Next Article in Journal
Friction Behavior of a Textured Surface against Several Materials under Dry and Lubricated Conditions
Next Article in Special Issue
The Multipole Structure and Symmetry Classification of Even-Type Deviators Decomposed from the Material Tensor
Previous Article in Journal
Cytotoxic and Genotoxic Effects of Composite Resins on Cultured Human Gingival Fibroblasts
Previous Article in Special Issue
Tailoring the Structural and Electronic Properties of Graphene through Ion Implantation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of TiB2 Particles on the Microstructure Evolution and Mechanical Properties of B4C/TiB2 Ceramic Composite

1
Science and Technology on Complex and System Simulation Laboratory, Beijing 100072, China
2
Beijing Institute of Aerospace Control Devices, Beijing 100039, China
3
School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
4
National Key Laboratory of Science and Technology on Materials under Shock and Impact, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(18), 5227; https://doi.org/10.3390/ma14185227
Submission received: 20 July 2021 / Revised: 31 August 2021 / Accepted: 6 September 2021 / Published: 11 September 2021
(This article belongs to the Special Issue Feature Paper Collection of Topical Advisory Members)

Abstract

:
B4C/TiB2 ceramic composites reinforced with three size scales (average particle size: 7 μm, 500 nm, and 50 nm) of TiB2 were prepared by using a pressureless sintering furnace at 2100 °C under Ar atmosphere for 60 min. The results demonstrated that during the sintering process, TiB2 located on the boundaries between different B4C grains could inhibit the grain growth which improved the mass transport mechanism and sintering driving force. A semi-coherent interface between B4C and SiC was found, which is supposed to help to reduce the interface energy and obtain good mechanical properties of the B4C/TiB2 ceramic composite. On sample cooling from sintering temperature to room temperature, the residual tensile stress fields formed at the TiB2 interfaces owning to the thermo-elastico properties mismatched, which might have contributed to increase the ability of the sample to resist crack propagation. The results showed that the relative density, Vickers hardness, and fracture toughness of the composite with 20 wt.% submicron and 10 wt.% nano-TiB2 were significantly improved, which were 98.6%, 30.2 GPa, and 5.47 MPa·m1/2, respectively.

