Next Article in Journal
Effect of Aluminum Ion Irradiation on Chemical and Phase Composition of Surface Layers of Rolled AISI 321 Stainless Steel
Next Article in Special Issue
Discrete One-Stage Mechanochemical Synthesis of Titanium-Nitride in a High-Energy Mill
Previous Article in Journal
Improving the Effectiveness of the Solid-Solution-Strengthening Elements Mo, Re, Ru and W in Single-Crystalline Nickel-Based Superalloys
Previous Article in Special Issue
Energetic Materials Based on W/PTFE/Al: Thermal and Shock-Wave Initiation of Exothermic Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Fe/Si Stoichiometry on Formation of Fe3Si/FeSi-Al2O3 Composites by Aluminothermic Combustion Synthesis

Department of Aerospace and Systems Engineering, Feng Chia University, Taichung 40724, Taiwan
*
Author to whom correspondence should be addressed.
Metals 2021, 11(11), 1709; https://doi.org/10.3390/met11111709
Submission received: 22 September 2021 / Revised: 18 October 2021 / Accepted: 23 October 2021 / Published: 26 October 2021
(This article belongs to the Special Issue Metallothermic Reactions)

Abstract

:
Aluminothermic combustion synthesis was conducted with Fe2O3–Al–Fe–Si reaction systems under Fe/Si stoichiometry from Fe-20 to Fe-50 at. % Si to investigate the formation Fe3Si/FeSi–Al2O3 composites. The solid-state combustion was sufficiently exothermic to sustain the overall reaction in the mode of self-propagating high-temperature synthesis (SHS). Dependence of iron silicide phases formed from SHS on Fe/Si stoichiometry was examined. Experimental evidence indicated that combustion exothermicity and flame-front velocity were affected by the Si percentage. According to the X-ray diffraction (XRD) analysis, Fe3Si–Al2O3 composites were synthesized from the reaction systems with Fe-20 and Fe-25 at.% Si. The increase of Si content led to the formation of both Fe3Si and FeSi in the final products of Fe-33.3 and Fe-40 at.% Si reaction systems, and the content of FeSi increased with Si percentage. Further increase of Si to Fe-50 at.% Si produced the FeSi–Al2O3 composite. Scanning electron microscopy (SEM) images revealed that the fracture surface morphology of the products featured micron-sized and nearly spherical Fe3Si and FeSi particles distributing over the dense and connecting substrate formed by Al2O3.

1. Introduction

Transition metal silicides have been promising for a wide range of applications including microelectronic transistors, high-temperature structural components, thin film coatings, thermoelectrics, spintronic devices, and catalysts [1,2,3,4,5,6]. Of many transition metal silicides, iron silicides exhibit magnetic, semiconducting, metallic, and insulating characteristics, depending on different crystalline structures [7,8]. According to the Fe-Si phase diagram, Fe3Si, FeSi, and β-FeSi2 are stable at room temperature, whereas Fe2Si, Fe5Si3, and α-FeSi2 are metastable [9]. Moreover, Fe3Si exists within a wide range of stoichiometry from 10 to 25 at.% Si, FeSi is stoichiometric in a very narrow composition range and both phases are coexistent in the range between 25 and 50 at.% Si [9,10].
Among various fabrication routes to prepare iron silicides in monolithic and composite forms, the reaction synthesis methods associated with mechanical alloying and combustion process have been of great interest. For example, FeSi and β-FeSi2 were produced by mechanically-activated combustion reaction of the Fe + 2Si powder mixture pretreated by shock-assisted ball milling for a long period of time [11,12]. In addition to mechanical activation, Gras et al. [13] adopted KNO3 of 20 wt.% to chemically promote self-sustaining combustion reaction of the Fe + 2Si mixture for the formation of FeSi and α-FeSi2 composites. Zakeri et al. [14] conducted mechanical alloying of SiO2 and Al powders with stainless steel balls for 45 h to fabricate FeSi–Al2O3 nanocomposite powders from induced reactions. According to Guan et al. [15], Fe3Si–Al2O3 nanocomposites were produced from Fe3O4, Al, and Si reactant powders through 4-h mechanical alloying followed by an annealing process at 900 °C for 1 h. Besides, Li et al. [16] produced iron silicide nanoparticles of various phases by thermal annealing of Fe–FeSi2 samples with a core–shell structure and identified the phase transformation from Fe–FeSi2 to FeSi and Fe3Si under an annealing time of 2 h and temperatures of 600 and 700 °C. Recently, an in situ fabrication approach combining the chemical interaction between Fe and Si with aluminothermic reduction of Fe2O3 and SiO2 has been attempted to produce FeSi–Al2O3 and α-FeSi2–Al2O3 composites in the mode of self-propagating high-temperature synthesis (SHS) [17,18]. With the advantages of energy efficiency, rapid reaction, simplicity, and high-purity products [19], the SHS scheme has been recognized as one of the most effective methods for preparing transition metal silicides in monolithic and composite forms [20,21,22,23].
As an extension of the previous studies [17,18], this work aims to investigate the production of Fe3Si/FeSi–Al2O3 composites by aluminothermite-based combustion synthesis in the SHS mode, with an emphasis on exploring the effect of Fe/Si stoichiometry on the formation of Fe3Si and FeSi. So far, no studies have been reported in the literature on combustion synthesis of Fe3Si which exists in a wide stoichiometric range from 10 to 25 at.% Si. In this work, the reactant mixtures with different Fe/Si stoichiometries were prepared for combustion experiments, reaction exothermicity and combustion wave kinetics of the SHS process were studied, and compositional and microstructural analyses of the final products were performed.

