Next Article in Journal
Effect of Polymer Addition on Performance of Portland Cement Mortar Exposed to Sulphate Attack
Next Article in Special Issue
Repeated Thermal Shock Behavior of ZrB2–SiC–Graphite Composite under Pre-Stress in Air and Ar Atmospheres
Previous Article in Journal
Research on the Rheological Characteristics and Aging Resistance of Asphalt Modified with Tourmaline
Previous Article in Special Issue
Tensile Fracture Behavior and Characterization of Ceramic Matrix Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low-Temperature Molten Salt Synthesis and the Characterisation of Submicron-Sized Al8B4C7 Powder

1
College of Engineering, Mathematics and Physical Sciences, University of Exeter, Exeter EX4 4QF, UK
2
The State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan 430081, China
3
College of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
4
Henan Key Laboratory of High Temperature Functional Ceramics, Zhengzhou University, Zhengzhou 450052, China
5
School of Energy Science and Engineering, Central South University, Changsha 410205, China
*
Authors to whom correspondence should be addressed.
Materials 2020, 13(1), 70; https://doi.org/10.3390/ma13010070
Submission received: 2 December 2019 / Revised: 16 December 2019 / Accepted: 18 December 2019 / Published: 22 December 2019
(This article belongs to the Special Issue High Temperature Ceramic Materials)

Abstract

:
Submicron-sized (~200 nm) aluminium boron carbide (Al8B4C7) particles were synthesised from Al, B4C and carbon black raw materials in a molten NaCl-based salt at a relatively low temperature. The effects of the salt type/assembly and the firing temperature on the synthesis process were examined, and the relevant reaction mechanisms discussed. The molten salt played an important role in the Al8B4C7 formation process. By using a combined salt of 95%NaCl + 5%NaF, an effective liquid reaction medium was formed, greatly facilitating the Al8B4C7 formation. As a result, essentially phase-pure Al8B4C7 was obtained after 6 h of firing at 1250 °C. This temperature was 350–550 °C lower than that required by the conventional direct reaction and thermal reduction methods.

1. Introduction

Like their binary counterparts, ternary carbides are also an important class of high-performance non-oxide ceramics that have attracted great research interest, especially during the past decade. Among them, aluminium boron carbide (Al8B4C7 or Al3BC3) is regarded as a promising candidate material potentially applicable to a variety of important areas, such as high-temperature structural ceramics, as an antioxidant for carbon-containing refractories [1,2,3,4,5], as a sintering agent for SiC, ZrB2 and B4C materials [6,7,8,9,10], and as an absorber for new-generation nuclear reactors [11]. This is because of its many superior properties, including high hardness (15.2 GPa), high melting point (1900 °C), relatively low density (2.69 g/cm3) and low thermal expansion coefficient (6.67 × 10−6/K), high thermal conductivity (29.2 W/m/K), and good oxidation/corrosion resistance [1,12,13,14].
To fabricate high-performance Al8B4C7-based bulk ceramics, high-quality Al8B4C7 powder often needs to be used. In this regard, several synthesis methods/techniques have been developed to date, among which the thermal reduction and direct reaction methods have been investigated most extensively. In the former, Al or carbon (C) is often used as a reducing agent [15,16,17,18], so inexpensive and readily available boron-containing oxides can be used as a boron source to replace the much more expensive element boron (B) or B4C. However, a high synthesis temperature (1700–1800 °C) is required to complete the formation reaction. Furthermore, some by-products/intermediate phases such as Al2O3, Al4O4C and Al2OC often remain in the final product powder. In addition, the product particles generally have relatively large sizes and suffer from heavy agglomeration. For example, Zhu et al. [15] and Deng et al. [16] prepared Al8B4C7 powder containing secondary phases of Al2O3, Al2OC and Al4O4C at 1700 °C by using B2O3 (or Na2B4O7·10H2O), Al and C as raw materials. By using similar raw materials and a higher temperature (1800 °C), Cui et al. [17] prepared hexagonal micro-platelets of Al8B4C7 containing minor Al2OC. On the other hand, Lee et al. [18] prepared Al3BC3 via a complex route using Al(OH)3, B2O3 and phenolic resin as raw materials. Despite the use of phenolic resin instead of solid C powder, the synthesis temperature still remained as high as 1725 °C.
In contrast to the thermal reduction method, no reducing agent is required for the direct reaction method. The raw material assemblies commonly used by this method include: (1) Al4C3 and B4C; (2) Al, B and C; and (3) Al, B4C and C. Unfortunately, this method also suffers from similar drawbacks to those of the thermal reduction method, i.e., high synthesis temperature (1600–1800 °C) [11,13,14,19,20,21], and relatively large sizes of product particles with heavy agglomeration among them. For example, Inoue et al. [22] synthesised Al8B4C7 powder via a respectively direct solid–solid reaction between Al4C3 and B4C at 1800 °C, and double stage reactions between Al, B and C initially at 1400 °C and then at 1830 °C. Several other researchers, e.g., Gao et al. [21], Hashimoto et al. [14] and Wang et al. [13], also synthesised Al8B4C7 powder at 1600–1800 °C by using Al, B4C and C as raw materials.
To overcome the drawbacks of the two main synthesis techniques stated above, it is necessary to develop alternative techniques. As a response to this, in the present work, a molten salt synthesis (MSS) method, used previously to prepare oxide and binary carbide powders [23,24,25], was further developed and extended to synthesise high-quality submicron-sized Al8B4C7 powder at a much lower temperature, from Al, B4C and C starting materials. As-prepared Al8B4C7 powder was characterised, and the effects of key processing factors such as firing temperature and salt type/assembly on MSS were investigated. Based on the experimental results, the synthesis/formation mechanism of Al8B4C7 was discussed.

