Next Article in Journal
Good Choice of Electrode Material as the Key to Creating Electrochemical Sensors—Characteristics of Carbon Materials and Transparent Conductive Oxides (TCO)
Previous Article in Journal
Bending and Buckling of Circular Sandwich Plates with a Hardened Core
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Pot Synthesis of Alumina-Titanium Diboride Composite Powder at Low Temperature

1
College of Civil Engineering and Architecture, Quzhou University, Quzhou 324000, China
2
College of Engineering, Mathematics and Physical Sciences, University of Exeter, Exeter EX4 4QF, UK
3
The State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan 430081, China
4
Henan Key Laboratory of High Temperature Functional Ceramics, Zhengzhou University, Zhengzhou 450052, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(16), 4742; https://doi.org/10.3390/ma14164742
Submission received: 16 June 2021 / Revised: 9 August 2021 / Accepted: 14 August 2021 / Published: 22 August 2021

Abstract

:
Alumina-titanium diboride (Al2O3-TiB2) composite powders were synthesised via aluminothermic reduction of TiO2 and B2O3, mediated by a molten chloride salt (NaCl, KCl, or MgCl2). The effects of salt type, initial batch composition, and firing temperature/time on the phase formation and overall reaction extent were examined. Based on the results and equilibrium thermodynamic calculations, the mechanisms underpinning the reaction/synthesis processes were clarified. Given their evaporation losses at test temperatures, appropriately excessive amounts of Al and B2O3 are needed to complete the synthesis reaction. Following this, phase-pure Al2O3-TiB2 composite powders composed of 0.3–0.6 μm Al2O3 and 30–60 nm TiB2 particles were successfully fabricated in NaCl after 5 h at 1050 °C. By increasing the firing temperature to 1150 °C, the time required to complete the synthesis reaction could be reduced to 4 h, although the sizes of Al2O3 and TiB2 particles in the resultant phase pure composite powder increased slightly to 1–2 μm and 100–200 nm, respectively.

1. Introduction

Alumina (Al2O3) is a representative high performance ceramic material possessing numerous superior properties. Apart from its high melting point, it exhibits high hardness and mechanical strength, excellent wear and chemical resistances, as well as good electrical/thermal insulation [1]. These, along with its ready availability, make Al2O3 an attractive ceramic material applicable to a wide range of structural and functional applications, in, e.g., wear-resistant components, high-speed cutting tools, armors, refractory linings/crucibles, electrical and chemical insulators, and artificial hip joints [2,3,4]. Unfortunately, it suffers from low fracture toughness and poor thermal shock resistance [5,6], negatively affecting its service performance and life. One of the commonly adopted strategies for overcoming this is to make a composite [7], by combining it with another reinforcement phase, e.g., particles/platelets, or fibrous phases such as carbon fibres and carbon nanotubes. Titanium diboride (TiB2) stood out as one of the reinforcement phases, owing to its good structural and thermodynamic compatibility with Al2O3 [8,9] and its excellent properties such as high melting point, high hardness/elastic modulus, relatively low density, and good thermal/electrical conductivities [10]. Previous studies found that incorporation of TiB2 particles into Al2O3 conferred much improved hardness, strength, fracture toughness and electrical conductivity on the resultant Al2O3-TiB2 composite materials [11,12,13,14,15,16], making them suitable for a variety of demanding applications, e.g., in electrodes, cutting tools, wear parts, lightweight armors, high-temperature/glow-plug heaters, and heat exchangers [14,17,18].
To fabricate high performance Al2O3-TiB2 composites, it is key to prepare and use high quality counterpart powders. Conventionally, such composite powders were prepared by directly mixing pre-synthesised TiB2 and Al2O3 powders in a ball mill [11,12,14,15,16,17,19,20]. Unfortunately, the final bulk composites exhibited unsatisfactory service performance arising mainly from poor dispersion and distribution of TiB2 in Al2O3. To overcome this, in-situ formation of the two constituent phases was suggested [21]. For example, a self-propagating high-temperature synthesis (SHS) method was attempted to form in-situ Al2O3 and TiB2 from powder mixtures of Al, TiO2/Ti, and B2O3/B/H3BO3 [18,22,23,24,25,26,27,28,29,30,31,32]. However, this method suffered from several weaknesses; in particular, difficulties to control the synthesis process and synthesise phase pure materials. Apart from SHS, mechanochemical synthesis [8,9,33,34,35], and milling assisted sol-gel process [36] were investigated. With these two techniques, a long processing time was required, and the resultant powders were often heavily agglomerated, and contaminated during prolonged milling.
In recent years, our group has successfully synthesised several ultra-high temperature boride powders via magnesiothermic reduction of oxide precursors in a molten salt [37,38,39]. In this work, such a molten salt synthesis (MSS) technique was further developed to synthesise high-quality Al2O3-TiB2 composite powders at relatively low temperatures. The effects of processing factors such as salt type, initial batch composition, and firing temperature/time on the synthesis process were examined, and the synthesis conditions optimised. Based on the results, and thermodynamic calculations, the mechanisms underpinning the reaction/synthesis process were proposed.

