Next Article in Journal
The COOL-Process—A Selective Approach for Recycling Lithium Batteries
Next Article in Special Issue
Mo–Nb–Si–B Alloy: Synthesis, Composition, and Structure
Previous Article in Journal
A Comparative Electrochemical and Morphological Investigation on the Behavior of NiCr and CoCr Dental Alloys at Various Temperatures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of FeSi-Al2O3 Composites by Autowave Combustion with Metallothermic Reduction

Department of Aerospace and Systems Engineering, Feng Chia University, Taichung 40724, Taiwan
*
Author to whom correspondence should be addressed.
Metals 2021, 11(2), 258; https://doi.org/10.3390/met11020258
Submission received: 15 January 2021 / Revised: 26 January 2021 / Accepted: 29 January 2021 / Published: 3 February 2021
(This article belongs to the Special Issue Metallothermic Reactions)

Abstract

:
Fabrication of FeSi-Al2O3 composites with a molar ratio of FeSi/Al2O3 ranging from 1.2 to 4.5 was conducted by the self-propagating high-temperature synthesis (SHS) method. The synthesis reaction involved metallothermic reduction of Fe2O3 and SiO2 by Al and the chemical interaction of Fe and Si. Two combustion systems were examined: one contained thermite reagents of 0.6Fe2O3 + 0.6SiO2 + 2Al, and the other had Fe2O3 + 2Al to mix with different amounts of Fe and Si powders. A thermodynamic analysis indicated that metallothermic reduction of oxide precursors was sufficiently exothermic to sustain the combustion reaction in a self-propagating mode. The SHS reaction carrying out co-reduction of Fe2O3 and SiO2 was less exothermic, and was applied to synthesize products with FeSi/Al2O3 = 1.2–2.5, while the reaction reducing only Fe2O3 was more energetic and was adopted for the composites with FeSi/Al2O3 = 2.5–4.5. Moreover, the former had a larger activation energy, i.e., Ea = 215.3 kJ/mol, than the latter, i.e., Ea = 180.4 kJ/mol. For both reaction systems, the combustion wave velocity and temperature decreased with increasing FeSi content. Formation of FeSi-Al2O3 in situ composites with different amounts of FeSi was achieved. Additionally, a trivial amount of aluminum silicate was detected in the products of high FeSi contents due to dissolution of Si into Al2O3 during the SHS process.