1. Introduction

Boron carbide (B4C) ceramics are interesting structural ceramics in view of their outstanding physical and mechanical properties, especially the combination of low density and extremely high hardness which make them superior anti-ballistic materials over other armor ceramics (such as Al2O3, SiC) [1,2,3]. However, the expensive costs of B4C ceramics fabricated through the hot isostatic pressing (HIP) method severely limits its wide application in the armor protection field [4,5,6]. In addition, its low self-diffusivity efficiency indicates that the sintered body could not achieve the goal of densification through the single solid-state sintering technique. Recently, numerous attempts have been made to overcome these disadvantages, such as the introduction of a second phase and sintering additives into the B4C matrix to fabricate composites. Transition metal borides, such as TiB2 [7,8], ZrB2 [9,10], and HfB2 [11], having high thermal expansion coefficients, and the residual stress fields between B4C and borides rising during the cooling process possibly enhance the fracture toughness of the fabricated composites [12,13].
Recently, the B4C/TiB2 ceramic composites have been the object of numerous works [14,15,16,17]. The additives of TiB2 to B4C phase can maintain the advantages of high Vickers hardness and low density of B4C and, in addition, inhibit the grain growth [16]. Additionally, the physical and mechanical properties of the B4C/TiB2 composites without additives prepared from the B4C and TiB2 powder are very low. The relative density of B4C-30 wt.% TiB2 composites without any additives prepared via pressureless sintering was lower than 90% [18,19]. The flexural strength of the B4C/TiB2 ceramic composite reached 717 MPa for the hot-pressured method, which was above two times higher than those (260–361 MPa) fabricated via pressureless sintering [16]. Many researchers have used the B4C-TiO2-C powder mixture to prepare the B4C/TiB2 composites in different ways, such as the reactive pressureless sintering, hot-pressing, and pulse electric-current sintering [20,21]. Since the introduction of fine TiB2 grains by in-situ reaction, the B4C and TiB2 grain size retained submicron sizes, and the mechanical property tests indicated that the prepared B4C/TiB2 ceramic composites achieved the excellent Vickers hardness ~39.3 GPa and flexural strength ~865 MPa, respectively [7]. B4C with various particle sizes was introduced to fabricate B4C/TiB2 ceramic composites under the condition of hot pressing, and both of the B4C and TiB2 grains were grown compared to the raw powders after hot pressing [22]. Many studies have shown that for B4C ceramic composites, C and Si are good sintering aids [23,24,25,26,27]. Carbon removes oxides (such as B2O3) in the B4C raw powder, and improves the interfacial tension by the way of solid solution of carbon atoms into the boron carbide lattice, which increases the sintering driving force [23,24]. A small amount of Si in the B4C ceramics tends to form a silicide phase, which could improve the sintering ability of B4C [25,26,27].
Although hot-pressing and pulse electric current sintering can obtain high-performance composites, the equipment and production costs are high, and the product size is small. The pressureless sintering is an efficient way to fabricate B4C/TiB2 composites with large sizes and low costs. At present, the research on improving the performance of the B4C/TiB2 ceramic composites prepared under pressureless sintering conditions is relatively scattered [18,28,29], and these reports indicate that the sintering temperature and TiB2 content have a great influence on the microstructure and density of the composite. Additionally, the research regarding the TiB2 particle size affecting the sintering behavior of B4C/TiB2 ceramic composites under pressureless sintering conditions is rarely reported. In our present work, the B4C/TiB2 ceramic composites with 30 wt.% TiB2 were fabricated via the pressureless sintering method from commercial B4C raw powder with the average size of 3 μm and TiB2 raw powder with three different size scales (7 μm, 500 nm, and 50 nm), and mixed in variable mass ratio. In addition, carbon black and silicon particles were used as sintering auxiliary components. Furthermore, the effect of TiB2 grains on the interfaces to optimize the microstructure of the B4C/TiB2 composites was thoroughly investigated. This research should be beneficial to fabricate the excellent performance of B4C/TiB2 ceramic composite.

2. Materials and Methods

Raw materials were B4C powder (3 μm, purity: >99.5%; Zhengzhou Songshan Boron Technology Co., Ltd., Zhengzhou, China), silicon raw powder and carbon black raw powder (submicron, purity: >99.8%, Shanghai ST-NANO Co. Ltd., Shanghai, China), and TiB2 powder (purity: >99%; Shanghai ST-NANO Co. Ltd., Shanghai, China). Figure 1a–c show the three type morphologies of the TiB2 raw powders. Figure 1a depicts the microtopography of micro-TiB2 powder with average size about 7 μm. Figure 1b,c show the microtopography of submicron TiB2 powder with average size about 500 nm and nano-TiB2 powder with average size about 50 nm, respectively. Table 1 lists the phase composition of the three mixtures. The mixed raw powders were ball-milled in ethyl alcohol absolute with ZrO2 balls and then dried using a rotary evaporator (R205B, Shanghai Shensheng Technology Co. Ltd., Shanghai, China). The powder mixture was pressed in a graphite die and then cold isostatic pressed (CIP, LDJ100/320–300, Sichuan Aviation Industry Chuanxi Machine Factory, Sichuan, China) to form a green body with a 50 mm diameter. The samples were processed by pressurelss sintering in a graphite crucible (FCT Systeme GmbH, Rauenstein, Germany) at 2100 °C for 60 min at a heating rate of 10 °C per minute under flowing Ar atmosphere.
The relative densities of the samples were determined through Archimedes’ principle in deionized water. The average grain size was estimated by intercept method and more than 200 grains on the surface after polishing and thermally etching were measured. The flexural strength of the prepared specimens which were cut into the bars of 3 × 4 × 35 mm3 was tested on an electromechanical universal testing machine (INSTRON-5566, Norwood, MA, USA) of which the crosshead speed was 0.5 mm per minute and the span was 30 mm. The fractural toughness of the composites tested on bars (the size of 3 × 6 × 35 mm3), and a notch depth of 3 mm, was measured by the single-edge notched beam (SENB) test of which the crosshead speed was 0.05 mm per minute and the span was 24 mm. Vickers hardness measurement applied a load of 1 kg for 15 s to the sample surfaces on a hardness testing device (AHVD-1000, Shanghai Jujing Precision Instrument Manufacturing Co., Ltd., Shanghai, China). The phases and components were characterized by X-ray diffraction (D8 Advance, Germany). The microstructure was analyzed by a scanning electron microscope (SEM, Hitachi-S3400N, Hitachi, Tokyo, Japan) and a transmission electron microscope (TEM, Oxford INCAX-ACT, Oxford Instruments, Oxford, UK). The TEM sample of a selected specimen was prepared through conventional mechanical thinning and finished with precision ion polishing system machine (PIPS, Gatan-691, Pleasanton, CA, USA).

