Next Article in Journal
Crystal Chemistry and Structural Complexity of the Uranyl Molybdate Minerals and Synthetic Compounds
Next Article in Special Issue
Advanced Aerospace Materials: Processing, Microstructure, Mechanical Properties and Applications
Previous Article in Journal
Highly Collimated Monochromatic X-rays Generated by Collision of High-Energy Electrons with Tightly Focused Linearly Polarized Laser Pulse
Previous Article in Special Issue
Comparative Analysis of Three Constitutive Models and Microstructure Characteristics of Nb521 during Hot Deformation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Preparation Method of (Ti,Zr,Nb,Mo,W)B2-SiC Composite Ceramic Based on Reactive Sintering of Pre-Alloyed Metals

1
Key Laboratory of Solidification Control and Digital Preparation Technology (Liaoning Province), School of Materials Science and Engineering, Dalian University of Technology, Dalian 116024, China
2
Key Laboratory of Advanced Technology for Aerospace Vehicles of Liaoning Province, School of Aeronautics and Astronautics, Dalian University of Technology, Dalian 116024, China
3
AVIC Manufacturing Technology Institute, Beijing 100024, China
4
College of Chemistry and Materials Engineering, Bohai University, Jinzhou 121013, China
5
State Key Laboratory of Structural Analysis for Industrial Equipment, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(1), 14; https://doi.org/10.3390/cryst14010014
Submission received: 13 November 2023 / Revised: 8 December 2023 / Accepted: 15 December 2023 / Published: 22 December 2023

Abstract

:
High-entropy diboride-based (MeB2-based) ceramics are promising high-temperature structural materials because of their excellent mechanical properties, high-temperature stability, and oxidation resistance. In order to achieve low-temperature sintering of the high-entropy ceramics, a novel preparation method of high-entropy (Ti,Zr,Nb,Mo,W)B2-SiC ceramics based on reactive sintering of pre-alloyed solid-solution metals and nonmetals of Si, C, B4C was conducted in the current work. Mechanical alloying behavior of the mixed metal powders, as well as the phase composition, microstructure, mechanical properties, and oxidation behavior of the as-sintered MeB2-SiC ceramic were investigated. The XRD, SEM, and EPMA results indicated that the primary MeB2 solid-solution and SiC phases could be successfully formed during reactive sintering at a relatively low temperature of 1650 °C. The as-sintered MeB2-SiC ceramics had a high relative density of 97.8% and high mechanical properties (hardness of 19.74 ± 0.8 GPa, flexure strength of 533 ± 38 MPa, and fracture toughness of 6.01 ± 0.77 MPa·m1/2). Combining the oxidation behavior and microstructure evolution of the oxidation layer, a continuous and relatively dense MeOx-SiO2 oxidation layer was gradually formed and covered on the external surface, leading to decelerating oxidation behavior after an oxidation exposure time of 10 min.

1. Introduction

With the rapid development of aerospace technology, the demand for materials suitable for elevated temperature applications has generated significant interest in ultra-high temperature ceramics (UHTCs). In order to better meet their application under harsh environments, further improving mechanical, physical, and chemical performances is an important development direction for UHTCs [1,2,3,4]. Due to their ultra-high melting point, high mechanical properties, and excellent thermal and chemical stability, UHTC-based composites of diboride-silicon carbide have been considered as potential thermal-structural materials, such as ZrB2-SiC, HfB2-SiC, and TiB2-SiC.
Since high-entropy alloys (HEAs) with superior mechanical and physical properties were reported by Yeh [5] and Cantor [6] in 2004, the design concept of high-entropy materials (HEM) has been considered as a new strategy to further improve the materials’ performance. In 2015, the HEM design concept was first applied to ceramic materials [7]. Since then, high-entropy ceramics, including high-entropy oxide [7,8,9], carbide [10,11], boride [12], nitride [13,14], fluoride [15], sulfides [16], and silicide [17,18], have attracted great interest as a new kind of ceramic material system with great potential. Among them, high-entropy diboride ceramic (MeB2) has both advantages of HEM and UHTC, which are more suitable for high-temperature applications under harsh environments [2]. In general, MeB2 is a hexagonal-structured solid-solution with multi-principal elements in the cation position. The reported research indicated that the hardness and oxidation resistance of some MeB2 was higher than individual diboride ceramics [12]. Therefore, MeB2-SiC composite ceramic is expected to be an ideal ultra-high-temperature thermal-structural material.
Based on the various starting materials, the preparation methods of MeB2 followed four principal routes: (1) direct sintering combined with dissolving five individual metal diborides into solid-solution; (2) densification of pre-synthesized MeB2 powder; (3) boro/carbothermal reduction of metal oxide, carbon, and B4C powders; (4) borothermal reaction through different metals and B. For the route (1), Gild et al. prepared six different MeB2 with a relative density of 92% via spark plasma sintering (SPS) at 2000 °C [12]. Li et al. prepared a (Zr0.2Ta0.2Nb0.2Hf0.2Mo0.2)B2 ceramic by oscillatory pressure sintering (OPS) at 1800–1900 °C; additionally, the single-phase solid-solution can be formed at 1900 °C [19]. For the route (2), the pre-synthesized MeB2 powders can be densified at high temperatures (>1900 °C), while some residual oxides [20,21] or other secondary phases [22] appeared in the final ceramics. For the routes (3) and (4), in order to achieve a complete chemical reaction, the synthesis of MeB2 always requires temperatures greater than 1950 °C [23,24]. In short, due to the limited diffusion coefficient of the diboride ceramics, the formation of MeB2 solid-solution usually needs an ultra-high temperature (>1900 °C) and long holding time. In order to reduce the fabricating cost of MeB2, a new preparation strategy that can prepare MeB2 at a relatively lower temperature needs to be developed.
In this paper, for preparing high-entropy MeB2-SiC ceramics at relatively low temperatures, a novel preparation method based on reactive sintering of pre-alloyed metals (Me) and nonmetals is proposed. According to the experimental results, the primary MeB2 solid-solution and SiC can be formed at 1650 °C. Moreover, the alloying behavior of metals, as well as microstructure, phase composition, and elemental distribution of the as-sintered MeB2-SiC were investigated. Finally, the oxidation behavior and microstructure evolution of MeB2-SiC ceramics at 1400 °C was studied.