2. Experimental Procedure

2.1. Raw Materials

Al (99.7% pure, <25 micron), B4C (99.98% pure), amorphous B (95% pure) and C (carbon black, ≥99% pure, <250 nm) powders were used as raw materials, and NaCl (≥99%) and NaF (≥98%) were used to form the desired liquid reaction medium. They were all purchased from Sigma-Aldrich (Gillingham, UK).

2.2. Sample Preparation

Al, B4C and C were mixed in the stoichiometric molar ratios of 8:1:6 (1.35:0.35:0.53 g in a powder batch) corresponding to Equation (1), and then they were further combined with 20 g binary salt of 95%NaCl + 5%NaF in an agate mortar. The mixed powder batch was contained in a graphite crucible covered with a graphite lid, and then it was placed in an alumina tube furnace protected by flowing argon (Ar). The furnace was heated to a target temperature between 1100 and 1250 °C (at 5 °C/min to 1000 °C, then 3 °C/min to 1200 °C and finally 1 °C/min to the target temperature) and held at the temperature for 6 h.
8Al + B4C + 6C = Al8B4C7
To study the effects of salt type/assembly on the formation of Al8B4C7, two other types of salts (NaCl, and 97.5%NaCl + 2.5%NaF) were used, as well as the binary salt NaCl-NaF stated above, to form the reaction media. They were then compared.
In addition, to assist in clarifying the relevant reaction/formation mechanisms, the following supplementary experiment was also carried out, and the resultant samples were similarly characterised (Section 2.3 below). In the first test, Al and B4C (1.35 and 0.35 g) in the molar ratio of 8:1 (referred to as Al-B4C sample) were heated in 20 g of 95%NaCl + 5%NaF at 1250 °C for 6 h. The reacted mass was further combined with 0.53 g C (so the molar ratio of Al:B4C:C = 8:1:6) and reheated at 1250 °C for 6 h in the identical salt. In the second test, Al4C3 (prepared via the reaction of stoichiometric amounts of Al and C in 20 g of 95%NaCl + 5%NaF at 1150 °C for 6 h) was combined with B and C (the molar ratio of Al/B4C/C = 8:1:6) (referred to as Al4C3-B-C sample) and fired at 1250 °C for 6 h.
Some of the samples after firing were placed immediately in a desiccator to avoid the hydration of Al4C3 in them prior to characterisation, whereas the other fired samples were subjected to repeated hot water washing to leach out the residual salt, followed by overnight oven-drying at 100 °C.