1. Introduction

Transition metal silicides have been the focus of many investigations due to their attractive high-temperature properties and a wide range of potential applications. Within the Fe-Si system, iron silicide compounds include Fe3Si, Fe2Si, Fe5Si3, FeSi, β-FeSi2, and α-FeSi2 [1]. According to the phase diagram, Fe3Si, FeSi, and β-FeSi2 are stable at room temperature, while Fe2Si, Fe5Si3, and α-FeSi2 are metastable. Depending on their crystal structures, these iron silicides exhibit magnetic, semiconducting, insulating, and metallic behavior [2,3,4]. For example, Fe3Si has a relatively high value of saturation magnetization and is a promising candidate as a ferromagnetic electrode [2]. FeSi has a high melting point, i.e., 1410 °C, good high-temperature structural stability, and excellent resistance to chemical corrosion, especially in harsh industrial environments [5]. Also, FeSi is a transition-metal Kondo insulator and acts as the host structure for the ferromagnetic semiconductor FexCo1−xSi [6]. Si-rich FeSi2 exists in two allotropic forms: orthorhombic semiconducting β-FeSi2 at low temperatures and tetragonal metallic α-FeSi2 at high temperatures. β-FeSi2 is a potential optoelectronic and thermoelectric material, owing to its direct band gap of 0.87 eV, with a wide range of absorption spectrum and relatively large Seebeck coefficient [7,8,9].
Among various methods for fabricating transition metal silicides, combustion synthesis in the mode of self-propagating high-temperature synthesis (SHS) has the advantages of energy efficiency, short processing time, low cost, and high purity of the product [10,11,12]. The SHS scheme has been extensively applied to synthesize many silicide compounds of the Ti-Si, Nb-Si, Zr-Si, Ta-Si, and Mo-Si systems from elemental powder compacts of their corresponding stoichoimetries [13,14,15,16,17]. However, weak combustion exothermicity restrains the preparation of iron silicides from the direct SHS reaction between Fe and Si powders. Despite the fact that the reactions are exothermic, Fe + Si → FeSi has an enthalpy of formation of ΔHf = −79.4 kJ/mol and an adiabatic temperature of Tad = 1650 K and Fe + 2Si → FeSi2 has ΔHf = −81.2 kJ/mol and Tad = 1310 K [18]. It has been empirically concluded that the autowave synthesis reaction will not be achievable unless the adiabatic combustion temperature is higher than 1800 K [19]. As a consequence, several activation approaches have been employed. By using high-energy ball milling to mechanically activate the self-sustaining combustion of Fe + 2Si powder mixtures, Gras et al. [20,21] obtained FeSi as the main product for the operation conditions of short duration milling (1–4 h) and moderate shock power (2.2 W), while β-FeSi2 was formed in the postcombustion zone for those of long duration milling (6 h) or high shock power (4.8 W). The role of 20 wt% KNO3 as an additive to chemically activate the reactivity of Fe + 2Si mixtures was examined [22]. It was found that decomposition of KNO3 generated enough thermal energy to assist the combustion reaction in a stable and self-sustaining mode, and that the resulting product consisted of FeSi and α-FeSi2 [22]. Moreover, the FeSi-Al2O3 composite was synthesized by the mechanical alloying of SiO2, Al, and Fe powders, and complete reduction of SiO2 by Al was reached after 45 h of milling [23].
As the first attempt, this study aims to fabricate FeSi-Al2O3 composites by the SHS process involving metallothermic reduction of metal oxides. By taking advantage of the highly-exothermic reaction of Fe2O3 with Al, this study employed two thermite reagents, Al-Fe2O3 and Al-Fe2O3-SiO2, blended with Fe and Si powders to serve as the reactant mixtures which were formulated to have a broad range of the starting composition. The effects of thermite mixtures and sample stoichiometries were investigated on the reaction exothermicity, combustion wave velocity and temperature, and phase compositions of the final products. In addition, the activation energy of the synthesis reaction was experimentally deduced from combustion wave kinetics.