2. Materials and Methods

The starting materials used by this study included Fe2O3 (Alfa Aesar Co., Ward Hill, MA, USA, <45 µm, 99.5%), Al (Showa Chemical, Tokyo, Japan, <45 µm, 99.9%), Fe (Alfa Aesar Co., <45 µm, 99.5%), and Si (Strem Chemicals, Newburyport, MA, USA, <45 µm, 99.5%). The stoichiometric composition of the reactants and products is expressed as Reaction (1) with three variables x, m, and n.
F e 2 O 3 + 2 A l + 2.5 F e + ( 1.5 x ) S i m F e 3 S i + n F e S i + A l 2 O 3
The stoichiometric coefficient x was varied between 0.75 and 3.0 to examine the influence of Fe/Si stoichiometry on the formation of silicide phases. Specifically, Reaction (1) was carried out with x = 0.75, 1.0, 1.5, 2.0, and 3.0, i.e., the test specimens with Fe/Si stoichiometries of Fe-20, 25, 33.3, 40, and 50 at.% Si were formulated. According to the value of x, the coefficients m and n were calculated and summarized in Table 1.
Note that with respect to Fe3Si, Reaction (1) with x = 0.75 represents an off-stoichiometric condition with a Si-lean mixture of Fe-20 at.% Si, under which the silicide phase to be produced is Fe3Si owing to its wide formation stoichiometry. The Fe/Si proportions of Reaction (1) with x = 1.0 and 3.0 match the exact stoichiometries of Fe3Si and FeSi, respectively; thus, they are expected to be the only silicide formed in the corresponding conditions. Based on the Fe-Si phase diagram, the products containing two silicide phases, Fe3Si and FeSi, are considered for Reaction (1) with x = 1.5 and 2.0 in Table 1.
Calculation of the adiabatic combustion temperature (Tad) was performed for Reaction (1) with x = 1.0, 1.5, 2.0, and 3.0 by the following energy balance equation [17,18] and thermochemical data taken from [24,25].
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 ΔHro 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.
Reactant powders were well mixed and compressed into cylindrical samples with 7 mm in diameter, 10 mm in length, and a relative density of 55%. The relative density of the test specimen is related to its 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 F e 2 O 3 ρ F e 2 O 3 + Y A l ρ A l + Y F e ρ F e + Y S i ρ S i
The SHS experiment was conducted in a stainless steel combustion chamber equipped with two quartz viewing windows and filled with high-purity argon (99.99%, Hochun Gas Co., Taichung, Taiwan) at 0.25 MPa. Based upon the time sequence series of combustion images, the flame-front trajectory as a function of time was constructed. The time derivative of the trajectory was determined as the combustion wave velocity (Vf). To facilitate the accurate measurement of instantaneous locations of the combustion front, a beam splitter (Rolyn Optics Co., Covina, CA, USA), with a mirror characteristic of 75% transmission and 25% reflection, was used to optically superimpose a scale onto the image of the test sample. The flame-front propagation velocity was slightly higher in the early stage right after the ignition, and then the linearity of time derivative of the trajectory implied that propagation of the flame front can be treated as a constant-velocity event. The relatively high propagation velocity in the beginning was attributed to the thermal energy supplied by the igniter, and the constant velocity in the later stage represents self-sustained propagation of the flame front.
The combustion temperature was measured by a fine-wire thermocouple with a bead diameter of 125 µm. R-type thermocouples (Omega Engineering Inc., Norwalk, CT, USA) with an alloy combination of Pt/Pt-13%Rh were used. The thermocouple bead was firmly attached on the sample surface at a position ~5 mm below the ignition plane. At this location, self-sustaining combustion was well developed so that measurement of the combustion front temperature (Tc) was justified. Phase constituents of the end products were identified by an X-ray diffractometer (Bruker D2 Phaser, Billerica, MA, USA) with CuKα radiation. Scanning electron microscopy (SEM, Hitachi, S3000H, Tokyo, Japan) examination and energy dispersive spectroscopy (EDS) analysis were performed to study the fracture surface microstructure and elemental composition of the final products. Details of the experimental methods were reported elsewhere [26].

