Next Article in Journal
Key Performance Analysis of Emulsified Asphalt Cold Recycling Mixtures of the Middle Layer of Pavement Structure
Next Article in Special Issue
A Comparative Study on Microstructural Characterization of Thick High Strength Low Alloy Steel Weld by Arc Welding and Laser Welding
Previous Article in Journal
Critical Roles of Impurities and Imperfections in Various Phases of Materials
Previous Article in Special Issue
Multipass Friction Stir Processing of Laser-Powder Bed Fusion AlSi10Mg: Microstructure and Mechanical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Formation of TiB2–MgAl2O4 Composites by SHS Metallurgy

Department of Aerospace and Systems Engineering, Feng Chia University, Taichung 40724, Taiwan
*
Author to whom correspondence should be addressed.
Materials 2023, 16(4), 1615; https://doi.org/10.3390/ma16041615
Submission received: 30 January 2023 / Revised: 13 February 2023 / Accepted: 13 February 2023 / Published: 15 February 2023
(This article belongs to the Special Issue Physical Metallurgy of Metals and Alloys)

Abstract

:
TiB2–MgAl2O4 composites were fabricated by combustion synthesis involving metallothermic reduction reactions. Thermite reagents contained Al and Mg as dual reductants and TiO2 or B2O3 as the oxidant. The reactant mixtures also comprised elemental Ti and boron, as well as a small amount of Al2O3 or MgO to serve as the combustion moderator. Four reaction systems were conducted and all of them were exothermic enough to proceed in the mode of self-propagating high-temperature synthesis (SHS). The reaction based on B2O3/Al/Mg thermite and diluted with MgO was the most exothermic, while that containing TiO2/Al/Mg thermite and Al2O3 as the diluent was the least. Depending on different thermites and diluents, the combustion front temperatures in a range from 1320 to 1720 °C, and combustion wave velocity from 3.9 to 5.7 mm/s were measured. The XRD spectra confirmed in situ formation of TiB2 and MgAl2O4. It is believed that MgAl2O4 was synthesized through a combination reaction between Al2O3 and MgO, both of which can be totally or partially produced from the metallothermic reduction of B2O3 or TiO2. The microstructure of the TiB2–MgAl2O4 composite exhibited fine TiB2 crystals surrounded by large densified MgAl2O4 grains. This study demonstrated an energy-saving and efficient route for fabricating MgAl2O4-containing composites.

1. Introduction

TiB2 has been one of the most studied ultra-high temperature ceramics (UHTCs) due to its unique properties, including a high melting point (3225 °C), high hardness (33 MPa), high Young’s modulus (530 MPa), excellent wear and oxidation resistance, thermal shock resistance, chemical inertness, and good electric conductivity [1,2,3]. Combination of these properties makes TiB2 an ideal candidate for use in ballistic armors, crucibles, metal evaporation boats, cutting tools, wear resistance parts, and cathodes for alumina smelting [4,5,6]. Many ceramic phases, such as Al2O3, SiC, B4C, and MgAl2O4, have been considered as the reinforcement to improve fracture toughness, oxidation resistance, heat resistance, and mechanical strength of the TiB2-based composites [7,8,9,10]. Moreover, a recent study showed that TiB2–Al2O3–MgAl2O4 composite possesses temperature insensitive and enhanced microwave absorption properties [11]. Magnesium aluminate spinel, MgAl2O4, as an additive has been rarely studied, possibly because the fabrication of MgAl2O4 via either the direct solid-state reaction of oxides or wet chemical methods requires multiple steps that are complicated and time-consuming [12,13,14]. However, MgAl2O4 is an attractive component due to its high melting point, chemical inertness, high hardness, corrosion resistance, high mechanical strength, and low cost [12].
Among various fabrication routes for preparing multiphasic ceramics, metallothermic reduction reactions (i.e., thermite reactions) combined with combustion synthesis have been recognized as a promising technique for in situ formation of MgAl2O4-containing composites [15]. Combustion synthesis in the mode of self-propagating high-temperature synthesis (SHS), which is based on strongly exothermic reactions, has merits of low energy consumption, short reaction time, simple equipment and operation, high-purity products, and in situ formation of composite components [16,17,18]. Moreover, aluminothermic and magnesiothermic reduction reactions have Al2O3 and MgO as their respective by-products, both of which are precursors for the formation of MgAl2O4. Consequently, Omran et al. [19] applied the reduction-based SHS technique to produce MgAl2O4–W–W2B composites through magnesiothermic reduction of B2O3 and WO3 in the presence of Al2O3. Zaki et al. [20] synthesized MoSi2– and Mo5Si3–MgAl2O4 composites by SHS with a reducing stage from raw materials consisting of MoO3, SiO2, and Al as aluminothermic reagents and MgO as a precursor. Similarly, MgO was added into the reactive mixture of TiO2, B2O3, and Al to fabricate TiB2–MgAl2O4 composites by thermitic combustion synthesis [21]. Generally, most of the previous studies on the formation of MgAl2O4-containing composites had prior addition of one of two precursors (Al2O3 or MgO) in the green samples.
This study represents the first attempt to prepare TiB2–MgAl2O4 composites from the SHS powder metallurgy simultaneously involving aluminothermic and magnesiothermic reduction of TiO2 or B2O3. Only a small amount of Al2O3 or MgO was included in the reactive mixture to serve as the combustion moderator and part of the precursors for the formation of MgAl2O4. Four SHS reaction systems formulated with different metallothermic reagents and combustion diluents were investigated. In this work, combustion exothermicity and kinetics of the combustion wave of the SHS process, as well as compositions and microstructures of the final products were explored.

