Next Article in Journal
A Review of Supercritical CO2 Fracturing Technology in Shale Gas Reservoirs
Previous Article in Journal
Review of Methods for Obtaining Rare Earth Elements from Recycling and Their Impact on the Environment and Human Health
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Ti/Al Ratio on Formation of Ti-Al Intermetallics/TiB2 Composites by SHS from Ti-Al-B Powder Mixtures

Department of Aerospace and Systems Engineering, Feng Chia University, Taichung 407102, Taiwan
*
Author to whom correspondence should be addressed.
Processes 2024, 12(6), 1237; https://doi.org/10.3390/pr12061237
Submission received: 24 April 2024 / Revised: 14 June 2024 / Accepted: 15 June 2024 / Published: 16 June 2024
(This article belongs to the Special Issue Composites by Metallurgy and Combustion Synthesis)

Abstract

:
Ti-Al intermetallics/TiB2 composites were prepared from elemental powder mixtures by the method of self-propagating high-temperature synthesis (SHS). Reactant mixtures were formulated to contain two parts; one group was (2Ti + 4B) to form 2TiB2 and the other group was (Ti + xAl) to produce Ti-Al intermetallic compounds. The content of Al ranged between x = 0.33 and 3.0, which was equivalent to the Ti/Al atomic ratio from Ti-25% Al to Ti-75% Al in the (Ti + xAl) group. The results showed that the increase of Al percentage reduced the overall combustion exothermicity and led to a slower self-sustaining combustion wave speed and a lower combustion temperature. Apparent activation energy of the Ti-Al-B solid-state combustion reaction was determined to be 114.7 kJ/mol by this study. Based on the XRD analysis, Ti-Al intermetallics/TiB2 composites featuring Ti3Al, TiAl, TiAl2, and TiAl3 as the dominant aluminide phase were respectively synthesized from the samples of Ti-25%~40% Al, Ti-50%~60% Al, Ti-71.4% Al, and Ti-75% Al. For the samples of Ti-25% Al and Ti-30% Al, Ti3Al was the only aluminide formed. The microstructure of the composites exhibited that TiB2 grains with a columnar shape of 2–3 μm in length were well distributed and embedded in the aluminide matrix. This study demonstrated an effective and energy-saving fabrication route for producing Ti-Al intermetallics/TiB2 composites with different dominant aluminide phases.