2. Materials and Experimental Procedures

2.1. Materials Preparation

Experiments were performed, starting with commercially-available Ti (Aladdin, Shanghai, China, particle size ≥ 48 μm, 99.99% purity), Zr (Aladdin, Shanghai, China, particle size ≥ 75 μm, 99.5% purity), Nb (Aladdin, Shanghai, China, particle size 25 μm, 99.95% purity), Mo (Aladdin, Shanghai, China, particle size 25 μm, 99.5% purity), W (Aladdin, Shanghai, China, particle size 45 μm, 99.99% purity), Si (Aladdin, Shanghai, China, particle size 75–380 μm, 99.9% purity), B4C (Aladdin, Shanghai, China, particle size 2–3 μm, 99.8% purity), and C (Aladdin, Shanghai, China, particle size 40 nm, 99% purity) powders.

2.2. Fabrication of MeB2-SiC

The detailed fabrication procedures of MeB2-SiC are described as follows, and schematically represented in Figure 1. The raw metals of Ti, Zr, Nb, Mo, and W were mechanically alloyed by high-energy ball milling. Equimolar Ti, Zr, Nb, Mo, and W were milled at 350 rpm for 48 h. Tungsten carbide (WC) balls (diameter of Φ5 and Φ8 mm) and WC pots were selected as the ball-milling media. The milling process was performed in a protective argon atmosphere and the mass ratio of WC ball and powder was fixed at 10:1. Then, Si, B4C, and C powders and 30 mL of anhydrous ethanol were added to the ball-mill pot and mixed with the pre-alloyed powders (Me) for 12 h. Anhydrous ethanol was used as the process control agent. The molar ratio of Me, Si, B4C, and C was 2:4:1.1:3. The nominal chemical reaction was:
2Me(s) + B4C(s) + 3C(s) + 4Si(s) = 2MeB2(s) + 4SiC(s)
Due to the inability to completely isolate the air during the experimental process, slight amounts of oxides are unavoidably present on the surface of the metal powders. The addition of a slight excess amount of B4C (ten mole percent) is typically necessary [25,26]. The reaction between carbide and oxide according to this reaction was:
MeOx(cr) + B4C(cr) → MeB2(cr) + CO(g) + B2O3(l)
Afterwards, the milled alloyed-Me, Si, B4C, and C powders were dried and grinded. Ultimately, these powders were reactive hot-pressed (HP) at 1650 °C for 30 min, with the heating rate of 10 °C/min under a pressure of 30 MPa in vacuum, followed by furnace cooling to room temperature.
The empirical criteria parameters were used to predict whether the five metals would form a stable solid-solution [27,28,29]. The empirical criteria parameters were the mixing enthalpy ΔHmix, related to the chemical affinity between the elements; the size parameter δ, linked to the atomic radii differences; the parameter Ω, which describes the relative contributions of the mixing enthalpy and the mixing entropy (ΔSmix) in the Gibbs free energy; and Valence electron concentration VEC, a useful parameter for predicting the stability of solid-solution phases. The empirical parameters of this experimental composition perfectly match the requirements for a BCC solid-solution formation: −15 kJ/mol < ΔHmix (−14.4) < 5 kJ/mol, Ω(2.492) < 6.6, δ(5.545) < 6%, and VEC(5) < 6.87. The parameters of the atomic radius and calculation process of empirical criterion parameters used in this paper are summarized in Supplement Tables S1 and S2, as well.
Thermodynamic calculations for the reactions between Me, Si, B4C, and C were performed using HSC 9 software. Due to the lack of thermodynamic data for the solid-solution of alloyed Me, the individual metals of Ti, Zr, Nb, Mo, and W were used instead of Me. The thermodynamic analysis of each metal in Me with Si, B4C, and C were performed in supplement (reaction S1–S5), and the corresponding results are shown in Figures S1–S5. The results indicated that the Gibbs free energy of the related reactions are negative at the sintering temperature; therefore, the reactions could occur from the perspective of thermodynamics.

2.3. Oxidation Tests

In order to study the oxidation processes of MeB2-SiC, oxidation tests were performed using the samples of 4 mm × 5 mm × 2 mm, cut from the as-sintered MeB2-SiC bulk. The MeB2-SiC was oxidated in a tube furnace (GSL-1800X, Hefei Kejing Materials Technology Co., Ltd., Hefei, China). The sample containing crucible was pushed into the furnace and heated to 1400 °C (10 °C/min), in an air atmosphere. After being oxidized in air at 1400 °C for the prescribed time and then being allowed to cool, the samples were reweighed to determine their mass gain due to the uptake of oxygen. The oxidation prescribed times of the samples at 1400 °C were 0, 5, 10, and 20 min.

2.4. Characterizations

Phase composition of the individual metals, alloyed-Me powders, and as-sintered MeB2-SiC bulk sample was characterized using an X-ray diffractometer (XRD, Shimadzu XRD-6000, Tokyo, Japan) with Cu-Kα radiation. Microstructures and elemental distribution of the as-received and oxidized samples were observed with a scanning electron microscope (SEM, NOVA NanoSEM 450, New York, NY, USA) and electron probe micro-analyzer (EPMA, JEOL JXA-8530F PLUS, Tokyo, Japan). Flexural strength was tested under four-point loading condition. Specimens with a size of 2 mm in width, 1.5 mm in thickness, and a length of at least 25 mm were used. The sample holder had a span between the two bearers of 20 mm. The distance between the two loading pistons was 10 mm. Supports and both loading pistons had steel knife edges, rounded to a radius of 1.5 mm. The flexural strength was calculated according to the equation:
σ = 3 F d 2 b h 2
where σ is the maximum center tensile stress (MPa), F is the load at fracture (N), d is the difference in the distance of the two supports and the distance of the two loading pistons (mm), b is the width of the specimen (mm), and h is the height of the specimen (mm). Fracture toughness was tested using the single edge notched beam (SENB) method under three-point loading condition. The specimens (2 mm × 4 mm × 25 mm) were used to measure fracture toughness. A blade was positioned in the center of the specimen to produce a 0.25 mm length notch. The fracture toughness was calculated according to the following equation:
K Ic   = P S B W 1.5 f a w
where P is the load of fracture (N), S is the spam between supports, B is the length, and W is the width (m). f(a/w) was calculated according to the standard test method [30]. Vickers microhardness (HVS-30, Shanghai Biman Instrument Co., Shanghai, China) was used to measure the hardness by indentation under a load of 9.8 N (Hv 1), with a dwelling time of 10 s. All indentations were carried out randomly during the hardness measurement. The distance between two neighboring indentations was kept at least four times the size of the indent, and an average of eight indentations were used as a hardness value.