2. Materials and Methods

The TiO2 (≥99%, 100–300 nm), B2O3 (99.98%) and Al (99.5%, ~44 μm) powders were used as the main starting materials, along with KCl (>99%), NaCl (>99%) and anhydrous MgCl2 (>98%). All of these materials except for Al (Alfa Aesar, Lancashire, UK) were from Sigma-Aldrich (Gillingham, UK). The three main starting materials were mixed in stoichiometric (indicated by the overall Reaction (1)) or nonstoichiometric proportions (with excess Al and B2O3), in a mortar and pestle, and further combined with 5 times (by weight) of each of the salts. The final powder mix was contained in an alumina crucible and heated in an argon protected tube furnace to a temperature between 850 and 1150 °C and held for 4–6 h. After cooling to room temperature, the reacted powder coexisting with the residual salt was subjected to repeated hot water washing. The remaining salt-free product powder was dried overnight at 80 °C before being subjected to detailed characterisations.
3TiO2 + 3B2O3 + 10Al = 3TiB2 + 5Al2O3
Phase formation and reaction extent in samples were evaluated based on their X-ray diffraction (XRD) spectra recorded by an X-ray diffractometer (D8 Advance, Bruker, Germany) operated at 40 mA/40 kV (using CuKα radiation), and at a scan rate of 2° (2θ) min−1, and the ICDD cards used are: TiB2 (35-741), Al2O3 (corundum) (10-173), Ti2O3 (43-1033), Al3Ti (37-1449), Al18B4O33 (32-3), and MgAl2O4 (21-1152). Samples were also gold- or carbon-coated for further microstructural examinations using a scanning electron microscope (SEM, Nova 600, FEI, Hillsboro, OR, USA) and a transmission electron microscope (TEM, JEM 2100, JEOL, Tokyo, Japan), linked with energy-dispersive X-ray spectroscopy (EDS, Oxford Instrument, Oxford, UK).
To assist in clarification of the mechanisms underpinning the whole reaction/synthesis process, thermodynamic calculations were carried out using the commercial FactSage package [40]. The Gibbs energy values for each of the intermediate reactions and the overall reaction (Reaction (1)), as a function of temperature, were calculated.

3. Results and Preliminary Discussion

3.1. Influence of Salt Type on Al2O3-TiB2 Formation

Figure 1 shows XRD patterns of stoichiometric samples after 4 h firing in different salts at 850 °C. After firing in KCl, α-Al2O3 and TiB2 were identified as the primary phases in the sample (Figure 1a). Some intermediate Al18B4O33 and Al3Ti were also detected, along with a trace of Ti2O3, indicating the low extents of Reaction (1) and Al2O3-TiB2 formation. Replacing KCl with NaCl (Figure 1b) led to an evident increase in the relative peak heights of both Al2O3 and TiB2, and a decrease in those of Al3Ti, while those of intermediate Al18B4O33 and Ti2O3 changed little, demonstrating the enhanced extents of Al2O3-TiB2 formation, and the better effect of NaCl than KCl on accelerating the overall reaction. As discussed previously [37,38,39], several processing parameters—in particular, the solubility and mobility of reaction species in the molten salt—play crucial roles in MSS. Unfortunately, these key processing parameters are not available, so it is difficult to figure out the reasons behind the better accelerating effect of NaCl in the present case. When NaCl was replaced with MgCl2, the peak heights of Al2O3 and TiB2 further increased (Figure 1c), and no Al18B4O33, Al3Ti or Ti2O3 was detected, suggesting much enhanced extents of Reaction (1). Nevertheless, an undesirable by-product phase, magnesium aluminate spinel (MgAl2O4), was formed in this case, due to likely the reaction between the formed Al2O3 and the MgCl2 (Reaction (2) [41]. In the three cases, except for α-Al2O3, no other transition aluminas were detected at such a low firing temperature. The reason for this was not clear, but it could be due to the locally enhanced salt bath temperature arising from the exothermic reactions, as suggested by the thermodynamic calculations presented later (see Section 4 below). Given the better accelerating effect of NaCl than KCl, and the formation of undesired by-product MgAl2O4 in the case of using MgCl2, NaCl was chosen specifically to form a liquid reaction medium for the following MSS investigations.
4Al2O3 + 3MgCl2 = 3MgAl2O4 + 2AlCl3