2. Materials and Methods

The starting materials of this study included Fe2O3 (Alfa Aesar Co., Ward Hill, MA, USA, <45 μm, 99.5%), SiO2 (Strem Chemicals, Newburyport, MA, USA, 99%), Al (Showa Chemical Co., Tokyo, Japan, <45 μm, 99.9%), Fe (Alfa Aesar Co., <45 μm, 99.5%), and Si (Strem Chemicals, <45 μm, 99.5%). Two thermite mixtures using Al as the reducing agent were prepared: one contains two oxidants at a molar proportion of Fe2O3:SiO2:Al = 0.6:0.6:2 and the other consists only of Fe2O3 and Al at a ratio of Fe2O3:Al = 1:2. Reactions (1) and (2) expressed below represent two combustion systems formulated with different thermite reagents and adjustable amounts of elemental Fe and Si powders for the synthesis of FeSi-Al2O3 composites.
0.6 Fe 2 O 3 + 0.6 SiO 2 + 2 Al + ( x 1.2 ) Fe + ( x 0.6 ) Si x FeSi + Al 2 O 3
Fe 2 O 3 + 2 Al + ( y 2 ) Fe + y Si y FeSi + Al 2 O 3
where stoichiometric coefficients x and y are related to the amounts of Fe and Si powders in the reactant mixture, and also signify the molar ratio of FeSi/Al2O3 formed in the final product. The metallothermic reaction of Fe2O3 + 2Al → 2Fe + Al2O3 is extremely exothermic with Tad = 3622 K, while exothermicity of aluminothermic reduction of SiO2 (Tad = 1760 K) is much weaker [24]. This suggests that Reaction (1) is less energetic than Reaction (2), since co-reduction of Fe2O3 and SiO2 occurs in Reaction (1). Moreover, because FeSi has low formation exothermicity (Tad = 1650 K), the increase of Fe and Si additions imposes a dilution effect on combustion. Accordingly, this study performed Reaction (1) with x varying from 1.2 to 2.5, within which stable and self-sustaining combustion was reached. It is apparent that the minimum value of x for Reaction (1) is 1.2, where no elemental Fe is added. When Reaction (1) was conducted with x > 2.5, it was found that, probably owing to inadequate reaction exothermicity, combustion was barely self-propagating and the phase conversion was incomplete.
On account of the reaction exothermicity, Reaction (2) is more favorable for the formation of higher contents of FeSi in the product. Stable and self-sustaining combustion for Reaction (2) occurs in the range of 2.5 ≤ y ≤ 4.5. For Reaction (2) with y = 2.0 (i.e., no addition of elemental Fe), violent combustion caused eruption of the reactants and massive melting of the sample, bringing about difficulty in recovering the end product. However, Reaction (2) with x > 4.5 appeared to be deficient in thermal energy to complete the phase conversion.
Calculation of Tad was performed for Reactions (1) and (2) under different x and y, based on the following energy balance equation [25,26] with thermochemical data taken from [18,27].
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.
The SHS experiment was conducted in a combustion chamber equipped with quartz viewing windows and filled with high-purity (99.99%) argon at 0.2 MPa. Reactant powders were well mixed in a ball mill, and then the mixture was uniaxially compressed in a stainless-steel mold at a pressure of 60–70 MPa to form cylindrical test specimens with 7 mm in diameter, 12 mm in height, and a relative density of 55%. The propagation velocity of combustion wave (Vf) was determined from the time series of recorded images. From the time derivative of combustion front trajectories, it was found the flame-front propagation velocity was slightly higher in the early stage right after the ignition, and then the velocity decelerated, finally reaching a nearly constant value. The relatively high propagation velocity in the beginning was attributed to the thermal energy supplied by the igniter. The constant velocity in the later stage represents the self-sustaining propagation of flame front and is reported in this study.
The thermocouple used in this study is R-type thermocouple (Omega Engineering Inc., Norwalk, CT, USA) with an alloy combination of Pt/Pt-13%Rh. The wire diameter is 62.5 μm, bead size 125 μm, and wire length 40 mm. The thermocouple bead was firmly attached on the sample surface at a position about 7 mm below the ignition plane. At this location, self-sustaining combustion was well developed, so that the measurement of combustion front temperature was justified. Details of the experimental setup were reported elsewhere [28]. Phase constituents of the final products were analyzed by X-ray diffraction (XRD) patterns from an X-ray diffractometer (Bruker D2 Phaser, Billerica, MA, USA) with CuKα radiation.
The temperature dependence of combustion wave velocity offers a relationship to determine the activation energy (Ea) of the solid-state combustion reaction. The combustion wave velocity derived from an energy equation with a heat source can be expressed as [29,30]
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, ρ the density, R the gas constant, Q the heat of reaction, Tc the combustion front temperature, and ko the Arrhenius rate constant. Thus, Ea can be obtained from the slope of a best-fitted linear line correlating ln(Vf/Tc)2 with 1/Tc [30].

3. Results and Discussion

3.1. Thermodynamic Analysis

The calculated ΔHr and Tad of Reactions (1) and (2) under different stoichiometric coefficients are presented in Figure 1a,b, respectively. With the increase of x from 1.2 to 2.5 for Reaction (1), Figure 1a shows an increase in ΔHr from −667.4 to −760.2 kJ, but a decrease in Tad from 3355 to 2847 K. The increase of total reaction enthalpy is justified, because both the aluminothermic reaction and formation of FeSi release heat. However, due to weak exothermicity of the FeSi formation, the value of Tad decreases with increasing content of FeSi. For Reaction (2) with y increasing from 2.5 to 4.5, Figure 1b shows a rise in ΔHr from −975.8 to −1118.6 kJ and a fall in Tad from 3418 to 2872 K. Similar trends are observed for both reaction systems and Reaction (2) is more energetic than Reaction (1). At the same coefficients of x and y equal to 2.5, it can be seen that Tad of Reaction (2) is higher than that of Reaction (1). Thermodynamic analysis confirms that both reaction systems are thermally satisfactory under the selected conditions. Because the calculated adiabatic temperatures are higher than the criterion proposed by Merzhanov [19], self-sustaining combustion can occur.
Besides thermodynamic considerations, the synthesis of FeSi-Al2O3 composite via autowave combustion with metallothermic reduction has to overcome the kinetic limitation of the reaction. Kinetic restraints are caused by incomplete reactivity owing to the presence of diffusion barriers. It is believed that the reduction of Fe2O3 by Al to produce Fe and Al2O3 acts as the initiation step, followed by aluminothermic reduction of SiO2 in the case of Reaction (1). Subsequently, FeSi is produced through the interaction between Fe and Si.

