AlO1.5
G°(TiO2) = G°(TiO2)φG°(Ti2O3) = G°(Ti2O3)δ + 45,185.94 + 5.18 TG°(Al2O3) = G°(Al2O3)β + 96, 232.00
Stoichiometric compounds
G°(Al2TiO5)α = – 2,620,616.52 – 5400.40 T0.5 + 1724.43 T – 249.29 TlnT – 24,03,050.06 T−1 – 86005000.00 T−2 (298.15 K < T < 2500 K)
G°(Al2O3)β = – 1,703,961.42 – 3313.55 T0.5 + 1099.96 T – 155.02 TlnT + 1,930,681.50 T−1 – 68,180,607.70 T−2 (298.15 K < T < 2327 K)
G°(Ti2O3)δ = – 1,547,548.01 – 3000.87 T0.5 + 1155.33 T – 169.96 TlnT – 8,04,824.47 T−1 – 260,920,167.33 T−2 (298.15 K < T < 2115 K)
     = – 1,589,513.65 + 1009.76 T – 156.90 TlnT (2115 K < T < 2500 K)
G°(TiO2)φ = – 2,33,505.41 + 155.85 T – 18.60 TlnT – 402,466.66 T−1 – 16,050,855.33 T−2 (298.15 K < T < 2130 K)
     = – 2,44,632.37 + 162.52 T – 24.00 TlnT (2130 K < T < 3000 K)
G°(Al6Ti2O13) = – 6,921,257.59 – 9940.64 T0.5 + 4183.74 T – 620.73 TlnT + 9,159,885.53 T−1 – 338,855,380.52 T−2 (298.15 K < T < 2115 K)

* Notations A, G, T are used for Al3+, Ti3+, and Ti4+, respectively.

a The Gibbs energies pure liquid AlO1.5 (= 0.5 Al2O3), TiO2 and TiO1.5 (= 0.5 Ti2O3) were taken from and Eriksson and Pelton.4,5)

b The Gibbs energy of pseudobrookite end-members in the Ti-O system were taken from Panda and Jung.6)

c The Gibbs energy of ilmenite end-members in the Ti–O system were taken from Panda et al.7)

β,δ,φ The Gibbs energies of pure solid AlO1.5 (= 0.5 Al2O3), TiO2 and TiO1.5 (= 0.5 Ti2O3) were taken from Eriksson and Pelton.4,5)

ψ The Gibbs energies of pure rutile TiO1.5 (= 0.5 Ti2O3) were taken from Kang et al.8)

2.1. Liquid Oxide Phase (Molten Slag)

The Modified Quasichemical Model (MQM)9,10,11,12) which takes into account short-range ordering of second–nearest–neighbor cations in the oxide melt, was used for describing the molten slag. For example, for the Al2O3–TiO2–Ti2O3 slag, the interactions between the cations such as Al3+, Ti3+, and Ti4+ are considered in the MQM with O2− as a common anion. The components of the slag are taken as AlO1.5–TiO1.5–TiO2. The brief description of the MQM is given elsewhere.4)

The binary liquid TiO2–Ti2O3 parameters8) were revised to reproduce the liquidus and melting point of the new Ti3O5-rich pseudobrookite solid solution. The Al2O3–TiO2 and Al2O3–Ti2O3 binary liquid parameters2) were modified to better reproduce all experimental data in air and reducing atmosphere for the Al2O3–TiO2–Ti2O3 system. The ‘coordination numbers’ of Al3+ and Ti3+ are identical to each other, which is ¾ of Ti4+ (see Table 1).

Once binary interactions for all the binary systems are determined, then parameters can be used to predict the Gibbs energy of ternary system. The Gibbs energy of ternary AlO1.5–TiO1.5–TiO2 melt was calculated using a Kohler type ‘symmetric’ interpolation method.13) A small ternary adjustable model parameter was added to destabilize the liquid slag within Al2TiO5–Ti3O5 system to reproduce the phase diagram, which will be discussed later. It should be noted that a Toop type ‘asymmetric’ interpolation method was used in the previous optimizations: Jung et al.2) adopted TiO2 as an asymmetric component without any ternary parameter, and Kang and Lee14) adopted AlO1.5 as an asymmetric component with a ternary parameter. Because the sub-binary liquid solutions within Al2O3–TiO2–Ti2O3 system have similar Gibbs energy of mixing values (about −12 to −20 kJ/mol at 1873 K) and show reasonably regular solution behavior, the choice of interpolation method would not influence largely to the Gibbs energy of ternary liquid solution and phase diagram involving liquid phase. So, the symmetric Kohler type interpolation method13) was chosen in the present study.

2.2. Pseudobrookite Solid Solution

As mentioned above, in the previous thermodynamic optimization by Jung et al.,2) the Al2TiO5 and Ti3O5 were treated as separate stoichiometric phases, which is obviously wrong. This mistake is corrected in this study.

In general, pseudobrookite oxides AB2O5 are thermodynamically stable at high temperature under 1 atm total pressure. Well known pseudobrookites are Fe2TiO5 (pseudobrookite), FeTi2O5 (ferropseudobrookite), MgTi2O5 (karrooite), MnTi2O5, Al2TiO5, (Mg,Fe)Ti2O5 (armalcolite), Ti3O5, etc. Pseudobrookite compounds have usually orthorhombic structure and belong to the Cmcm space group.15) A two-sublattice pseudobrookite model7) in the framework of the Compound Energy Formalism (CEF) (Hillert et al., 198816)) was developed in the present study to describe the Gibbs energy (Gm) of pseudobrookite solution:   

G m = i j Y i 4c Y j 8f G ij -T S config + G excess (1)
where Y i 4c and Y j 8f represent the site fractions of constituents ‘i’ and ‘j’ on the 4c and 8f octahedral sublattices, respectively (see above for the cations in each sublattice); Gij is the Gibbs energy of an “end member [i]4c[j]28fO5” in which 4c and 8f sites are occupied only by i and j cations, respectively; Gexcess is the excess Gibbs energy; and Sconfig is the configurational entropy which takes into account the random mixing of cations on each sublattice. The detail description of the Gibbs energy of pseudobrookite solution can be found elsewhere.7) The model parameters are listed in Table 1 where the notation A, G and T are used for Al3+, Ti3+ and Ti4+, respectively.