3. Results and Discussion

The phase compositions of the sintered B4C/TiB2 ceramic composites with various raw powders are shown in Figure 2. All samples contained B4C, TiB2, SiC, and graphite. The X-ray characteristic peak patterns of the BM30 and BM10S20 were the same. With the introduction of TiB2 nanoparticle powders, the characteristic peaks of TiB2 changed significantly. For the BS20N20, the 2θ = 68.206° characteristic peak of the TiB2 was higher than the characteristic peak intensity of the sample BM30 and BM10S20. The well-defined peaks in the as-prepared B4C/TiB2 composite suggests that the TiB2 phase has a preferred orientation in (102) and (111).
Figure 3 and Figure 4 show the SEM pictures of the fractured surface of the microstructure of the B4C/TiB2 ceramic composites. It could be clearly seen that due to the fact that the BM30 raw material powder particles are coarse and the sintering driving forcing is small, as shown in Figure 3a,b, there were a large number of interconnected open pores, and the coarsened particles were connected in an island chain. A large amount of sinter-necks with clear contours among the grains in the BM30 sample were still visible. With the size of TiB2 powder decreasing, the pore content and pore size decreased, and the dense areas increased significantly, as shown in Figure 3c,f. In the BS20N10 sample containing both 500 nm and 50 nm particle sizes of TiB2 powder, the shapes of the pores were relatively regular, tending to form regular polygon or nearly circular shapes, as shown in Figure 4a,b. Additionally, it can be inferred that these small particles belong to TiB2.
The TiB2 grains on the grain boundaries can pin the movement of the B4C grain boundaries and hinder the grain growth, thus increasing the content of grain boundaries and increasing the sintering rate [29]. In the BS20N10 sample, the interfaces between TiB2 and B4C phases are well distributed, indicating that TiB2 and other phases achieved good wetting during the sintering process as shown in Figure 5. The interface between B4C and TiB2 is jagged, which indicates that the interface feature helps to improve the ability to resist external loads.
The SEM pictures of the polished surfaces of the specimens sintered with various TiB2 powders are shown in Figure 6 and Figure 7. In Figure 6, SiC grains were dispersed and distributed on the B4C substrate in sample BS20N10, which acts as a pinning to prevent the grain boundary and inhibit grain growth. Additionally, the compound reaction of Si and C generated SiC exotherm, which helps the sintering process. Figure 7 show that the average grain sizes of the TiB2 in the prepared specimens with various raw TiB2 particles obtained under pressureless sintering conditions at 2100 °C for 60 min dwell were similar. B4C grains ranged from 2 μm to 10 μm, and comparing with the TiB2 raw powder with an average grain size of 50 nm, the grain sizes of the ceramic composites increased by a maximum of 200 times. With the size of TiB2 raw powder decreasing, the amount and size of the pores in the samples decreased significantly. The B4C average grain size of the BM30 sample to which the 7 μm-sized TiB2 powder was added was 3.01 μm, but many large pores were present in Figure 7a. B4C average grain size of the BM10S20 sample, with the TiB2 addition of 7 μm and 500 nm, was consistent with BM30, but the amount of the pores decreased, and the densification area of the BM10S20 sample increased. B4C average grain size of the BS20N10 sample with the TiB2 addition of the 500 nm and 50 nm remained close to the starting raw powder, about 2.63 μm, and the amount and size of the pores in the BS20N10 sample was significantly reduced. The relative density of the BS20N10 sample was also increased to 98.6%, as shown in Table 2.
Figure 8a–b are TEM images of the interface structure between B4C and SiC in the BS20N10 sample. According to the selected area-electron diffraction (SAD) result in Figure 8b, the unit cell structure parameter of B4C was a = b = 0.56 nm, c = 1.21 nm, α = β = 90°, γ = 120°, and the unit cell structure parameter of SiC was a = b = c = 0.44 nm, α = β = γ = 90°. The zone axis of the two phases of B4C and SiC satisfies the relationship: [ 120 ] B 4 C / / [ 1 ¯ 12 ] S i C , and a group of crystal planes satisfies the relationship: ( 303 ) B 4 C / / ( 311 ) S i C . Additionally the interface between crystal plane ( 303 ) B 4 C and crystal plane ( 311 ) S i C satisfies:
δ = d ( 303 ) d ( 311 ) d ( 303 ) = 0.151 nm 0.137 nm 0.151 nm = 9.27 %
According to the calculation Formula (1), the mismatch degree between the crystal plane ( 303 ) B 4 C and crystal plane ( 311 ) S i C is 9.27%, which could form a semi-coherent interface, and it helps to reduce the interface energy and obtains a bonding strong interface [30].
Figure 9 and Figure 10 are the bright field and high-resolution TEM images of the interfaces between TiB2 and B4C, SiC, respectively. As can be seen in the bright field images of Figure 9a and Figure 10a, the interfaces were clean and straight, and there were no other new phases. There were lattice distortion regions with a wide range of 2~3 nm at the interfaces, which were also the transition regions of the lattice structures between TiB2 and the other two phases, shown in Figure 9b and Figure 10b. The main reason for the formation of these transition zones may be attributed to the unit cell structure parameters of TiB2: a = b = 0.30 nm, c = 0.32 nm, α = β = 90°, γ = 120°. Additionally, the difference of the unit cell structure between B4C, SiC, and TiB2 was huge. During the sintering process, the transition zones were created to coordinate the arrangement of atoms at the interfaces.
The mechanical and physical properties of the prepared ceramic composites with different TiB2 particle sizes are presented in Table 2. With the size of TiB2 raw powder decreasing, the relative density and mechanical properties of the prepared ceramic composites all showed a significantly increasing trend. The relative density of the prepared specimens increased, which helped to achieve the excellent mechanical properties of the prepared specimens. The relative density of the BS20N10 sample reached 98.6%, which is the first major requirement to obtain competitive B4C/TiB2 ceramic composites. The optimized flexural strength, Vickers hardness, and fracture toughness of the BS20N10 sample reached 364 MPa, 30.2 GPa, and 5.47 MPa·m1/2, respectively.
The relative density of the BS20N10 sample was high (98.6%) and the grain sizes were fine (about 2.63 μm), which were mainly due to the following four aspects: (1) the 50 nm-sized TiB2 particles filled the pores of the green body and increased the density of the green body, being conducive to higher densification upon sintering; (2) with the size of the TiB2 powder decreasing, the specific surface energy of the green body was higher than in analogous compositions with coarser grain sizes, which provides a strong driving force for sintering; (3) with the size of TiB2 powder decreasing, the amount of the grain boundary increased, resulting in enhanced grain boundary diffusion during the sintering process; (4) TiB2 grains on the grain boundaries hindered the movement of the grain boundaries and helped to preserve a fine B4C grain size.
The improvement of the relative density and reduction of the grain size of the samples both contributed to obtain high flexural strength. In addition, the shapes of the pores in the BS20N10 were regular polygons or near circles, as shown in Figure 4. According to the fracture mechanics of ceramic materials [31,32,33], these types of the pores could significantly increase the critical value of fracture failure caused by the stress concentration in the sample, and the sample could achieve a high flexural strength.
The thermal expansion coefficients of TiB2 (8.1 × 10−6/°C), B4C (4.5 × 10−6/°C), and SiC (4.7 × 10−6/°C) are quite different [34,35]. During the cooling process, the residual tensile stress fields rise at the interfaces between TiB2 and another phase (such as B4C or SiC). When the crack enters the residual stress field zone, the crack propagated proceeds in the direction perpendicular to the tensile stress as shown in Figure 11, so that the crack propagation directions can be deflected. The crack deflections and the crack propagation paths are extended, which increase the energy consumption and increase the fracture toughness of the prepared ceramic composite.