2.3. Sample Characterisation

Phases in fired samples were identified by powder X-ray diffraction (XRD) analysis (Bruker D8 advance reflection diffractometer, Karlsruhe, Germany). The diffractometer was operated at 40 mA and 40 kV using Ni-filtered CuKa radiation. The scan rate was 2.4° (2θ)/min with a step size of 0.04°. The ICDD cards used for identification were Al8B4C7 (35-1216), Al3BC3 (88-1267), Al4C3 (35-0799), AlOOH (21-1307), Al (65-2869), Al2O3 (46-1212), Al3BC (50-1470), AlB2 (65-9698) and NaCl (05-0628). The microstructure and morphology of the as-prepared product powder were observed using a scanning electron microscope (SEM Nova Nanolab 600, FEI Company, Hillsboro, OR, USA) and a JEM 2100 transmission electron microscope (TEM, 200 kV).

3. Results and Preliminary Discussion

3.1. Effect of Firing Temperature on the Formation of Al8B4C7

Figure 1 and Figure 2 show XRD patterns of samples resulting from 6 h of firing at different temperatures in 95%NaCl + 5%NaF, before and after water washing, respectively. The formation of Al8B4C7 was already evident after 6 h at 1100 °C, but the intermediate phase of Al4C3 was also detected (Figure 1a). With an increase in temperature to 1150 and then 1200 °C, Al8B4C7 increased, whereas Al4C3 decreased (Figure 1b,c). Upon increasing the temperature to 1250 °C, Al4C3 disappeared, and only Al8B4C7 was identified (though a minor peak from an unknown phase appeared at 2θ = 23.3°), i.e., essentially phase-pure Al8B4C7 was formed (Figure 1d). AlOOH was detected in some of the washed samples (Figure 2a–c) due to the partial hydration of Al4C3 in the original fired samples (Figure 1a–c) during the repeated water-washing process (Equation (2)). As shown in Figure 2, AlOOH, i.e., Al4C3, decreased with the increase in the corresponding firing temperature. It disappeared upon increasing the firing temperature to 1250 °C, verifying the completion of the formation reaction at this temperature (Figure 2d).
Al4C3 + 8H2O = 4AlOOH + 3CH4

3.2. Supplementary Experiment for Mechanism Clarification

Shown in Figure 3 are the XRD patterns of the Al-B4C sample after the first-stage firing, and the patterns after subsequent re-firing with C in the 95%NaCl + 5%NaF salt. After the first-stage firing (Figure 3a), Al3BC was formed as the main phase, along with some AlB2. However, after the second-stage firing with C, Al8B4C7 became the primary phase (Figure 3b), suggesting that the Al3BC that formed in the sample after the first-stage firing was converted into Al8B4C7. Figure 4 further presents the XRD pattern of the Al4C3-B-C sample after 6 h of firing in the 95%NaCl + 5%NaF salt at 1250 °C, revealing the formation of the primary phase of Al8B4C7, as well as minor residual C and Al2O3. The minor Al2O3 detected in this case (also in Figure 3b) was likely a result of the decomposition of AlOOH formed from the quick hydration of Al4C3 by the moisture in the atmosphere during the sample processing.

3.3. Effect of Salt Type/Assembly on the Formation of Al8B4C7

Figure 5 demonstrates the effect of salt type/assembly on the Al8B4C7 formation. In the case of using NaCl (Figure 5a), only minor Al8B4C7 was formed, but large amounts of AlOOH were detected in the water-washed sample, indicating the presence of large amounts of intermediate Al4C3 in the original fired sample. This implied the limited accelerating effect of NaCl on the Al8B4C7 formation. However, when small amounts (0.5 g, i.e., 2.5%) of NaF were combined with NaCl, Al8B4C7 became the main phase, although some AlOOH (i.e., Al4C3 in the original fired sample) was still detected (Figure 5b). This indicated the great accelerating effect of the NaF addition on the overall synthesis process. Upon further increasing the NaF amount to 1 g (i.e., 5%), AlOOH (i.e., Al4C3) disappeared and essentially phase-pure Al8B4C7 was formed (Figure 5c). The above results indicated that the optimal salt type/assembly in the present work was 95%NaCl + 5%NaF.

3.4. Microstructure of As-Prepared Al8B4C7 Powder

Figure 6 presents SEM and TEM images of Al8B4C7 particles synthesised in 95%NaCl + 5%NaF at 1250 °C for 6 h, revealing their irregular morphologies and average size of about 200 nm. The particles overall were dispersed well, though some were agglomerated together. The average size of the particles was much smaller, and their dispersion was much better than it was when the conventional synthesis techniques were used [15,16,17,18]. The lattice interlayer spacing (one of the insets in Figure 6) was measured as around 0.29 nm, which corresponds to the (111) plane of hexagonal Al8B4C7. This, in addition to the selected area electron diffraction (SAED) pattern (the other inset in Figure 6) and the XRD results in Figure 1 and Figure 2, verified that the synthesised particles were Al8B4C7.