3.2. Self-Sustaining Combustion Wave Kinetics

Figure 2a,b illustrate two typical combustion sequences observed in this study and are respectively recorded from Reaction (1) of x = 1.5 and Reaction (2) of y = 4.5. Firstly of all, Figure 2 confirms the establishment of self-propagating combustion for Reactions (1) and (2). It is evident that for both SHS processes, a well-defined combustion front forms upon ignition and propagates along the sample compact in a self-sustaining manner. Secondly, Figure 2a,b signify different combustion behavior. As shown in Figure 2a, the flame propagation was fast and it took about 3.33 s for the combustion wave to traverse the sample. Moreover, combustion was in companion with obvious melting of the sample, suggestive of strong exothermicity of this test condition. On the other hand, Figure 2b exhibits a relatively slow combustion wave with a long spreading time of about 6 s and that the melting of the sample is considerably alleviated. Namely, a comparison between Figure 2a,b indicated that the increase of Fe and Si additions changed the combustion behavior from a rapid/molten mode to a moderate style and lowered the flame-front speed, due to a decrease in combustion exothermicity.
Figure 3 reveals the variations of flame-front propagation velocity (Vf) of Reactions (1) and (2) with their corresponding stoichiometric coefficients, x and y. A significant decline in the combustion wave speed from about 3.9 to 1.1 mm/s was observed for Reaction (1) with increasing x from 1.2 to 2.5. The decrease of combustion front velocity was ascribed to the dilution effect of Fe and Si additions, because the formation of FeSi was less energetic than aluminothermic reduction of oxide reagents. Similarly, the combustion velocity of Reaction (2) decreased from 4.5 mm/s at y = 2.5 to 1.5 mm/s at y = 4.5. The combustion propagation rate is mostly governed by the layer-by-layer heat transfer from the reaction zone to unburned region. Namely, the reaction zone temperature has a great influence on combustion wave velocity.
Typical combustion temperature profiles measured from the samples of Reactions (1) and (2) are depicted in Figure 4. The profile features a sharp rise to the highest point, followed almost immediately by a considerable decline owing to heat losses to the surroundings. This implies a thin reaction zone. The sudden increase signifies the rapid arrival of the combustion wave. The peak value represents the combustion front temperature (Tc). A decrease in Tc with stoichiometric coefficient was detected for both reaction systems. Figure 4a indicates that Tc of Reaction (1) decreases from 1710 °C at x = 1.2 to 1413 °C at x = 2.5, confirming the dilution effect on combustion by increasing Fe and Si additions. Likewise, as shown in Figure 4b, the increase of y from 2.5 to 4.5 lowers the Tc of Reaction (2) from 1826 to 1506 °C. Due to the fact that the thermite mixture of Fe2O3 + 2Al is more exothermic than that of 0.6Fe2O3 + 0.6SiO2 + 2Al, the value of Tc of Reaction (2) is higher than that of Reaction (1) under the same stoichiometric coefficient of 2.5. Moreover, the dependence of combustion front temperature on sample stoichiometry is in a manner consistent with that of combustion wave velocity.
The accuracy of fine-wire thermocouple measurement is essentially affected by the conduction cooling and radiation loss. According to Bradley and Matthews [31], the conduction loss from the thermocouple depends on the length of wire between the junction and the support. The dimension of the thermocouple used in this study explains why the measurement was practically unaffected by the conduction cooling [31]. The radiation loss from the thermocouple bead is another major source of error in the thermocouple measurement. An estimate of the radiation correction was performed, assuming that a steady state existed between the convective heat transfer to, and radiation loss from, the thermocouple bead [32,33]. Radiation corrections for the combustion front temperatures measured in this study are in the range from 28 to 36 °C, which is more significant in the regions of higher temperatures. After the radiation correction, it is believed that the accuracy of the combustion temperature measurement is within ±10 °C. When compared with the calculated Tad, however, the measured Tc was much lower. This suggests that the burning samples experienced substantial heat losses to the surrounding gas by conduction and convection and to the chamber wall by radiation.
According to combustion wave kinetics [29,30], as mentioned above, the activation energy of solid-state combustion can be determined from the correlation between ln(Vf/Tc)2 and 1/Tc. Figure 5 plots two sets of experimental data with their respective best-fitted linear lines. For Reaction (1), five data points from left to right were according calculated from Vf and Tc of x = 1.2, 1.5, 1.75, 2.0, and 2.5. Five data points of Reaction (2) from left to right are orderly associated with y = 2.5, 3.0, 3.5, 4.0, and 4.5. From the slopes of two straight lines, the values of Ea = 215.3 and 180.4 kJ/mol were deduced for Reactions (1) and (2), respectively. A smaller Ea for Reaction (2) means a lower kinetic barrier for the synthesis reaction to occur. Because the aluminothermic reduction is considered as the initiation step of the SHS process, co-reduction of Fe2O3 and SiO2 by Al is probably responsible for the large Ea obtained for Reaction (1). When compared with Ea = 145 kJ/mol reported for the Al-Fe2O3 thermite reaction [34], Reactions (1) and (2) of this study have higher kinetic barriers to conquer. This could be most likely caused by the fact that in addition to aluminothermic reduction, Reactions (1) and (2) involve chemical interactions between Fe and Si to form FeSi. These interactions could be complex and occur in both sequential and parallel manners during the SHS process.