The Gibbs energies of nine end-members are required for the present model. Among them, the Gibbs energies of four end-members (GGG, GTT, GGT, and GTG) were previously determined in the Ti–O system.9) The Gibbs energies of the remaining five end-members (GAT, GAA, GTA, GAG, GGA) were obtained in the present study (see Table 1). The Gibbs energy of stoichiometric end-member, TA (TiAl2O5) was optimized by slightly adjusting the ΔH°298K and S°298K of G°(Al2TiO5) to reproduce the experimental cation distribution data. The Gibbs energy GGA was set positive to form a miscibility gap between Al2TiO5–Ti3O5 (for detail, see the discussion in section (3). In the present study, no excess Gibbs energy parameter, Gexcess, was required.

2.3. Corundum and Ilmenite Solid Solutions

The corundum phase is an Al2O3-rich solid solution, with a trigonal structure and belongs to R3c space group.18) Ti2O3 is iso-structural with corundum Al2O3 with approximate hexagonal close packing of the oxygens and metal ions in 2/3 of the octahedral sites.19) Therefore, ideally Al2O3-rich corundum and Ti2O3-rich should be considered as one solid solution. However, Ti2O3 forms a complete solid solution with ilmenite compounds in the Fe–Ti–O (FeTiO3),6) Mn–Ti–O (MnTiO3)7) and Mg–Ti–O (MgTiO3)20) systems at high temperatures, while Al2O3 has no solid solution with such ilmenite compounds. Therefore, in the present study, two different solid solutions were considered: Ti2O3-rich ilmenite solid solution (space group: R3) and Al2O3-rich corundum solid solution (space group: R3c).

No TiO2 solubility in corundum solution was detected by previous authors.23,24) The Gibbs energy of the corundum was described using the Bragg-Williams random mixing model11) considering Al2O3 and Ti2O3 as components.   

G m = (X A G° A + X B G° B )+ nRT(X A lnX A + X B lnX B ) + i,j0 (q AB ij X A i X B j )X A X B (2)
where A and B are Al2O3 and Ti2O3 (pseudo-end member), and their Gibbs energies G ° i are listed in Table 1. As two moles of Al and Fe are assumed to be randomly mixed per mole of formula solution, ‘n’ is ‘2’ in this case. One binary model parameter was used to reproduce the Ti2O3 solubility in corundum solution in reducing atmosphere.

Similar to pseudobrookite solutions, the Gibbs energy of the ilmenite solution is expressed using the CEF considering their cation structure (see above). The model is detailed elsewhere,7) and their model parameters are listed in Table 1.

2.4. Rutile Solid Solution

TiO2 rutile has a body-centered tetragonal structure which has non-stoichiometry toward the Ti2O3 direction and is usually expressed as TiO2−δ. Eriksson and Pelton4) was first to model the rutile solid solution using the simple Henrian solution model. Then, Kang et al.8) further modified the solution by assigning a temperature dependent term to the Henrian activity coefficient of TiO1.5 in order to reproduce all experimental data over wide range of temperatures. In the present study, the solubility of Al2O3 in the rutile solution was also considered for the first time.

Like the corundum solid solution, the rutile solution is modeled using the Bragg-Williams random mixing model11) considering TiO2, TiO1.5 and AlO1.5 as the components. In the cation sites, random mixing of Ti4+ (TiO2), Ti3+ (TiO1.5) and Al3+ (AlO1.5) was considered, and cation vacancies were assumed to remain associated with Ti3+ and Al3+ to maintain electrical neutrality and so do not contribute to the configurational entropy. Table 1 shows the assembly of Gibbs energy of rutile solution. The Gibbs energy of AlO1.5 (pseudo-endmember) was modified to reproduce Al solubility in rutile from the present experimental data, and the Gibbs energy of ternary solution was calculated using the Kohler interpolation technique.13) No ternary excess ternary parameter was required.

2.5. Stoichiometric Compounds, Metallic Phases and Gas Phases

In this study, the Gibbs energies of several stochiometric Magnéli phases (TinO2n−1, n ≥ 4) were taken from previous studies by Kang et al.8) (now stored in FACT pure substance database) and Gibbs energies of all metallic phases and gasses species were taken from the FACT pure substance database.3) The Gibbs energy of Al6Ti2O13 stoichiometric compound is evaluated in the present study.

3. Experimental Procedure and Results

Although phase equilibria in the oxidizing atmosphere (mostly in the Al2O3–TiO2 system) are reasonably well known, the solubility of Al2O3 in TiO2-rich rutile solution at high temperature is unexplored. Therefore, phase diagram study in air was conducted in this study mainly to obtain such solubility data.

Reagent grade powders of Al2O3 from Sigma Aldrich (99.99%) and TiO2 from Sigma Aldrich (99.99%) were used to prepare the starting materials. These oxides were employed to guarantee the purity of the starting materials, which can exist in multivalent oxidation states (for example, Ti2O3). To remove hydroxide impurities or Ti2O3 from the TiO2 reagent, Al2O3 and TiO2 powders were heated overnight (16 hours) at 873 K (600°C) in air in an ST–1700C box furnace (Sentro Tech, USA; inner dimensions: 10 cm × 10 cm × 20 cm) equipped with MoSi2 heating elements. The powders were then removed from the ST–1700C furnace and allowed to cool down in a drying oven set at 393 K (120°C). To confirm the purities and oxidation states of starting materials, X-ray diffraction (XRD) analyses of pre-dried powders were conducted using Bruker Discover D8 X-ray diffractometer with a Co-Kα source (λ = 1.79 Å) equipped with HiSTAR area detector at McGill University.