4. Further Discussion and Reaction/Synthesis Mechanisms

Upon increasing the firing temperature above their melting/eutectic points, NaCl (melting point: ~714 °C) and NaF (melting point: ~743 °C) interacted with each other, forming a liquid medium in which Al slightly dissolved [26,27]. The dissolved Al diffused rapidly through the liquid medium onto the surfaces of C and B4C, and then reacted with them to form Al4C3, and Al3BC + AlB2, according to Equations (3) and (4), respectively.
4Al + 3C= Al4C3
9Al + 2B4C = 2Al3BC + 3AlB2
AlB2 = Al + 2 B
Since AlB2 is not thermodynamically stable at >1000 °C [28], the AlB2 formed from Equation (4) decomposed, forming Al and B (Equation (5)) in the molten salt [28]. The detection of Al4C3/AlOOH indicated the occurrence of Equation (3) at the test temperatures (Figure 1a–c and Figure 2a–c), and the detection of Al3BC and AlB2 in the Al-B4C sample (Figure 3) indicated the occurrence of Equation (4). The AlB2 phase detected in this case is believed to be formed upon cooling from the Equation between the residual Al and B in the salt. The B formed from Equation (5) at the test temperatures also slightly dissolved in the molten salt [29,30] and then diffused through the molten salt onto the surface of the Al4C3 formed from Equation (3), forming Al8B4C7 according to Equation (6). As shown in Figure 4, Al8B4C7 was formed in the fired Al4C3-B-C sample, indicating that the original Al4C3 reacted directly with the B dissolved in the salt, to form Al8B4C7.
7Al4C3 + 12 [B] = 3Al8B4C7 + 4 [Al]
According to Figure 3a, if no carbon was present, the intermediate Al3BC formed from Equation (4) appeared to be stable. However, when C was present, it was readily converted into more stable Al8B4C7 (Figure 4). This also explained why no Al3BC was found in the samples whose XRD patterns are shown in Figure 1 and Figure 2. The mechanism by which it was transformed into Al8B4C7 in the molten salt was not clear, but a plausible mechanism could be considered as follows: when C was present, it reacted with the Al in the molten salt to form Al4C3, which further reacted with the B in the molten salt to form Al8B4C7. The consumption of Al and B in the molten salt might have led to the decomposition of Al3BC and thus the additional formation of Al8B4C7 according to Equation (7).
7Al3BC = 13 [Al] + Al8B4C7 + 3 [B]
The overall reaction processes/mechanisms described above can also be used to explain the effects of firing temperature and salt type/assembly on the MSS process. With an increase in the firing temperature, the solubilities of Al and B in the molten salt were increased, and their diffusions in the molten salt accelerated. Consequently, Equations (3)–(7) were greatly facilitated. Therefore, the overall formation reaction (Equation (1)) was considerably accelerated (Figure 1, Figure 2, Figure 3 and Figure 4). When a single salt of NaCl was used, there was only limited formation of Al8B4C7 in the sample after 6 h of firing at 1250 °C (Figure 5a). However, when small amounts of NaF (2.5%) were added to NaCl, much more Al8B4C7 was formed (Figure 5b). Upon further increasing NaF to 5%, the formation reaction was completed, and essentially phase-pure Al8B4C7 was obtained (Figure 5c). This can be explained as follows. Al and B have very limited solubility in molten NaCl [31], so Equations (3)–(7) proceeded very slowly in it. However, when NaF was added to NaCl, the solubilities of Al and B in the binary salt were increased significantly, which led to great acceleration of Equations (3)–(7), i.e., the overall formation reaction (Equation (1)).
Thanks to the strong accelerating effect of the NaCl-NaF binary salt discussed above, essentially phase-pure Al8B4C7 particles were successfully prepared at 1250 °C. This synthesis temperature was 350–550 °C lower than that required by the conventional synthesis routes [11,13,14,15,16,17,18,19,20,21], demonstrating the great advantage and feasibility of the MSS technique developed in this work.