3.2. Influence of Firing Temperature on Al2O3-TiB2 Formation

Figure 2 presents XRD patterns of stoichiometric samples after 4 h firing in NaCl at different temperatures. At 850 °C, as already described above, Al2O3 and TiB2 were formed as the primary phases, but some intermediate Al18B4O33, Al3Ti and Ti2O3 were detected (Figure 2a/Figure 1b). Upon increasing the temperature to 950 (Figure 2b) and 1050 °C (Figure 2c), the relative peak heights of Al2O3 and TiB2 increased significantly. Meanwhile, Al3Ti disappeared, and Ti2O3 changed little, but the peak heights of Al18B4O33 increased. These results implied that increasing the firing temperature favoured not only the Al2O3-TiB2 formation but also the Al18B4O33 formation. Further increasing the temperature to 1150 °C led to a further increase in the peak heights of Al18B4O33 and Ti2O3 but decrease in those of Al2O3 and TiB2 (Figure 2d), indicating the adverse effect of such further temperature increase on the Al2O3-TiB2 formation. This, according to our previous studies on MSS of transition metal borides [37,38,39], was due to probably evaporation loss of Al at relatively high reaction temperatures [42]. To confirm this, the influence of using excess Al on the Al2O3-TiB2 formation was further investigated, as described and discussed next.

3.3. Influence of Excess Al on Al2O3-TiB2 Formation

Figure 3 illustrates the influence of excess Al on phase evolution in samples after 4 h firing in NaCl at 1150 °C. The use of 20 wt% more Al resulted in considerable increase in the peak heights of Al2O3 and TiB2 and decrease in those of Al18B4O33 and Ti2O3 (Figure 3a,b). By further increasing excess Al to 25 wt%, the peak heights of Al2O3 and TiB2 were further increased. Meanwhile, Ti2O3 was completely eliminated, and only minor Al18B4O33 remained (Figure 3c), illustrating positive effects from the Al compensation. Nevertheless, the formation of phase-pure Al2O3-TiB2 was still not achieved.
To eliminate Al18B4O33, the excess Al was further increased to 30 wt%. Unfortunately, some Al3Ti was formed instead (Figure 3d), implying that there was no sufficient supply of B in the system, due to probably evaporation loss of B2O3 at such a relatively high temperature [43,44].

3.4. Influence of Excess B2O3 on Al2O3-TiB2 Formation

To confirm and address the issue with B2O3 evaporation loss, the influence of using excess B2O3 (along with 30 wt% excess Al) on the Al2O3-B2O3 formation in the sample fired at 1150 °C for 4 h was further investigated as an example. When 10 wt% excess B2O3 was used, the peak heights of both Al2O3 and TiB2 increased, whereas those of Al3Ti decreased considerably (Figure 4b), compared with in the case of stoichiometric sample (Figure 4a). Further increasing the excess B2O3 to 20 wt% led to complete elimination of Al3Ti, and the formation of phase pure Al2O3-TiB2 (Figure 4c).

3.5. Influence of Firing Time on Al2O3-TiB2 Formation and Further Optimisation of Synthesis Condition

To reveal the influence of firing time on the Al2O3-TiB2 formation and further optimise the synthesis condition, samples containing 30 wt% excess Al and 20 wt% excess B2O3 were fired in NaCl at 1050 °C for different time periods, and then analysed by XRD (Figure 5). As shown in Figure 5a,b, with increasing the time at 1050 °C from 4 to 5 h, Al18B4O33 disappeared, whereas the peak heights of Al3Ti increased. Further extending the time to 6 h resulted in little change in the XRD pattern (Figure 5c), and Al3Ti was still detected. This, as mentioned above (Section 3.4), indicated that 20 wt% excess B2O3 might not be sufficient to compensate for the evaporation loss of B2O3 under this condition, in other words, more excess B2O3 had to be used. As verified by Figure 6, increasing excess B2O3 from 20 wt% to 25 wt% led to significant decrease in the peak heights of Al3Ti, and further increasing the excess amount to 30 wt%, all the intermediate phases including Al3Ti disappeared, and phase pure Al2O3-TiB2 powder was finally obtained.