3. Results and Discussion

3.1. Combustion Exothermicity and Combustion Wave Kinetics

Calculated ΔHro and Tad of Reactions (1) with different values of x are presented in Figure 1. The calculation was based on the products with phase compositions specified by the values of m and n listed in Table 1. Results showed that the increase of x from 1.0 to 3.0 led to an increase of ΔHro from −881.9 to −1118.6 kJ for Reaction (1), but a decrease of Tad from 3059 to 2872 K. The increase of ΔHro is caused by the fact that both the aluminothermic reduction of Fe2O3 and formation of iron silicides are heat-releasing reactions. Moreover, the number of moles of iron silicides (i.e., the sum of m and n) formed in the final product increases with increasing x value. Specifically, the reaction of Fe2O3 + 2Al → 2Fe + Al2O3 is extremely exothermic with ΔHro = −852.3 kJ and the formation enthalpies of Fe3Si and FeSi are −79.4 and −73.1 kJ/mol, respectively [24]. On the other hand, the decrease of Tad was because the reaction exothermicity (i.e., ΔHro/Cp) of aluminothermic reduction of Fe2O3 is much higher than that of the formation of Fe3Si and FeSi. Consequently, the combustion temperature of the overall reaction decreased as the molar fraction of iron silicides in the product increased.
As calculated adiabatic temperatures are higher than the criterion proposed by Merzhanov [19], combustion synthesis based on Reaction (1) is thermally satisfactory to be self-sustaining. Besides thermodynamic considerations, the SHS process must overcome the kinetic limitation of the reaction. Kinetic restraints are caused by inadequate 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 the interaction between Fe and Si to generate FeSi and/or Fe3Si [27].
Figure 2 illustrates a representative sequence of combustion images recorded from Reaction (1) with x = 1.5. It is evident that upon ignition a well-defined combustion front forms and propagates along the powder compact in a self-sustaining style. The progression of combustion wave arrived at a nearly constant rate after the ignition energy was faded out and the high exothermicity of combustion caused partial melting of the burned sample.
The measured combustion wave velocity reported in Figure 3a decreases from about 2.7 to 1.6 mm/s when the value of x increases from 0.75 to 3.0. The decline of combustion front speed was mainly ascribed to the dilution effect of Fe and Si additions and to the increase of iron silicides formed in the product. The combustion wave propagation rate is mostly governed by the layer-by-layer heat transfer from the reacting region to unburned zone and is likely affected by the combustion front temperature. Typical temperature profiles of Reaction (1) under different stoichiometries are depicted in Figure 3b. The temperature profile suggested that the sample experienced a steep thermal gradient and a rapid cooling rate, both of which are SHS characteristics. The peak value of the profile was defined as the combustion front temperature (Tc). As shown in Figure 3b, the value of Tc decreased from 1730 °C at x = 0.75 to 1506 °C at x = 3.0, confirming the dilution effect on combustion with Fe and Si additions. It is important to note that the dependence of combustion front temperature on Fe/Si stoichiometry was in a manner consistent with that of combustion wave velocity. The descending trend of Tc with x value is in agreement with that of Tad. However, because burning samples suffered considerable heat losses to surrounding argon by conduction and convection and to the inner wall of the chamber by radiation, the measured Tc was lower than the calculated Tad.