5. Conclusions

A low-temperature molten salt synthesis technique was developed to synthesise high-quality Al8B4C7 particles. The main conclusions can be drawn as follows.
  • Al8B4C7 particles with an average size of about 200 nm were successfully synthesised after 6 h of firing in NaCl-NaF at 1250 °C, from Al, B4C and C starting powders. They were essentially phase-pure and generally well-dispersed.
  • Compared with the temperature required by a conventional synthesis technique, the synthesis temperature (1250 °C) in the present work was significantly lower (350–500 °C lower), owing to the great accelerating effect of NaCl-NaF salt.
  • Al8B4C7 particles were formed via the following mechanisms: at the test temperatures, NaCl and NaF interacted with each other, forming a liquid medium in which Al slightly dissolved. The dissolved Al diffused rapidly through the molten salt onto the surfaces of C and B4C, reacting with them to form Al4C3, and Al3BC + AlB2, respectively. AlB2 is not stable at >1000 °C, so at the test temperatures, it decomposed into B and Al. The newly formed B also slightly dissolved in the salt, diffused onto the surface of the Al4C3 formed earlier, and reacted with it to form Al8B4C7, which consumed Al and B in the salt, making the Al3BC formed earlier decompose into additional Al8B4C7, Al and B.

Author Contributions

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

Funding

X.L., Q.J., Z.H. and B.C. would like to acknowledge the financial support from the State Key Laboratory of Refractories and Metallurgy (Wuhan University of Science and Technology), No. G201810, and visiting scholarships from China Scholarship Council.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, T.; Yamaguchi, A. Oxidation protection of MgO–C refractories by means of Al8B4C7. J. Am. Ceram. Soc. 2001, 84, 577–582. [Google Scholar] [CrossRef]
  2. Wang, T.; Yamaguchi, A. Antioxidation behavior and effect of ZrB2–Al3BC3 composites prepared using Al–B4C–C additives and spark plasma sintering added to carbon-containing refractories. J. Ceram. Soc. Jpn. 2000, 108, 818–822. [Google Scholar] [CrossRef] [Green Version]
  3. Yamaguchi, A.; Nakano, Y.; Wang, T. Effect and behaviour of Al-B-C system antioxidants added to MgO-C refractories. Can. Metall. Q. 2000, 39, 381–386. [Google Scholar] [CrossRef]
  4. Zhang, S.; Marriott, N.J.; Lee, W.E. Thermochemistry and microstructures of MgO–C refractories containing various antioxidants. J. Eur. Ceram. Soc. 2000, 21, 1037–1047. [Google Scholar] [CrossRef]
  5. Zhang, S.; Lee, W.E. Influence of additives on corrosion resistance and corroded microstructures of MgO–C refractories. J. Eur. Ceram. Soc. 2001, 21, 2393–2405. [Google Scholar] [CrossRef]
  6. Lee, S.H.; Kim, B.N.; Tanaka, H. Low temperature sintering of nano-SiC using a novel Al8B4C7 additive. J. Mater. Res. 2010, 25, 471–475. [Google Scholar] [CrossRef]
  7. Lee, S.H.; Sakka, Y.; Tanaka, H.; Kagawa, Y. Wet Processing and low-temperature pressureless sintering of SiC using a novel Al3BC3 sintering additive. J. Am. Ceram. Soc. 2009, 92, 2888–2893. [Google Scholar] [CrossRef]
  8. Wang, H.; Feng, L.; Lee, S.H.; Chen, J.; Fan, B.; Chen, D.; Lu, H.; Xu, H.; Zhang, R. ZrB2–Al3BC3 composites prepared using Al–B4C–C additives and spark plasma sintering. Ceram. Int. 2013, 39, 897–901. [Google Scholar] [CrossRef]
  9. Jung, J.Y.; Jung, S.H.; Oh, H.C.; Lee, S.H.; Choi, S.C. Spark plasma sintering of ZrB2 using Al3BC3 as an additive. J. Ceram. Process. Res. 2012, 13, 641–645. [Google Scholar]