1. Introduction

Titanium aluminide-based alloys and composites have attracted much attention for high-temperature structural applications in the automotive, aerospace, and nuclear industries [1,2]. Ti-Al intermetallics possess many unique properties, such as low density, high strength, excellent oxidation resistance, and creep resistance at elevated temperatures [1,2,3]. Alloying additions of 1–3 at.% of Cr, Nb, Ta, Mn, V, Mo, and Zr led to an increase in the low-temperature ductility and high-temperature strength as well as oxidation resistance of the Ti-Al intermetallics [3,4]. Formation of TiAl-based composites by adding ceramic particles has been shown to improve the mechanical and tribological properties of the Ti-Al intermetallics [5]. Ceramic compounds, for example TiB2, Al2O3, B4C, SiC, Ti5Si3, and Ti2AlC, are thermally and chemically stable with the Ti-Al matrix and have been considered as the effective reinforcements [6,7,8,9,10,11,12,13,14,15]. Among most of the ceramic reinforcements, the coefficient of thermal expansion (CTE) of TiB2 is similar to that of γ-TiAl alloys. Moreover, TiB2 possesses high elastic modulus and outstanding wear resistance, and shows equal density to, and excellent chemical compatibility with, γ-TiAl alloys [6]. Fine TiB2 particles with special morphology and induced bridging effect contributed to the notable fracture toughness of the TiAl/TiB2 composite [7,8]. Therefore, in situ TiB2 particles were found to significantly enhance the strength of TiAl alloys and have been adopted as a promising reinforcement for γ-TiAl based composites. For the TiB2/TiAl composites prepared by the method of hot isostatic pressing, Li et al. [7] confirmed the strengthening effect of TiB2 particulates and obtained the tensile stress of 653.4 ± 39.2 MPa and strain of 2.1 ± 0.33% for the TiB2/TiAl composites. Han et al. [8] fabricated TiB2-strengthened TiAl-based alloys by the induction melting method and reported that such composites at 700 °C showed fracture toughness (KIC) of 21.89 ± 0.84 MPa m1/2, ultimate tensile strength (UTS) of 545 ± 19 MPa, and elongation-to-failure (εf) of 3.9% ± 0.18.
TiAl-based composites composed of in situ TiB2 particles and 3 at.% Cr were prepared by hot pressing (HP) sintering from mixed Ti, Al, B, and Cr powders at 10 MPa and 1300 °C, and composites with a significant improvement in the yield strength and compressive strength were attained [6]. Li et al. [7] fabricated TiAl/TiB2 composites alloyed with Nb, Cr, and C by hot isostatic pressing (HIP) at 1150 °C and 120 MPa for 3 h and showed that the tensile strength and strain of the composites increased with TiB2 content. Lazurenko et al. [9] utilized spark plasma sintering (SPS) with Ti and Al foils and TiB2 and TiC particles as the starting materials at 1250 °C to produce Ti3Al/TiAl matrix composites reinforced by TiB2 and Ti2AlC. The composites displayed an improved compressive strength and better creep resistance [8]. With the use of the HP method operating at 1280 °C and 30 MPa for 2 h, a TiAl/Ti3Al/Al2O3 composite was produced from pre-alloyed Ti, Al, Nb2O5, Nb, and B powder mixture [10]. In situ synthesis of a TiAl/Al2O3 composite from Ti, Al, Nb, W, and TiO2 mixture was conducted by mechanical milling with subsequent spark plasma sintering [11]. Tan et al. [12] prepared Ti3Al/TiAl composites with added SiC fibers by vacuum arc melting. The addition of SiC led to the formation of Ti5Si3 and Ti2AlC, both of which contributed to better strength and ductility of the products [12]. Wang et al. [13] fabricated Ti2AlC/TiAl composites with excellent room-temperature compression strength and fracture strain by the SPS technique using pre-alloyed Ti-48Al-2Cr-2Nb powders and graphene nanosheets as raw materials and operating at different temperatures of 1200–1320 °C for 10 min. TiAl/Ti5Si3 composites characterized by superplastic deformation were produced by high-energy ball milling of Ti, Al, and Si powders, followed by HIP at 1050–1150 °C and 300 MPa for 2 h [14]. Shu et al. [15] synthesized TiB2- and Ti5Si3-reinforced TiAl composites from elemental powder blends by the thermal explosion method. TiB2 particles were found to enhance the ultimate compression strength and Ti5Si3 particles to increase the strength and ductility of the composite [15]. Through combustion synthesis in the mode of self-propagating high-temperature synthesis (SHS), Yeh and Li [16] studied the formation of TiAl/Ti5Si3 and TiAl/Al2O3 composites. From elemental powder compacts, TiAl/Ti5Si3 composites were produced and the increase of the Ti5Si3 content facilitated sustainability of a steady combustion process. TiAl/Al2O3 composites were prepared from Ti, Al, and TiO2 powders by the SHS process involving a thermite reaction in which TiO2 was reduced by Al. The increase of Al2O3 lowered combustion exothermicity and was responsible for combustion wave propagation in a pulsating mode [16].
Of various fabrication methods, the SHS technique takes advantage of the highly exothermic reaction and is an energy-efficient, time-saving, facile, and low-cost approach. Moreover, the SHS method is an in situ fabrication route for preparation of composite materials [16,17]. The objective of this study was to investigate in situ formation of Ti-Al intermetallics/TiB2 composites by the SHS process from elemental powder mixtures consisting of Ti, Al, and boron. According to the Ti-Al phase diagram [18,19,20], Ti-Al intermetallic compounds include α2-Ti3Al, γ-TiAl, TiAl2, and TiAl3. Ti3Al has a wide homogeneity range from 22 at.% Al to approximately 35 at.% Al. Depending on temperature, the TiAl phase has a composition of 49–66 at.% Al. Two-phase alloys (γ-TiAl + α2-Ti3Al) exist between 35 at.% and 49 at.% Al. For the Al-rich compounds, TiAl2 has a narrow homogeneity range of 64.5–67 at.% Al and TiAl3 is a line compound. In this study, the effects on the combustion wave temperature and velocity of the Ti/Al atomic ratio from Ti-25 at.% Al to Ti-75 at.% Al were explored, as well as those on the Ti-Al intermetallic phases formed in the composite. Unlike most of the previous studies focusing only on the γ-TiAl phase, this work represents the first attempt to investigate the formation of Ti-Al intermetallics/TiB2 composites featuring four different dominant aluminide phases. The activation energy of solid-state combustion of the Ti-Al-B system was deduced from combustion kinetics. Microstructure and composition of the synthesized composites were examined.