3.2. Composition and Microstructure of Synthesized Products

Figure 4a,b plots XRD patterns of two products respectively produced from Reaction (1) with x = 0.75 and 1.0. As indicated in Figure 4a, the formation of Fe3Si and Al2O3 with almost no other phases was obtained from an off-stoichiometric and Si-lean sample of Fe-20 at.% Si (i.e., x = 0.75). For Reaction (1) with Fe/Si stoichiometry of Fe-25 at.% Si (i.e., x = 1.0), Figure 4b reveals that in addition to two target compounds, Fe3Si and Al2O3, a minor compound, aluminum silicate or mullite, was found. Mullite is a stable solid solution in the Al2O3–SiO2 system and refers to Al4+2zSi2−2zO10−z with z varying between around 0.2 and 0.9 [28]. Formation of mullite was essentially due to dissolution of a small amount of Si into Al2O3 during the SHS process [29,30]. The yield of Fe3Si as the only silicide phase under these two test conditions is justified by the fact that Fe3Si exists in a composition range from 10 to 25 at.% Si. In view of no impurities in the final product, the Si-lean mixture of Fe-20 at.% Si was more favorable for the formation of Fe3Si–Al2O3 composite.
On account of its large negative Gibbs free energy change (ΔGo = −840.8 kJ) [24], the reaction pathways of Reaction (1) were proposed to be initiated by the aluminothermic reaction of Fe2O3 + 2Al → 2Fe + Al2O3. In addition to the kinetic aspect, the reduction of Fe2O3 by Al was also a thermodynamically favored reaction which supplied a large amount of heat (ΔHro = −852.3 kJ) to maintain the synthesis reaction in the SHS mode. After the initiation step, the chemical interaction between Fe and Si proceeded to produce Fe3Si.
For Reaction (1) with x = 1.5 and 2.0 (i.e., Fe-33.3% Si and Fe-40% Si), Figure 5a,b shows the presence of two silicide phases—Fe3Si and FeSi—along with Al2O3 and aluminum silicate. The content of aluminum silicate appeared to be larger in the case of Fe-40% Si. Formation of Fe3Si and FeSi was justified because of coexistence of both phases in the composition range from Fe-25 to Fe-50 at.% Si. Figure 5a,b indicates that the content of FeSi relative to Fe3Si increases with increasing Si percentage. The strongest XRD peaks associated with FeSi and Fe3Si are, respectively, located at 2θ = 45.062° (Inorganic Crystal Structure Database (ICSD) card number: 88-1298) and 45.337° (ICSD card number: 65-0994). It is believed that the addition of Si transformed a part of the Si-lean phase Fe3Si to monosilicide FeSi, thus resulting in a decrease of Fe3Si but an increase of FeSi. As proposed by Guan et al. [15], the reaction of Fe3Si with additional Si generated an intermediate phase Fe5Si3 which further converted into the stable FeSi via the reaction of Fe5Si3 + 2Si → 5FeSi.
Figure 6 presents the XRD spectrum of the product synthesized from Reaction (1) with x = 3.0 (i.e., Fe-50% Si). Apparently, monosilicide FeSi was the only silicide formed in the end product. Besides, Al2O3 and a small amount of aluminum silicate were detected. It should be noted that in spite of a high Si percentage, aluminum silicide is at a much less quantity in Figure 6 when compared with that in Figure 5. This could be explained by the fact that FeSi possesses a very narrow composition range. As a result, few Si particles dissolved in Al2O3.
Figure 7 presents an SEM image illustrating the fracture surface microstructure of the product synthesized from Reaction (1) with x = 0.75. The morphology features lots of granular particles distributing over the dense and connecting substrate. The EDS spectrum (a) indicates that the granular product is composed primarily of Fe and Si and has an atomic ratio of Fe/Si = 73.92/26.08 which is close to Fe:Si = 3:1. On the other hand, the EDS spectrum (b) shows that the dense substrate is made up of Al and O at a ratio of Al/O = 42.05/57.95. Therefore, it is believed that granular particles are Fe3Si and the dense substrate is formed by Al2O3. Figure 7 also exhibits the particle size of Fe3Si ranging approximately from 1 to 5 µm.
The microstructure and elemental compositions of the product from Reaction (1) with x = 1.5 are shown in Figure 8. A similar morphology to that of Figure 7 was observed. The EDS analysis of two silicide grains selected in Figure 8 indicated that the intensity of Si peaks relative to that of Fe signals is stronger in the spectrum (a) than (b) and their respective atomic ratios are Fe/Si = 51.45/48.55 and 77.16/22.84. The former certainly is FeSi and the latter Fe3Si. This verified the formation of both FeSi and Fe3Si in the product of Reaction (1) with x = 1.5.
The SEM image of Figure 9 was associated with the FeSi–Al2O3 composite synthesized from Reaction (1) with x = 3.0. Nearly spherical FeSi particles were identified with a size from submicron to around 2 µm. An atomic ratio of Fe/Si = 48.48/51.52 deduced from EDS spectrum (a) matched closely with the stoichiometry of FeSi. All four constituent elements were specified in EDS spectrum (b), among which the ratio of Al/O = 38.82/61.18 signifying Al2O3 was obtained.