  10. Da Rocha, R.M.; de Melo, F.C.L. Sintering of B4C by pressureless liquid phase sintering. Mater. Sci. Forum 2010, 660, 170–175. [Google Scholar] [CrossRef]
  11. Li, F.; Zhou, Y.; He, L.; Liu, B.; Wang, J. Synthesis, microstructure, and mechanical properties of Al3BC3. J. Am. Ceram. Soc. 2008, 91, 2343–2348. [Google Scholar] [CrossRef]
  12. Wang, T.; Yamaguchi, A. Some properties of sintered Al8B4C7. J. Mater. Sci. Lett. 2000, 19, 1045–1046. [Google Scholar] [CrossRef]
  13. Wang, T.; Yamaguchi, A. Synthesis of Al8B4C7 and its oxidation properties in air. J. Ceram. Soc. Jpn. 2000, 108, 375–380. [Google Scholar] [CrossRef] [Green Version]
  14. Hashimoto, S.; Ishihara, T.; Inoue, K.; Honda, S.; Iwamoto, Y.; Zhang, S. Synthesis and mechanical properties of Al8B4C7. J. Ceram. Soc. Jpn. 2009, 117, 18–21. [Google Scholar] [CrossRef] [Green Version]
  15. Zhu, H.; Pan, C.; Deng, C.; Yuan, W. Preparation of Al8B4C7 composite materials by using oxide raw materials. IOP Conf. Ser. Mater. Sci. Eng. 2011, 18, 222008. [Google Scholar] [CrossRef] [Green Version]
  16. Deng, C.; Zhou, W.; Zhu, H.; Yu, Y. Synthesis of Al8B4C7 from Al/Na2B4O7·10H2O/C Mixed Powders. Adv. Mater. Res. 2009, 79, 1375–1378. [Google Scholar] [CrossRef]
  17. Cui, P.; Yuan, W.; Deng, C.; Zhu, H.; Li, J. Synthesis of Al8B4C7 from Aluminum, Boron Trioxide and Activated Carbon Mixed Powders. Adv. Mater. Res. 2013, 634, 2383–2387. [Google Scholar] [CrossRef]
  18. Lee, S.H.; Yin, J.; Feng, L.; Lee, J.S. Synthesis of Al3BC3 particulates by carbo-thermal reduction process—Parameter optimization and mechanism analysis. J. Ceram. Soc. Jpn. 2014, 122, 772–776. [Google Scholar] [CrossRef] [Green Version]
  19. Lee, S.H.; Kim, H.D.; Choi, S.C.; Nishimura, T.; Lee, J.S.; Tanaka, H. Chemical composition and microstructure of Al3BC3 prepared by different densification methods. J. Eur. Ceram. Soc. 2010, 30, 1015–1020. [Google Scholar] [CrossRef]
  20. Che, Q.; Ma, Q.; Lu, J. Fabrication of Al3BC3-based porous ceramics by pressureless sintering. Adv. Mater. Res. 2012, 412, 340–343. [Google Scholar] [CrossRef]
  21. Gao, Y.; Huang, Z.; Fang, M.; Liu, Y.; Huang, S.; Ouyang, X. Synthesis of Al8B4C7 ceramic powder from Al/B4C/C mixtures. Powder Technol. 2012, 226, 269–273. [Google Scholar] [CrossRef]
  22. Inoue, Z.; Tanaka, H.; Inomata, Y. Synthesis and X-ray crystallography of aluminium boron carbide, Al8B4C7. J. Mater. Sci. 1980, 15, 3036–3040. [Google Scholar] [CrossRef]
  23. Li, Z.; Zhang, S.; Lee, W.E. Molten salt synthesis of LaAlO3 powder at low temperatures. J. Eur. Ceram. Soc. 2007, 27, 3201–3205. [Google Scholar] [CrossRef]
  24. Jayaseelan, D.D.; Zhang, S.; Hashimoto, S.; Lee, W.E. Template formation of magnesium aluminate (MgAl2O4) spinel microplatelets in molten salt. J. Eur. Ceram. Soc. 2007, 27, 4745–4749. [Google Scholar] [CrossRef]
  25. Xie, W.; Möbus, G.; Zhang, S. Molten salt synthesis of silicon carbide nanorods using carbon nanotubes as templates. J. Mater. Chem. 2011, 21, 18325. [Google Scholar] [CrossRef]
  26. Yoshida, K.; Dewing, E.W. The Solubility of Aluminum in Cryolite Melts. Essential Readings in Light Metals: Aluminum Reduction Technology; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2016; pp. 12–18. [Google Scholar]
  27. Yoshida, K.; Dewing, E.W. The apparent solubility of aluminum in cryolite melts. Metall. Mater. Trans. 1972, 3, 1817–1821. [Google Scholar] [CrossRef]