2. Materials and Methods

The materials used in this study included Ti (Alfa Aesar, Ward Hill, MA, USA, <45 μm, 99.8%), Al (Alfa Aesar, <45 μm, 99.8%), and amorphous boron (B) (Noah Technologies, San Antonio, TX, USA, < 1 μm, 93.5%). Reactant mixtures were formulated as in Equation (1):
2 T i + 4 B + T i + x A l 2 T i B 2 + ( T i 3 A l , T i A l , T i A l 2 , a n d / o r T i A l 3 )
As expressed in Equation (1), the amount of Ti in the reactant blend was divided into two groups. One part of Ti is to react with boron (i.e., 2Ti + 4B) to produce TiB2, which is a highly exothermic phase. The other part of Ti is to react with Al to produce Ti-Al intermetallic compounds. The phase of titanium aluminide depends on the content of Al in the (Ti + xAl) group. As shown in Table 1, the parameter x varies from 0.33 to 3.0, which corresponds to the Ti/Al atomic ratio from Ti-25 at.% Al to Ti-75 at.% Al. The expression of Ti-25% Al means that the atomic percentage of Al in the (Ti + xAl) group with x = 0.33 is 25%. Likewise, Ti-75% Al represents 75% of Al in the (Ti + xAl) group for the case of x = 3.0. This type of expression for the Ti/Al atomic ratio was used in the following text. Based on stoichiometric balance under different values of x, Table 1 lists either sole or mixed Ti-Al intermetallic compounds to be formed as the products. It should be noted that the sample of x = 0.43 (Ti-30% Al) is an Al-rich condition to produce Ti3Al, because Ti3Al exists in a wide homogeneity range from 22 to 35 at.% Al.
The combustion exothermicity of Equation (1) was evaluated by calculating the enthalpy of reaction (∆Hr) at 298 K and adiabatic combustion temperature (Tad) based on the intermetallic compositions presented in Table 1. Equation (2) is an energy balance equation for computing Tad [21]. ∆Hr was calculated from the heats of formation (∆Hf) of the reactants and products. The required thermochemical data including ∆Hf and heat capacity cp(Pi) of the product species are listed in Table 2 [22].
H r + 298 T a d n j c p P j d T + 298 T a d n j L P j = 0
where nj is the stoichiometric coefficient of the product and L(Pj) is the latent heat of the product.
The SHS experiments were performed in a windowed combustion chamber. The chamber was first purged with high-purity (99.99%) argon for 3 min and then filled with argon at 2 bars. Well-mixed powders were uniaxially compressed to form cylindrical test samples with a diameter of 7 mm, a height of 12 mm, and a relative density of 55%. The ignition of the powder compact was achieved by a heated tungsten coil with a voltage of 120 V and a current of 5 A. Based on the time series of recorded combustion images, the propagation velocity of the self-sustaining combustion wave (Vf) was determined from the time derivative of the combustion front trajectory. The exposure time of each recorded image was set at 0.1 ms. A beam splitter with a mirror characteristic of 75% transmission and 25% reflection was used to optically superimpose a scale onto the image of the sample in order to accurately measure the instantaneous locations of the combustion front. The combustion temperature was measured by a bare Pt/Pt-13%Rh (R-type) thermocouple with a bead diameter of 125 μm and the thermocouple was mounted at a position about 6–7 mm from the ignited top plane of the sample. Details of the experimental setup and approaches are presented elsewhere [23]. For the synthesized products, the phase composition was identified by a Bruker D2 X-ray diffractometer (XRD Phaser, Karlsruhe, Germany). Microstructure and elemental analysis of the final products were examined by scanning electron microscopy (Hitachi S-3400N, Tokyo, Japan) and energy-dispersive X-ray spectroscopy.

3. Results and Discussion

3.1. Analysis of Combustion Exothermicity

Variations in ∆Hr and Tad of the test samples containing different amounts of Al in the (Ti + xAl) group of Equation (1) are presented in Figure 1. The reaction of Equation (1) is exothermic and the enthalpy of reaction increases from 664.4 to 778.2 kJ with an increase in Al percentage from 25% to 75%. Because TiB2 and four different Ti-Al compounds were formed as exothermic phases, the increase of the total number of mole of Ti-Al intermetallic compounds produced in the composites was responsible for the increase of ∆Hr from Ti-25% Al to Ti-50% Al. For the samples containing more than Ti-50% Al, because an equal amount of Ti-Al compounds was produced, the increase of ∆Hr is ascribed to the formation of Al-rich phases, TiAl2 and TiAl3. As indicated in Table 2, the heat of formation between TiAl, TiAl2, and TiAl3 is in an ascending order.
In contrast, Figure 1 shows that the Tad of Equation (1) decreased from 3329 to 2954 K with increasing Al content in the samples from Ti-25% Al to Ti-75% Al. The high adiabatic combustion temperatures provide sufficient exothermicity for a self-sustaining reaction. Based on the heat of formation and specific heat of each compound listed in Table 2, TiB2 was the most exothermic phase to form in the composite. Within Equation (1), the reaction of Ti with boron is considered as not only the major heat-releasing step but also the initiation step. Although the interaction of Ti with Al liberates heat in producing Ti-Al intermetallics, its exothermicity is less than the formation of TiB2. As a result, the overall reaction exothermicity is weakened when the mole fraction of Ti-Al compounds augmented in the composite. This explains the descent of Tad for the samples containing Al from Ti-25% Al to Ti-50% Al. Further decline in Tad for the samples with Al percentage from Ti-50% Al to Ti-75% Al was mainly caused by the formation of Al-rich phases, TiAl2 and TiAl3, both of which have large specific heats.