3. Results

3.1. Pre-Alloying Behavior of Multiple Metallic Elements

Figure 2 shows the XRD patterns of the Ti-Zr-Nb-W-Mo mixture at different milling times. As can be seen in Figure 2, the diffraction peaks of the five individual metals (Ti, Zr, Nb, Mo, and W) were clearly detected before ball milling. Then, with the milling time increased, the peak intensity of the individual metals gradually decreased. Especially after 24 h of ball-milling, some diffraction peaks of the individual metals, e.g., Ti and Zr, even become invisible, and meanwhile, the obvious broadening of the residual peaks can also be observed. Such phenomena indicated that the various metals were forming solid-solution. During the ball milling process, the metallic powders were collided and squeezed by the WC grinding balls in the milling tank, which resulted in severe plastic deformation, fracture, and cold welding in the metallic powders. During this process, the powders were easily formed to a BCC solid-solution, due to the shorter diffusion distance and larger contact area. As the milling time reached up to 48 h, only three major diffraction peaks of the BCC solid-solution were detected, those being 2θ = 40.14°, 2θ = 58.02°, and 2θ = 73.08°, which correspond to crystal planes {110}, {200}, and {211}. Due to the lack of symmetry in the peak of {110}, the software XPSPEAK4.1 was used to resolve the issue. The results indicate that this diffraction peak may be a superposition of two peaks (2θ = 38.5° and 2θ = 40.14°), as shown in the Supplementary Figure S6. According to the standard card, the diffraction peaks should be BCC solid-solution and a slight amount of Nb (PDF # 89-5291). Although a minor amount of unsolidified undissolved Nb is present, the pre-alloying of the five individual metallic elements was basically completed. Therefore, the pre-alloying metallic powders milled at 48 h were selected for further high-entropy diboride ceramic preparation.

3.2. Phase Composition and Microstructure of the As-Sintered MeB2-SiC Ceramic

The pre-alloying metallic powders and the nonmetallic raw materials of Si, C, and B4C were mixed, and then reactive sintered at 1650 °C by the hot-pressing method. The measured density and calculated theoretical density are 4.857 g/cm3 and 4.966 g/cm3, respectively. Thus, the relative density of the as-sintered MeB2-SiC ceramic is 97.8%. Figure 3 shows the XRD pattern of the MeB2-SiC high-entropy diboride-based ceramics prepared by HP sintering. As can be seen in Figure 3, the diffraction peaks of the primary MeB2 solid-solution, SiC, and undesired WB2 phases were detected. Table 1 shows the space group and crystal parameter of the five individual MeB2 phases (ZrB2, NbB2, TiB2, MoB2, and WB2) reported in the standard PDF cards and the primary MeB2 solid-solution phase detected in this study. Compared to the individual MeB2 phases, the primary MeB2 solid-solution phase detected in this study has the same space group (P6/mmm (191)) and a similar crystal parameter (a = b = 3.078 Å, c = 3.384 Å), which are close to the average crystal parameter (a = b = 3.066 Å, c = 3.262 Å). The average crystal parameter of MeB2 is averaged from the individual diboride of the P6/mmm (191) space group. It should be noted that besides primary MeB2 solid-solution and SiC phases, a minor amount of WB2 phase with a P63/mmc (194) crystal structure was also detected in the XRD pattern (Figure 3). As the previous studies reported [31,32,33], for WB2 ceramics, the P63/mmc (194) crystal structure is more stable than the P6/mmm (191) crystal structure at an ambient condition. Generally, a WB2 ceramic with a P6/mmm (191) crystal structure is thermodynamically unstable and difficult to synthesize [34,35]. Therefore, in the current study, massive WB2 is difficult to dissolve completely into the MeB2 solid-solution phase with a P6/mmm (191) crystal structure, and consequently, minor WB2 with a P63/mmc (194) crystal structure was precipitated in the as-sintered MeB2-SiC ceramic.
In order to further understanding of the microstructure of the as-sintered MeB2-SiC ceramic, the typical morphology and elemental distribution on the polished surface are shown in Figure 4 and Figure 5, respectively. The ceramic consists of three distinctly different phases, named gray, black, and bright phases. As can be seen in the low-magnification morphology (Figure 4a), gray and black phases are the main phases, which were well dispersed with each other. Moreover, the bright phase has an extremely low volume fraction, indicating that it is not the major product after the reactive sintering. The elemental composition of each phase was determined by EPMA mapping (Figure 5) and spot (Table 2) analysis. Figure 5a,b show the BSE and SE2 images, while Figure 5c–j are the individual elemental distribution maps of Ti, Zr, Mo, Nb, W, B, C, and Si, respectively. The gray phase is enriched with Ti (6.39%), Zr (6.43%), Nb (6.65%), Mo (5.61%), W (5.08%), and B (69.84%). Combined with XRD, SEM, and EPMA results, the gray phase is the primary MeB2 solid-solution phase. For the black phase, it is enriched with Si (46.11%) and C (53.89%). The atomic ratio of Si:C is approximately equal to 1:1, indicating that the black phase is identified as the primary SiC phase. For the minor amount of bright phase, it is enriched of W (36.77%), B (50.57%), and Mo (12.66%). Combined with XRD, SEM, and EPMA results, it is probably a WB2 phase with P63/mmc (194) crystal structure, which contains some Mo solute.