Batch of 1 gram of starting material was prepared by mixing in appropriate proportions the reagents in an alumina mortar filled with isopropyl alcohol (<0.02 vol.% H2O) for 30 minutes. The alcohol was driven off under a lamp and the starting materials were stored in the drying oven at 393 K (120°C). Prior to usage, the starting material was taken out from the drying oven and allowed to reach room temperature in a desiccator. To prevent any contamination from crucible materials, the experiments were carried out using pure Pt capsules (length = 10 mm, inner diameter = 2.87 mm, outer diameter = 3 mm). The capsules were sealed from one end with a three-corner weld using a PUK 04 micro-welder (Lampert, Germany) and their integrities were checked using an optical microscope. About 40–50 mg of starting material was tightly packed into each Pt capsule. The other side of the Pt capsule was crimped slightly to prevent any spilling of the starting material during the run and to make sure the equilibration under air atmosphere.

The quenching experiments were conducted in a vertical tube furnace (DelTech®, USA) equipped with a dense alumina reaction tube. In order to make the samples be equilibrated in air, the end of the reaction tube was open. A Pt30Rh–Pt6Rh (type B) thermocouple connected to a PID (Proportional-Integral-Differential) controller was used to keep the temperature within ± 1 K. Another Pt30Rh–Pt6Rh (type B) thermocouple was inserted inside the alumina tube from the top of the furnace to determine the temperature immediately above the Pt capsules. The Pt capsules were placed in a small porous alumina boat, suspended in the hot zone using a Pt–Rh wire and kept at the target temperature (1640 and 1660°C) for enough time to fully equilibrate the samples. After equilibration, the capsules were quenched in a bath filled with ice–cold water.

The quenched samples were mounted in epoxy and polished using an oil–based diamond suspension to avoid any moisture pick up. To remove any oil or dust particles on the surface of the polished sections, samples were cleaned 3 times in an ultrasonic bath using isopropyl alcohol (<0.02 vol.% H2O) for about 60 seconds each time. Phase identification and composition analysis were conducted by EPMA with the help of the JEOL 8900 (Tokyo, Japan) superprobe at McGill University using an accelerating voltage of 15 kV and a 20 nA beam current. Phase identification was conducted using the backscattered electron (BSE) images produced by the scanning electron microscope (SEM) of the EPMA and phase analysis using wavelength dispersive spectrometry (WDS). Spot analysis were performed with beam diameter kept at 3 μm. Raw data were reduced with the ZAF correction using corundum (Al) and rutile (Ti) standards. The EPMA provides information only on the metal content in a phase and does not distinguish between multivalent cations. In the current study, all titanium was safely assumed in quadrivalent state, Ti4+ as the experiments were performed in air.

The experimental results for the key experiments at 1640 and 1660°C are summarized in Table 2 and the BSE images of one equilibrated sample (AT1) from EPMA are presented in Fig. 1. The small standard deviation (2σ) in the composition of each phase determined by EPMA indicates the samples were well equilibrated. No Al and Ti solubility in Pt crucible was detected and no Pt was detected in oxide phases.

Table 2. Result of the present quenching experiment for Al2O3–TiO2 in air.
SampleStarting mixture mole fractionTemperatureAnnealing time hoursPhases*
(#analysis)
Phase composition mole fraction
Al2O3TiO2°CAl2O3TiO2
AT10.3340.667164024Psb (7)0.495 ± 0.0040.505 ± 0.004
Rut (10)0.981 ± 0.0090.019 ± 0.009
AT20.3340.667166024Psb (6)0.499 ± 0.0020.501 ± 0.002
Rut (6)0.981 ± 0.0010.019 ± 0.001
*  Psb: Pseudobrookite, Rut: Rutile

Fig. 1.

Backscattered electron images of the equilibrated sample, AT1. (a) overview, and (b) close-up view.

4. Critical Evaluation/optimization of Experimental Data

All the thermodynamic property and phase diagram data of the Al2O3–Ti2O3–TiO2 system available in the literature were first reviewed. New phase diagram experimental data from this study and all reliable experimental data (with known atmospheric conditions) from the literature23,24,25,26,27,28,29,30,31,32,33,34) were then simultaneously considered to obtain a set of Gibbs energy functions for all the phases. The optimized model parameters in this study are listed in Table 1.

4.1. Ti2O3–TiO2 Binary System

Figure 2 shows the calculated phase diagram of Ti–O (TiO2–Ti2O3) system. The liquid parameters of the TiO2–Ti2O3 system were revised from the previous study,8) where the liquid solution was described by old Modified Quasichemical Model using equivalent fraction expression. In the current optimization, it was replaced with new Modified Quasichemical Model using pair fraction expression. This helped in better prediction of liquidus along the Ti–O region with less parameters and increase the predictive ability of Ti–containing multicomponent oxide system. The description of rutile solution for Ti–O system was taken directly from the previous study.8) The Ti3O5 was initially treated as stoichiometric compound,2,4,8) which is replaced in this study with pseudobrookite solid solution.6) The details of the thermodynamic modeling and model parameters of Ti3O5 pseudobrookite solid solution can be found elsewhere.6) Iso-oxygen partial pressure lines in the liquid phase are also calculated and plotted in Fig. 3. As can be seen, the equilibrium amount of Ti2O3 and TiO2 in the liquid phase varies significantly with oxygen partial pressure and temperature.