3.2. Self-Propagating Combustion Wave Kinetics

Figure 2a–c illustrate typical time sequences of the SHS processes, which were respectively recorded from powder compacts formulated with Ti-25% Al, Ti-50% Al, and Ti-75% Al. It is evident that the combustion reaction started upon ignition and established a distinct combustion wave that propagated throughout the entire sample in a self-sustaining manner. Self-sustaining combustion was achieved for all samples conducted in this study. With the increase of Al percentage in the reactant mixture, as shown in Figure 2a–c, the burning luminosity decreased to some extent and the total spreading time of combustion wave was notably prolonged from about 1.50 s to 1.83 s and then to 3.97 s. These observations reflect the fact that combustion exothermicity decreased with increasing Al content in the reactant mixture.
The variation of combustion front velocity (Vf) with Al percentage in the (Ti + xAl) group of the reactant mixture is presented in Figure 3. The velocity decreased moderately from 6.6 to 5.6 mm/s as the Al content increased from Ti-20% Al to Ti-50% Al, beyond which a significant drop in combustion velocity was observed. The lowest flame velocity of about 2.2 mm/s was detected for the sample of Ti-75% Al. It is believed that energy transfer by heat conduction from the thin reaction zone to its adjacent unburned region is of critical importance to maintain a self-sustaining combustion process. As a result, the flame-front propagation velocity, in large part, is governed by the combustion wave temperature.
Figure 4 depicts four combustion temperature profiles measured from the powder compacts formulated with different Al percentages. Temperature profiles featured a sharp rise to a peak value followed by a continuous decline. This could be caused by the fact that the combustion propagation was speedy and synthesis reaction was normally confined in a hot thin layer. Moreover, the product was formed and cooled down after the passage of the combustion front. The peak value is considered as the combustion front temperature (Tc). As revealed in Figure 4, Tc reached 1655 °C for the sample of Ti-25% Al and decreased to about 1574 °C and 1442 °C with an increase in Al content to Ti-50% Al and Ti-66.7% Al, respectively. The sample of Ti-75% Al exhibited a less steep curve with Tc of about 1278 °C. More importantly, the dependency of combustion temperature on sample stoichiometry was consistent with that of combustion wave velocity.
According to combustion wave kinetics [23,24], a modified Arrhenius rate equation, Equation (3), was employed to deduce the apparent activation energy (Ea) of solid-state combustion reaction.
V f 2 = 2 λ ρ Q R T c 2 E a k o exp ( E a / R T c )
where λ is the thermal conductivity, R is the universal gas constant, Q is the heat of the reaction, and ko is a constant. Based on Equation (3), the relationship of ln(Vf/Tc)2 with 1/Tc can be established by the measured data of Vf and Tc, and the slope of a linear line correlating ln(Vf/Tc)2 and 1/Tc represents Ea/R. Figure 5 plots calculated values of −ln(Vf/Tc)2 with respect to 1/Tc and a best-fitted linear line with the slope. Activation energy Ea equal to 114.7 kJ/mol was deduced for the Ti-Al-B combustion reaction. This value is much lower than the activation energy, 539 kJ/mol, obtained in the Ti + 2B combustion reaction [25]. It is believed that due to the presence of liquid-phase Al during combustion synthesis, the activation energy of the ternary Ti-Al-B combustion system was reduced. Combustion temperatures of this study shown in Figure 4 are higher than the melting point of Al. The existence of molten species could have facilitated the diffusion of solid reactants and substantially reduced the kinetic barrier of the reaction system. A similar example confirming the aid of molten species is that the activation energy of 102.5 kJ/mol was determined for the Ti-B4C-Al-Ni combustion system to produce NiAl/TiB2/TiC composites [26].

3.3. Phase Composition and Microstructure of Synthesized Products

Figure 6a–d present XRD patterns of Ti-Al intermetallics/TiB2 composites synthesized from powder compacts with different Al percentages in the (Ti + xAl) group of reactant mixtures. The other mixture group in Equation (1) generated TiB2. The XRD analysis revealed that TiB2 was well produced from the elemental reaction between Ti and boron for all samples and the phase of the intermetallic compound was subject to the Al percentage in the (Ti + xAl) group. The sample of Ti-30% Al, shown in Figure 6a, produced only Ti3Al, which was also the only aluminide produced from the sample of Ti-25% Al. This agrees with the fact that Ti3Al has a wide homogeneity range from 22 to 35 at.% Al.
Figure 6b depicts the XRD spectrum of the product synthesized from the sample of Ti-50% Al. Formation of TiB2 along with two aluminides, TiAl and Ti3Al, was identified. The dominant intermetallic phase in Figure 6b is TiAl. The presence of Ti3Al could be due to the evaporation loss of a small amount of Al during the SHS process, or because of TiAl having a composition range from 49 at.% Al to about 66 at.% Al. Both TiAl and Ti3Al were also formed in the product from the sample with Ti-40% Al, which was consistent with the coexistence of TiAl and Ti3Al in the composition between Ti-35 at.% Al and Ti-49 at.% Al. However, the governing aluminide phase was Ti3Al in the product of Ti-40% Al.
The product composed of TiB2, TiAl, and TiAl2, as displayed in Figure 6c, is associated with the sample of Ti-66.7% Al. Since TiAl2 has a narrow homogeneity range of 64.5–67 at.% Al and TiAl has a Al-rich composition of about 60–66 at.% Al at temperatures between 1200 °C and 1400 °C, a small evaporation loss of Al could result in the formation of TiAl. It is useful to mention that the formation of TiAl was more pronounced in the product from the sample of Ti-60% Al than that revealed in Figure 6c. Namely, TiAl dominated over TiAl2 in the TiAl/TiAl2/TiB2 composite synthesized from the sample of Ti-60% Al, while TiAl and TiAl2 were nearly comparable in the product from the sample of Ti-66.7% Al.
The XRD pattern of Figure 6d represents the product obtained from the sample of Ti-75% Al. The yield of TiAl3 as the major aluminide is identified. The presence of a small amount of TiAl2 in the product was because TiAl3 is a line compound. Therefore, a slight deficiency of Al might lead to the formation of TiAl2. It should be noted that a composite consisting of TiB2, TiAl2, and TiAl3 was also synthesized from the sample of Ti-71.4% Al and the content of TiAl2 relative to TiAl3 was more than that observed in Figure 6d. Table 3 summarizes the major and minor phases of titanium aluminides formed in the TiB2-containing composites from combustion of Equation (1) under different contents of Al. It is evident that Ti-Al intermetallics/TiB2 composites featuring four different dominant aluminide phases were produced.
Figure 7 presents the SEM image and EDS spectrum, showing fracture surface microstructure and elemental composition of the product synthesized from the sample of Ti-30% Al. Most of the TiB2 grains were formed in a columnar shape with a length of 2–3 μm and were embedded in the Ti-Al aluminide matrix. According to the formation mechanism of Ti-Al intermetallics [3,27], the stages involve the initial melting of Al, subsequent spreading of molten Al through channels of the capillary-porous medium, and diffusion of aluminum atoms into the titanium lattice. This resulted in the formation of the TiAl3 compound, which can transform into TiAl2, TiAl, and Ti3Al as it is saturated with titanium. The formation of TiB2 is governed by a dissolution/precipitation mechanism including solid-phase diffusion of boron into titanium and the precipitation of TiB2 particles [28,29]. The EDS spectrum of Figure 7 confirms three constituent elements, from which an atomic ratio of B:Al:Ti = 55.84:8.94:35.22 was obtained. Low Al proportion was ascribed to this composite containing a Ti-rich aluminide Ti3Al at a small mole fraction.
The SEM image and EDS spectrum presented in Figure 8 are associated with the product obtained from the sample of Ti-50% Al. The morphology clearly exhibited columnar TiB2 grains well distributed in the Ti-Al aluminide matrix. The EDS intensity of the Al peak relative to that of Ti in Figure 8 is noticeably stronger than that observed in Figure 7. The atomic ratio of the scanned region was B:Al:Ti = 45.89:15.22:38.89, which is close to a composite consisting of TiB2 and TiAl. It should be noted that the peak at around 2.1 keV is Pt (platinum). In this study, Pt was used to sputter coat SEM samples as a conductive layer.
For the sample of Ti-75% Al, the microstructural and elemental analyses of its resulting product are unveiled in Figure 9. TiB2 grains were formed in a smaller size of about 1 μm and mixed closely with the Ti-Al aluminide matrix. When compared with that in Figure 8, an Al peak with a further increase in intensity was detected in the EDS spectrum of Figure 9. An atomic ratio of B:Al:Ti = 40.54:28.64:30.82 was obtained, which suggested TiAl3 as the dominant aluminide in the Ti-Al intermetallics/TiB2 composite. It is important to note that the major titanium aluminide revealed by the atomic ratio of constituent elements from the EDS analysis of Figure 7, Figure 8 and Figure 9 was in good agreement with the intermetallic phase identified by the XRD pattern.
The green powder compacts conducted by this work had a relative density of 55%. Due to volume expansion of the burned samples after the SHS process, the final products were porous and had an estimated porosity of about 50%. As a result, composite powders were easy to obtain. In view of practical applications, the size range and external morphology of Ti-Al intermetallics/TiB2 composite powders are suitable as feedstock materials for powder metallurgy or in different thermal spraying systems for depositing protective coatings.