3.3. Mechanical and Oxidation Properties of the MeB2-SiC Ceramic

The mechanical properties of the as-sintered MeB2-SiC ceramic are shown in Table 3. The hardness, flexure strength, and fracture toughness were 19.74 ± 0.8 GPa, 533 ± 38 MPa, and 6.01 ± 0.77 MPa·m1/2, respectively. Table 4 presents the Vickers hardness of the five individual diborides (TiB2, ZrB2, NbB2, MoB2, and WB2) and β-SiC reported in the previous studies. According to the rule-of-mixture law, the theoretical hardness of the MeB2-SiC ceramics was estimated as 21.26 GPa. Generally, due to solid-solution strengthening, the hardness of the high-entropy ceramic is usually slightly higher than the estimated one calculated by rule-of-mixture law. Moreover, a minor amount of P63/mmc (194)-WB2 grains with high-hardness (26.8 GPa) [36] were precipitated in the as-sintered MeB2-SiC ceramic, which may also increase the hardness of the ceramic. However, in the current study, the measured hardness (19.74 ± 0.8 GPa) is slightly lower than the estimated one (21.26 GPa). This may be caused by the as-sintered MeB2-SiC ceramic not being dense enough. Some mechanical properties of the reported diborides are also listed in Table 3. The hardness value of MeB2-SiC in this work is in the same range of individual metal diborides and some high-entropy diborides reported in the previous literature (17.9–23.8 GPa). However, there are also similar materials with higher hardness (26.3–27 GPa). The difference in hardness is not only due to the difference in density, but also to the elemental composition of the material. Therefore, in order to further improve the mechanical properties, the optimization of elemental composition and processing parameters need to be investigated in further works.
Isothermal oxidation tests at 1400 °C with different exposure times were conducted in order to investigate the oxidation property of as-sintered MeB2-SiC ceramics. The mass change per unit surface area (Δm/S) with respect to oxidation exposure time is presented in Figure 6. The results depicted in Figure 6 show that the Δm/S increased rapidly before 10 min, and then increased slowly, indicating an oxide passivation stage after 10 min. In order to study the oxidation process, the oxidation behavior was discussed combining with the microstructure evolution of the oxidation layer. Figure 7 shows the surface (Figure 7a,c,e,g) and cross-sectional (Figure 7b,d,f,h) morphologies of the MeB2-SiC ceramic after oxidation with different exposure times. As can be seen in Figure 7, with the increase of exposure time, the oxidation layer gradually thickened, meanwhile the oxidated surface morphologies changed significantly. When the oxidation time is 0 min, the oxidated surface is rough and porous, as shown in Figure 7a,b. During the heating process, the oxidation of MeB2 preceded that of SiC; therefore, since the temperature had just risen to 1400 °C, the oxidation products of B2O3 and MeOx were more than the oxidation products of SiO2. Due to the evaporation temperature of B2O3 being lower than 1400 °C, a massive amount of micron pores appeared on the oxidated surface because of the volatilization of B2O3. With the oxidation exposure time increasing to 5 min, the oxidation product of B2O3 continued to volatilize, leading to lots of large-size pores appearing on the oxidated surface (Figure 7c,d). As the exposure time continued to increase, the SiO2 oxidation products gradually accumulated on the oxidated surface. Due to the viscous flow characteristic of SiO2 at 1400 °C, massive accumulated SiO2 filled the pores. This made the oxidation products become a continuous and relatively dense SiO2-MeOx oxidation layer to cover the external surface of the MeB2-SiC ceramic, as shown in Figure 7e,f (10 min) and Figure 7g,h (20 min). Such a continuous and relatively dense oxidation layer could prevent a further oxidation process, which leads to decelerating oxidation behavior after the exposure time of 10 min.
Figure 8 shows the elemental distribution mapping results on the cross-sectional of the MeB2-SiC ceramic after oxidation at 1400 °C for 20 min. As can be seen in Figure 8, the external region II was enriched with O, Si, and five individual metallic elements, implying that the oxidation layer is mainly composed of SiO2 and MeOx. In this region, the dense SiO2-MeOx oxidation layer played a dominant role in the oxidation resistance of MeB2-SiC ceramic. The oxygen element could only enter into the interior of the ceramic by anionic oxygen diffusion through the oxidation layer. In other words, after the formation of the dense and continuous SiO2-MeOx oxidation layer, the oxidation reactions are controlled by diffusion of oxygen (atomic or anionic oxygen) from the air-ceramic interface to the oxide-ceramic interface. Therefore, only trace amounts of oxygenelements was detected in the inner region, indicating that the dense SiO2-MeOx oxidation layer can effectively slow down the entry of oxygen.

4. Conclusions

In summary, a high-entropy diboride-based ultra-high temperature ceramic (Ti,Zr,Nb,Mo,W)B2-SiC (MeB2-SiC), has been successfully prepared at a relatively low sintering temperature of 1650 °C by using reactive hot-press sintering of pre-alloyed Me solid-solution metals and nonmetals of Si, C, and B4C. The as-sintered MeB2-SiC ceramic showed a high relative density of 97.8% and high mechanical properties (hardness of 19.74 ± 0.8 GPa, flexure strength of 533 ± 38 MPa, and fracture toughness of 6.01 ± 0.77 MPa·m1/2). SEM and EPMA results showed that the as-sintered ceramic had a fine and homogeneous microstructure without significant elemental aggregation and grain coarsening. Aside from the primary high-entropy P6/mmm (191)-MeB2 and SiC phases, a minor amount of P63/mmc (194)-WB2 was also detected in the MeB2-SiC ceramics. In addition, the oxidation tests at 1400 °C with different exposure times were conducted to investigate the oxidation behavior and microstructure evolution of the oxidation layer. With the oxidation exposure time increasing to 10 min, a continuous and relatively dense SiO2-MeOx oxidation layer was gradually formed. After that, the oxygen (atomic or anionic oxygen) diffusion mechanism was changed from air-ceramic interface diffusion-controlled to the oxide-ceramic interface diffusion-controlled, leading to decelerating oxidation behavior.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14010014/s1, Figures S1–S5: Standard reaction Gibbs free energy of each component at different temperatures; Figure S6: The results of resolving the diffraction peaks by XPSPEAK 4.1. Table S1: Parameters of the five individual metals used in this article; Table S2: Thermodynamic parameters of TiZrNbMoW calculated by empirical criteria.