4. Conclusions

B4C/TiB2 ceramic composites containing different proportions of submicron and nano TiB2 powders were prepared by pressureless sintering at 2100 °C. With the decrease of the particle size of TiB2 raw powders, the surface energy of the powder increased significantly and the density of the sintered body increased. During the sintering process, nano-TiB2 inhibited the grain growth, increased the number of the grain boundaries, and promoted the densification of the material to 98%. With the size of TiB2 powders decreasing, the average grain sizes of the B4C/TiB2 ceramic composites decreased, and the interfaces between the different phases were strongly bonded, which helped to obtain good mechanical properties. As a result, the B4C/TiB2 ceramic composite with 20 wt.% submicron and 10 wt.% nano-TiB2 addition had a significant improved in mechanical and physical properties. The optimized relative density, grain size, Vickers hardness, flexural strength, and fracture toughness of the sample were 98.6%, 2.63 μm, 30.2 GPa, 364 MPa, and 5.47 MPa·m1/2, respectively. Finally, it was illustrated that the sub-fine TiB2 powder could control the grain growth in the preparation of the B4C/TiB2 ceramic composite under the pressureless sintering condition, and was confirmed to be an effective approach to enhance the mechanical properties of B4C ceramics.

Author Contributions

H.N. and Y.Z. conducted experiments and data analysis works, and H.N. and Y.Z. wrote and revised the manuscript; Y.W. and H.C. contributed to the conception of the study; N.Y., D.L., M.T. and J.Z. helped perform the pressureless sintering tests and analysis of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to acknowledge the Beijing Centre for Physical & Chemical Analysis and Fuzhou Qiyue Micro Powder Co., Ltd. for their help in experiment work and analysis technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thevent, F. Boron Carbide—A Comprehensive Review. J. Eur. Ceram. Soc. 1990, 6, 205–225. [Google Scholar] [CrossRef]
  2. Jiang, Z.G.; Zeng, S.Y.; Shen, Z.Q. Research progress on lightweight ceramic composite armor structure. Acta Armamentarii 2010, 31, 603–610. [Google Scholar] [CrossRef]
  3. Karandikar, P.G.; Evans, G.; Wong, S.; Aghajanian, M.K.; Sennett, M. A review of ceramics for armor applications. Adv. Ceram. Armor IV 2009, 29, 163–175. [Google Scholar] [CrossRef]
  4. Savio, S.G.; Ramanjaneyulu, K.; Madhu, V.; Bhat, T.B. An experimental study on ballistic performance of boron carbide tiles. Int. J. Impact Eng. 2011, 38, 535–541. [Google Scholar] [CrossRef]
  5. Da Rocha, R.M.; de Melo, F.C.L. Pressureless Sintering of B4C-SiC Composites for Armor Applications. Ceram. Eng. Sci. Proc. 2009, 30, 113. [Google Scholar] [CrossRef]
  6. Wu, C.; Li, Y.K.; Wan, C.L. Reactive-sintering B4C matrix composite for Armor Applications. Rare Met. 2020, 39, 529–544. [Google Scholar] [CrossRef]
  7. Huang, S.G.; Vanmeensel, K.; van der Biest, O.; Vleugels, J. In situ synthesis and densification of submicormeter-grained B4C-TiB2 composites by pulsed electric current sintering. J. Eur. Ceram. Soc. 2011, 31, 637–644. [Google Scholar] [CrossRef]