3.6. Microstructure of Al2O3-TiB2 Product Powder

As demonstrated in Figure 4c and Figure 6c and discussed above, phase pure Al2O3-TiB2 powder could be formed in NaCl after 5 h firing at 1050 °C (using 30 wt% excess Al and 30 wt% excess B2O3), or 4 h firing at 1150 °C (using 30 wt% excess Al and 20 wt% excess B2O3), which was further confirmed by microstructural characterisation. Figure 7, as an example, gives TEM images of the product powder prepared under the former condition, along with the corresponding EDS, showing the coexistence of submicron-sized Al2O3 particles (0.3–0.6 μm) and nanosized TiB2 particles (30–60 nm). Except these two phases, no other impurity phases were seen, as already revealed by XRD (Figure 6c). Figure 8 further displays SEM images of the product powder formed under the latter condition mentioned above. The formation of phase pure Al2O3-TiB2 in this case was also confirmed by EDS, along with XRD (Figure 4c). However, due to the higher synthesis temperature, Al2O3 and TiB2 particles in the composite powder became slightly bigger (1–2 μm Al2O3 and 100–200 nm TiB2).

4. Further Discussion and Reaction/Synthesis Mechanism

Based on the results presented in Figure 1, Figure 2, Figure 3, Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8 and the preliminary discussion above, as well as on the thermodynamic calculations (Figure 9), the reaction/synthesis mechanisms (taking the case of using NaCl as an example) could be proposed; they are discussed as follows.
The firing temperatures (850–1150 °C) were above the melting points of NaCl, B2O3 and Al. So, at these temperatures, they all melted, forming a liquid NaCl pool, and B2O3 and Al liquid droplets. Then, the latter two would start to react with each other in the molten NaCl medium, forming B and Al2O3, according to Reaction (3).
On the other hand, as discussed in our previous work on MSS of TiB2 [39], TiO2 also would be reduced by the Al droplets in the molten NaCl medium, forming Ti and Al2O3 (Reaction (4)). The Al2O3 formed from Reactions (3) and (4) would partially react with the unreduced B2O3, in the NaCl medium, forming the intermediate Al18B4O33 (Reaction (5)), as detected by XRD (Figure 1, Figure 2 and Figure 3 and Figure 5).
B2O3 + 2Al = 2B + Al2O3
3TiO2 + 4Al = 3Ti + 2Al2O3
9Al2O3 + 2B2O3 = Al18B4O33
With the optimisation of synthesis condition, the intermediate Al18B4O33 could be further reduced by Al in the molten NaCl, forming additional B and Al2O3 according to Reaction (6), as verified by XRD (Figure 3 and Figure 5).
Apart from Al18B4O33, another main intermediate phase, Al3Ti, was formed from the reaction between the Ti formed from Reaction (4), and Al (Reaction (7)), as detected by XRD (Figure 1, Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6).
As discussed, and confirmed in our previous studies [37,38,39], the formed B and Ti (Reactions (3), (4) and (6)) were slightly dissoluble in the molten chloride salts. So, the dissolved B and Ti would diffuse through the molten salt medium and reacted with each other, forming the other desirable phase of TiB2 (Reaction (8)).
Al18B4O33 + 4Al = 4B + 11Al2O3
Ti + 3Al = Al3Ti
Ti + 2B = TiB2
B dissolved in the molten salt would also diffuse through the salt medium and react respectively with the Al3Ti produced from Reaction (7), and the unreacted raw materialTiO2, forming additional TiB2 (Reactions (9) and (10)). Reaction (9) was mainly responsible for the complete elimination of this intermediate phase under the optimal conditions (Figure 4 and Figure 6).
Al3Ti + 2B = TiB2 + 3Al
10B + 3TiO2 = 3TiB2 + 2B2O3
As shown in Figure 9, the Gibbs energy values corresponding to Reactions (3)–(9) and the overall Reaction (1) are also negative, at test temperatures (and even at lower temperatures), suggesting that all of these exothermic Reactions are thermodynamically favourable, which further supports the plausibility of the reaction/synthesis mechanisms proposed above.