4. Conclusions

In situ formation of Ti-Al intermetallics/TiB2 composites was investigated by the SHS technique from Ti-Al-B powder mixtures. The reactant mixtures comprised Ti and boron in a configuration of (2Ti + 4B) to produce 2TiB2, as well as Ti and Al in a composition of (Ti + xAl) to generate titanium aluminide. The phase of aluminide compounds formed depended on the Al content in the (Ti + xAl) group. A value of x ranging from 0.33 to 3.0 was conducted, which corresponded to the Ti/Al atomic ratio from Ti-25% Al to Ti-75% Al for studying the formation of four different aluminides, Ti3Al, TiAl, TiAl2, and TiAl3.
Combustion exothermicity of the reaction was sufficient to ensure progression of the synthesis process in the SHS manner. However, combustion exothermicity was weakened by an increasing Al percentage in the reactant mixture due to the increase of aluminide compounds formed and the high specific heats of Al-rich aluminide phases. As a result, the combustion wave velocity decreased from 6.6 to 2.2 mm/s and combustion temperature from 1655 °C to 1278 °C when the Al percentage in the sample increased from Ti-25% Al to Ti-75% Al. Based on combustion wave kinetics, the apparent activation energy of 114.7 kJ/mol was deduced for the Ti-Al-B combustion reaction. The XRD analysis of the final products indicated that Ti3Al was the only intermetallic compound synthesized from samples of Ti-25% Al and Ti-30% Al. A composite containing TiB2, Ti3Al, and TiAl was produced from the sample of Ti-40% Al. TiAl was the dominant aluminide in the products from samples of Ti-50% Al and Ti-60% Al. For the samples with Ti-66.7% Al and Ti-75% Al, their resulting products were TiAl2/TiAl/TiB2 and TiAl3/TiAl2/TiB2 composites, respectively. The transformation of the dominant aluminide from the Ti-rich to Al-rich phase was a consequence of the increase of Al in the samples, and the synthesized intermetallic compounds with respect to the Ti/Al ratio in the samples were essentially in good agreement with titanium aluminides in the Ti-Al phase diagram. The microstructure of the as-synthesized Ti-Al intermetallics/TiB2 composites illustrated that columnar TiB2 grains with a length of 2–3 μm were formed and well distributed and closely engaged in the aluminide matrix. The EDS analysis confirmed constituent elements and provided rational atomic ratios between Ti, Al, and B. In summary, this study demonstrated an effective synthesis route for fabricating Ti-Al intermetallics/TiB2 composites featuring different dominant aluminide phases.