Fig. 2.

Calculated optimized phase diagram of the Ti–O (Ti2O3–TiO2) system along with iso oxygen pressure lines in the liquid slag.

Fig. 3.

Optimized heat capacity of Al2TiO5 from pseudobrookite solid solution along with the experimental data by King.35)

4.2. Al2O3–Ti2O3–TiO2 System

4.2.1. Thermodynamic Data

The optimized heat capacity of Al2TiO5 from the present pseudobrookite solution is shown in Fig. 3 along with the experimental data between 52.74 and 298.15 K measured by King35) using low temperature adiabatic calorimetry (Nernst Method46)). Using Simpson rule integrations and Debye & Einstein function37) (below 52.74 K), King35) determined the entropy S°298.15 K to be 109.42 ± 0.84 J mol−1 K−1. However, the current optimized entropy was set to be 121.82 J mol−1K−1 which takes into consideration the configurational entropy, ΔSmix = – 2R (0.5ln0.5 + 0.5ln0.5) = 11.526 J mol−1K−1 of Al2TiO5 at 298.15 K, generated due to random mixing of cations (Al3+, Ti4+) in the second sublattice (8f) of Al2TiO5 at room temperature, i.e. (Al3+)4c(Al3+, Ti4+)28fO5. Figure 4 shows the calculated heat content (HT – H298.15) of Al2TiO5 from the current pseudobrookite solution along with the experimental data by Bonnickson38) between 411 and 1803.2 K using drop calorimetry. The enthalpy of Al2TiO5 at 298.15 K (ΔH°298.15) was estimated using the Neumann–Kopp (N–K) rule from its constituent oxides (Al2O3 and TiO2) and modified to reproduce its phase stability region in air atmosphere (all values listed in Table 1).

Fig. 4.

Calculated heat content (HT – H298.15) of Al2TiO5 from the present pseudobrookite solid solution along with the experimental data by Bonnickson.38)

The thermodynamic properties of stoichiometric Al6Ti2O13 have been rarely studied. No low-temperature and high-temperature heat capacities were measured and, therefore, the enthalpy and entropy at 298.15 K (ΔH°298.15, S°298.15) and heat capacity were estimated from its constituent oxides (Al2O3 and TiO2) using the Neumann–Kopp (N–K) rule17. The estimated H°298.15 and S°298.15 were then revised (ΔH°298.15 = – 677.20 kJ mol−1 and S°298.15 = 339.00 J mol−1K−1) during the optimization to fix the high-temperature stability suggested by different authors.39,40,41)

4.2.2. Structural Data

The calculated cation distribution of Al2TiO5 from the present study is shown in Fig. 5. Recently, Ohya et al.42) determined the cation distribution in Al2TiO5 pseudobrookite compound at 600–1500°C using XRD Rietveld analysis. They found partial decomposition of Al2TiO5 to corundum (Al2O3) and rutile (TiO2) when annealed at 1100 and 1200°C for 1 hour and complete decomposition at 1200°C after 10 hours of annealing time. No decomposition was found at 600°C after 100 hours of annealing time. After water quenching, Rietveld analysis were performed on Al2TiO5 XRD pattern for all temperatures. It is well-known from Jung et al.43,44) that the cation distribution can be frozen at low temperature due to the slow kinetic of cation-exchange reactions. Therefore, the experimental data collected at 600°C can most probably be in non-equilibrium state. Similarly, Skala et al.45) synthesized Al2TiO5 from Al2O3 and TiO2 at 1550°C and then performed XRD Rietveld analysis for the furnace-cooled sample. They reported about 0.4 occupancy of Ti4+ in 4c site. It should be noted that the experimental data from Ohya et al.42) and Skala et al.45) are consistent each other, and they are reasonably well reproduced in the present study. As discussed above in section 2, the end member Gibbs energy of GTA was determined to reproduce this cation distribution data.

Fig. 5.

Calculated cation distribution in Al2TiO5 pseudobrookite solution along with the experimental data.42,45)

4.2.3. Phase Diagram Data

The phase equilibria of the Al2O3–TiO2–Ti2O3 system strongly dependent on temperature and oxygen partial pressure.

(1) Air Atmosphere

The calculated phase diagram of the Al2O3–TiO2 system in air atmosphere is shown in Fig. 6 along with the experimental data23,24,28,29,30,34) and current experimental data discussed below. The Al2O3–TiO2 was first studied by von Wartenberg and Reusch23) and by Bunting24) using thermal analysis (TA) technique and petrographic microscope (PM). The former gave the intermediate compound to be Al2Ti2O7, while the later reported Al2TiO5 (tialite). Later, Gulamova and Sarkisova30) conducted similar experiments using TA signifying Al2TiO5 as intermediate compound. Goldberg28) preformed phase equilibrium experiments using quenching technique (QT) and optical microscopic (OM) analysis and reported the eutectic at 1703°C and 80.13 mole% TiO2. Slepetys and Vaughan29) determined the Al2O3 solubility in TiO2 from 1200 to 1426°C ranging from 0.49 mol% to 1.55 mol%. They provided a temperature error range of 30–60°C. Recently, Ilatovskaia et al.34) performed thermal analysis (TA–DSC/DTA) and quenching experiments along with XRD and SEM/EDX for phase analysis between 1060 to 1840°C.

Fig. 6.

Calculated phase diagram of the Al2O3–TiO2 in air along with the experimental data from the literature23,24,28,29,30,34) and current experimental data.