2. Materials and Methods

The starting materials adopted by this study included TiO2 (Acros Organics, Geel, Belgium, 99.5%), B2O3 (Acros Organics, 99%), Al2O3 (Alfa Aesar, Haverhill, MA, USA, 99%), MgO (Acros Organics, 99.5%), Al (Showa Chemical Co., Tokyo, Japan, <45 μm, 99.9%), Mg (Alfa Aesar, <45 μm, 99.8%), Ti (Alfa Aesar, <45 μm, 99.8%), and amorphous boron (Noah Technologies, San Antonio, TX, USA, <1 μm, 93.5%). Four SHS reactions were formulated for the synthesis of 3TiB2–MgAl2O4 composites. Two metallothermic reagents (i.e., thermites) were considered; one is composed of TiO2, Al, and Mg, as shown in Equations (1) and (2), and the other comprises B2O3, Al, and Mg, as in Equations (3) and (4). Due to strong exothermicity of combustion, Al2O3 with an amount of 0.3 mol. was included in Equations (1) and (3) as the combustion moderator (or combustion diluent) in order to attain stable propagation of the combustion wave. The pre-added Al2O3 also acted as part of the precursor for the synthesis of MgAl2O3. Likewise, an equal amount of MgO was adopted by Equations (2) and (4) and MgO played the same role as Al2O3 in Equations (1) and (3).
( 1.55 T i O 2 + 1.4 A l + M g ) + 0.3 A l 2 O 3 + 1.45 T i + 6 B 3 T i B 2 + M g A l 2 O 4
( 1.85 T i O 2 + 2 A l + 0.7 M g ) + 0.3 M g O + 1.15 T i + 6 B 3 T i B 2 + M g A l 2 O 4
( 1.033 B 2 O 3 + 1.4 A l + M g ) + 0.3 A l 2 O 3 + 3 T i + 3.934 B 3 T i B 2 + M g A l 2 O 4
( 1.233 B 2 O 3 + 2 A l + 0.7 M g ) + 0.3 M g O + 3 T i + 3.534 B 3 T i B 2 + M g A l 2 O 4
Combustion exothermicity of the above four reactions, Equations (1)–(4), was evaluated by calculating their adiabatic combustion temperatures (Tad) from the following energy balance equation [17,22] with thermochemical data taken from [23].
Δ 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 ∆Hr is the reaction enthalpy at 298 K, cp and L are the heat capacity and latent heat, nj is the stoichiometric coefficient, and Pj refers to the product component.
The SHS experiments were conducted in a windowed combustion chamber filled with Ar at 0.25 MPa. Reactant powders were well mixed in a tubular ball mill and then uniaxially compressed into cylindrical test specimens with a diameter of 7 mm, a height of 12 mm, and a relative density of 55%. The sample compact was ignited by an electrically heated tungsten coil. An R-type bare-wire thermocouple (Pt/Pt-13%Rh) with a bead size of 125 μm was used to measure the combustion temperature. The propagation velocity of combustion wave (Vf) was determined by the time derivative of the flame-front trajectory constructed from the recorded series of combustion images. Phase compositions of the products were identified by an X-ray diffractometer (XRD, Bruker D2 Phaser, Karlsruhe, Germany). Microstructures and constituent elements of the products were examined by the scanning electron microscopy (SEM, Hitachi, Tokyo, Japan, S3000H) and energy dispersive spectroscopy (EDS). Details of the experimental methods were reported elsewhere [24].

3. Results and Discussion

3.1. Combustion Exothermicity of Reduction-Based SHS Reactions

Figure 1 presents the calculated ∆Hr and Tad of reactions Equations (1)–(4) and shows that Equation (4) has the highest values while Equation (1) has the lowest ones. Both ∆Hr and Tad increase from Equations (1)–(4). Specifically, the values of Tad are 2530 K, 2595 K, 2783 K, 2897 K for Equations (1)–(4), respectively. A comparison between Equations (1) and (2) revealed that the combustion moderator Al2O3 appeared to impose a stronger dilution effect on combustion than MgO, which led to a lower Tad for Equation (1) than Equation (2). Similar results were observed in Equations (3) and (4). These findings could also be explained by the fact that metallothemic reduction of TiO2 or B2O3 by Al is more exothermic than that by Mg [15,25].
On the other hand, because the B2O3-based thermite is more energetic than the one using TiO2 [25], Equation (3) has a higher Tad than Equation (1). Similarly, Equation (4) has a higher Tad than Equation (2). According to the calculated Tad, it is realized that the thermite oxidants (i.e., B2O3 versus TiO2) have a more pronounced influence on combustion exothermicity than the diluent oxides (i.e., Al2O3 versus MgO).

3.2. Combustion Temperature and Self-Propagating Velocity

Two series of the SHS processes recorded from reactions Equations (1) and (3) are illustrated in Figure 2a,b, respectively. It is apparent that upon ignition, the reaction was initiated and characterized by a self-sustaining combustion wave. More intense combustion accompanied with a faster combustion wave was observed in Figure 2b, when compared with that in Figure 2a. Combustion luminosity and flame spreading speed reflected the degree of reaction exothermicity. As mentioned above, B2O3/Al/Mg-based Equation (3) is more energetic than TiO2/Al/Mg-based Equation (1). Similar combustion behavior was also noticed in Equations (2) and (4).
Figure 3 depicts typical combustion temperature profiles measured from four different reactions. All profiles exhibit a steep temperature rise followed by a rapid descent, which is characteristic of the SHS reaction that features a fast combustion wave and a thin reaction zone. The peak value is considered as the combustion front temperature (Tc). When compared with pinnacles in the contours of Equations (1) and (2), sharper peaks were detected in the profiles of Equations (3) and (4). This implied a faster combustion wave in Equations (3) and (4). As shown in Figure 3, the values of Tc from Equation (1)–(4) in ascending order are 1348 °C, 1445 °C, 1660 °C, and 1736 °C. It should be noted that the measured combustion front temperatures are in agreement with the calculated reaction exothermicity.
It is useful to note in Figure 3 that the curves of Equations (3) and (4) have a shape peak with a pronounced shoulder. The shape peak was a result of the fast combustion wave. The pronounced shoulder could be caused by the occurrence of volumetric synthesis reactions after the passage of the rapid combustion wave.
Figure 4 plots the measured combustion wave propagation velocities (Vf) and temperatures (Tc) of four reactions. The rising trend of Vf from Equations (1)–(4) is consistent with that of Tc. This can be understood by the fact that the propagation of combustion wave is essentially governed by layer-by-layer heat transfer from the reaction zone to unreacted region, and therefore, is subject to the combustion front temperature. As presented in Figure 4, the average combustion velocities are 3.9, 4.7, 5.1, and 5.7 mm/s for Equation (1)–(4), respectively. It is worth noting that the measured combustion temperature not only justified the reaction exothermic analysis, but confirmed the temperature dependence of combustion wave velocity.

3.3. Composition and Microstructure Analyses of Synthesized Products

The XRD spectra of the final products synthesized from Equations (1) and (2) are shown in Figure 5a,b, respectively. Both indicated the formation of TiB2 and MgAl2O4 along with two minor phases, magnesium titanate (MgTiO3) and MgO. It is believed that MgAl2O4 was synthesized through a combination reaction between Al2O3 and MgO. Equation (1) and (2) were formulated with the same thermite reagents of TiO2, Al, and Mg, but diluted by different metal oxides. That is, Al2O3 was partly pre-added and partly thermite-produced, while MgO was completely generated from the reduction of TiO2 by Mg in Equation (1). In contrast, the required Al2O3 in Equation (2) was entirely produced from the reduction of TiO2 by Al, but MgO was supplied in part from prior addition and in part from the reduction of TiO2 by Mg. For both Equations (1) and (2), TiB2 was synthesized from the reaction of elemental boron with reduced and metallic Ti.
Traces of MgO suggested an incomplete reaction due probably to the relatively low reaction temperatures of Equations (1) and (2). The presence of MgTiO3 in the final products of Equations (1) and (2) could be attributed to the reaction of MgO with the thermite oxidant TiO2 [26,27]. The formation of MgTiO3 in the SHS-produced TiB2–MgAl2O4 composites was also observed by Radishevskaya et al. [10] using Ti, boron, and MgAl2O4 as their starting materials and a partial decomposition of MgAl2O4 during combustion synthesis was considered as a possible route resulting in the formation of MgTiO3.
The presence of MgO along with no detection of Al2O3 in the final products of Equations (1) and (2) suggested that the as-synthesized MgAl2O4 is an Al2O3-rich spinel. The formation of MgTiO3 could also result in the production of Al2O3-rich spinel. According to Naghizadeh et al. [28], magnesium titanate compounds (MgTiO3 and Mg2TiO4) were identified in the phase evolution of MgAl2O4 produced from TiO2-containing samples and stoichiometric MgAl2O4 spinel shifted toward the Al2O3-rich type. Due to the formation of MgTiO3, the amount of TiB2 formed in the composite should be less than the stoichiometric amount.
Figure 6a,b exhibits the XRD patterns of the synthesized composites from Equations (3) and (4), respectively. In addition to TiB2 and MgAl2O4, a small amount of MgTiO3 was identified. The formation of MgAl2O4 from a combination reaction between Al2O3 and MgO was proved. Both Al2O3 and MgO can be totally or partially produced from the reduction of B2O3 by Al and Mg. For Equations (3) and (4) containing B2O3/Al/Mg-based thermite, TiB2 was produced from the reaction of metallic Ti with reduced and elemental boron. Moreover, the formation of MgTiO3 in Equations (3) and (4) might involve some interaction of Ti with B2O3 to form TiO2 which further reacted with MgO. Unlike that in Equations (1) and (2), MgO was no longer detected in the final products of Equations (3) and (4).
MgTiO3 ceramic has been proved to be an excellent dielectric material, owing to its high dielectric constant, low dielectric loss, high value of quality factors, and good temperature stability [29,30]. It is believed that a trace amount of MgTiO3 as a minor phase existed in the as-synthesized TiB2-MgAl2O4 products has no effect on the refractory properties of the composites. However, removal of MgTiO3 from the TiB2-MgAl2O4 composite would be difficult, since it could combine with MgAl2O4 in a solid solution form [28].
The SEM image shown in Figure 7 illustrates the microstructure of fracture surface of the product synthesized from Equation (1) which contains a TiO2/Al/Mg-based thermite. The morphology displays several large and solidified MgAl2O4 aggregates of 5–15 μm surrounded by fine-grain TiB2 crystals with a particle size of about 1–2 μm. Moreover, EDS analysis of two characteristic regions in the product surface indicates that the atomic ratios of Ti:B = 35.2:64.8 and Mg:Al:O = 13.1:28.1:58.8 match well with the stoichiometries of TiB2 and MgAl2O4, respectively.
For the final product of B2O3/Al/Mg-based Equation (4), the microstructure and elemental ratios of the components are presented in Figure 8. As can be seen, MgAl2O4 was formed as large densified aggregates of around 20 μm and TiB2 crystals were in a short-rod form with a length of 2–4 μm or in a shape of fine grains of 1–2 μm. Based on the EDS analysis, the atomic ratio of the selected area in an aggregate is Mg:Al:O = 15.1:30.6:54.3 that is reasonably close to MgAl2O4. Short-rod crystals have a composition of Ti:B = 34.3:65.7, which certainly is TiB2.
In summary, the addition of MgAl2O4 into TiB2 enhanced the refractory properties, such as the high-temperature oxidation and corrosion resistance and thermal shock resistance [10,31,32]. Like many sintering aids, MgAl2O4 as an additive could improve densification of TiB2 ceramics and reduce sintering temperatures [33,34]. The abnormal grain growth could be efficiently prevented during the sintering process. As a result, it is more likely to obtain a uniform grain distribution.
Moreover, the TiB2-MgAl2O4 composite is a promising high-temperature microwave absorption material with a reflection loss less than –5 dB at 8.2–18.0 GHz in the temperature range of 25 °C to 1100 °C [11]. The composite also exhibits an extremely high tolerance against intense irradiation in harsh environments [35,36]. Therefore, the potential uses of the TiB2-MgAl2O4 composite might include heat-resistant coatings, nozzles and nose cones of supersonic jets, microwave absorption components, diagnostic or detector windows in fusion devices, target materials in the nuclear applications, etc., [11,35,36].

4. Conclusions

In situ formation of 3TiB2–MgAl2O4 composites was conducted by combustion synthesis combined with metallothermic reduction reactions involving Al and Mg as dual reductants. Thermite reagents with different oxidants were considered; one utilized TiO2 and the other B2O3. The reactant mixtures also contained elemental Ti and boron. This study in total completed four SHS reactions, within which a small amount of Al2O3 or MgO was included in the reactive mixture to serve as the combustion moderator and part of the precursors for the formation of MgAl2O4. The overall synthesis reaction was exothermic enough to proceed in the SHS mode. An energy-saving and efficient fabrication route for the formation MgAl2O4-containing composites was demonstrated.
The analysis of combustion exothermicity indicated that the SHS reaction containing B2O3/Al/Mg-based thermites was more energetic than that adopting TiO2 as the oxidant. Prior addition of Al2O3 had a greater cooling effect on combustion than that of MgO. Depending on different thermites and diluents, the measured combustion front temperatures ranged from 1320 to 1720 °C, and combustion wave velocity from 3.9 to 5.7 mm/s. The temperature dependence of combustion wave velocity was justified. The XRD analysis confirmed in situ formation of TiB2 and MgAl2O4. A small amount of MgTiO3 was found as the impurity. It is believed that MgAl2O4 was synthesized through a combination reaction between Al2O3 and MgO, both of which can be totally or partially produced from the metallothermic reduction of B2O3 or TiO2. The microstructure of the synthesized composite exhibited that MgAl2O4 was surrounded by closely packed TiB2 grains. MgAl2O4 was formed as densified aggregates with a size of 5–20 μm. TiB2 crystals were produced in a shape of short rods of 2–4 μm and fine grains of 1–2 μm.

Author Contributions

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

Funding

This research work was funded by the National Science and Technology Council of Taiwan under the grant of NSTC 110-2221-E-035-042-MY2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wilkins, J.M.L. Boron and Refractory Borides; Matkovich, V.I., Ed.; Springer: New York, NY, USA, 1977; p. 633. [Google Scholar]
  2. Balcı, Ö.; Burkhardt, U.; Schmidt, M.; Hennicke, J.; Yağcı, M.B.; Somer, M. Densification, microstructure and properties of TiB2 ceramics fabricated by spark plasma sintering. Mater. Charact. 2018, 145, 435–443. [Google Scholar] [CrossRef]
  3. Ramesh, B.; Showman, E.; Abraar, S.A.M.; Saxena, K.K.; Tharwan, M.Y.; Alsaadi, N.; Al Sofyani, S.; Elsheikh, A.H. Microstructure, mechanical characteristics, and wear performance of spark plasma sintered TiB2–Si3N4 as affected by B4N doping. Materials 2022, 15, 7096. [Google Scholar] [CrossRef] [PubMed]
  4. Munro, R.G. Material properties of titanium diboride. J. Res. Natl. Inst. Stand. Technol. 2000, 105, 709–720. [Google Scholar] [CrossRef] [PubMed]
  5. Vaferi, K.; Nekahi, S.; Vajdi, M.; Moghanlou, F.S.; Shokouhimehr, M.; Motallebzadeh, A.; Sha, J.; Asl, M.S. Heat transfer, thermal stress and failure analyses in a TiB2 gas turbine stator blade. Ceram. Int. 2019, 45, 19331–19339. [Google Scholar] [CrossRef]
  6. Sulima, I.; Hyjek, P.; Podsiadło, M.; Boczkal, S. Effect of zirconium diboride and titanium diboride on the structure and properties of 316L steel-based composites. Materials 2023, 16, 439. [Google Scholar] [CrossRef]
  7. Popov, A.Y.; Sivak, A.A.; Borodianska, H.Y.; Shabalin, I.L. High toughness TiB2–Al2O3 composite ceramics produced by reactive hot pressing with fusible components. Adv. Appl. Ceram. 2015, 114, 178–182. [Google Scholar] [CrossRef]
  8. Kozień, D.; Czekaj, I.; Gancarz, P.; Ziąbka, M.; Wieczorek, W.; Pasiut, K.; Zientara, D.; Pędzich, Z. Ceramic matrix composites obtained by the reactive sintering of boron carbide with intermetallic compounds from the Ti-Si system. Materials 2022, 15, 8657. [Google Scholar] [CrossRef]
  9. Li, Z.; Chen, Q.; Jiao, B.; Wang, Q.; Zhang, H.; Jia, Q.; Zhang, S.; Liu, J. Direct laser powder-bed fusion additive manufacturing of complex-shaped TiB2-B4C composite with ultra-fine eutectic microstructure and outstanding mechanical performances. J. Eur. Ceram. Soc. 2023, 43, 1230–1236. [Google Scholar] [CrossRef]
  10. Radishevskaya, N.; Lepakova, O.; Karakchieva, N.; Nazarova, A.; Afanasiev, N.; Godymchuk, A.; Gusev, A. Self-propagating high temperature synthesis of TiB2–MgAl2O4 composites. Metals 2017, 7, 295. [Google Scholar] [CrossRef] [Green Version]
  11. Yang, J.; Liu, X.; Gong, W.; Wang, T.; Wang, X.; Gong, R. Temperature-insensitive and enhanced microwave absorption of TiB2/Al2O3/MgAl2O4 composites: Design, fabrication, and characterization. J. Alloys Compd. 2022, 894, 162144. [Google Scholar] [CrossRef]
  12. Ganesh, I. A review on magnesium aluminate (MgAl2O4) spinel: Synthesis, processing and applications. Int. Mater. Rev. 2013, 58, 63–112. [Google Scholar] [CrossRef]
  13. Corlett, C.A.; Frontzek, M.D.; Obradovic, N.; Watts, J.L.; Fahrenholtz, W.G. Mechanical activation and cation site disorder in MgAl2O4. Materials 2022, 15, 6422. [Google Scholar] [CrossRef]
  14. Yuan, L.; Tian, C.; Yan, X.; He, X.; Liu, Z.; Wen, T.; Jin, E.; Yu, J. Preparation of porous MgAl2O4 ceramics by a novel pectin gel-casting process. J. Aust. Ceram. Soc. 2021, 57, 1049–1055. [Google Scholar] [CrossRef]
  15. Yeh, C.L.; Chen, M.C.; Shieh, T.H. Formation of silicide/spinel ceramic composites via Al- and Mg-based thermitic combustion synthesis. J. Aust. Ceram. Soc. 2022, 58, 1275–1282. [Google Scholar] [CrossRef]
  16. 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]
  17. Xu, J.; Ma, P.; Zou, B.; Yang, X. Reaction behavior and formation mechanism of ZrB2 and ZrC from the Ni-Zr-B4C system during self-propagating high-temperature synthesis. Materials 2023, 16, 354. [Google Scholar] [CrossRef]
  18. Zakaryan, M.K.; Zurnachyan, A.R.; Amirkhanyan, N.H.; Kirakosyan, H.V.; Antonov, M.; Rodriguez, M.A.; Aydinyan, S.V. Novel pathway for the combustion synthesis and consolidation of boron carbide. Materials 2022, 15, 5042. [Google Scholar] [CrossRef]
  19. Omran, J.G.; Afarani, M.S.; Sharifitabar, M. Fast synthesis of MgAl2O4–W and MgAl2O4–W–W2B composite powders by self-propagating high-temperature synthesis reactions. Ceram. Int. 2018, 44, 6508–6513. [Google Scholar] [CrossRef]
  20. Zaki, Z.I.; Mostafa, N.Y.; Rashad, M.M. High pressure synthesis of magnesium aluminate composites with MoSi2 and Mo5Si3 in a self-sustaining manner. Ceram. Int. 2012, 38, 5231–5237. [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. Liang, Y.H.; Wang, H.Y.; Yang, Y.F.; Zhao, R.Y.; Jiang, Q.C. Effect of Cu content on the reaction behaviors of self-propagating high-temperature synthesis in Cu-Ti-B4C system. J. Alloys Compd. 2008, 462, 113–118. [Google Scholar] [CrossRef]
  23. Binnewies, M.; Milke, E. Thermochemical Data of Elements and Compounds; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2002. [Google Scholar]
  24. Yeh, C.L.; Lin, J.Z. Combustion synthesis of Cr-Al and Cr-Si intermetallics with Al2O3 additions from Cr2O3-Al and Cr2O3-Al-Si reaction systems. Intermetallics 2013, 33, 126–133. [Google Scholar] [CrossRef]
  25. Wang, L.L.; Munir, Z.A.; Maximov, Y.M. Thermite reactions: Their utilization in the synthesis and processing of materials. J. Mater. Sci. 1993, 28, 3693–3708. [Google Scholar] [CrossRef]
  26. Singh, J.; Bahel, S. Synthesis of single phase MgTiO3 and influence of Sn4+ substitution on its structural, dielectric and electrical properties. J. Alloys. Compd. 2020, 816, 152679. [Google Scholar] [CrossRef]
  27. Cheng, L.; Liu, P.; Qu, S.X.; Cheng, L.; Zhang, H. Microwave dielectric properties of Mg2TiO4 ceramics synthesized via high energy ball milling method. J. Alloys. Compd. 2015, 623, 238–242. [Google Scholar] [CrossRef]
  28. Naghizadeh, R.; Rezaie, H.R.; Golestani-Fard, F. Effect of TiO2 on phase evolution and microstructure of MgAl2O4 spinel in different atmospheres. Ceram. Int. 2011, 37, 349–354. [Google Scholar] [CrossRef]
  29. Jo, H.J.; Kim, J.S.; Kim, E.S. Microwave dielectric properties of MgTiO3-based ceramics. Ceram. Int. 2015, 41, S530–S536. [Google Scholar] [CrossRef]
  30. Thatikonda, S.K.; Goswami, D.; Dobbidi, P. Effects of CeO2 nanoparticles and annealing temperature on the microwave dielectric properties of MgTiO3 ceramics. Ceram. Int. 2014, 40, 1125–1131. [Google Scholar] [CrossRef]
  31. Li, Y.; Nan, L.; Ruan, G.; Li, X. Reaction path in the aluminothermic reduction nitridation reaction to synthesize MgAl2O4/TiN composite. Ceram. Int. 2005, 31, 825–829. [Google Scholar] [CrossRef]
  32. Liu, X.; Li, K.; Bao, K.; Chen, J.; Zhang, H.; Zhang, S. In-situ synthesis of magnesium aluminate spinel–Zirconium diboride composite powder in magnesium chloride melt. Ceram. Int. 2022, 48, 11869–11871. [Google Scholar] [CrossRef]
  33. Ma, L.; Yu, J.; Guo, X.; Xie, B.; Gong, H.; Zhang, Y.; Zhaie, Y.; Wu, X. Preparation and sintering of ultrafine TiB2 powders. Ceram. Int. 2018, 44, 4491–4495. [Google Scholar] [CrossRef]
  34. Sarkar, R. Additives for magnesium aluminate spinel: A review. Interceram Refract. Manual 2011, 1, 28–32. [Google Scholar]
  35. Lushchik, A.; Feldbach, E.; Kotomin, E.A.; Kudryavtseva, I.; Kuzovkov, V.N.; Popov, A.I.; Seeman, V.; Shablonin, E. Distinctive features of diffusion-controlled radiation defect recombination in stoichiometric magnesium aluminate spinel single crystals and transparent polycrystalline ceramics. Sci. Rep. 2020, 10, 7810. [Google Scholar] [CrossRef] [PubMed]
  36. Ananchenko, D.V.; Nikiforov, S.V.; Kuzovkov, V.N.; Popov, A.I.; Ramazanova, G.R.; Batalov, R.I.; Bayazitovc, R.M.; Novikov, H.A. Radiation-induced defects in sapphire single crystals irradiated by a pulsed ion beam. Nucl. Instrum. Methods Phys. Res. B 2020, 466, 1–7. [Google Scholar] [CrossRef]
Figure 1. Enthalpies of reaction (∆Hr) and adiabatic combustion temperatures (Tad) of Equation (1)–(4) for the synthesis of 3TiB2–MgAl2O4 composites.
Figure 1. Enthalpies of reaction (∆Hr) and adiabatic combustion temperatures (Tad) of Equation (1)–(4) for the synthesis of 3TiB2–MgAl2O4 composites.
Materials 16 01615 g001
Figure 2. Time sequences of recorded SHS images illustrating self-sustaining combustion wave of (a) Equation (1) and (b) Equation (3).
Figure 2. Time sequences of recorded SHS images illustrating self-sustaining combustion wave of (a) Equation (1) and (b) Equation (3).
Materials 16 01615 g002
Figure 3. Typical combustion temperature profiles measured from Equation (1)–(4) for the synthesis of 3TiB2–MgAl2O4 composites.
Figure 3. Typical combustion temperature profiles measured from Equation (1)–(4) for the synthesis of 3TiB2–MgAl2O4 composites.
Materials 16 01615 g003
Figure 4. Combustion wave velocity (Vf) and combustion front temperature (Tc) measured from Equation (1)–(4) for the synthesis of 3TiB2–MgAl2O4 composites.
Figure 4. Combustion wave velocity (Vf) and combustion front temperature (Tc) measured from Equation (1)–(4) for the synthesis of 3TiB2–MgAl2O4 composites.
Materials 16 01615 g004
Figure 5. XRD patterns of TiB2/MgAl2O4 composites obtained from SHS reactions of (a) Equation (1) and (b) Equation (2).
Figure 5. XRD patterns of TiB2/MgAl2O4 composites obtained from SHS reactions of (a) Equation (1) and (b) Equation (2).
Materials 16 01615 g005
Figure 6. XRD patterns of TiB2/MgAl2O4 composites obtained from SHS reactions of (a) Equation (3) and (b) Equation (4).
Figure 6. XRD patterns of TiB2/MgAl2O4 composites obtained from SHS reactions of (a) Equation (3) and (b) Equation (4).
Materials 16 01615 g006
Figure 7. SEM image and EDS spectra of TiB2/MgAl2O4 composite obtained from the SHS reaction of Equation (1).
Figure 7. SEM image and EDS spectra of TiB2/MgAl2O4 composite obtained from the SHS reaction of Equation (1).
Materials 16 01615 g007
Figure 8. SEM image and EDS spectra of TiB2/MgAl2O4 composite obtained from the SHS reaction of Equation (4).
Figure 8. SEM image and EDS spectra of TiB2/MgAl2O4 composite obtained from the SHS reaction of Equation (4).
Materials 16 01615 g008
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.; Zheng, F.-Y. Formation of TiB2–MgAl2O4 Composites by SHS Metallurgy. Materials 2023, 16, 1615. https://doi.org/10.3390/ma16041615

AMA Style

Yeh C-L, Zheng F-Y. Formation of TiB2–MgAl2O4 Composites by SHS Metallurgy. Materials. 2023; 16(4):1615. https://doi.org/10.3390/ma16041615

Chicago/Turabian Style

Yeh, Chun-Liang, and Fu-You Zheng. 2023. "Formation of TiB2–MgAl2O4 Composites by SHS Metallurgy" Materials 16, no. 4: 1615. https://doi.org/10.3390/ma16041615

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