3.3. Phase Constituents of Synthesized Products

Figure 6a–c presents the XRD patterns of SHS-produced composites from Reaction (1) with x = 1.2, 1.5, and 2.0, respectively. Formation of FeSi (ICSD card number: 88-1298) and Al2O3 (ICSD card number: 88-0826) with almost no other phases was achieved, suggesting that both the aluminothermic reduction and chemical interaction were complete for Reaction (1). The intensity of the signature peaks of FeSi is obviously stronger in Figure 6c than Figure 6a and Figure 6b. This confirms an increase in the ratio of FeSi/Al2O3 in the end product. The weak peaks between the signals of main phases in Figure 6 might reflect the presence of a tiny amount of impurities. The impurity phase possibly existing in the final products is considered to be aluminum silicate (or called mullite). Mullite is a stable solid solution compound in the Al2O3-SiO2 system. The stoichiometry of mullite refers to Al4+2mSi2−2mO10−m with m varying between about 0.2 and 0.9 and the composition ranges from relatively silica-rich 3Al2O3⋅2SiO2 (3:2 mullite) to alumina-rich 2Al2O3⋅SiO2 (2:1 mullite) [35]. Formation of aluminum silicate was attributable to dissolution of a small amount of Si into Al2O3 during the SHS process. To be shown below in Figure 7, the signature peaks of aluminum silicate are detectable, because the amount of Si is increased in the sample.
The XRD spectra of the final products synthesized from Reaction (2) with y = 2.5, 3.5, and 4.0 are shown in Figure 7a–c. Both FeSi and Al2O3 were identified and the peak intensity of FeSi relative to Al2O3 was augmented in the case of y = 4.0 when compared to those of y = 2.5 and 3.5. Besides, as indicated in Figure 7, aluminum silicate was detected in the final products. As can be seen in Figure 7, the quantity of aluminum silicate slightly increased with increasing FeSi content formed in the product, because the sample contained more Si.