5. Conclusions

Low temperature synthesis of Al2O3-TiB2 composite powders via aluminothermic reduction of TiO2 and B2O3 in a molten chloride salt (KCl, NaCl or MgCl2) was investigated. Processing parameters including salt type, initial batch composition, and firing temperature/time played important roles in the synthesis process. Phase pure Al2O3-TiB2 nanocomposite powder could be prepared in NaCl after 5 h firing at 1050 °C, when 30 wt% excess Al and 30 wt% excess B2O3 were used. The particle sizes of Al2O3 and TiB2 coexisting in the as-prepared composite powder were estimated to be 0.3–0.6 μm and 30–60 nm, respectively. By increasing the temperature to 1150 °C and using 30 wt% excess Al and 20 wt% excess B2O3, phase pure Al2O3-TiB2 composite powder also could be prepared in NaCl after only 4 h, although sizes of Al2O3 and TiB2 particles coexisting in the resultant composite power increased slightly to 1–2 μm and 100–200 nm, respectively.

Author Contributions

Experiment and initial draft, K.B. and X.L.; conception, experiment design and thermodynamic calculation, K.B., X.L., J.C. and Q.J.; review and editing, X.L., Q.J. and S.Z.; supervision, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 51902180), and the EPSRC Centre for Doctoral Training in Electromagnetic Metamaterials (EP/L015331/1) at the University of Exeter.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Munro, M. Evaluated Material Properties for a Sintered alpha-Alumina. J. Am. Ceram. Soc. 2005, 80, 1919–1928. [Google Scholar] [CrossRef]
  2. Basu, B.; Balani, K. (Eds.) Bonding, Structure, and Physical Properties. In Advanced Structural Ceramics; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2011; pp. 28–29. [Google Scholar]
  3. Williams, D. Contemporary and Future Biomaterials; Cambridge University Press (CUP): Cambridge, UK, 2018; pp. 475–589. [Google Scholar]
  4. Yazdani, B.; Xia, Y.; Ahmad, I.; Zhu, Y. Graphene and carbon nanotube (GNT)-reinforced alumina nanocomposites. J. Eur. Ceram. Soc. 2015, 35, 179–186. [Google Scholar] [CrossRef]
  5. Niihara, K. New Design Concept of Structural Ceramics. J. Ceram. Soc. Jpn. 1991, 99, 974–982. [Google Scholar] [CrossRef] [Green Version]
  6. Rainforth, W.M. The wear behaviour of oxide ceramics-A Review. J. Mater. Sci. 2004, 39, 6705–6721. [Google Scholar] [CrossRef]
  7. Chen, H.; Rainforth, W.M.; Lee, W.E. The wear behaviour of Al2O3-SiC ceramic nanocomposites. Scr. Mater. 2000, 42, 555–560. [Google Scholar] [CrossRef]
  8. Sharifi, E.M.; Karimzadeh, F.; Enayati, M.H. Synthesis of titanium diboride reinforced alumina matrix nanocomposite by mechanochemical reaction of Al-TiO2-B2O3. J. Alloys Compd. 2010, 502, 508–512. [Google Scholar] [CrossRef]
  9. Sharifi, E.M.; Karimzadeh, F.; Enayati, M.H. Preparation of Al2O3-TiB2 nanocomposite powder by mechanochemical reaction between Al, B2O3 and Ti. Adv. Powder Technol. 2011, 22, 526–531. [Google Scholar] [CrossRef]
  10. Golla, B.R.; Bhandari, T.; Mukhopadhyay, A.; Basu, B. Titanium Diboride. In Ultra-High Temperature Ceramics: Materials for Extreme Environment Applications; Fahrenholtz, W.G., Wuchina, E.J., Lee, W.E., Zhou, Y., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014; pp. 316–360. [Google Scholar]
  11. Kimura, I.; Hotta, N.; Hiraoka, Y.; Saito, N.; Yokota, Y. Sintering and characterization of Al2O3-TiB2 composites. J. Eur. Ceram. Soc. 1989, 5, 23–27. [Google Scholar] [CrossRef]