  8. Xu, C.M.; Cai, Y.B.; Flodstrom, K.; Li, Z.; Esmaeilzadeh, S.; Zhang, G.J. Spark plasma sintering of B4C ceramics: The effects of milling medium and TiB2 addition. Int. Refract. Met. Hard Mater. 2012, 30, 139–144. [Google Scholar] [CrossRef]
  9. Goldstein, A.; Geffen, Y.; Goldenberg, A. Boron Carbide–Zirconium Boride In-Situ Composites by the Reactive Pressureless Sintering of Boron Carbide–Zirconia Mixtures. J. Am. Ceram. Soc. 2001, 84, 642–644. [Google Scholar] [CrossRef]
  10. Subramanian, C.; Roy, T.; Murthy, T.S.; Sengupta, P.; Kale, G.B.; Krishnaiah, M.V.; Suri, A.K. Effect of zirconia addition on pressureless sintering of boron carbide. Ceram. Int. 2008, 34, 1543–1549. [Google Scholar] [CrossRef]
  11. Sairam, K.; Sonber, J.K.; Murthy, T.S.; Subramanian, C.; Hubli, R.C.; Suri, A.K. Development of B4C–HfB2 composites by reaction hot pressing. Int. Refract. Met. Hard Mater. 2012, 35, 32–40. [Google Scholar] [CrossRef]
  12. Tang, J.; Tan, S.H.; Chen, Z.M.; Jiang, D.L. Strengthening and Toughening of B4C-TiB2 Multiphase Ceramics. J. Inorg. Mater. 1997, 12, 169–174. [Google Scholar] [CrossRef]
  13. Nils, C.; Bernhard, M.; Michael, V.S. Grain-Size dependence of fracture energy in ceramics. J. Am. Ceram. Soc. 1981, 64, 345–350. [Google Scholar] [CrossRef]
  14. Liu, Z.; Wang, D.; Li, J.; Huang, Q.; Ran, S. Densification of high-strength B4C–TiB2 composites fabricated by pulsed electric current sintering of TiC–B mixture. Scr. Mater. 2017, 135, 15–18. [Google Scholar] [CrossRef]
  15. Wang, D.; Ran, S.; Shen, L.; Sun, H.; Huang, Q. Fast synthesis of B4C–TiB2 composite powders by pulsed electric current heating TiC–B mixture. J. Eur. Ceram. Soc. 2015, 35, 1107–1112. [Google Scholar] [CrossRef]
  16. Huang, S.G.; Vanmeensel, K.; Malek, O.J.A.; Van der Biest, O.; Vleugels, J. Microstructure and mechanical properties of pulsed electric current sintered B4C–TiB2 composites. Mater. Sci. Eng. A 2011, 528, 1302–1309. [Google Scholar] [CrossRef]
  17. Wu, C.; Li, Y.K. AlN-induced reinforcement of nano-amorphous B-C-N compound for TiB2–B4C ceramic composite. J. Alloys Compd. 2020, 831, 154074. [Google Scholar] [CrossRef]
  18. Mashhadi, M.; Taheri-Nassaj, E.; Mashhadi, M.; Sglavo, V.M. Pressureless sintering of B4C–TiB2 composites with al additions. Ceram. Int. 2011, 37, 3229–3235. [Google Scholar] [CrossRef]
  19. Malek, O.; Vleugels, J.; Vanmeensel, K.; Huang, S.; Liu, J.; Van den Berghe, S.; Datye, A.; Wu, K.H.; Lauwers, B. Electrical discharge machining of B4C–TiB2 composites. J. Eur. Ceram. Soc. 2011, 31, 2023–2030. [Google Scholar] [CrossRef]
  20. Goldstein, A.; Yeshurun, Y.; Goldenberg, A. B4C/metal boride composites derived from B4C/metal oxide mixtures. J. Eur. Ceram. Soc. 2007, 27, 695–700. [Google Scholar] [CrossRef]
  21. Yamada, S.; Hirao, K.; Yamauchi, Y.; Kanzaki, S. High strength B4C–TiB2 composites fabricated by reaction hot-pressing. J. Eur. Ceram. Soc. 2003, 23, 1123–1130. [Google Scholar] [CrossRef]
  22. Liu, Z.T.; Deng, X.G.; Li, J.M.; Sun, Y.; Ran, S. Effects of B4C particle size on the microstructures and mechanical properties of hot-pressed B4C-TiB2 composites. Ceram. Int. 2018, 44, 21415–21420. [Google Scholar] [CrossRef]
  23. Li, P.; Yang, J.F. Microstructure and Mechanical Properites of B4C-C Ceramic. Rare Met. Mater. Eng. 1999, 28, 151–154. [Google Scholar]