Author Contributions

Conceptualization, Y.Z.; methodology, H.T. and Z.W.; analysis, H.T., Z.W., F.W., L.F. and J.D.; investigation, H.T. and Z.W.; resources, Y.Z., J.D. and J.S.; data curation, H.T. and Z.W.; writing—original draft preparation, H.T. and Z.W.; writing—review and editing, Y.Z.; visualization, Y.Z., H.T. and Z.W.; supervision, J.S.; project administration, Y.Z.; funding acquisition, Y.Z. and L.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China [No.: 51805069], Fundamental Research Funds for the Central Universities, China [No.: DUT20JC52], and Provincial Doctoral Research Start-up Fund Program [No.: 2020-BS-235].

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors would like to thank Fengyun Yu (Dalian University of Technology) for assistance with EPMA experiments and related analysis.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Fahrenholtz, W.G.; Hilmas, G.E. Ultra-high temperature ceramics: Materials for extreme environments. Scr. Mater. 2017, 129, 94–99. [Google Scholar] [CrossRef]
  2. Golla, B.R.; Mukhopadhyay, A.; Basu, B.; Thimmappa, S.K. Review on ultra-high temperature boride ceramics. Prog. Mater. Sci. 2020, 111, 100651. [Google Scholar] [CrossRef]
  3. He, L.; Pan, L.; Zhou, W.; Niu, Z.; Chen, X.; Chen, M.; Zhang, Q.; Pan, W.; Xiao, P.; Li, Y. Thermal corrosion behavior of Yb4Hf3O12 ceramics exposed to calcium-ferrum-alumina-silicate (CFAS) at 1400 °C. J. Eur. Ceram. Soc. 2023, 43, 4114–4123. [Google Scholar] [CrossRef]
  4. Pan, L.; He, L.; Niu, Z.; Xiao, P.; Zhou, W.; Li, Y. Corrosion behavior of ytterbium hafnate exposed to water-vapor with Al(OH)3 impurities. J. Eur. Ceram. Soc. 2023, 43, 612–620. [Google Scholar] [CrossRef]
  5. Yeh, J.W.; Chen, S.K.; Lin, S.J.; Gan, J.Y.; Chin, T.S.; Shun, T.T.; Tsau, C.H.; Chang, S.Y. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  6. Cantor, B.; Chang, I.T.; Knight, P.; Vincent, A.J.B. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375, 213–218. [Google Scholar] [CrossRef]
  7. Rost, C.M.; Sachet, E.; Borman, T.; Moballegh, A.; Dickey, E.C.; Hou, D.; Jones, J.L.; Curtarolo, S.; Maria, J.P. Entropy-stabilized oxides. Nat. Commun. 2015, 6, 8485. [Google Scholar] [CrossRef]
  8. Vinnik, D.A.; Trofimov, E.A.; Zhivulin, V.E.; Zaitseva, O.V.; Gudkova, S.A.; Starikov, A.Y.; Zherebtsov, D.A.; Kirsanova, A.A.; Häßner, M.; Niewa, R. High-entropy oxide phases with magnetoplumbite structure. Ceram. Int. 2019, 45, 12942–12948. [Google Scholar] [CrossRef]
  9. Anandkumar, M.; Trofimov, E. Synthesis, properties, and applications of high-entropy oxide ceramics: Current progress and future perspectives. J. Alloys Compd. 2023, 960, 170690. [Google Scholar] [CrossRef]
  10. Castle, E.; Csanádi, T.; Grasso, S.; Dusza, J.; Reece, M. Processing and properties of high-entropy ultra-high temperature carbides. Sci. Rep. 2018, 8, 8609. [Google Scholar] [CrossRef]
  11. Feng, L.; Fahrenholtz, W.G.; Hilmas, G.E.; Zhou, Y. Synthesis of single-phase high-entropy carbide powders. Scr. Mater. 2019, 162, 90–93. [Google Scholar] [CrossRef]
  12. Gild, J.; Zhang, Y.; Harrington, T.; Jiang, S.; Hu, T.; Quinn, M.C.; Mellor, W.M.; Zhou, N.; Vecchio, K.; Luo, J. High-entropy metal diborides: A new class of high-entropy materials and a new type of ultrahigh temperature ceramics. Sci. Rep. 2016, 6, 37946. [Google Scholar] [CrossRef]
  13. Li, B.; Zhang, L.; Xu, Y.; Liu, Z.; Qian, B.; Xuan, F. Selective laser melting of CoCrFeNiMn high entropy alloy powder modified with nano-TiN particles for additive manufacturing and strength enhancement: Process, particle behavior and effects. Powder Technol. 2020, 360, 509–521. [Google Scholar] [CrossRef]
  14. Chen, T.K.; Shun, T.T.; Yeh, J.W.; Wong, M.S. Nanostructured nitride films of multi-element high-entropy alloys by reactive DC sputtering. Surf. Coat. Technol. 2004, 188, 193–200. [Google Scholar] [CrossRef]
  15. Chen, X.; Wu, Y. High-entropy transparent fluoride laser ceramics. J. Am. Ceram. Soc. 2020, 103, 750–756. [Google Scholar] [CrossRef]
  16. Zhang, R.Z.; Gucci, F.; Zhu, H.; Chen, K.; Reece, M.J. Data-driven design of ecofriendly thermoelectric high-entropy sulfides. Inorg. Chem. 2018, 57, 13027–13033. [Google Scholar] [CrossRef]
  17. Gild, J.; Braun, J.; Kaufmann, K.; Marin, E.; Harrington, T.; Hopkins, P.; Vecchio, K.; Luo, J. A high-entropy silicide: (Mo0.2Nb0.2Ta0.2Ti0.2W0.2)Si2. J. Mater. 2019, 5, 337–343. [Google Scholar] [CrossRef]