  12. Liu, J.; Ownby, P.D. Enhanced mechanical properties of alumina by dispersed titanium diboride particulate inclusions. J. Am. Ceram. Soc. 1991, 74, 241–243. [Google Scholar] [CrossRef]
  13. Ranganath, S.; Subrahmanyam, J. Impact response of Al2O3 and Al2O3-TiB2 ceramic composites. J. Mater. Sci. Lett. 1991, 10, 1297–1298. [Google Scholar] [CrossRef]
  14. Bellosi, A.; De Portu, G.; Guicciardi, S. Preparation and properties of electroconductive Al2O3-based composites. J. Eur. Ceram. Soc. 1992, 10, 307–315. [Google Scholar] [CrossRef]
  15. Jianxin, D.; Xing, A.; Zhaoqian, L. Friction and wear behavior of Al2O3/TiB2 composite against cemented carbide in various atmospheres at elevated temperature. Wear 1996, 195, 128–132. [Google Scholar] [CrossRef]
  16. Jianxin, D.; Tongkun, C.; Lili, L. Self-lubricating behaviors of Al2O3/TiB2 ceramic tools in dry high-speed machining of hardened steel. J. Eur. Ceram. Soc. 2005, 25, 1073–1079. [Google Scholar] [CrossRef]
  17. Tampieri, A.; Bellosi, A. Oxidation of monolithic TiB2 and of Al2O3-TiB2 composite. J. Mater. Sci. 1993, 28, 649–653. [Google Scholar] [CrossRef]
  18. Plovnick, R.H.; Richards, E. New combustion synthesis route to TiB2-Al2O3. Mater. Res. Bull. 2001, 36, 1487–1493. [Google Scholar] [CrossRef]
  19. Stadlbauer, W.; Kladnig, W.; Gritzner, G. Al2O3-TiB2 composite ceramics. J. Mater. Sci. Lett. 1989, 8, 1217–1220. [Google Scholar] [CrossRef]
  20. Liu, G.; Yan, D.; Zhang, J. Microstructure and mechanical properties of TiB2-Al2O3 composites. J. Wuhan Univ. Technol. Sci. Ed. 2011, 26, 696–699. [Google Scholar] [CrossRef]
  21. Li, J.; Gao, L.; Guo, J. Mechanical properties and electrical conductivity of TiN-Al2O3 nanocomposites. J. Eur. Ceram. Soc. 2003, 23, 69–74. [Google Scholar] [CrossRef]
  22. Logan, K.V. Shaped Refractory Products and Method of Making Same. United States Patent No. 4, 891, 337, 2 January 1990. [Google Scholar]
  23. Ray, S.P. Boride-alumina composites: Synthesis and Fabrication. Metall. Trans. A 1992, 23, 2381–2385. [Google Scholar] [CrossRef]
  24. Kecskes, L.J.; Niiler, A.; Kottke, T.; Logan, K.V.; Villalobos, G.R. Dynamic consolidation of combustion-synthesized alumina-titanium diboride composite ceramics. J. Am. Ceram. Soc. 2005, 79, 2687–2695. [Google Scholar] [CrossRef]
  25. Ma, Z.Y.; Tjong, S.C. In Situ ceramic particle-reinforced aluminum matrix composites fabricated by reaction pressing in the TiO2 (Ti)-Al-B (B2O3) systems. Met. Mater. Trans. A 1997, 28, 1931–1942. [Google Scholar] [CrossRef]
  26. Xia, T.D.; Liu, T.Z.; Zhao, W.J.; Munir, Z.A.; Wang, T.M. Photo- and cathodoluminescence of the combustion-synthesized Al2O3–TiB2 composites. J. Mater. Res. 2000, 15, 1622–1629. [Google Scholar] [CrossRef]
  27. Meyers, M.; Olevsky, E.; Ma, J.; Jamet, M. Combustion synthesis/densification of an Al2O3-TiB2 composite. Mater. Sci. Eng. A 2001, 311, 83–99. [Google Scholar] [CrossRef]
  28. Zhu, H.G.; Wang, H.Z.; Ge, L.Q.; Chen, S.; Wu, S.Q. Formation of composites fabricated by exothermic dispersion reaction in Al-TiO2-B2O3 system. Trans. Nonferrous Met. Soc. China 2007, 17, 590–594. [Google Scholar] [CrossRef]
  29. Zhu, H.; Wang, H.; Ge, L. Wear properties of the composites fabricated by exothermic dispersion reaction synthesis in an Al-TiO2-B2O3 system. Wear 2008, 264, 967–972. [Google Scholar] [CrossRef]
  30. Yeh, C.; Li, R. Formation of TiB2-Al2O3 and NbB2-Al2O3 composites by combustion synthesis involving thermite reactions. Chem. Eng. J. 2009, 147, 405–411. [Google Scholar] [CrossRef]