  24. Suri, A.K.; Subramanian, C.; Sonber, J.K.; Murthy, T.C. Synthesis and consolidation of boron carbide: A review. Int. Mater. Rev. 2010, 55, 4–40. [Google Scholar] [CrossRef]
  25. Hayun, S.; Frage, N.; Dariel, M.P. The morphology of ceramic phases in BxC-SiC-Si infiltrated composites. J. Solid State Chem. 2006, 179, 2875–2879. [Google Scholar] [CrossRef]
  26. Feng, Y.; Hou, Z.; Zhang, H.; Liu, L. Densification and Mechanical Properties of Spark Plasma Sintered B4C with Si as a Sintering Aid. J. Am. Ceram. Soc. 2010, 93, 2956–2959. [Google Scholar] [CrossRef]
  27. Zhang, Z.; Du, X.; Li, Z.; Wang, W.; Zhang, J.; Fu, Z. Microstructures and mechanical properties of B4C-SiC intergranular/intragranular nanocomposite ceramics fabricated from B4C, Si, and graphite powders. J. Eur. Ceram. Soc. 2014, 34, 2153–2161. [Google Scholar] [CrossRef]
  28. Skorokhod, V.V., Jr.; Vlajic, M.D.; Kristic, V.D. Pressureless Sintering of B4C-TiB2 Ceramic Composites. Mater. Sci. Forum 1998, 282–283, 219–224. [Google Scholar] [CrossRef]
  29. Baharvandi, H.R.; Hadian, A.M. Pressureless Sintering of TiB2-B4C Ceramic Matrix Composite. J. Mater. Eng. Perform. 2008, 17, 838–841. [Google Scholar] [CrossRef]
  30. Pan, J.S.; Tong, J.M.; Tian, M.B. Fundamentals of Materials Science; Tsinghua University Press: Beijing, China, 2011. [Google Scholar]
  31. Mu, B.C. Strengthening and Toughening of Ceramic Materials; Metallurgical Industry Press Co., Ltd.: Beijing, China, 2002. [Google Scholar]
  32. Choi, S.R.; Sanders, W.A.; Salem, J.A.; Tikare, V. Young’s modulus, strength and fracture toughness as a function of density of in situ toughened silicon nitride with 4 wt.% Scandia. J. Mater. Sci. Lett. 1995, 14, 276–278. [Google Scholar] [CrossRef]
  33. Li, Z.N. Applied Fracture Mechanics; Bei Hang University Press: Beijing, China, 2012. [Google Scholar]
  34. Wu, Q.R.; Wen, B.X. Study on temperature dependence of thermal conductivity and linear expansion for SiC material. J. South China Univ. Techno. Nat. Sci. Ed. 1996, 24, 11–15. [Google Scholar]
  35. Einarsrud, M.A.; Hagen, E.; Pettersen, G.; Grande, T. Pressureless sintering of titanium diboride with nickel, nickel boride, and iron additives. J. Am. Ceram. Soc. 2005, 80, 3013–3020. [Google Scholar] [CrossRef]
Figure 1. Nominal particle size and SEM images of commercial TiB2 raw powders (a) TiB2-7 μm (Micron), (b) TiB2-500 nm (Submicron), (c) TiB2-50 nm (Nano).
Figure 1. Nominal particle size and SEM images of commercial TiB2 raw powders (a) TiB2-7 μm (Micron), (b) TiB2-500 nm (Submicron), (c) TiB2-50 nm (Nano).
Materials 14 05227 g001
Figure 2. XRD patterns of B4C/TiB2 ceramic composites obtained with different grades of TiB2 powders.
Figure 2. XRD patterns of B4C/TiB2 ceramic composites obtained with different grades of TiB2 powders.
Materials 14 05227 g002
Figure 3. Fracture morphology SEM images of the B4C/TiB2 ceramic composites with various TiB2 raw powder mixtures. (a,b) BM30, (c,d) BM10S20 and (e,f) BS20N10.
Figure 3. Fracture morphology SEM images of the B4C/TiB2 ceramic composites with various TiB2 raw powder mixtures. (a,b) BM30, (c,d) BM10S20 and (e,f) BS20N10.
Materials 14 05227 g003
Figure 4. Small TiB2 grains in the B4C/TiB2 ceramic composite processed with a mixture of sub-micron and nano-sized TiB2 particles. (a) SEM image and (b) corresponding BSE image.