  18. Qin, Y.; Liu, J.X.; Li, F.; Wei, X.F.; Wu, H.Z.; Zhang, G.J. A high entropy silicide by reactive spark plasma sintering. J. Adv. Ceram. 2019, 8, 148–152. [Google Scholar] [CrossRef]
  19. Li, M.; Zhao, X.; Shao, G.; Wang, H.; Zhu, J.; Liu, W.; Fan, B.; Xu, H.; Lu, H.; Zhou, Y.; et al. Oscillatory pressure sintering of high entropy (Zr0.2Ta0.2Nb0.2Hf0.2Mo0.2)B2 ceramic. Ceram. Int. 2021, 47, 8707–8710. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Guo, W.M.; Jiang, Z.B.; Zhu, Q.Q.; Sun, S.K.; You, Y.; Plucknett, K.; Lin, H.T. Dense high-entropy boride ceramics with ultra-high hardness. Scr. Mater. 2019, 164, 135–139. [Google Scholar] [CrossRef]
  21. Liu, J.; Yang, Q.Q.; Zou, J.; Wang, W.M.; Wang, X.G.; Fu, Z.Y. Strong high-entropy diboride ceramics with oxide impurities at 1800 °C. Sci. China Mater. 2023, 66, 2061–2070. [Google Scholar] [CrossRef]
  22. Yang, Y.; Bi, J.Q.; Sun, K.N.; Qiao, L.J.; Liang, G.D.; Wang, H.Y.; Yuan, J.L.; Chen, Y.G. The effect of chemical element on hardness in high-entropy transition metal diboride ceramics. J. Eur. Ceram. Soc. 2023, 43, 5774–5781. [Google Scholar] [CrossRef]
  23. Feng, L.; Fahrenholtz, W.G.; Hilmas, G.E. Processing of dense high-entropy boride ceramics. J. Eur. Ceram. Soc. 2020, 40, 3815–3823. [Google Scholar] [CrossRef]
  24. Hoque, M.S.B.; Milich, M.; Akhanda, M.S.; Shivakumar, S.; Hoglund, E.R.; Staicu, D.; Qin, M.; Quiambao-Tomko, K.F.; Tomko, J.A.; Braun, J.L. Thermal and ablation properties of a high-entropy metal diboride: (Hf0.2Zr0.2Ti0.2Ta0.2Nb0.2)B2. J. Eur. Ceram. Soc. 2023, 43, 4581–4587. [Google Scholar] [CrossRef]
  25. Zhang, S.C.; Hilmas, G.E.; Fahrenholtz, W.G. pressure less Densification of Zirconium Diboride with Boron Carbide Additions. J. Am. Ceram. Soc. 2006, 89, 1544–1550. [Google Scholar] [CrossRef]
  26. Fahrenholtz, W.G.; Binner, J.; Zou, J. Synthesis of ultra-refractory transition metal diboride compounds. J. Mater. Res. 2016, 31, 2757–2772. [Google Scholar] [CrossRef]
  27. Yang, X.; Zhang, Y. Prediction of high-entropy stabilized solid-solution in multi-component alloys. Mater. Chem. Phys. 2012, 132, 233–238. [Google Scholar] [CrossRef]
  28. Guo, S.; Ng, C.; Lu, J.; Liu, C.T. Effect of valence electron concentration on stability of fcc or bcc phase in high entropy alloys. J. Appl. Phys. 2011, 109, 103505. [Google Scholar] [CrossRef]
  29. Chen, S.Y.; Tong, Y.; Tseng, K.K.; Yeh, J.W.; Poplawsky, J.D.; Wen, J.G.; Gao, M.C.; Kim, G.; Chen, W.; Ren, Y.; et al. Phase transformations of HfNbTaTiZr high-entropy alloy at intermediate temperatures. Scr. Mater. 2019, 158, 50–56. [Google Scholar] [CrossRef]
  30. ASTM E-399-90; Standard Test Method for Plane-Strain Fracture Toughness of Metallic Materials. Annual Book of ASTM Standards, Section 3, Metal Test Methods and Analytical Procedures, 1990; ASTM: Philadelphia, PA, USA, 1992; pp. 519–554.
  31. Wang, D.Y.; Li, X.; Wang, Y.X. First-principles investigation of hexagonal WB2 surfaces. J. Phys. Soc. Jpn. 2012, 81, 044712. [Google Scholar] [CrossRef]
  32. Zhao, E.J.; Meng, J.A.; Ma, Y.M.; Wu, Z.J. Phase stability and mechanical properties of tungsten borides from first principles calculations. Phys. Chem. Chem. Phys. 2010, 12, 13158–13165. [Google Scholar] [CrossRef] [PubMed]
  33. Hao, X.F.; Xu, Y.H.; Wu, Z.J.; Zhou, D.F.; Liu, X.J.; Cao, X.Q.; Meng, J. Low-compressibility and hard materials ReB2 and WB2: Prediction from first-principles study. Phys. Rev. B 2006, 74, 224112. [Google Scholar] [CrossRef]
  34. Qin, M.D.; Gild, J.; Wang, H.R.; Harrington, T.; Vecchio, K.S.; Luo, J. Dissolving and stabilizing soft WB2 and MoB2 phases into high-entropy borides via boron-metals reactive sintering to attain higher hardness. J. Eur. Ceram. Soc. 2020, 40, 4348–4353. [Google Scholar] [CrossRef]
  35. Chen, X.Q.; Fu, C.L.; Krcmar, M.; Painter, G.S. Electronic and structural origin of ultraincompressibility of 5d transition-metal diborides MB2 (M=W, Re, Os). Phys. Rev. Lett. 2008, 100, 196403. [Google Scholar] [CrossRef] [PubMed]
  36. Zhu, H.; Shi, L.; Li, S.; Duan, Y.; Xia, W.; Wang, Y. Influence of uniaxial strains on the mechanical properties of transition metal borides X2B, XB and XB2 (X=Cr, Mo, W). Phys. B 2018, 550, 100–111. [Google Scholar] [CrossRef]