  28. Ren, D.; Deng, Q.; Wang, J.; Li, Y.; Li, M.; Ran, S.; Du, S.; Huang, Q. Densification and mechanical properties of pulsed electric current sintered B4C with in situ synthesized Al3BC obtained by the molten-salt method. J. Eur. Ceram. Soc. 2017, 37, 4524–4531. [Google Scholar] [CrossRef]
  29. 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, 100, 2266–2272. [Google Scholar] [CrossRef]
  30. 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]
  31. Viala, J.C.; Bouix, J.; Gonzalez, G.; Esnouf, C. Chemical reactivity of aluminium with boron carbide. J. Mater. Sci. 1997, 32, 4559–4573. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of samples resulting from 6 h of firing in 95%NaCl + 5%NaF salt at: (a) 1100, (b) 1150, (c) 1200, and (d) 1250 °C, respectively (prior to water washing).
Figure 1. X-ray diffraction (XRD) patterns of samples resulting from 6 h of firing in 95%NaCl + 5%NaF salt at: (a) 1100, (b) 1150, (c) 1200, and (d) 1250 °C, respectively (prior to water washing).
Materials 13 00070 g001
Figure 2. XRD patterns of samples resulting from 6 h of firing at: (a) 1100, (b) 1150, (c) 1200, and (d) 1250 °C, respectively (after water washing).
Figure 2. XRD patterns of samples resulting from 6 h of firing at: (a) 1100, (b) 1150, (c) 1200, and (d) 1250 °C, respectively (after water washing).
Materials 13 00070 g002
Figure 3. XRD patterns of: (a) the Al-B4C sample after 6 h of firing in 95%NaCl + 5%NaF at 1250 °C, and (b) the sample resulting from further firing of (a) with C in 95%NaCl + 5%NaF at 1250 °C for 6 h.
Figure 3. XRD patterns of: (a) the Al-B4C sample after 6 h of firing in 95%NaCl + 5%NaF at 1250 °C, and (b) the sample resulting from further firing of (a) with C in 95%NaCl + 5%NaF at 1250 °C for 6 h.
Materials 13 00070 g003
Figure 4. XRD pattern of the Al4C3-B-C sample after 6 h of firing in 95%NaCl + 5%NaF at 1250 °C.
Figure 4. XRD pattern of the Al4C3-B-C sample after 6 h of firing in 95%NaCl + 5%NaF at 1250 °C.
Materials 13 00070 g004
Figure 5. XRD patterns of samples resulting from 6 h of firing at 1250 °C in 20 g of (a) NaCl, (b) 97.5%NaCl+2.5%NaF, and (c) 95%NaCl + 5%NaF, respectively (after water washing).
Figure 5. XRD patterns of samples resulting from 6 h of firing at 1250 °C in 20 g of (a) NaCl, (b) 97.5%NaCl+2.5%NaF, and (c) 95%NaCl + 5%NaF, respectively (after water washing).
Materials 13 00070 g005
Figure 6. (a) SEM and (b) TEM images of Al8B4C7 product powder prepared in 95%NaCl + 5%NaF at 1250 °C for 6 h.
Figure 6. (a) SEM and (b) TEM images of Al8B4C7 product powder prepared in 95%NaCl + 5%NaF at 1250 °C for 6 h.
Materials 13 00070 g006aMaterials 13 00070 g006b

Share and Cite

MDPI and ACS Style

Liu, C.; Liu, X.; Hou, Z.; Jia, Q.; Cheng, B.; Zhang, S. Low-Temperature Molten Salt Synthesis and the Characterisation of Submicron-Sized Al8B4C7 Powder. Materials 2020, 13, 70. https://doi.org/10.3390/ma13010070

AMA Style

Liu C, Liu X, Hou Z, Jia Q, Cheng B, Zhang S. Low-Temperature Molten Salt Synthesis and the Characterisation of Submicron-Sized Al8B4C7 Powder. Materials. 2020; 13(1):70. https://doi.org/10.3390/ma13010070

Chicago/Turabian Style

Liu, Cheng, Xueyin Liu, Zhaoping Hou, Quanli Jia, Benjun Cheng, and Shaowei Zhang. 2020. "Low-Temperature Molten Salt Synthesis and the Characterisation of Submicron-Sized Al8B4C7 Powder" Materials 13, no. 1: 70. https://doi.org/10.3390/ma13010070

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