Figure 4. Small TiB2 grains in the B4C/TiB2 ceramic composite processed with a mixture of sub-micron and nano-sized TiB2 particles. (a) SEM image and (b) corresponding BSE image.
Materials 14 05227 g004
Figure 5. Interface between B4C and TiB2. (a) SEM image and (b) corresponding BSE image.
Figure 5. Interface between B4C and TiB2. (a) SEM image and (b) corresponding BSE image.
Materials 14 05227 g005
Figure 6. SEM images of the polished surfaces (a) and corresponding BSE images (b) in the sample.
Figure 6. SEM images of the polished surfaces (a) and corresponding BSE images (b) in the sample.
Materials 14 05227 g006
Figure 7. (ac) SEM images of the polished surfaces and (d) plot of the average grain sizes in the samples.
Figure 7. (ac) SEM images of the polished surfaces and (d) plot of the average grain sizes in the samples.
Materials 14 05227 g007
Figure 8. Interface between B4C and SiC. (a) Bright field TEM image, (b) SAED pattern.
Figure 8. Interface between B4C and SiC. (a) Bright field TEM image, (b) SAED pattern.
Materials 14 05227 g008
Figure 9. Interface structure between B4C and TiB2. (a) Bright field TEM image, (b) HRTEM image.
Figure 9. Interface structure between B4C and TiB2. (a) Bright field TEM image, (b) HRTEM image.
Materials 14 05227 g009
Figure 10. Interface structure between SiC and TiB2. (a) Bright field TEM image, (b) HRTEM image.
Figure 10. Interface structure between SiC and TiB2. (a) Bright field TEM image, (b) HRTEM image.
Materials 14 05227 g010
Figure 11. Schematic diagrams of the toughing mechanism by residual stress.
Figure 11. Schematic diagrams of the toughing mechanism by residual stress.
Materials 14 05227 g011
Table 1. Starting composition of BM30, BM10S20, and BS20N10 ceramic composites.
Table 1. Starting composition of BM30, BM10S20, and BS20N10 ceramic composites.
GradeB4C (wt.%)C (wt.%)Si (wt.%)Micron TiB2 (wt.%)Submicron TiB2 (wt.%)Nano TiB2 (wt.%)
BM30607330//
BM10S2060731020/
BS20N106073/2010
Table 2. Mechanical properties of the B4C/TiB2 ceramic composites.
Table 2. Mechanical properties of the B4C/TiB2 ceramic composites.
SampleRelative Density (%)Flexural Strength (MPa)Fracture Toughness (MPa·m1/2)Vickers Hardness (GPa)
BM3090.1 ± 0.2217 ± 133.70 ± 0.198.3 ± 0.6
BM10S2092.6 ± 0.1288 ± 124.46 ± 0.1212.5 ± 1.1
BS20N1098.6 ± 0.1364 ± 95.47 ± 0.1230.2 ± 2.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Niu, H.; Zhu, Y.; You, N.; Wang, Y.; Cheng, H.; Luo, D.; Tang, M.; Zhang, J. Effects of TiB2 Particles on the Microstructure Evolution and Mechanical Properties of B4C/TiB2 Ceramic Composite. Materials 2021, 14, 5227. https://doi.org/10.3390/ma14185227

AMA Style

Niu H, Zhu Y, You N, Wang Y, Cheng H, Luo D, Tang M, Zhang J. Effects of TiB2 Particles on the Microstructure Evolution and Mechanical Properties of B4C/TiB2 Ceramic Composite. Materials. 2021; 14(18):5227. https://doi.org/10.3390/ma14185227

Chicago/Turabian Style

Niu, Haiyan, Yu Zhu, Ning You, Yangwei Wang, Huanwu Cheng, Dujun Luo, Mengying Tang, and Jiamin Zhang. 2021. "Effects of TiB2 Particles on the Microstructure Evolution and Mechanical Properties of B4C/TiB2 Ceramic Composite" Materials 14, no. 18: 5227. https://doi.org/10.3390/ma14185227

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