Unfortunately, as the solubility of Al2O3 in TiO2 rutile solution is not available near the eutectic temperature between Al2TiO5 and TiO2 at about 1710°C, we performed the phase equilibrium experiments at the present study to constrain the solubility limit of Al2O3 in TiO2 rutile solution. Two samples were prepared for equilibration at 1640 and 1660°C. From the current experimental data, 2 mol% Al2O3 solubility in rutile solid solution was found at this high temperature. As Ti3O5 is not stable in air, no measurable inhomogeneity range of Al2TiO5 pseudobrookite phase was detected at 1640 and 1660°C. Wartenburg and Reusch,23) Goldberg28) and Ilatovskaia et al.34) experimental measurements were used to fix the liquidus between Al2O3 and TiO2. To fix the Al2O3 solubility in TiO2, measurements performed from the current study and from Sleptys and Vaughan29) were used. Ilatovskaia et al.34) measurements were used to fix the eutectic temperature and decomposition temperature of Al2TiO5. The optimized eutectic (L → Al2TiO5 + TiO2) temperature and composition are 1710°C and 81.80 mole% TiO2, respectively which is close to the experimental data of Goldberg28) (1703°C and 81.13 mole% TiO2) and Ilatovskaia et al.34) (1715°C and 83.70 mole% TiO2), as shown in Fig. 6.

The melting temperature (between 1850 and 1863°C) and decomposition temperature (between 1240 and 1295°C) of Al2TiO5 pseudobrookite solution in air atmosphere were determined by many authors using different techniques,24,25,28,29,33,46,47,48,49,50) as shown in Table 3. There was a dispute on the presence of a high temperature compound or a high allotropic form of Al2TiO5 close to its decomposition temperature. Lang et al.46) performed experiments by visually measuring the melting temperatures for different compositions combined with petrographic and XRD studies of unquenched samples and suggested the presence of high allotropic form of Al2TiO5 between 1820 to 1860 ± 10°C. Due to the extreme experimental conditions of high temperature and high melt viscosity, they propose two versions of their phase diagrams with distinct invariant points: one with a eutectic point (L → Al2TiO5 + Al2O3), the other with a peritectic point (L + Al2O3 → Al2TiO5). The optimized melting and decomposition temperatures of Al2TiO5 are 1860 and 1279°C, respectively, in the present study.

Table 3. Optimized melting and decomposition temperature of Al2TiO5 (tialite) compared with experimental data.24,25,28,29,33,50,51,52,53,54)
Temp., °CTechniqueReferenceTemp., °CTechniqueReference
Al2TiO5 → Liq (Melting)AlTi2O5 → Al2O3 + TiO2 (Decomposition)
1860TA, PM24)1240QM, XRD29)
1850 ± 30TA25)1280QM, XRD33)
1860QM, OM28)1262 ± 7QM, PM, XRD47)
1850DTA29)1295IH48)
1863VT, OM, XRD46)1280QM, XRD49)
1860OptimizedPresent Study1280In-situ ND50)
1279OptimizedPresent Study

QM: Quenching Method, OM: Optical Microscopy, DTA: Differential Thermal Analysis, VT: Visual Technique, XRD: X-Ray Diffraction, TA: Thermal analysis, PM: Petrographic Microscopy, IH: Isothermal Heating, ND: Neutron Diffraction

Apart from Al2TiO5, Lejus et al.51) found two new compounds with compositions between 33 and 40 mol% TiO2 which are stable in the approximate temperature range between 1800 and 1900°C. In the previous optimization, Jung et al.2) adopted Al4TiO8 as the high temperature compound. However, Hoffman et al.,39) Norberg et al.,40) and Berger et al.41) suggested the high temperature phase to be Al6Ti2O13. Norberg et al.40) was first to synthesize single crystal Al6Ti2O13 and confirmed by energy-dispersive XRD. They proposed a refined structure of Al6Ti2O13 and indicated the crystal structure to be orthorhombic and space group Cm2m which is similar to mullite, Al6Si2O13. Later on, Berger et al.41) perform directional solidification experiments at three compositions (0.11, 0.26, 0.439 mole fraction of TiO2) at Al2O3 rich side using laser heat floating zone method along with microstructural characterization using XRD, WDX, SEM, HRTEM and STEM-EDX. From their observations, they suggested the formation of high temperature phase, Al6Ti2O13 and the invariant reaction to be eutectic (L → Al2TiO5 + Al6Ti2O13) and peritectic (L + Al2O3 → Al6Ti2O13), which confirms the suggestion made by Lang50) but with different invariant reactions. Berger et al.41) could not able to measure the exact temperature of the eutectoid or solidification reaction. In the present study, Al6Ti2O13 is believed to be stable at high temperature between 1797°C to 1857°C as shown in Fig. 6.

(2) Reducing Atmosphere

The calculated phase diagram of the Al2O3–Ti2O3 system in reducing atmosphere (pO2 = 10-16 atm) is shown in Fig. 7 along with the experimental data discussed below. McKee and Aleshin25) determined the solubility of Ti2O3 in the Al2O3 corundum solution in H2 gas atmosphere at 1400, 1600, 1700°C. The used lattice parameter data from X-ray diffraction measurements of quenched samples to measure the Ti2O3 solubility in corundum. No solubility of Al2O3 in Ti2O3 was found in their study. Later, Belon and Forestier27) investigated the phase diagram under vacuum using both X-ray diffraction technique and optical pyrometer. They reported a maximum solubility of Ti2O3 in Al2O3 of 12.5 mol% near 1700°C. The eutectic temperature was found to be about 1695°C from high temperature optical pyrometer, which was well reproduced in our present optimization within experimental error limits. Horibe and Kuwabara26) used experimental technique to determine the solubility of Ti2O3 in Al2O3 corundum solution under high purity Ar gas atmosphere along with liquidus measurements throughout the whole composition range. They prepared the Ti2O3 by reducing TiO2 under CO gas at 1400°C. They reported the solubility of Ti2O3 in Al2O3 to be 2.7 mol % at 1710°C. Yasuda et al.31) performed XRD measurements on Al2O3-3 mol% Ti2O3 quenched samples in order to determine the phases and calculated the Ti2O3 solubility in Al2O3 corundum solid solution using lattice constant measurements. Equilibrium experiments were performed at Ar-5% H2 gas (reducing atmosphere) for 1400, 1500, 1600 and 1700°C. As, the oxygen partial pressure is not accurately known for pure gases (H2, Ar, Ar-5%H2) and vacuum for above experimental data, therefore, are being compared at reducing atmosphere (pO2 = 10-16 atm) as shown in Fig. 7. Recently, Ohta and Morita32) determined the reciprocal solid solubility of Ti2O3 and Al2O3 in corundum and ilmenite solid solution at 1600°C in oxygen partial pressure, respectively at (pO2 = 10-16 atm). The XRD lattice parameter data (d-spacing) of quenched sample (using Ar gas) at 50 mass% TiO2 was used to determine the reciprocal solubility. They reported solubility of 4.57 mol% Ti2O3 in Al2O3 corundum solid solution and 5.07 mol% Al2O3 in Ti2O3-rich ilmenite solid solution. Mizoguchi and Ueshima33) performed ternary phase equilibrations in the CaO–Ti2O3–Al2O3 system at pO2 = 10−14 – 10−15 atm and reported 21.3 and 17.6 mol% Al2O3 solubility in the Ti2O3 ilmenite solid solution at 1550 and 1600°C, respectively. The samples were quenched using Ar gas and phases were analyzed by SEM-EDS.

Fig. 7.

Calculated phase diagram of the Al2O3–Ti2O3 in reducing atmosphere (pO2 = 10-16 atm) along with the experimental data.26,27,28,31,32,33)

In Fig. 7, the phase diagram was calculated at pO2 = 10−16 atm. In the present evaluation, more preference was given to solubility data by Ohta and Morita32) due to its well-controlled oxygen partial pressure in the experiments. The solubility data by Mizoguchi and Ueshima33) are strangely much larger than other studies.25,26,31,32) The eutectic temperature and liquidus in the present study are consistent with the experimental data by Belon and Forestier27) and Horibe and Kuwabara,26) respectively.

(3) Al2TiO5–Ti3O5 Section

As pointed out in the thermodynamic model, both Al2TiO5 and Ti3O5 belong to pseudobrookite solid solution. Matsuura1) equilibrated liquid Fe–Al–Ti metal with Al2TiO5 pellet in Al2O3 crucible at 1600°C for 3 hours under Ar-3%H2 atmosphere. Then, the interface between metal and Al2TiO5 was analyzed using SEM-EDS, which showed the formation of Ti oxide phase containing about 10 mol% Al and Al2O3 phase containing about 3 mol%Ti. This Ti oxide phase is believed to be Ti3O5 phase containing a considerable amount of Al2TiO5. In fact, this is the only literature data which revealed the solubility of Al in Ti3O5 phase. Figure 8 shows the calculated isopleth of Al2TiO5–Ti3O5 in this study. The result of Matsuura1) is plotted in the figure. In order to reproduce the experimental data, a large miscibility gap between Al2TiO5 and Ti3O5 was considered in the pseudobrookite solid solution.

Fig. 8.

Calculated isopleth of the Al2TiO5–Ti3O5 along with the experimental data by Matsuura.1)

(4) Liquidus Projection

The calculated liquidus projection of Al2O3–TiO2–Ti2O3 system from the present study is shown in Fig. 9 and the invariant reaction points are listed in Table 4. A small adjustable ternary model parameter is used in the Al2O3–Ti2O3–TiO2 liquid oxide phase in the present study to make sure that there is no liquid phase forming between Al2TiO5 and Ti3O5 at 1600°C (see Fig. 8). There is no phase diagram measurement available in the literature for this ternary system. No ternary solid phases have been reported in the literature. Pseudobrookite solid solutions containing Al2TiO5-rich and Ti3O5-rich phase can be clearly seen in the liquidus projection.

Fig. 9.

Calculated (predicted) liquidus projection of the Al2O3–TiO2–Ti2O3 system.

Table 4. Invariant reactions in the Al2O3–TiO2–Ti2O3 system involving liquid phase, calculated from the present optimization.
No.aPhases in equilibrium with liquid*Composition
(mole fraction)
Temp., (°C)
Al2O3Ti2O3TiO2
1Cor + Psb + Al6Ti2O130.5660.0280.4061829
2Rut + Ti10O19 + Ti20O390.0050.2100.7851714
3Rut + Ti9O17 + Ti10O190.04402130.7431675
4Rut + Ti8O15 + Ti9O170.1010.2050.6931639
5Ilm + Cor + Psb0.2420.4400.3181611
6Psb + Rut + Ti8O150.1530.1960.6511610
7Psb + Ti7O13 + Ti8O150.1530.2030.6441607
8Cor + 2 Psb0.2570.2870.4561602
9Psb + Ti6O11 + Ti7O130.1540.2220.6241599
10Psb + Ti5O9 + Ti6O110.1550.2430.6021591
a  The number index of the invariant reaction points are the same as the ones in Fig. 8.