  37. Liu, J.X.; Shen, X.Q.; Wu, Y.; Li, F.; Liang, Y.; Zhang, G.J. Mechanical properties of hot-pressed high-entropy diboride-based ceramics. J. Adv. Ceram. 2020, 9, 503–510. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Jiang, Z.B.; Sun, S.K.; Guo, W.M.; Chen, Q.S.; Qiu, J.X.; Plucknett, K.; Lin, H.T. Microstructure and mechanical properties of high-entropy borides derived from boro/carbothermal reduction. J. Eur. Ceram. Soc. 2019, 39, 3920–3924. [Google Scholar] [CrossRef]
  39. Kovalčíková, A.; Tatarko, P.; Sedlák, R.; Medveď, D.; Chlup, Z.; Múdra, E.; Dusza, J. Mechanical and tribological properties of TiB2-SiC and TiB2-SiC-GNPs ceramic composites. J. Eur. Ceram. Soc. 2020, 40, 4860–4871. [Google Scholar] [CrossRef]
  40. Orlovskaya, N.; Stadelmann, R.; Lugovy, M.; Subbotin, V.; Subhash, G.; Neubert, M.; Aneziris, C.G.; Graule, T.; Kuebler, J. Mechanical properties of ZrB2–SiC ceramic composites: Room temperature instantaneous behaviour. Adv. Appl. Ceram. 2013, 112, 9–16. [Google Scholar] [CrossRef]
  41. Akin, I.; Ocak, B.C.; Sahin, F.; Goller, G. Effects of SiC and SiC-GNP additions on the mechanical properties and oxidation behavior of NbB2. J. Asian. Ceram. Soc. 2019, 7, 170–182. [Google Scholar] [CrossRef]
  42. Bhaumik, S.K.; Divakar, C.; Singh, A.K.; Upadhyaya, G.S. Synthesis and sintering of TiB2 and TiB2–TiC composite under high pressure. Mater. Sci. Eng. A 2000, 279, 275–281. [Google Scholar] [CrossRef]
  43. Zhang, G.J.; Deng, Z.Y.; Kondo, N.; Yang, J.F.; Ohji, T. Reactive hot pressing of ZrB2-SiC composites. J. Am. Ceram. Soc. 2000, 83, 2330–2332. [Google Scholar] [CrossRef]
  44. Sairam, K.; Sonber, J.K.; Murthy, T.S.R.C.; Subramanian, C.; Fotedar, R.K.; Hubli, R.C. Reaction spark plasma sintering of niobium diboride. Int. J. Refract. Met. Hard Mat. 2014, 43, 259–262. [Google Scholar] [CrossRef]
  45. Tao, Q.; Zhao, X.P.; Chen, Y.L.; Li, J.; Li, Q.; Ma, Y.M.; Li, J.J.; Cui, T.; Zhu, P.W.; Wang, X. Enhanced Vickers hardness by quasi-3D boron network in MoB2. RSC. Adv. 2013, 3, 18317–18322. [Google Scholar] [CrossRef]
  46. Nastic, A.; Merati, A.; Bielawski, M.; Bolduc, M.; Fakolujo, O.; Nganbe, M. Instrumented and Vickers indentation for the characterization of stiffness, hardness and toughness of zirconia toughened Al2O3 and SiC armor. J. Mater. Sci. Technol. 2015, 31, 773–783. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the processing route adopted for the low-temperature reactive sintering of high-entropy MeB2-SiC ceramics.
Figure 1. Schematic representation of the processing route adopted for the low-temperature reactive sintering of high-entropy MeB2-SiC ceramics.
Crystals 14 00014 g001
Figure 2. XRD patterns of Me powders under different milling times.
Figure 2. XRD patterns of Me powders under different milling times.
Crystals 14 00014 g002
Figure 3. XRD patterns of the as-sintered high-entropy MeB2-SiC ceramic.
Figure 3. XRD patterns of the as-sintered high-entropy MeB2-SiC ceramic.
Crystals 14 00014 g003
Figure 4. Back-scattered electron (BSE) morphology of the as-sintered MeB2-SiC composite materials: (a) low-magnification image and (b) high-magnification image. WB2 is indicated by the yellow circles.
Figure 4. Back-scattered electron (BSE) morphology of the as-sintered MeB2-SiC composite materials: (a) low-magnification image and (b) high-magnification image. WB2 is indicated by the yellow circles.
Crystals 14 00014 g004
Figure 5. (a,b) Back-scattered electron (BSE) and secondary electron (SEI) images of the microstructure and (cj) individual elemental mapping (Ti, Zr, Mo, Nb, W, B, C, and Si) of the as-sintered MeB2-SiC composite ceramics.
Figure 5. (a,b) Back-scattered electron (BSE) and secondary electron (SEI) images of the microstructure and (cj) individual elemental mapping (Ti, Zr, Mo, Nb, W, B, C, and Si) of the as-sintered MeB2-SiC composite ceramics.
Crystals 14 00014 g005
Figure 6. Mass change per unit surface area after oxidation tests at 1400 °C as a function of oxidation exposure time.
Figure 6. Mass change per unit surface area after oxidation tests at 1400 °C as a function of oxidation exposure time.
Crystals 14 00014 g006
Figure 7. SEM images of the surface (a,c,e,g) and cross-sectional (b,d,f,h) of MeB2-SiC after oxidation tests at 1400 °C for different exposure time: (a,b) 0 min; (c,d) 5 min; (e,f) 10 min; and (g,h) 20 min. Dashed lines are used as markers to distinguish the oxide layer (right of the dashed line) from the matrix(left of the dashed line).