  31. Mousavian, R.T.; Sharafi, S.; Shariat, M.H. Microwave-assisted combustion synthesis in a mechanically activated Al-TiO2-H3BO3 system. Int. J. Refract. Met. Hard Mater. 2011, 29, 281–288. [Google Scholar] [CrossRef]
  32. Mishra, S.; Gokuul, V.; Paswan, S. Alumina-titanium diboride in situ composite by self-propagating high-temperature synthesis (SHS) dynamic compaction: Effect of compaction pressure during synthesis. Int. J. Refract. Met. Hard Mater. 2014, 43, 19–24. [Google Scholar] [CrossRef]
  33. Khaghani-Dehaghani, M.A.; Ebrahimi-Kahrizsangi, R.; Setoudeh, N.; Nasiri-Tabrizi, B. Mechanochemical synthesis of Al2O3-TiB2 nanocomposite powder from Al-TiO2-H3BO3 mixture. Int. J. Refract. Met. Hard Mater. 2011, 29, 244–249. [Google Scholar] [CrossRef]
  34. Abdellahi, M.; Heidari, J.; Sabouhi, R. Influence of B source materials on the synthesis of TiB2-Al2O3 nanocomposite powders by mechanical alloying. Int. J. Miner. Met. Mater. 2013, 20, 1214–1220. [Google Scholar] [CrossRef]
  35. Yang, W.; Dong, S.; Luo, P.; Yangli, A.; Liu, Q.; Xie, Z. Effect of Ni addition on the preparation of Al2O3-TiB2 composites using high-energy ball milling. J. Asian Ceram. Soc. 2014, 2, 399–402. [Google Scholar] [CrossRef] [Green Version]
  36. Rabiezadeh, A.; Hadian, A.; Ataie, A. Preparation of alumina/titanium diboride nano-composite powder by milling assisted sol-gel method. Int. J. Refract. Met. Hard Mater. 2012, 31, 121–124. [Google Scholar] [CrossRef]
  37. Zhang, S.; Khangkhamano, M.; Zhang, H.; Yeprem, H.A. Novel Synthesis of ZrB2 powder via molten-salt-mediated magnesiothermic Reduction. J. Am. Ceram. Soc. 2014, 97, 1686–1688. [Google Scholar] [CrossRef]
  38. Bao, K.; Liu, C.; Yazdani, B.; Damavandi, Y.; Zhang, S. Low-Temperature preparation of lanthanum hexaboride fine powder via magnesiothermic reduction in molten salt. J. Ceram. Sci. Technol. 2016, 7, 403–408. [Google Scholar]
  39. Bao, K.; Wen, Y.; Khangkhamano, M.; Zhang, S. Low-temperature preparation of titanium diboride fine powder via magnesiothermic reduction in molten salt. J. Am. Ceram. Soc. 2017, 51, 257–2272. [Google Scholar] [CrossRef]
  40. Li, T.; Chen, J.; Xiao, J.; Wei, Y.; Li, G.; Zhang, S.; Li, N. Formation of liquid-phase isolation layer on the corroded interface of MgO/Al2O3-SiC-C refractory and molten steel: Role of SiC. J. Am. Ceram. Soc. 2021, 104, 2366–2377. [Google Scholar] [CrossRef]
  41. Hashimoto, S.; Yamaguchi, A. Synthesis of MgAl2O4 (Spinel) Powder Using MgCl2. J. Ceram. Soc. Jpn. 2001, 109, 894–896. [Google Scholar] [CrossRef] [Green Version]
  42. Horlait, D.; Grasso, S.; Chroneos, A.; Lee, W. Attempts to synthesise quaternary MAX phases (Zr,M)2AlC and Zr2(Al,A)C as a way to approach Zr2AlC. Mater. Res. Lett. 2016, 4, 137–144. [Google Scholar] [CrossRef] [Green Version]
  43. Mousavian, R.T.; Sharafi, S.; Roshan, M.R.; Shariat, M.H. Effect of mechanical activation of reagents’ mixture on the high-temperature synthesis of Al2O3-TiB2 composite powder. J. Therm. Anal. Calorim. 2011, 104, 1063–1070. [Google Scholar] [CrossRef]
  44. Jacobson, N.S.; Myers, D.L. High-temperature vaporization of B2O3(l) under reducing conditions. J. Phys. Chem. B 2011, 115, 13253–13260. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of stoichiometric samples after 4 h firing at 850 °C in: (a) KCl, (b) NaCl, and (c) MgCl2.