*  Cor: Corundum, Psb: Pseudobrookite, Ilm: Ilmenite, and Rut: Rutile phase

Figure 10 predicts the isothermal phase diagram between Al2O3–TiO2–Ti2O3 at 1600°C. Although the phase diagram between Al2O3–TiO2 at air (see Fig. 6) and Al2O3–Ti2O3 at pO2 = 10−16 atm (see Fig. 7) shows no stable liquid phase at 1600 °C, the ternary liquid interpolation technique (without any parameter) used in current modeling predicts the stability of liquid phase up to 1590°C. Therefore, a small positive ternary liquid parameter was used for reproducing the experimental data of Matsuura1) at 1600°C, which shows immiscibility within Al2TiO5–Ti3O5 system (see Fig. 9). As, there are no direct phase equilibrium experiments to confirm the phase equilibria between Al2O3–Ti2O3–TiO2 at 1600°C (shown in Fig. 10), few indirect experiments confirmed the stability of oxides. Doo et al.52) reported the inclusions in Al-killed Ti-bearing steel in the RH process at 1600°C. They found rugged-surface spherical complex oxide inclusions, composed of Al2O3 corundum and Al–Ti–O phase with Ti/Al ratio of 1.0 to 2.3 found using EPMA. The inclusion data support the existence of pseudobrookite phase complex phase (Ti3−xAlxO5, 0 ≤ x ≤ 2) in equilibrium with Al2O3-rich corundum phase, as shown as shaded area in Fig. 10.

Fig. 10.

Calculated (predicted) isothermal phase diagram of the Al2O3–TiO2–Ti2O3 system at 1600°C.

5. Inclusion Stability Diagram and Nozzle Clogging of Al-killed Ti Bearing Steel

During the continuous casting of Al-killed Ti bearing steel, the steel consisting of non-metallic inclusions (Al2O3, Ti2O3, Al2TiO5, Ti3O5) and solidified liquid oxide (Al2O3–TiOx) causes severe clogging of submerged entry nozzle.53,54) Numerous researchers51,54,55) have investigated the inclusions causing nozzle clogging in Al-killed Ti-bearing steel and Al-killed Ti-free steel and found the difference to be the existence of complex Al–Ti–O inclusion covering Al2O3. They reported the reason behind the nozzle clogging is due to reoxidation of molten steel (Al and Ti) by liquid tundish slag containing high SiO2 content, or the reducible oxide (FeO and MnO) contained in ladle slag after Al-killing. Another important reason of nozzle clogging found by various authors54,56,57) is the carbothermic reduction of impurities (SiO2, Na2O) present within Al2O3–C SEN refractories at continuous casting temperature (< 1600°C). It releases SiO and CO gases which can oxidize Al and Ti in molten steel, and formin-situ Al2O3 corundum and Al2TiO5 pseudobrookite phase on SEN surface during casting process. Therefore, accurate knowledge on thermodynamics and phase equilibria in the Al–Ti–O and Fe–Al–Ti–O system is a crucial step to predict the oxide phases in equilibrium with liquid steel for different steel compositions.

Ruby-Meyer et al.58) was first to propose the “inclusion stability diagram” in the Fe–Al–Ti–O system at 1580°C using the CEQSI code based on slag model.59) They showed the formation of Al2TiO5 (stoichiometric compound), and liquid (TiOx–Al2O3) phase between Al2O3 and Ti2O3 stable phase. Although Ti3O5 pseudobrookite should exist in low Al region, it was shown in the calculated diagram. Jung et al.60) proposed a new stability diagram at 1600°C by using oxide database (FToxid) from FactSage 5.3 version, to perform similar calculations and found Ti3O5 phase and Al2TiO5 phase stable between Al2O3 and Ti2O3 stable phases, and no liquid phase was calculated. Later, Jung et al.2) revised the thermodynamic database of the Al2O3–Ti2O3–TiO2 system (FactSage 6.3–7.2 version) and calculated the inclusion diagram again. Their new diagram indicated the formation of liquid phase between Ti3O5 and Al2O3 corundum phase, but no Al2TiO5 phase was calculated. Kim et al.61) also reported the phase stability diagram of oxides by equilibrating with Fe–Al–Ti melt at 1600°C in which the stable oxide phases are Al2O3, Ti2O3 and Ti3O5. Matsuura1) and Choi et al.62) experimentally measured the inclusion in Al killed and Ti bearing steel at 1600°C and proposed the formation Al2O3 phase containing TiOx and TiOx phase containing Al2O3 as stable phase for steels containing O < 60 ppm. Jo63) found out the formation of only Ti3O5 and Al2O3 stable phases for steels containing O < 40 ppm. He equilibrated the liquid Fe–Al–Ti alloy in an Al2O3 crucible at 1600°C and used ICP for analysis of Al and Ti concentrations in the liquid alloy, LECO for O concentrations, and SEM and EBSD for interface between the liquid alloy and Al2O3 crucible. Recently, Kang and Lee12) slightly revised the thermodynamic database of the Al2O3–Ti2O3–TiO2 system from Jung et al.1) and made the liquid oxide phase slightly less stable. As result, their inclusion stability diagram showed the formation of liquid oxide phase only at low Al and Ti region where [O] > 50 ppm and the formation of Al2O3 and Ti3O5 oxides at high Al and Ti region where [O] < 50 ppm. In general, most of the available inclusion diagrams in the Fe–Al–Ti–O system can be summarized into two groups: (i) diagram containing only solid Al2O3, Ti2O3, Ti3O5, and Al2TiO5 phases, and (ii) diagram containing Al2O3, Ti2O3, Ti3O5 and liquid phases.

Based on the present optimization results of the Al–Ti–O oxide system, the inclusion stability diagram for Fe–Al–Ti–O system has been re-calculated in Fig. 11. In order to construct this diagram, both oxide and liquid steel databases containing all the elements are necessary. For the oxide database, the present model parameters for Al–Ti–O system were combined with the newly optimized parameters for Fe–Ti–O system6) (Fe2O3–Ti2O3–TiO2) and Fe–Al–O (FeO–Fe2O3–Al2O3)64) systems. For the molten steel phase, the FactSage FTmisc (FeLQ) database3) was used.

Fig. 11.