4. Conclusions

The SHS process involving metallothermic reduction of Fe2O3 and SiO2 by Al and the elemental reaction of Fe with Si was conducted to fabricate FeSi-Al2O3 composites with a broad composition range. Thermite reagents made up of 0.6Fe2O3 + 0.6SiO2 + 2Al and Fe2O3 + 2Al were incorporated into the Fe‒Si reaction system. Aluminothermic reduction of metal oxides generated Al2O3 and released a large amount of the reaction enthalpy which was adequate to sustain the combustion reaction in a self-propagating manner. The SHS reaction including the mixture of 0.6Fe2O3 + 0.6SiO2 + 2Al was less exothermic than that containing the reagent of Fe2O3 + 2Al. Moreover, the increase of Fe and Si additions for the production of a higher content of FeSi reduced the overall reaction exothermicity. Consequently, the former was adopted to produce the composites with a molar ratio of FeSi/Al2O3 from 1.2 to 2.5, within which a decrease in both the combustion wave velocity from 3.9 to 1.1 mm/s and reaction front temperature from 1710 to 1413 °C was observed. The latter was applied to fabricate the composites with FeSi/Al2O3 from 2.5 to 4.5; within this range, the increase of FeSi/Al2O3 lowered the reaction temperature from 1826 to 1506 °C and decelerated the combustion wave from 4.5 to 1.5 mm/s. Correlation of combustion wave kinetics indicated that the SHS reaction based on 0.6Fe2O3 + 0.6SiO2 + 2Al had a larger Ea of 215.3 kJ/mol in comparison to Ea = 180.4 kJ/mol for the reaction constructed out of Fe2O3 + 2Al. The difference in Ea could be ascribed to co-reduction of Fe2O3 and SiO2 by Al, in contrast to the sole reduction of Fe2O3. The formation of FeSi and Al2O3 in the final products was identified by XRD analysis. The increase of the FeSi content with increasing Fe and Si additions was confirmed. Phase conversion from the reactants to products was completely achieved, except for a trivial amount of aluminum silicate present in the end products with a high FeSi content.

Author Contributions

Conceptualization, C.-L.Y.; methodology, C.-L.Y. and K.-T.C.; validation, C.-L.Y. and K.-T.C.; formal analysis, C.-L.Y. and K.-T.C.; investigation, C.-L.Y. and K.-T.C.; resources, C.-L.Y.; data curation, C.-L.Y. and K.-T.C.; writing—original draft preparation, C.-L.Y. and K.-T.C.; writing—review and editing, C.-L.Y. and K.-T.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 109-2221-E-035-037.

Data Availability Statement

Data presented in this study are available in the article.

