Next Article in Journal
Pharmaceutical Cocrystal Development of TAK-020 with Enhanced Oral Absorption
Previous Article in Journal
Determination and Data Correlation of Solubility of Sofosbuvir Polymorphs in Ethyl Acetate + Toluene and Methyl tert-Butyl Ether Binary Solvents at the Temperature Range from 268.15 to 308.15 K
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combustion Synthesis of NbB2–Spinel MgAl2O4 Composites from MgO-Added Thermite-Based Reactants with Excess Boron

Department of Aerospace and Systems Engineering, Feng Chia University, Taichung 40724, Taiwan
*
Author to whom correspondence should be addressed.
Submission received: 24 February 2020 / Revised: 13 March 2020 / Accepted: 17 March 2020 / Published: 18 March 2020
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
The formation of NbB2–MgAl2O4 composites from the MgO-added thermite-based reaction systems was investigated by self-propagating high-temperature synthesis (SHS). Two thermite mixtures, Nb2O5/B2O3/Al and Nb2O5/Al, were, respectively, adopted in Reactions (1) and (2). The XRD analysis confirmed the combination of Al2O3 with MgO to form MgAl2O4 during the SHS process and that excess boron of 30 atom.% was required to yield NbB2–MgAl2O4 composites with negligible NbB and Nb3B4. The microstructure of the composite reveals that rod-shaped MgAl2O4 crystals are closely interlocked and granular NbB2 are embedded in or scattered over MgAl2O4. With the addition of MgAl2O4, the fracture toughness (KIC) of 4.37–4.82 MPa m1/2 was obtained for the composites. The activation energies Ea = 219.5 ± 16 and 167.9 ± 13 kJ/mol for Reactions (1) and (2) were determined from combustion wave kinetics.

1. Introduction

Transition metal (IVB and VB) diborides, such as TiB2, ZrB2, NbB2, and TaB2, have been referred to as ultra-high-temperature ceramics (UHTCs). Transition metal diborides crystallize in the hexagonal A1B2 type structure (space group P6/mmm) with c/a ratio close to unity. In this arrangement, the hexagonal nets of metal atoms and triangle nets of pure boron atoms are alternately stacked along the c axis [1]. Besides their melting points exceeding 3000 °C, they possess a unique combination of high hardness, thermal conductivity, electrical conductivity, excellent chemical stability, corrosion resistance, and thermal shock resistance [2,3,4,5]. They have found a broad range of applications in the mechanical, automobile, aerospace industries, etc. [5]. To improve the refractory properties of metal borides and carbides, Al2O3 or MgAl2O4 (magnesium aluminate spinel) has been considered as an additive. Extensive studies have been conducted on understanding the production and characteristics of different metal borides reinforced by Al2O3 [6,7,8,9,10], but relatively few studies have focused on the boride–MgAl2O4 composite [11].
MgAl2O4 is of particular interest due to its exceptional mechanical, thermal, and optical properties, such as high melting point (2135 °C), high hardness (16 GPa), relatively low density (3.58 g/cm3), high mechanical strength (135–216 MPa), good transmittance in the wavelength range of 0.25 to 5.0 μm, high thermal shock resistance, high chemical inertness, low dielectric constant, and low thermal expansion coefficient [12,13]. MgAl2O4 spinel has a close-packed face-centered cubic (fcc) structure of space group Fd3m (number 227). There are eight MgAl2O4 units per cubic cell. Mg and Al cations occupy 1/8 of the tetrahedral sites and 1/2 of the octahedral sites and there are 32 oxygen ions in the unit cell. In the normal spinel, all Al3+ ions are in octahedral coordination with local symmetry D3d and all Mg2+ ions are in tetrahedral coordination with point group symmetry Td. The anion sublattice is arranged in a pseudo-cubic close-packed (ccp) spatial configuration. There are four layers of AlO6 octahedral chains along the c axis in one unit cell [12]. Fabrication of MgAl2O4 and other spinel compounds like NiFe2O4 and CoFe2O4 by either solid-state or wet-chemistry reaction methods usually requires long processing time and complicated steps [14,15,16,17,18].
With the merits of energy and time savings, simple operation, and high-purity products, combustion synthesis in the mode of self-propagating high-temperature synthesis (SHS) has been recognized as an alternative route to material synthesis and processing [19,20]. When the SHS process is combined with the aluminothermic or magnesiothermic reduction of metal oxides to generate Al2O3 or MgO, the multistage combustion reaction represents an in situ approach to producing MgAl2O4-containing composites [21,22,23,24,25]. Horvitz and Gotman [21] obtained TiAl–Ti3Al–MgAl2O4 composites from the powder compacts composed of 2TiO2–Mg–4Al via the SHS reaction and thermal explosion. Magnesiothermic reduction of WO3 and B2O3 in the presence of Al2O3 under the SHS manner was conducted by Omran et al. [22] to prepare MgAl2O4–W and MgAl2O4–W–W2B composites. With the use of MoO3, SiO2, and Al as the thermite reagents mixing with Mo, Si, and MgO powders for the SHS reaction, the formation of various molybdenum silicides (MoSi2, Mo5Si3, and Mo3Si) and MgAl2O4 composites was achieved [23,24,25].
This study made an attempt to investigate the preparation of NbB2–MgAl2O4 in situ composites via a reduction-based SHS process. The multistage combustion process involves the aluminothermic reduction of Nb2O5 and B2O3 in the presence of MgO, elemental interaction of Nb with B, and the combination of MgO with Al2O3. To resolve the loss of boron in the form of gas-phase B2O2 and BO during the SHS process [8,9], this study examined test specimens with excess boron. The effects of thermite reagents and excess boron on the reaction exothermicity and combustion wave kinetics were explored. Moreover, the phase composition, microstructure, and mechanical properties of the final products were characterized.