Figure 1. XRD patterns of stoichiometric samples after 4 h firing at 850 °C in: (a) KCl, (b) NaCl, and (c) MgCl2.
Materials 14 04742 g001
Figure 2. XRD patterns of stoichiometric samples after 4 h firing in NaCl at: (a) 850, (b) 950, (c) 1050, and (d) 1150 °C.
Figure 2. XRD patterns of stoichiometric samples after 4 h firing in NaCl at: (a) 850, (b) 950, (c) 1050, and (d) 1150 °C.
Materials 14 04742 g002
Figure 3. XRD patterns of samples using (a) 0, (b) 20, (c) 25, and (d) 30 wt% excess Al, after 4 h firing in NaCl at 1150 °C.
Figure 3. XRD patterns of samples using (a) 0, (b) 20, (c) 25, and (d) 30 wt% excess Al, after 4 h firing in NaCl at 1150 °C.
Materials 14 04742 g003
Figure 4. Influence of excess B2O3 on phase formation in samples using 30 wt% excess Al, after 4 h firing in NaCl at 1150 °C: (a) 0 (stoichiometric), (b) 10, and (c) 20 wt% excess B2O3.
Figure 4. Influence of excess B2O3 on phase formation in samples using 30 wt% excess Al, after 4 h firing in NaCl at 1150 °C: (a) 0 (stoichiometric), (b) 10, and (c) 20 wt% excess B2O3.
Materials 14 04742 g004
Figure 5. XRD patterns of samples using 30 wt% excess Al and 20 wt% excess B2O3 after firing in NaCl at 1050 °C for: (a) 4, (b) 5, and (c) 6 h.
Figure 5. XRD patterns of samples using 30 wt% excess Al and 20 wt% excess B2O3 after firing in NaCl at 1050 °C for: (a) 4, (b) 5, and (c) 6 h.
Materials 14 04742 g005
Figure 6. XRD patterns of samples after firing in NaCl at 1050 °C for 5 h using 30 wt% excess Al, and (a) 20, (b) 25, and (c) 30 wt% excess B2O3.
Figure 6. XRD patterns of samples after firing in NaCl at 1050 °C for 5 h using 30 wt% excess Al, and (a) 20, (b) 25, and (c) 30 wt% excess B2O3.
Materials 14 04742 g006
Figure 7. TEM images (a,b) and corresponding EDS (c,d) of the Al2O3-TiB2 composite powder resultant from 5 h firing in NaCl at 1050 °C (the small C and Cu peaks arose from the carbon coating and the Cu grid).
Figure 7. TEM images (a,b) and corresponding EDS (c,d) of the Al2O3-TiB2 composite powder resultant from 5 h firing in NaCl at 1050 °C (the small C and Cu peaks arose from the carbon coating and the Cu grid).
Materials 14 04742 g007
Figure 8. SEM images (a,b) and corresponding EDS (c,d) of the Al2O3-TiB2 composite powder resultant from 4 h firing in NaCl at 1150 °C.
Figure 8. SEM images (a,b) and corresponding EDS (c,d) of the Al2O3-TiB2 composite powder resultant from 4 h firing in NaCl at 1150 °C.
Materials 14 04742 g008
Figure 9. The standard Gibbs energy values corresponding to Reactions (1) and (3)–(10), as a function of temperature.
Figure 9. The standard Gibbs energy values corresponding to Reactions (1) and (3)–(10), as a function of temperature.
Materials 14 04742 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, X.; Bao, K.; Chen, J.; Jia, Q.; Zhang, S. One-Pot Synthesis of Alumina-Titanium Diboride Composite Powder at Low Temperature. Materials 2021, 14, 4742. https://doi.org/10.3390/ma14164742

AMA Style

Liu X, Bao K, Chen J, Jia Q, Zhang S. One-Pot Synthesis of Alumina-Titanium Diboride Composite Powder at Low Temperature. Materials. 2021; 14(16):4742. https://doi.org/10.3390/ma14164742

Chicago/Turabian Style

Liu, Xueyin, Ke Bao, Junfeng Chen, Quanli Jia, and Shaowei Zhang. 2021. "One-Pot Synthesis of Alumina-Titanium Diboride Composite Powder at Low Temperature" Materials 14, no. 16: 4742. https://doi.org/10.3390/ma14164742

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