Acknowledgments

The authors thank the Precision Instrument Support Center of Feng Chia University for providing facilities for materials analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Okamoto, H. Desk Handbook: Phase Diagrams for Binary Alloys; ASM International: Materials Park, OH, USA, 2000. [Google Scholar]
  2. Dahal, N.; Chikan, V. Phase-Controlled synthesis of iron silicide (Fe3Si and FeSi2) nanoparticles in solution. Chem. Mater. 2010, 22, 2892–2897. [Google Scholar] [CrossRef]
  3. Liang, Y.F.; Shang, S.L.; Wang, J.; Wang, Y.; Ye, F.; Lin, J.P.; Chen, G.L.; Liu, Z.K. First-Principles calculations of photon and thermodynamic properties of Fe-Si compounds. Intermetallics 2011, 19, 1374–1384. [Google Scholar] [CrossRef]
  4. Wu, J.; Chong, X.Y.; Jiang, Y.H.; Feng, J. Stability, electronic structure, mechanical and thermodynamic properties of Fe-Si binary compounds. J. Alloys Compd. 2017, 693, 859–870. [Google Scholar] [CrossRef]
  5. Shen, B.; He, Y.; Wang, Z.; Yu, L.; Jiang, Y.; Gao, H. Reactive synthesis of porous FeSi intermetallic compound. J. Alloys Compd. 2020, 826, 154227. [Google Scholar] [CrossRef]
  6. Schmitt, A.L.; Bierman, M.J.; Schmeisser, D.; Himpsel, F.L.; Jin, S. Synthesis and properties of single-crystal FeSi nanowires. Nano Lett. 2006, 6, 1617–1621. [Google Scholar] [CrossRef]
  7. Sen, S.; Gogurla, N.; Banerji, P.; Guha, P.K.; Pramanik, P. Synthesis and characterization of β-phase iron silicide nano-particles by chemical reduction. Mater. Sci. Eng. B 2015, 200, 28–39. [Google Scholar] [CrossRef]
  8. Meng, Q.S.; Fan, W.H.; Chen, R.X.; Munir, Z.A. Thermoelectric properties of nanostructured FeSi2 prepared by field-activated and pressure-assisted reactive sintering. J. Alloys Compd. 2010, 492, 303–306. [Google Scholar] [CrossRef]
  9. Qu, X.; Lü, S.; Hu, J.; Meng, Q. Microstructure and thermoelectric properties of β-FeSi2 ceramics fabricated by hot-pressing and spark plasma sintering. J. Alloys Compd. 2011, 509, 10217–10221. [Google Scholar] [CrossRef]
  10. Merzhanov, A.G. Combustion processes that synthesize materials. J. Mater. Process. Technol. 1996, 56, 222–241. [Google Scholar] [CrossRef]
  11. Mossino, P. Some aspects in self-propagating high-temperature synthesis. Ceram. Int. 2004, 30, 311–332. [Google Scholar] [CrossRef]
  12. 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]
  13. Yeh, C.L.; Wang, H.J.; Chen, W.H. A comparative study on combustion synthesis of Ti-Si compounds. J. Alloys Compd. 2008, 450, 200–207. [Google Scholar] [CrossRef]
  14. Bertolino, N.; Anselmi-Tamburini, U.; Maglia, F.; Spinolo, G.; Munir, Z.A. Combustion synthesis of Zr-Si intermetallic compounds. J. Alloys Compd. 1999, 288, 238–248. [Google Scholar] [CrossRef]
  15. Yeh, C.L.; Chen, W.H. A comparative study on combustion synthesis of Nb-Si compounds. J. Alloys Compd. 2006, 425, 216–222. [Google Scholar] [CrossRef]
  16. Yeh, C.L.; Wang, H.J. A comparative study on combustion synthesis of Ta-Si compounds. Intermetallics 2007, 15, 1277–1284. [Google Scholar] [CrossRef]
  17. Yeh, C.L.; Chen, W.H. Combustion synthesis of MoSi2 and MoSi2-Mo5Si3 composites. J. Alloys Compd. 2007, 438, 165–170. [Google Scholar] [CrossRef]
  18. Binnewies, M.; Milke, E. Thermochemical Data of Elements and Compounds; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2002. [Google Scholar]
  19. Merzhanov, A.G. Solid flames: Discoveries, concepts, and horizons of cognition. Combust. Sci. Technol. 1994, 98, 307–336. [Google Scholar] [CrossRef]
  20. Gras, C.; Gaffet, E.; Bernard, F.; Niepce, J.C. Enhancement of self-sustaining reaction by mechanical activation: Case of an Fe-Si system. Mater. Sci. Eng. A 1999, 264, 94–107. [Google Scholar] [CrossRef]
  21. Gras, C.; Bernsten, N.; Bernard, F.; Gaffet, E. The mechanically activated combustion reaction in the Fe–Si system: In situ time-resolved synchrotron investigations. Intermetallics 2002, 10, 271–282. [Google Scholar] [CrossRef]
  22. Gras, C.; Zink, N.; Bernard, F.; Gaffet, E. Assisted self-sustaining combustion reaction in the Fe–Si system: Mechanical and chemical activation. Mater. Sci. Eng. A 2007, 456, 270–277. [Google Scholar] [CrossRef]