2. Materials and Methods

The starting materials utilized by this study included niobium (V) oxide (Nb2O5) (Strem Chemicals, <45 μm, 99.9%, Newburyport, MA, US), B2O3 (Strem Chemicals, 99.9%), Al (Showa Chemical Co., <45 μm, 99.9%, Tokyo, Japan), Nb (Strem Chemicals, <45 μm, 99.8%), amorphous boron (Noah Technologies Corp., <1 μm, 92%, San Antonio, TX, US), and MgO (Alfa Aesar, 99%, Ward Hill, MA, US). Two MgO-added combustion systems with different thermite reagents were studied. As expressed below, Reaction (1) comprises Nb2O5 and B2O3 as thermite oxidants, while Reaction (2) has Nb2O5 only. For both reaction systems, Al acts as the thermite reductant.
[ x 2 Nb 2 O 5 + ( 1 5 6 x ) B 2 O 3 + 2 Al ] + ( 11 3 x 2 ) B + MgO x NbB 2 + MgAl 2 O 4
( 3 5 Nb 2 O 5 + 2 Al ) + ( y 6 5 ) Nb + 2 y B + MgO y NbB 2 + MgAl 2 O 4
where x and y are stoichiometric coefficients signifying the mole number of NbB2 formed per unit mole of MgAl2O4 in Reactions (1) and (2), respectively.
The experiments of Reaction (1) were performed with x = 0.6–1.0. The increase of x raises the amount of Nb2O5 but reduces that of B2O3 in the reactant mixture. Under this condition, more amorphous boron is added to make up for the decrease of boron provided from B2O3. Samples of Reaction (2) were formulated with y = 1.2–1.8. The increase of y augments elemental Nb and B for the production of a larger amount of NbB2, but has no change in the molar quantity of thermite reagents, Nb2O5 and Al, in Reaction (2). A combination of Reactions (1) and (2) renders this study feasible to obtain products with a molar proportion of NbB2/MgAl2O4 from 0.6 to 1.8. In addition, the SHS reactions with test specimens containing excess boron of 20 and 30 atom.% were conducted to examine the extent of boron loss during combustion and to compensate for the relatively low purity (92%) of amorphous boron used in this study. It should be noted that boron has high hardness, great stability in extreme environments, and good resistance to heat. It has several forms, the most common of which is amorphous boron, and is unreactive to oxygen, water, acids, and alkalis. While barely reactive at room temperature, boron reacts strongly at high temperature with metals to form borides. Moreover, its reducing properties allow it to react with numerous compounds and, as in the case of oxygenated or halogenated compounds, in a violent manner [26].
In Reaction (1), the aluminothermic reduction of Nb2O5 (the reaction enthalpy, ΔHr = −536 kJ/mol of Al2O3 and the adiabatic temperature, Tad = 2756 K) is more energetic than that of B2O3Hr = −403.8 kJ/mol of Al2O3 and Tad = 2315 K) [27,28]. The intermetallic reaction of Nb + 2B to yield NbB2 with formation enthalpy ΔHf = −175.3 kJ/mol and Tad = 2315 K is exothermically comparable to the B2O3 + 2Al reaction [29]. The combination reaction between MgO and Al2O3 to form MgAl2O4 is weakly exothermic with ΔHr = −35.6 kJ [23]. This means that with the increase of NbB2 content (the x value), two opposing effects govern the combustion exothermicity of Reaction (1). As far as Reaction (2) is concerned, the increase of y for the formation of a larger NbB2 content has a dilution effect on combustion. To elucidate the combustion exothermicity, calculation of Tad of Reactions (1) and (2) under different stoichiometric coefficients was performed according to the following equation [23,30] with thermochemical data taken from [29].
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 ΔHr is the reaction enthalpy at 298 K, nj is the stoichiometric constant, Cp and L are the heat capacity and latent heat, and Pj refers to the product. Calculations of the adiabatic temperature were based upon the final products of stoichiometric reactions described in R(1) and R(2).
Reactant powders were mixed in a ball mill and then uniaxially compressed in a stainless-steel mold at a pressure of 70–80 MPa to form cylindrical test samples with 7 mm in diameter, 12 mm in height, and a relative density of 60%. The relative density of the test specimen is related to the initial components. The theoretical density (ρTD) of the test specimen is calculated from the mass fraction (Y) and density (ρ) of each component through the following equation.
1 ρ T D = Y N b 2 O 5 ρ N b 2 O 5 + Y B 2 O 3 ρ B 2 O 3 + Y A l ρ A l + Y B ρ B + Y M g O ρ M g O
The SHS experiments were conducted in a windowed combustion chamber filled with high-purity (99.99%) argon. The propagation velocity of the combustion wave (Vf) was determined from the time sequence of recorded pictures. The reaction temperature was measured by the Pt/Pt-13%Rh bare wire thermocouple with a bead diameter of 125 μm. A thin ceramic (SiO2) coating is usually to prevent the catalytic effect on the thermocouple in the measurement of gas-phase flame temperature of a combustion mixture involving hydrogen, methane, or propane as the fuel. Because solid-state combustion with argon as the surrounding gas is under investigation, the thermocouple without an inert coating was used by this work. Details of the experimental setup and scheme were described elsewhere [25,31]. After the SHS process, phase constituents of the products were analyzed by an X-ray diffractometer (Bruker D2, Billerica, MA, US) using CuKα radiation. The microstructure of the final product was examined by a scanning electron microscope (Hitachi S3000H, Tokyo, Japan) and elemental proportion was deduced from the energy dispersive spectroscopy (EDS). For Vickers hardness (Hv) and fracture toughness (KIC) measurement, selected experiments with the reactant compact placing in a steel mold were conducted. Upon the completion of the self-sustaining combustion reaction, densification of the product was carried out by a hydraulic press machine [25].

3. Results and Discussion

3.1. Combustion Wave Kinetics and Reaction Temperature

Figure 1 illustrates a typical series of combustion images obtained by this study, which was recorded from the powder compact of Reaction (2) with y = 1.4. As can be seen in Figure 1, a well-defined combustion wave is established upon ignition and propagates throughout the entire sample in a self-sustaining manner. This demonstrates sufficient reaction exothermicity of the reactant mixture. After combustion, the burned sample essentially retained its original shape.
The influence of the stoichiometric coefficient and excess boron on the flame-front velocity of Reactions (1) and (2) is presented in Figure 2. It was found that for the samples without excess boron, the combustion wave velocity of Reaction (1) increased from 2.5 to 6.6 mm/s with x increasing from 0.6 to 1.0, while that of Reaction (2) decreased from 8.4 to 4.8 mm/s with y in the range from 1.2 to 1.8. To be presented lately, it is believed that the variation of combustion front velocity with the reaction stoichiometry depends mainly on the exothermicity of the SHS process. As also indicated in Figure 2, samples with excess boron of 30 atom.% exhibit lower combustion wave speeds when compared to those without additional boron. This could be attributable to a prolonged sequence of phase evolution of borides in response to the increase of boron.
Figure 3 plots several measured sample temperature profiles, which depict a sharp rise signifying the rapid arrival of the combustion wave and a peak value corresponding to the combustion front temperature (Tc). After the progression of the combustion wave, a substantial temperature decline is a consequence of heat loss to the surroundings. As revealed in Figure 3, the peak temperature (Tc = 1516 °C) of Reaction (1) with x = 0.9 is higher than that of x = 0.6 (Tc = 1317 °C) and the combustion front temperature of Reaction (2) decreases from 1606 °C to 1416 °C with an increase of y from 1.2 to 1.8. Based on the experimental measurement, the stoichiometric dependence of the combustion wave temperature is in agreement with that of flame-front velocity.
A comparison between the calculated adiabatic temperature and measured combustion front temperature of Reactions (1) and (2) is presented in Figure 4, indicative of a consistent variation of Tad and Tc with the reaction stoichiometry. The increase of the combustion temperature of Reaction (1) with increasing NbB2/MgAl2O3 ratio is ascribed to a larger proportion of Nb2O5 to B2O3 in the thermite mixture, since the aluminothermic reduction of Nb2O5 is more exothermic. However, a decline of the combustion temperature of Reaction (2) with NbB2/MgAl2O3 molar ratio confirms the cooling effect on combustion by increasing Nb and B, because the elemental reaction of Nb with B is less energetic than the thermite reaction of Nb2O5 and Al. As shown in Figure 4, the values of Tad are higher by about 350–400 °C than those of Tc. The discrepancy between Tc and Tad might result from considerable heat loss mostly by radiation, substantial boron elimination from the reaction zone, and formation of boride phases different from the stoichiometric composition.
The activation energy (Ea) of solid-state combustion was deduced from combustion wave kinetics by constructing a correlation between ln(Vf/Tc)2 and 1/Tc in a form of linear relationship [32,33]. Figure 5 depicts two sets of experimental data with best-fitted straight lines. From the slopes of straight lines, Ea = 219.5 ± 16 and 167.9 ± 13 kJ/mol were deduced for Reactions (1) and (2), respectively. A larger Ea for Reaction (1) means a higher kinetic barrier in comparison to Reaction (2). This could be caused most likely by the fact that the co-reduction of Nb2O5 and B2O3 by Al is required in Reaction (1) for the synthesis sequence to proceed, but Reaction (2) has only Nb2O5 to be reduced. According to Arrhenius kinetics, the activation energy of the solid-state reaction is governed by the reaction mechanism. In this study, aluminothermic reduction of metal oxides is considered as a first step of the SHS process, which is followed by a combination of Al2O3 and MgO to form MgAl2O4 and elemental interactions between Nb and B to produce NbB2.

3.2. Phase Composition and Microstructure of As-Synthesized Products

Figure 6a–c displays the XRD patterns of SHS-derived products from Reaction (1) of x = 1.0 without and with excess boron. As shown in Figure 6, MgAl2O4 is identified and niobium borides exist in three phases including NbB2, Nb3B4, and NbB. The formation of MgAl2O4 confirms a combination reaction between pre-added MgO and thermite-produced Al2O3. The presence of NbB and Nb3B4 denotes that the amount of boron is inadequate to transform all the borides into NbB2. For the sample without extra boron, as shown in Figure 6a, NbB is the dominant boride phase. Besides, there is NbO2 detected in the final product, indicative of an incomplete reduction of Nb2O5. For the sample with excess boron of 20 atom.%, Figure 6b indicates that the yield of Nb3B4 and NbB2 is enhanced and NbO2 is no longer detectable. Furthermore, Figure 6c reveals that more NbB2 is formed in the final product obtained from the sample with extra boron of 30 atom.%. Although NbB2 is the dominant boride in Figure 6c, the other two borides, NbB and Nb3B4, are not trivial. This means a substantial boron loss from the samples of Reaction (1) possibly through two different paths. One is the reduction of B2O3 and the other is the borothermal reaction between Nb2O5 and boron. Both reactions could generate gaseous B2O2 and BO, and they might expel from the sample compact.
The effect of excess boron on the formation of borides for Reaction (2) is presented in Figure 7a–c. Likewise, excess boron enhanced the production of NbB2. It was found that the improvement is more effective for Reaction (2) than Reaction (1). Figure 7b unveils that NbB2 prevails over NbB and Nb3B4 for the sample with excess boron of 20 atom.%. Moreover, as shown in Figure 7c, NbB and Nb3B4 become negligible in the resulting product from the sample with excess boron of 30 atom.%. This is because Reaction (2) contains no B2O3, only borothermal reduction of Nb2O5 could result in the loss of boron.
Typical microstructures of the fracture surface of the NbB2–MgAl2O4 composites synthesized from Reactions (1) and (2) with 30 atom.% extra boron are illustrated in Figure 8a,b, respectively. As displayed in the micrographs of Figure 8a,b; the long rod-shaped MgAl2O4 crystals form a dense matrix with an interlocking structure and small NbB2 grains are embedded in MgAl2O4 or distributed over the surface. Furthermore, MgAl2O4 and NbB2 phases in Figure 8a are confirmed by the atomic ratios of Mg:Al:O = 14.1:28.9:57.0 and Nb:B = 32.7:67.3 deduced from the EDS spectrum. Similarly, in Figure 8b, MgAl2O4 and NbB2 are determined to have Mg:Al:O = 13.7:29.8:56.5 and Nb:B = 33.1:66.9, both of which match well with their exact stoichiometries.
Densified products have a relative density of about 92%–95%. For the NbB2–MgAl2O4 composite synthesized from Reaction (1) of x = 1.0, Hv = 15.2 GPa and KIC = 4.82 MPa m1/2 were determined. The composite obtained from Reaction (2) of y = 1.8 exhibits Hv = 16.9 GPa and KIC = 4.37 MPa m1/2. When compared with pure NbB2 (KIC = 3.76 MPa m1/2) [34], the NbB2–MgAl2O4 composite is toughened. Similar materials like TiB2–Al2O3 composites fabricated by hot pressing at a sintering temperature of 1700 °C showed a relative density of about 96.2%, Vickers hardness of 24.8 GPa, and fracture toughness of 4.56 MPa m1/2 [35]. This confirms that satisfactory toughness is obtained for the NbB2–MgAl2O4 composite synthesized by this study.

4. Conclusions

This study prepared NbB2–MgAl2O4 in situ composites with a molar ratio of NbB2/MgAl2O4 from 0.6 to 1.8 by the SHS process with reducing stages. Within the scope of experimental variables, combustion front velocity and temperature increased with NbB2 content for Reaction (1), because the proportion of Nb2O5 to B2O3 in the thermite mixture increased. On the other hand, Reaction (2) showed a decrease in combustion velocity and temperature as NbB2 content increased, because of the dilution effect of additional Nb and B on combustion. The activation energies, Ea = 219.5 ± 16 and 167.9 ± 13 kJ/mol, were, respectively, deduced for Reactions (1) and (2), suggesting a higher kinetic barrier for Reaction (1).
The XRD analysis of the final product confirms the formation of MgAl2O4 from thermite-produced Al2O3 and pre-added MgO. For the samples without extra boron, three boride phases, NbB, Nb3B4, and NbB2, were formed and dominated by NbB. Excess boron up to 30 atom.% effectively compensated for the loss of boron during combustion and promoted the formation of NbB2 as the major boride compound. The microstructure of as-synthesized NbB2–MgAl2O4 composites is characterized by interlocking rod-shaped MgAl2O4 crystals and small NbB2 grains embedded in MgAl2O4 or distributed over the surface. Fracture toughness of the NbB2–MgAl2O4 composite was improved up to KIC = 4.37–4.82 MPa m1/2.

Author Contributions

Conceptualization, C.-L.Y. and Y.-C.C.; methodology, C.-L.Y. and Y.-C.C.; validation, C.-L.Y. and Y.-C.C.; formal analysis, C.-L.Y. and Y.-C.C.; investigation, C.-L.Y. and Y.-C.C.; resources, C.-L.Y.; data curation, 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.; 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 Ministry of Science and Technology of Taiwan under the grant of MOST 108-2221-E-035-026.

Acknowledgments

Authors are thankful for Precision Instrument Support Center of Feng Chia University in providing materials analytical facilities.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Okamoto, N.L.; Kusakari, M.; Tanaka, K.; Inui, H.; Otani, S. Anisotropic elastic constants and thermal expansivities in monocrystal CrB2, TiB2, and ZrB2. Acta Mater. 2010, 58, 76–84. [Google Scholar] [CrossRef] [Green Version]
  2. Kurbatkina, V.V.; Patsera, E.I.; Levashov, E.A.; Timofeev, A.N. Self-propagating high-temperature synthesis of refractory boride ceramics (Zr, Ta) B2 with superior properties. J. Eur. Ceram. Soc. 2018, 38, 1118–1127. [Google Scholar] [CrossRef]
  3. Alsawat, M.; Altalhi, T.; Alotaibi, N.F.; Zaki, Z.I. Titanium carbide—Titanium boride composites by self propagating high temperature synthesis approach: Influence of zirconia additives on the mechanical properties. Results Phys. 2019, 13, 102292. [Google Scholar] [CrossRef]
  4. Demirskyi, D.; Vasylkiv, O. Mechanical properties of SiC–NbB2 eutectic composites by in situ spark plasma sintering. Ceram. Int. 2016, 42, 19372–19385. [Google Scholar] [CrossRef]
  5. Wuchina, E.; Opila, E.; Opeka, M.; Fahrenholtz, W.; Talmy, I. UHTCs: Ultra-high temperature ceramic materials for extreme environment applications. Electrochem. Soc. Interface 2007, 16, 30–36. [Google Scholar]
  6. Yan, C.; Liu, R.; Zhang, C.; Cao, Y.; Long, X. A solid-state precursor route to MB2–Al2O3 composite powders. Powder Technol. 2016, 301, 596–600. [Google Scholar] [CrossRef]
  7. Mohammad Sharifi, E.; Karimzadeh, F.; Enayati, M.H. Synthesis of titanium diboride reinforced alumina matrix nanocomposite by mechanochemical reaction of Al–TiO2–B2O3. J. Alloy. Compd. 2010, 502, 508–512. [Google Scholar] [CrossRef]
  8. Xiao, G.Q.; Fu, Y.L.; Zhang, Z.W.; Hou, A.D. Mechanism and microstructural evolution of combustion synthesis of ZrB2–Al2O3 composite powders. Ceram. Int. 2015, 41, 5790–5797. [Google Scholar] [CrossRef]
  9. Yeh, C.L.; Chong, M.H. Effects of B4C and BN additions on formation of NbB2–Al2O3 composites from reduction-based combustion synthesis. Ceram. Int. 2017, 43, 7560–7564. [Google Scholar] [CrossRef]
  10. Yeh, C.L.; Huang, Y.S. Effects of excess boron on combustion synthesis of alumina-tantalum boride composites. Ceram. Int. 2014, 40, 2593–2598. [Google Scholar] [CrossRef]
  11. 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]
  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. Ma, Y.; Liu, X. Kinetics and thermodynamics of Mg-Al disorder in MgAl2O4-spinel: A review. Molecules 2019, 24, 1704. [Google Scholar] [CrossRef] [Green Version]
  14. Mouyane, M.; Jaber, B.; Bendjemil, B.; Bernard, J.; Houivet, D.; Noudem, J.G. Sintering behavior of magnesium aluminate spinel MgAl2O4 synthesized by different methods. Int. J. Appl. Ceram. Technol. 2019, 16, 1138–1149. [Google Scholar] [CrossRef]
  15. Padmaraj, O.; Venkateswarlu, M.; Satyanarayana, N. Structural, electrical and dielectric properties of spinel type MgAl2O4 nanocrystalline ceramic particles synthesized by the gel-combustion method. Ceram. Int. 2015, 41, 3178–3185. [Google Scholar] [CrossRef]
  16. Zegadi, A.; Kolli, M.; Hamidouche, M.; Fantozzi, G. Transparent MgAl2O4 spinel fabricated by spark plasma sintering from commercial powders. Ceram. Int. 2018, 44, 18828–18835. [Google Scholar] [CrossRef]
  17. Wen, Y.; Liu, X.; Chen, X.; Jia, Q.; Yu, R.; Ma, T. Effect of heat treatment conditions on the growth of MgAl2O4 nanoparticles obtained by sol-gel method. Ceram. Int. 2017, 43, 15246–15253. [Google Scholar] [CrossRef]
  18. Lin, J.; He, Y.; Du, X.; Lin, Q.; Yang, H.; Shen, H. Structural and magnetic studies of Cr3+ substituted nickel ferrite nanomaterials prepared by sol-gel auto-combustion. Crystals 2018, 8, 384. [Google Scholar] [CrossRef] [Green Version]
  19. Merzhanov, A.G. Combustion processes that synthesize materials. J. Mater. Process. Technol. 1996, 56, 222–241. [Google Scholar] [CrossRef]
  20. 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]
  21. Horvitz, D.; Gotman, I. Pressure-assisted SHS synthesis of MgAl2O4–TiAl in situ composites with interpenetrating networks. Acta Mater. 2002, 50, 1961–1971. [Google Scholar] [CrossRef]
  22. 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]
  23. 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]
  24. Yeh, C.L.; Chen, Y.C. Fabrication of MoSi2–MgAl2O4 in situ composites by combustion synthesis involving intermetallic and aluminothermic reactions. Vacuum 2019, 167, 207–213. [Google Scholar] [CrossRef]
  25. Yeh, C.L.; Chen, Y.C. Formation of Mo5Si3/Mo3Si–MgAl2O4 composites via self-propagating high-temperature synthesis. Molecules 2020, 25, 83. [Google Scholar] [CrossRef] [Green Version]
  26. Cueilleron, J.; Thevenot, F. Chemical properties of boron. In Boron and Refractory Borides; Matkovich, V.I., Ed.; Springer: Berlin/Heidelberg, Germany, 1977; pp. 203–213. [Google Scholar]
  27. 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]
  28. Yeh, C.L.; Ke, C.Y. Combustion synthesis of FeAl-based composites from thermitic and intermetallic reactions. Crystals 2019, 9, 127. [Google Scholar] [CrossRef] [Green Version]
  29. Binnewies, M.; Milke, E. Thermochemical Data of Elements and Compounds; Wiley-VCH Verlag GmbH: Weinheim, NY, USA, 2002. [Google Scholar]
  30. 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. Alloy. Compd. 2008, 462, 113–118. [Google Scholar] [CrossRef]
  31. Yeh, C.L.; Chen, Y.L. An experimental study on self-propagating high-temperature synthesis in the Ta-B4C system. J. Alloy. Compd. 2009, 478, 163–167. [Google Scholar] [CrossRef]
  32. 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]
  33. Yeh, C.L.; Chen, W.H. A comparative study on combustion synthesis of Nb-B compounds. J. Alloy. Compd. 2006, 422, 78–85. [Google Scholar] [CrossRef]
  34. Akin, I.; Ocak, B.C.; Sahin, F.; Goller, G. Effects of SiC and SiC-GNP additions on the mechanical properties and oxidation behavior of NbB2. J. Asian Ceram. Soc. 2019, 7, 170–182. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, G.; Yan, D.; Zhang, J. Microstructure and mechanical properties of TiB2–Al2O3 composites. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2011, 26, 696–699. [Google Scholar] [CrossRef]
Figure 1. Typical sequence of self-sustaining combustion images recorded from Reaction (2) with y = 1.4.