4. Conclusions

Fabrication of Fe3Si/FeSi–Al2O3 composites was conducted by the aluminothermic SHS reactions with Fe2O3–Al–Fe–Si systems under Fe/Si stoichiometry from Fe-20 to Fe-50 at.% Si. Experimental results showed that self-propagating combustion was achieved upon ignition. The increase of Si percentage lowered combustion exothermicity and thus decreased not only the combustion front temperature from 1730 to 1506 °C, but flame-front velocity from 2.7 to 1.6 mm/s. Both Fe-20 and Fe-25 at.% Si reaction systems produced Fe3Si–Al2O3 composites and the off-stoichiometric case with a Si-lean composition of Fe-20 at.% Si yielded almost no impurity, aluminum silicate. Two silicide phases—FeSi and Fe3Si—were present in the final products of reaction systems with Fe-33.3 and Fe-40 at.% Si and the content of FeSi increased with increasing Si percentage. For the reaction with Fe-50 at.% Si, the phase conversion was almost completed, and the product was FeSi–Al2O3 composite. SEM micrographs of the products revealed that Al2O3 formed a dense and connecting substrate, over which granular Fe3Si and FeSi particles were uniformly distributed. The size of iron silicide particles varied in the range from submicron to 5 µm.

Author Contributions

Conceptualization, C.-L.Y.; methodology, C.-L.Y., K.-T.C. and T.-H.S.; validation, C.-L.Y., K.-T.C. and T.-H.S.; formal analysis, C.-L.Y., K.-T.C. and T.-H.S.; 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., K.-T.C. and T.-H.S.; 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 Ministry of Science and Technology of Taiwan under the grant of MOST 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. Murarka, S.P. Silicide thin films and their applications in microelectronics. Intermetallics 1995, 3, 173–186. [Google Scholar] [CrossRef]
  2. Petrovic, J.J.; Vasudevan, A.K. Key developments in high temperature structural silicides. Mater. Sci. Eng. A 1999, 261, 1–5. [Google Scholar] [CrossRef]
  3. Viennois, R.; Tao, X.; Jund, P.; Tedenac, J.C. Stability and thermoelectric properties of transition-metal silicides. J. Electron. Mater. 2011, 40, 597–600. [Google Scholar] [CrossRef] [Green Version]
  4. Chirkin, A.D.; Lavrenko, V.O.; Talash, V.M. High-temperature and electrochemical oxidation of transition metal silicides. Powder Metall. Metal. Ceram. 2009, 48, 330–345. [Google Scholar] [CrossRef]
  5. Chen, X.; Liang, C. Transition metal silicides: Fundamentals, preparation and catalytic applications. Catal. Sci. Technol. 2019, 9, 4785–4820. [Google Scholar] [CrossRef]
  6. Kauss, O.; Obert, S.; Bogomol, I.; Wablat, T.; Siemensmeyer, N.; Naumenko, K.; Krüger, M. Temperature Resistance of Mo3Si: Phase stability, microhardness, and creep properties. Metals 2021, 11, 564. [Google Scholar] [CrossRef]
  7. 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]
  8. 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]
  9. Kubaschewski von Goldbeck, O. Iron-Binary Phase Diagrams; Springer-Verlag: Berlin/Heidelberg, Germany, 1982; pp. 136–139. [Google Scholar]
  10. Zhang, Y.; Ivey, D.G. Fe3Si formation in Fe-Si diffusion couples. J. Mater. Sci. 1998, 33, 3131–3135. [Google Scholar] [CrossRef]
  11. Gras, C.; Gaffet, E.; Bernard, F.; Niepce, C.J. 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]
  12. 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]
  13. 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]
  14. 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]
  15. Guan, J.; Chen, X.; Zhang, L.; Wang, J.; Liang, C. Rapid preparation and magnetic properties of Fe3Si-Al2O3 nanocomposite by mechanical alloying and heat treatment. Phys. Status Solidi A 2016, 213, 1585–1591. [Google Scholar] [CrossRef]
  16. Li, M.; Chen, X.; Guan, J.; Wang, J.; Liang, C. Thermally induced phase transition and magnetic properties of Fe–FeSi2 with core–shell structure. Phys. Status Solidi A 2013, 210, 2710–2715. [Google Scholar] [CrossRef]
  17. Yeh, C.L.; Chen, K.T. Synthesis of FeSi–Al2O3 composites by autowave combustion with metallothermic reduction. Metals 2021, 11, 258. [Google Scholar] [CrossRef]
  18. Yeh, C.L.; Chen, K.T. Fabrication of FeSi/α-FeSi2–based composites by metallothermically assisted combustion synthesis. J. Aust. Ceram. Soc. 2021. [Google Scholar] [CrossRef]
  19. Merzhanov, A.G. Combustion processes that synthesize materials. J. Mater. Process. Technol. 1996, 56, 222–241. [Google Scholar] [CrossRef]
  20. 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]
  21. Yeh, C.L.; Chen, W.H. Combustion synthesis of MoSi2 and MoSi2-Mo5Si3 composites. J. Alloys Compd. 2007, 438, 165–170. [Google Scholar] [CrossRef]
  22. Yeh, C.L.; Peng, J.A. Combustion synthesis of MoSi2-Al2O3 composites from thermite-based reagents. Metals 2016, 6, 235. [Google Scholar] [CrossRef] [Green Version]
  23. Maung, S.T.M.; Chanadee, T.; Niyomwas, S. Two reactant systems for self-propagating high-temperature synthesis. J. Aust. Ceram. Soc. 2019, 55, 873–882. [Google Scholar] [CrossRef]
  24. Binnewies, M.; Milke, E. Thermochemical Data of Elements and Compounds; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2002. [Google Scholar]
  25. 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]
  26. 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]
  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. Schneider, H.; Schreuer, J.; Hildmann, B. Structure and properties of mullite—A review. J. Eur. Ceram. Soc. 2008, 28, 329–344. [Google Scholar] [CrossRef]
  29. Zaki, Z.I.; Mostafa, N.Y.; Ahmed, Y.M.Z. Synthesis of dense mullite/MoSi2 composite for high temperature applications. Int. J. Refract. Met. Hard Mater. 2014, 45, 23–30. [Google Scholar] [CrossRef]
  30. Yeh, C.L.; Chou, C.C. Formation of NbB2/mullite composites by combustion synthesis involving metallothermic reduction of Nb2O5 and B2O3. Ceram. Int. 2015, 41, 14592–14596. [Google Scholar] [CrossRef]
Figure 1. Enthalpy of reaction (ΔHro) and adiabatic combustion temperature (Tad) of Reaction (1) with different stoichiometric coefficients, x.
Figure 1. Enthalpy of reaction (ΔHro) and adiabatic combustion temperature (Tad) of Reaction (1) with different stoichiometric coefficients, x.
Metals 11 01709 g001
Figure 2. A time series of recorded images illustrating self-propagating combustion wave of Reaction (1) with x = 1.5 (unit of scale bar: 1 mm).
Figure 2. A time series of recorded images illustrating self-propagating combustion wave of Reaction (1) with x = 1.5 (unit of scale bar: 1 mm).
Metals 11 01709 g002
Figure 3. Effects of stoichiometric coefficient x on (a) flame-front propagation velocity and (b) combustion temperature.
Figure 3. Effects of stoichiometric coefficient x on (a) flame-front propagation velocity and (b) combustion temperature.
Metals 11 01709 g003
Figure 4. XRD patterns of Fe3Si-Al2O3 composites synthesized from Reaction (1) with (a) x = 0.75 and (b) x = 1.0.
Figure 4. XRD patterns of Fe3Si-Al2O3 composites synthesized from Reaction (1) with (a) x = 0.75 and (b) x = 1.0.
Metals 11 01709 g004
Figure 5. XRD patterns of Fe3Si/FeSi-Al2O3 composites synthesized from Reaction (1) with (a) x = 1.5 and (b) x = 2.0.
Figure 5. XRD patterns of Fe3Si/FeSi-Al2O3 composites synthesized from Reaction (1) with (a) x = 1.5 and (b) x = 2.0.
Metals 11 01709 g005
Figure 6. XRD pattern of FeSi-Al2O3 composite synthesized from Reaction (1) with x = 3.0.
Figure 6. XRD pattern of FeSi-Al2O3 composite synthesized from Reaction (1) with x = 3.0.
Metals 11 01709 g006
Figure 7. SEM image and EDS spectra associated with (a) Fe3Si and (b) Al2O3 of SHS-derived product from Reaction (1) with x = 0.75.
Figure 7. SEM image and EDS spectra associated with (a) Fe3Si and (b) Al2O3 of SHS-derived product from Reaction (1) with x = 0.75.
Metals 11 01709 g007
Figure 8. SEM image and EDS spectra associated with (a) FeSi and (b) Fe3Si of SHS-derived product from Reaction (1) with x = 1.5.
Figure 8. SEM image and EDS spectra associated with (a) FeSi and (b) Fe3Si of SHS-derived product from Reaction (1) with x = 1.5.
Metals 11 01709 g008
Figure 9. SEM image and EDS spectra associated with (a) FeSi and (b) Al2O3 of SHS-derived product from Reaction (1) with x = 3.0.
Figure 9. SEM image and EDS spectra associated with (a) FeSi and (b) Al2O3 of SHS-derived product from Reaction (1) with x = 3.0.
Metals 11 01709 g009
Table 1. Stoichiometric coefficients (x, m, and n) of Reaction (1).
Table 1. Stoichiometric coefficients (x, m, and n) of Reaction (1).
Stoichiometric Coefficients
xmn
0.751.50
1.01.50
1.51.1251.125
2.00.752.25
3.004.5
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.; Shieh, T.-H. Effects of Fe/Si Stoichiometry on Formation of Fe3Si/FeSi-Al2O3 Composites by Aluminothermic Combustion Synthesis. Metals 2021, 11, 1709. https://doi.org/10.3390/met11111709

AMA Style

Yeh C-L, Chen K-T, Shieh T-H. Effects of Fe/Si Stoichiometry on Formation of Fe3Si/FeSi-Al2O3 Composites by Aluminothermic Combustion Synthesis. Metals. 2021; 11(11):1709. https://doi.org/10.3390/met11111709

Chicago/Turabian Style

Yeh, Chun-Liang, Kuan-Ting Chen, and Tzong-Hann Shieh. 2021. "Effects of Fe/Si Stoichiometry on Formation of Fe3Si/FeSi-Al2O3 Composites by Aluminothermic Combustion Synthesis" Metals 11, no. 11: 1709. https://doi.org/10.3390/met11111709

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