Figure 7. SEM images of the surface (a,c,e,g) and cross-sectional (b,d,f,h) of MeB2-SiC after oxidation tests at 1400 °C for different exposure time: (a,b) 0 min; (c,d) 5 min; (e,f) 10 min; and (g,h) 20 min. Dashed lines are used as markers to distinguish the oxide layer (right of the dashed line) from the matrix(left of the dashed line).
Crystals 14 00014 g007
Figure 8. MeB2-SiC ceramic after oxidation for 20 min at 1400 °C: (a) cross sectional microstructure; (b) individual elemental mapping (O, Si, Ti, Zr, Nb, Mo, and W, respectively) detected in the pink region marked in figure (a).
Figure 8. MeB2-SiC ceramic after oxidation for 20 min at 1400 °C: (a) cross sectional microstructure; (b) individual elemental mapping (O, Si, Ti, Zr, Nb, Mo, and W, respectively) detected in the pink region marked in figure (a).
Crystals 14 00014 g008
Table 1. Crystal structures and parameters of the five individual diborides and the MeB2 solid-solution.
Table 1. Crystal structures and parameters of the five individual diborides and the MeB2 solid-solution.
DiboridesPearson SymbolSpace Groupa = b (Å)c (Å)Standard Card (PDF #)
TiB2hP3P6/mmm (191)3.0363.23865-8698
ZrB2hP3P6/mmm (191)3.1653.53089-3930
NbB2hP3P6/mmm (191)3.1023.32175-1048
MoB2hP3P6/mmm (191) *3.0053.17386-1097
hR6R-3m (166)3.01520.97173-0704
WB2hP3P6/mmm (191) *3.0203.05089-3928
hP12P63/mmc (194)2.98313.87943-1386
Calculation of MeB2 by average methodhP3P6/mmm (191)3.0663.262This work
Experimental results of MeB2 from XRDhP3P6/mmm (191)3.0783.384This work
* Unstable phase at ambient temperature.
Table 2. Chemical composition (at.%) of the three phases measured by EPMA spot analysis in the as-sintered MeB2-SiC ceramic.
Table 2. Chemical composition (at.%) of the three phases measured by EPMA spot analysis in the as-sintered MeB2-SiC ceramic.
PhasesTiZrNbMoWBCSi
MeB2 (Gray)6.396.436.655.615.0869.84--
SiC (Black)------53.8946.11
WB2 (White)---12.6636.7750.57--
Table 3. Summary of mechanical properties of the as-sintered high-entropy MeB2-SiC ceramics and other similarly reported ceramics.
Table 3. Summary of mechanical properties of the as-sintered high-entropy MeB2-SiC ceramics and other similarly reported ceramics.
MaterialRelative Density (%)Vickers Hardness (GPa)Flexure Strength
(MPa)
Fracture Toughness
(MPa·m½)
Ref.
(Ti,Zr,Nb,Mo,W)B2-SiC97.819.74 ± 0.8533 ± 386.01 ± 0.77This work
(Ti0.2Zr0.2Hf0.2Nb0.2Ta0.2)B299.822.0 ± 0.9339 ± 173.81 ± 0.40[37]
(Hf0.2Zr0.2Ta0.2Nb0.2Ti0.2)B296.321.7  ±  1.1-4.06  ±  0.35[38]
(Ti0.2Zr0.2Hf0.2Nb0.2Ta0.2)B2-SiC9922.4 ± 0.7447 ± 454.85 ± 0.33[37]
(Hf0.2Zr0.2Mo0.2Nb0.2Ti0.2)B298.126.3  ±  1.8-3.64  ±  0.36[38]
(Hf0.2Mo0.2Ta0.2Nb0.2Ti0.2)B298.527.0  ±  0.4-4.47  ±  0.40[38]
(Ti0.2Zr0.2Hf0.2Ta0.2Mo0.2)B292.419.1 ± 1.8--[12]
(Ti0.2Zr0.2Hf0.2Ta0.2Cr0.2)B292.219.9 ± 2.6--[12]
TiB2-SiC98.1–98.723.7–23.8349–6014.8–5.3[39]
ZrB2–SiC99.120–21-3.8  ±  0.1[40]
NbB2–SiC97.517.9–18.9-4.16–4.62[41]
MoB2–SiC *-----
WB2–SiC *-----
* Unstable phase at ambient temperature.
Table 4. Vickers hardness of metal diborides, SiC and MeB2-SiC.
Table 4. Vickers hardness of metal diborides, SiC and MeB2-SiC.
MaterialsVickers Hardness (GPa)Ref.
TiB2 (191)23.6[42]
ZrB2 (191)22.0[43]
NbB2 (191)20.25[44]
MoB2 (191)15.2[45]
WB2 (191)11.5[36]
Calculated value of MeB2 (191) *18.6/
SiC23.0[46]
WB2 (194)26.8[33]
Calculated value of MeB2-SiC *21.26/
Measured value of MeB2-SiC19.74 ± 0.8This work
* Estimated by the rule-of-mixture law.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zu, Y.; Wang, Z.; Tian, H.; Wu, F.; Fu, L.; Dai, J.; Sha, J. A Novel Preparation Method of (Ti,Zr,Nb,Mo,W)B2-SiC Composite Ceramic Based on Reactive Sintering of Pre-Alloyed Metals. Crystals 2024, 14, 14. https://doi.org/10.3390/cryst14010014

AMA Style

Zu Y, Wang Z, Tian H, Wu F, Fu L, Dai J, Sha J. A Novel Preparation Method of (Ti,Zr,Nb,Mo,W)B2-SiC Composite Ceramic Based on Reactive Sintering of Pre-Alloyed Metals. Crystals. 2024; 14(1):14. https://doi.org/10.3390/cryst14010014

Chicago/Turabian Style

Zu, Yufei, Zi Wang, Hongliang Tian, Fan Wu, Lianshen Fu, Jixiang Dai, and Jianjun Sha. 2024. "A Novel Preparation Method of (Ti,Zr,Nb,Mo,W)B2-SiC Composite Ceramic Based on Reactive Sintering of Pre-Alloyed Metals" Crystals 14, no. 1: 14. https://doi.org/10.3390/cryst14010014

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