  23. Zakeri, M.; Rahimipour, M.R.; Sadrnezhad, S.K. In situ synthesis of FeSi-Al2O3 nanocomposite powder by mechanical alloying. J. Alloys Compd. 2010, 492, 226–230. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. Yeh, C.L.; Ke, C.Y. Intermetallic/ceramic composites synthesized from Al–Ni–Ti combustion with B4C addition. Metals 2020, 10, 873. [Google Scholar] [CrossRef]
  27. Acker, J.; Bohmhammel, K.; van den Berg, G.J.K.; van Miltenburg, J.C.; Kloc, C. Thermodynamic properties of iron silicides FeSi and α-FeSi2. J. Chem. Thermodyn. 1999, 31, 1523–1536. [Google Scholar] [CrossRef]
  28. 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]
  29. 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]
  30. Yeh, C.L.; Ke, C.Y. Combustion synthesis of FeAl-Al2O3 composites with TiB2 and TiC additions via metallothermic reduction of Fe2O3 and TiO2. Trans. Nonferrous Met. Soc. China 2020, 30, 2510–2517. [Google Scholar] [CrossRef]
  31. Bradley, D.; Matthews, K.J. Measurement of high gas temperatures with fine wire thermocouples. J. Mech. Eng. Sci. 1968, 10, 299–305. [Google Scholar] [CrossRef]
  32. Holman, J.P. Experimental Methods for Engineers, 7th ed.; McGraw-Hill: New York, NY, USA, 2001. [Google Scholar]
  33. 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]
  34. Fan, R.H.; Lü, H.L.; Sun, K.N.; Wang, W.X.; Yi, X.B. Kinetics of thermite reaction in Al-Fe2O3 system. Thermochim. Acta 2006, 440, 129–131. [Google Scholar] [CrossRef]
  35. Schneider, H.; Schreuer, J.; Hildmann, B. Structure and properties of mullite—A review. J. Eur. Ceram. Soc. 2008, 28, 329–344. [Google Scholar] [CrossRef]
Figure 1. Enthalpies of reaction (ΔHr) and adiabatic combustion temperatures (Tad) of (a) Reaction (1) and (b) Reaction (2) with different stoichiometric coefficients, x and y.
Figure 1. Enthalpies of reaction (ΔHr) and adiabatic combustion temperatures (Tad) of (a) Reaction (1) and (b) Reaction (2) with different stoichiometric coefficients, x and y.
Metals 11 00258 g001aMetals 11 00258 g001b
Figure 2. Self-propagating combustion sequences associated with formation of FeSi-Al2O3 composites: (a) Reaction (1) with x = 1.5 and (b) Reaction (2) with y = 4.5 (Unit of scale bar: 1 mm).
Figure 2. Self-propagating combustion sequences associated with formation of FeSi-Al2O3 composites: (a) Reaction (1) with x = 1.5 and (b) Reaction (2) with y = 4.5 (Unit of scale bar: 1 mm).
Metals 11 00258 g002aMetals 11 00258 g002b
Figure 3. Effects of stoichiometric coefficients (x and y) on flame-front propagation velocities of Reactions (1) and (2).
Figure 3. Effects of stoichiometric coefficients (x and y) on flame-front propagation velocities of Reactions (1) and (2).
Metals 11 00258 g003
Figure 4. Variations of combustion temperature with stoichiometric coefficients (x and y): (a) Reaction (1) and (b) Reaction (2).
Figure 4. Variations of combustion temperature with stoichiometric coefficients (x and y): (a) Reaction (1) and (b) Reaction (2).
Metals 11 00258 g004
Figure 5. Correlation of Vf and Tc for determination of activation energies (Ea) of Reactions (1) and (2).
Figure 5. Correlation of Vf and Tc for determination of activation energies (Ea) of Reactions (1) and (2).
Metals 11 00258 g005
Figure 6. XRD patterns of FeSi-Al2O3 composites synthesized from Reaction (1) with (a) x = 1.2, (b) x = 1.5, and (c) x = 2.0.
Figure 6. XRD patterns of FeSi-Al2O3 composites synthesized from Reaction (1) with (a) x = 1.2, (b) x = 1.5, and (c) x = 2.0.
Metals 11 00258 g006
Figure 7. XRD patterns of FeSi-Al2O3 composites synthesized from Reaction (2) with (a) y = 2.5, (b) y = 3.5, and (c) y = 4.0.
Figure 7. XRD patterns of FeSi-Al2O3 composites synthesized from Reaction (2) with (a) y = 2.5, (b) y = 3.5, and (c) y = 4.0.
Metals 11 00258 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yeh, C.-L.; Chen, K.-T. Synthesis of FeSi-Al2O3 Composites by Autowave Combustion with Metallothermic Reduction. Metals 2021, 11, 258. https://doi.org/10.3390/met11020258

AMA Style

Yeh C-L, Chen K-T. Synthesis of FeSi-Al2O3 Composites by Autowave Combustion with Metallothermic Reduction. Metals. 2021; 11(2):258. https://doi.org/10.3390/met11020258

Chicago/Turabian Style

Yeh, Chun-Liang, and Kuan-Ting Chen. 2021. "Synthesis of FeSi-Al2O3 Composites by Autowave Combustion with Metallothermic Reduction" Metals 11, no. 2: 258. https://doi.org/10.3390/met11020258

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