Calculated (predicted) oxide stability diagram of the Fe–Al–Ti–O system at 1600°C in the present study along with the experimental data.1,58,61,63) Thin lines are iso-O (wt. ppm) lines. (Online version in color.)

Figure 11 show inclusion stability diagram of Fe–Al–Ti–O system at 1600°C along with the experimental data.1,58,61,63) Thin lines denote the calculated iso-oxygen line (ppm) and thick lines denote the phase boundaries between two different oxide inclusion phases. The most stable oxide inclusions formed during the production Al-killed Ti bearing steel are Al2O3-rich corundum phase and Ti3O5-rich pseudobrookite phase. A small stability region of Al2TiO5-rich pseudobrookite is calculated in low Al and Ti region. Most of the low Al and Ti region is occupied by liquid oxide phase mainly of FeO–Al2O3–Ti2O3–TiO2. Compared to the previous stability diagram by the present author2) the stability region of liquid oxide phase is shrunk a lot because the liquid slag phase optimized in this study is less stable than the previous one by about 50°C, as mentioned in Fig. 9. A small region of the Al2TiO5-rich pseudobrookite is calculated because this phase is described as a pseudobrookite solution phase rather than a stoichiometric Al2TiO5 solid phase as in the previous study, and its Gibbs energy is relatively more stable than the previous study. Both Ti3O5-rich pseudobrookite and liquid oxide phase contain noticeable amount of FeO in low Al and Ti region of the diagram.

Compared to the inclusion diagrams in the previous works, one of the main differences of the present diagram is the existence of both liquid and Al2TiO5 phases. The only diagram containing both phases was drawn by Ruby-Meyer et al.,58) but Ti3O5 phase was not considered in their diagram, and the liquid oxide phase always existed between Ti2O3 and Al2O3. In many Al/Ti complex deoxidation experiments,65,66,67) Al2O3/TiOx complex inclusions were frequently found in high Al and Ti content region of molten steel, and the Al–Ti–O inclusions (cluster-type Al–Ti–O inclusions (solid Al2TiO5 inclusions), and random compositional Al–Ti–O inclusions (liquid oxide inclusions)) were observed in low Al and Ti region of molten steel. The occurrence of such inclusions can be more easily explained by the present inclusion diagram in Fig. 11, than the previous diagrams. That is, the formation of complex Al2O3/TiOx (TiO2 or Ti3O5) inclusions where TiOx inclusion covered by Al2O3 layer tells that the inclusion diagram should have Al2O3 and TiOx (TiO2 or Ti3O5) phase boundary without liquid phase in-between. This is well explained by the present diagram but not from the one by Ruby-Meyer et al.58) The liquid inclusion of Al2O3–Ti2O3–TiO2(–FeO) and solid Al2TiO5 phase calculated in the present diagram are also well matching with Al–Ti–O inclusion in low Al and Ti region.

In order to simulate the reoxidation of Al-killed and Ti bearing steel by CO gas in SEN at 1550°C, thermodynamic calculations were performed, and the results are given in Fig. 12. The partial pressure of CO gas was varied from 0.1 to 1 atm, and Ti content and Al content are fixed at 400 and 100 wt. ppm, respectively. As can be seen in Fig. 12, CO gas produced by carbothermic reduction process can form liquid slag, Al2TiO5 pseudobrookite, Ti3O5 pseudobrookite and Al2O3 phase depending on the degree of reoxidation of molten steel. That is, the original Al-killed and Ti bearing steel where Al = about 400 wt. ppm and Ti = about 800 wt. ppm can contain Al2O3 inclusions, but the chemistry of inclusions can be changed with decreasing soluble Al and Ti by reaction with CO gas (as shown in Fig. 12). This can be also understood from the inclusion diagram in Fig. 11. Such oxide products from reoxidation can be deposited on SEN which results in nozzle clogging.

Fig. 12.

Reoxidation of Al-killed Ti bearing steel by CO gas at 1550°C, (a) Ti = 400 wt. ppm and (b) Al = 100 wt. ppm.

6. Conclusions

A complete review and critical evaluation of all available phase diagrams, thermodynamic and structural data for the Al2O3–TiO2–Ti2O3 system at a total pressure of 1 atm has been carried out. The equilibrium experiments of Al2O3–TiO2 in air were also performed and revealed about 2 mol% solubility of Al2O3 in TiO2 rutile solution. The thermodynamic model with optimized parameters can reproduce most of reliable experimental data from room temperature to above liquidus under the oxygen partial pressures ranging from 10−24 to 1 atm within ± 10°C and ± 1 mol% compositional range. The unexplored phase equilibria and thermodynamic properties of the system are also predicted. Compared to the previous optimization, the major update of this study is the modeling of Al2TiO5 and Ti3O5 as part of pseudobrookite solid solution. The stability diagram of Fe–Al–Ti–O system was calculated using the new optimization results. In addition, the possible reoxidation phenomenon of Al-killed and Ti bearing steel by CO gas is calculated. The present database containing optimized parameters along with general thermodynamic software, such as FactSage, can calculate phase equilibria and thermodynamic properties at any given set of conditions under 1 atm total pressure.

Acknowledgements

Financial support from Hyundai Steel, JFE Steel, Nippon Steel and Sumitomo Metals Corp., Nucor Steel, POSCO, RHI, RioTinto Iron and Titanium, RIST, Schott A. G., Tata Steel Europe, Voestalpine Stahl, and the Natural Sciences and Engineering Research Council of Canada are gratefully acknowledged. This work was also supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. NRF- 2015R1A5A1037627). One of the authors (S. K. Panda) would like to thank the McGill Engineering Doctorate Award (MEDA) program.

References
 
© 2020 by The Iron and Steel Institute of Japan
feedback
Top