Author Contributions

Conceptualization, C.-L.Y.; validation, C.-L.Y. and Y.-C.C.; methodology, C.-L.Y. and Y.-C.C.; data curation, C.-L.Y. and Y.-C.C.; formal analysis, C.-L.Y. and Y.-C.C.; resources, C.-L.Y.; investigation, C.-L.Y. and Y.-C.C.; writing—original draft preparation, C.-L.Y. and Y.-C.C.; writing—review and editing, C.-L.Y. and Y.-C.C.; project administration, C.-L.Y.; supervision, C.-L.Y.; funding acquisition, C.-L.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science and Technology Council of Taiwan, and the grant number was NSTC 112-2221-E-035-041-MY2.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Genc, O.; Unal, R. Development of gamma titanium aluminide (γ-TiAl) alloys: A review. J. Alloys Compd. 2022, 929, 167262. [Google Scholar] [CrossRef]
  2. Liu, P.; Xie, J.; Wang, A. Recent research progress in TiAl matrix composites: A review. J. Mater. Sci. 2022, 57, 16147–16174. [Google Scholar] [CrossRef]
  3. Avdeeva, V.; Bazhina, A.; Antipov, M.; Stolin, A.; Bazhin, P. Relationship between structure and properties of intermetallic materials based on γ-TiAl hardened in situ with Ti3Al. Metals 2023, 13, 1002. [Google Scholar] [CrossRef]
  4. Musi, M.; Graf, G.; Clemens, H.; Spoerk-Erdely, P. Alloying elements in intermetallic γ-TiAl based alloys–a review on their influence on phase equilibria and phase transformations. Adv. Eng. Mater. 2023, 26, 2300610. [Google Scholar] [CrossRef]
  5. Chen, Z.; Liu, J.; Wang, Y.; Ma, T.; Zhu, D.; Xing, Q.; Feng, H.; Chen, R. High-temperature oxidation behavior of ceramic particles-reinforced TiAl composites with multilayered structure. Ceram. Int. 2024, 50, 2233–2241. [Google Scholar] [CrossRef]
  6. Wei, Y.; Qiu, F.; Shu, S.; Tong, H.; Yang, H.; Jiang, Q. Microstructure manipulation and strengthening mechanism of TiAl composites reinforced by Cr solid solution and in-situ nanometer-sized TiB2 particles. Mater. Sci. Eng. A 2022, 845, 143214. [Google Scholar]
  7. Li, W.; Yang, Y.; Li, M.; Liu, J.; Cai, D.; Wei, Q.; Yan, C.; Shi, Y. Enhanced mechanical property with refined microstructure of a novel γ-TiAl/TiB2 metal matrix composite (MMC) processed via hot isostatic press. Mater. Design 2018, 141, 57–66. [Google Scholar] [CrossRef]
  8. Han, J.; Xiao, S.; Tian, J.; Chen, Y.; Xu, L.; Wang, X.; Jia, Y.; Rahoma, H.K.S.; Du, Z.; Cao, S. Microstructure characterization, mechanical properties and toughening mechanism of TiB2-containing conventional cast TiAl-based alloy. Mater. Sci. Eng. A 2015, 645, 8–19. [Google Scholar] [CrossRef]
  9. Lazurenko, D.V.; Stark, A.; Esikov, M.A.; Paul, J.; Bataev, I.A.; Kashimbetova, A.A.; Mali, V.I.; Lorenz, U.; Pyczak, F. Ceramic-reinforced γ-TiAl-based composites: Synthesis, structure, and properties. Materials 2019, 12, 629. [Google Scholar] [CrossRef]
  10. Lu, X.; Li, J.; Chen, X.; Qiu, J.; Wang, Y.; Liu, B.; Liu, Y.; Rashad, M.; Pan, F. Mechanical, tribological and electrochemical corrosion properties of in-situ synthesized Al2O3/TiAl composites. Intermetallics 2020, 120, 106758. [Google Scholar] [CrossRef]
  11. Lu, X.; Li, J.; Chen, X.; Ran, C.; Wang, Y.; Liu, B.; Liu, Y.; Rashad, M.; Pan, F. Grinding mechanism and mechanical properties of the in-situ synthesized Al2O3/TiAl composites. Ceram. Int. 2019, 45, 12113–12121. [Google Scholar] [CrossRef]
  12. Tan, Y.; Chen, R.; Fang, H.; Liu, Y.; Ding, H.; Su, Y.; Guo, J.; Fu, H. Microstructure evolution and mechanical properties of TiAl binary alloys added with SiC fibers. Intermetallics 2018, 98, 69–78. [Google Scholar] [CrossRef]
  13. Wang, Z.; Liu, P.; Wang, A.; Xie, J.; Hou, B. Effect of spark plasma sintering temperature on the multi-scale microstructure evolution and mechanical properties of Ti2AlC/TiAl composites with network architecture. J. Mater. Res. Technol. 2023, 25, 6209–6223. [Google Scholar] [CrossRef]
  14. Suryanarayana, C.; Behn, R.; Klassen, T.; Bormann, R. Mechanical characterization of mechanically alloyed ultrafine-grained Ti5Si3+ 40 vol% γ-TiAl composites. Mater. Sci. Eng. A 2013, 579, 18–25. [Google Scholar] [CrossRef]