Figure 1. Typical sequence of self-sustaining combustion images recorded from Reaction (2) with y = 1.4.
Crystals 10 00210 g001
Figure 2. Effects of NbB2/MgAl2O4 ratio and excess boron on flame-front propagation velocity of Reactions (1) and (2).
Figure 2. Effects of NbB2/MgAl2O4 ratio and excess boron on flame-front propagation velocity of Reactions (1) and (2).
Crystals 10 00210 g002
Figure 3. Combustion temperature profiles measured from samples of Reactions (1) and (2) with different stoichiometric coefficients.
Figure 3. Combustion temperature profiles measured from samples of Reactions (1) and (2) with different stoichiometric coefficients.
Crystals 10 00210 g003
Figure 4. Calculated adiabatic temperatures and measured combustion temperatures of Reactions (1) and (2) as a function of the NbB2/MgAl2O4 molar ratio.
Figure 4. Calculated adiabatic temperatures and measured combustion temperatures of Reactions (1) and (2) as a function of the NbB2/MgAl2O4 molar ratio.
Crystals 10 00210 g004
Figure 5. Correlation between combustion wave velocity and temperature for determination of activation energies (Ea) of Reactions (1) and (2).
Figure 5. Correlation between combustion wave velocity and temperature for determination of activation energies (Ea) of Reactions (1) and (2).
Crystals 10 00210 g005
Figure 6. XRD patterns of NbB2–MgAl2O4 composites synthesized from Reaction (1) of x = 1.0 under conditions: (a) without extra boron, (b) with extra boron 20%, and (c) with extra boron 30%.
Figure 6. XRD patterns of NbB2–MgAl2O4 composites synthesized from Reaction (1) of x = 1.0 under conditions: (a) without extra boron, (b) with extra boron 20%, and (c) with extra boron 30%.
Crystals 10 00210 g006
Figure 7. XRD patterns of NbB2–MgAl2O4 composites synthesized from Reaction (2) of y = 1.6 under conditions: (a) without extra boron, (b) with extra boron 20%, and (c) with extra boron 30%.
Figure 7. XRD patterns of NbB2–MgAl2O4 composites synthesized from Reaction (2) of y = 1.6 under conditions: (a) without extra boron, (b) with extra boron 20%, and (c) with extra boron 30%.
Crystals 10 00210 g007
Figure 8. SEM micrographs and EDS spectra of NbB2–MgAl2O4 composites of (a) Reaction (1) of x = 1.0 with 30 atom.% excess boron and (b) Reaction (2) of y = 1.6 with 30 atom.% excess boron.
Figure 8. SEM micrographs and EDS spectra of NbB2–MgAl2O4 composites of (a) Reaction (1) of x = 1.0 with 30 atom.% excess boron and (b) Reaction (2) of y = 1.6 with 30 atom.% excess boron.
Crystals 10 00210 g008

Share and Cite

MDPI and ACS Style

Yeh, C.-L.; Chen, Y.-C. Combustion Synthesis of NbB2–Spinel MgAl2O4 Composites from MgO-Added Thermite-Based Reactants with Excess Boron. Crystals 2020, 10, 210. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10030210

AMA Style

Yeh C-L, Chen Y-C. Combustion Synthesis of NbB2–Spinel MgAl2O4 Composites from MgO-Added Thermite-Based Reactants with Excess Boron. Crystals. 2020; 10(3):210. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10030210

Chicago/Turabian Style

Yeh, Chun-Liang, and Yin-Chien Chen. 2020. "Combustion Synthesis of NbB2–Spinel MgAl2O4 Composites from MgO-Added Thermite-Based Reactants with Excess Boron" Crystals 10, no. 3: 210. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10030210

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