  15. Shu, S.; Xing, B.; Qiu, F.; Jin, S.; Jiang, Q. Comparative study of the compression properties of TiAl matrix composites reinforced with nano-TiB2 and nano-Ti5Si3 particles. Mater. Sci. Eng. A 2013, 560, 596–600. [Google Scholar] [CrossRef]
  16. Yeh, C.L.; Li, R.F. Formation of TiAl–Ti5Si3 and TiAl–Al2O3 in situ composites by combustion synthesis. Intermetallics 2008, 16, 64–70. [Google Scholar] [CrossRef]
  17. Levashov, E.A.; Mukasyan, A.S.; Rogachev, A.S.; Shtansky, D.V. Self-propagating high-temperature synthesis of advanced materials and coatings. Int. Mater. Rev. 2017, 62, 203–239. [Google Scholar] [CrossRef]
  18. Cobbinah, P.V.; Matizamhuka, W.R. Solid-state processing route, mechanical behaviour, and oxidation resistance of TiAl alloys. Adv. Mater. Sci. Eng. 2019, 2019, 1–21. [Google Scholar] [CrossRef]
  19. Lei, C.; Xu, Q.; Sun, Y.Q. Phase orientation relationships in the TiAl–TiAl2 region. Mater. Sci. Eng. A 2001, 313, 227–236. [Google Scholar] [CrossRef]
  20. Peng, M.; Shou, H.; Cao, Y. First-principles calculations of structural, elastic and thermodynamic properties of (h, r)-TiAl2. Phys. B Condens. Matter 2019, 561, 29–36. [Google Scholar] [CrossRef]
  21. Zaki, Z.I.; Ahmed, Y.M.Z.; Abdel-Gawad, S.R. In-situ synthesis of porous magnesia spinel/TiB2 composite by combustion technique. J. Ceram. Soc. Jpn. 2009, 117, 719–723. [Google Scholar] [CrossRef]
  22. Binnewies, M.; Milke, E. Thermochemical Data of Elements and Compounds; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2002. [Google Scholar]
  23. Yeh, C.L.; Chen, C. In situ formation of titanium diboride/magnesium titanate composites by magnesiothermic-based combustion synthesis. Processes 2024, 12, 459. [Google Scholar] [CrossRef]
  24. Varma, A.; Rogachev, A.S.; Mukasyan, A.S.; Hwang, S. Combustion synthesis of advanced materials: Principals and applications. Adv. Chem. Eng. 1998, 24, 79–225. [Google Scholar]
  25. Holt, J.B.; Kingman, D.D.; Bianchini, G.M. Kinetics of the combustion synthesis of TiB2. Mater. Sci. Eng. 1985, 71, 321–327. [Google Scholar] [CrossRef]
  26. Yeh, C.L.; Ke, C.Y.; Chen, Y.C. In situ formation of TiB2/TiC and TiB2/TiN reinforced NiAl by self-propagating combustion synthesis. Vacuum 2018, 151, 185–188. [Google Scholar] [CrossRef]
  27. Kurbatkina, V.V. Titanium Aluminides. In Concise Encyclopedia of Self-Propagating High-Temperature Synthesis; Borovinskaya, I.P., Gromov, A.A., Levashov, E.A., Maksimov, Y.M., Mukasyan, A.S., Rogachev, A.S., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 392–393. [Google Scholar]
  28. Shen, P.; Zou, B.; Jin, S.; Jiang, Q. Reaction mechanism in self-propagating high temperature synthesis of TiC-TiB2/Al composites from an Al-Ti-B4C system. Mater. Sci. Eng. A 2007, 454, 300–309. [Google Scholar] [CrossRef]
  29. Fjellstedt, J.; Jarfors, A.E. On the precipitation of TiB2 in aluminum melts from the reaction with KBF4 and K2TiF6. Mater. Sci. Eng. A 2005, 413, 527–532. [Google Scholar] [CrossRef]
Figure 1. Variations of enthalpy of reaction (∆Hr) and adiabatic combustion temperature (Tad) with Al percentage in the (Ti + xAl) group of reactant mixtures.
Figure 1. Variations of enthalpy of reaction (∆Hr) and adiabatic combustion temperature (Tad) with Al percentage in the (Ti + xAl) group of reactant mixtures.
Processes 12 01237 g001
Figure 2. Series of self-sustaining combustion images recorded from powder compacts of Equation (1) with (a) Ti-25% Al, (b) Ti-50% Al, and (c) Ti-75% Al (unit of scale bar: 1 mm).
Figure 2. Series of self-sustaining combustion images recorded from powder compacts of Equation (1) with (a) Ti-25% Al, (b) Ti-50% Al, and (c) Ti-75% Al (unit of scale bar: 1 mm).
Processes 12 01237 g002
Figure 3. Variation of combustion wave propagation velocity with Al percentage in the (Ti + xAl) group of reactant mixtures.
Figure 3. Variation of combustion wave propagation velocity with Al percentage in the (Ti + xAl) group of reactant mixtures.
Processes 12 01237 g003
Figure 4. Combustion temperature profiles of compacted samples formulated with Ti-25% Al, Ti-50% Al, Ti-66.7% Al, and Ti-75% Al.
Figure 4. Combustion temperature profiles of compacted samples formulated with Ti-25% Al, Ti-50% Al, Ti-66.7% Al, and Ti-75% Al.
Processes 12 01237 g004
Figure 5. Linear correlation between ln(Vf/Tc)2 and 1/Tc and activation energy (Ea) of the Ti-Al-B combustion reaction.
Figure 5. Linear correlation between ln(Vf/Tc)2 and 1/Tc and activation energy (Ea) of the Ti-Al-B combustion reaction.
Processes 12 01237 g005
Figure 6. XRD patterns of Ti-Al intermetallics/TiB2 composites synthesized from the samples containing different percentages of Al in the (Ti + xAl) group of reactant mixtures: (a) Ti-30% Al, (b) Ti-50% Al, (c) Ti-66.7% Al, and (d) Ti-75% Al.
Figure 6. XRD patterns of Ti-Al intermetallics/TiB2 composites synthesized from the samples containing different percentages of Al in the (Ti + xAl) group of reactant mixtures: (a) Ti-30% Al, (b) Ti-50% Al, (c) Ti-66.7% Al, and (d) Ti-75% Al.
Processes 12 01237 g006
Figure 7. SEM micrograph and EDS spectrum of Ti3Al/TiB2 composite synthesized from the sample of Ti-30% Al.
Figure 7. SEM micrograph and EDS spectrum of Ti3Al/TiB2 composite synthesized from the sample of Ti-30% Al.
Processes 12 01237 g007
Figure 8. SEM micrograph and EDS spectrum of TiAl/Ti3Al/TiB2 composite synthesized from the sample of Ti-50% Al.
Figure 8. SEM micrograph and EDS spectrum of TiAl/Ti3Al/TiB2 composite synthesized from the sample of Ti-50% Al.
Processes 12 01237 g008
Figure 9. SEM micrograph and EDS spectrum of TiAl3/TiAl2/TiB2 composite synthesized from the sample of Ti-75% Al.
Figure 9. SEM micrograph and EDS spectrum of TiAl3/TiAl2/TiB2 composite synthesized from the sample of Ti-75% Al.
Processes 12 01237 g009
Table 1. Ti/Al atomic ratio and Ti-Al intermetallics of Equation (1) under different values of x.
Table 1. Ti/Al atomic ratio and Ti-Al intermetallics of Equation (1) under different values of x.
xTi/Al Atomic RatioTi-Al Intermetallics
0.33Ti-25% Al1/3 Ti3Al
0.43Ti-30% Al1/3 Ti3Al
0.67Ti-40% Al1/6 Ti3Al + 1/2 TiAl
1.0Ti-50% AlTiAl
1.5Ti-60% Al1/2 TiAl + 1/2 TiAl2
2.0Ti-66.7% AlTiAl2
2.5Ti-71.4% Al1/2 TiAl2 + 1/2 TiAl3
3.0Ti-75% AlTiAl3
Table 2. Heat of formation (∆Hf) and specific heat (cp) of product species.
Table 2. Heat of formation (∆Hf) and specific heat (cp) of product species.
Product
Species
Heat of Formation ∆Hf (kJ/mol)Specific Heat, cp (J/mol·K)
TiB2–315.9 56.38 + 25.86 · 10 3 · T 1.75 · 10 6 · T 2
Ti3Al–97.9 103.3 + 15.90 · 10 3 · T 1.73 · 10 6 · T 2
TiAl–75.3 49.2 + 8.85 · 10 3 · T 0.75 · 10 6 · T 2
TiAl2–90.4 62.2 + 10.33 · 10 3 · T 0.96 · 10 6 · T 2
TiAl3–146.4 98.9 + 19.34 · 10 3 · T 1.66 · 10 6 · T 2
Table 3. Major and minor titanium aluminides of Ti-Al intermetallics/TiB2 composites synthesized from Equation (1) with different Al percentages in the (Ti + xAl) group.
Table 3. Major and minor titanium aluminides of Ti-Al intermetallics/TiB2 composites synthesized from Equation (1) with different Al percentages in the (Ti + xAl) group.
Al Percentage
in Ti + xAl
Major AluminideMinor Aluminide
Ti-25% AlTi3Al
Ti-30% AlTi3Al
Ti-40% AlTi3AlTiAl
Ti-50% AlTiAlTi3Al
Ti-60% AlTiAlTiAl2
Ti-66.7% AlTiAl2, TiAl
Ti-71.4% AlTiAl2TiAl3
Ti-75% AlTiAl3TiAl2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yeh, C.-L.; Chan, Y.-C. Effects of Ti/Al Ratio on Formation of Ti-Al Intermetallics/TiB2 Composites by SHS from Ti-Al-B Powder Mixtures. Processes 2024, 12, 1237. https://doi.org/10.3390/pr12061237

AMA Style

Yeh C-L, Chan Y-C. Effects of Ti/Al Ratio on Formation of Ti-Al Intermetallics/TiB2 Composites by SHS from Ti-Al-B Powder Mixtures. Processes. 2024; 12(6):1237. https://doi.org/10.3390/pr12061237

Chicago/Turabian Style

Yeh, Chun-Liang, and Yi-Cheng Chan. 2024. "Effects of Ti/Al Ratio on Formation of Ti-Al Intermetallics/TiB2 Composites by SHS from Ti-Al-B Powder Mixtures" Processes 12, no. 6: 1237. https://doi.org/10.3390/pr12061237

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop