Next Article in Journal
Molecular Dynamics Study on Crack Propagation in Al Containing Mg–Si Clusters Formed during Natural Aging
Next Article in Special Issue
Improved Mechanical Properties of Alumina Ceramics Using Plasma-Assisted Milling Technique
Previous Article in Journal
Temperature during Repetitive Short-Term Operation of a Brake with Functionally Graded Friction Element
Previous Article in Special Issue
Corrosion Behavior of Cobalt Oxide and Lithium Carbonate on Mullite–Cordierite Saggar Used for Lithium Battery Cathode Material Sintering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement of Mechanical Properties of Composites with Surface Modified B4C for Precision Machining

The State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan 430081, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(2), 882; https://doi.org/10.3390/ma16020882
Submission received: 15 December 2022 / Revised: 7 January 2023 / Accepted: 13 January 2023 / Published: 16 January 2023

Abstract

:
In order to solve the problem of difficult sintering and high brittleness of B4C-based ceramics, B4C@ZrB2-TiB2 composite powder was synthesized by molten salt method, and B4C–(Zr, Ti)B2 composite ceramics were successfully prepared by spark plasma sintering. The effects of different raw material ratios on the composition, microstructure, and mechanical properties of the prepared composite ceramics were characterized by XRD, XPS, SEM, and TEM. The results show that ZrB2 and TiB2 were grown on the surface of B4C by template mechanism to form a dense nanocrystalline coating, and the original surface of B4C was exposed gradually with the decrease of the ratio of metal powder. When the composite powders were sintered at 1700 °C, ZrB2 and TiB2 formed a solid solution, which can refine grains and improve strength. When the raw material ratio is n(B4C): n(Zr): n(Ti) = 12:1:1, the composite ceramics have excellent comprehensive properties, the Vickers hardness reaches 41.2 GPa.

1. Introduction

Given the higher grinding requirements of rough wafers grinding in the integrated circuit field, it is urgent to develop ultra-hard ceramics with excellent mechanical properties to meet the new growth requirements. The hardness of boron carbide ceramics is third only to diamond and cubic boron nitride, and the lower cost makes it have great application potential. Boron carbide has the characteristics of high hardness, high melting point, high modulus of elasticity, low thermal expansion coefficient and good chemical stability [1,2], and is widely used in mechanical equipment, abrasive abrasives, microwave absorption, and refractory antioxidants and other fields [3,4,5]. Zamora et al. [6] prepared dense boron carbide composites at low temperatures and they showed good wear resistance with only slight wear after linear sliding over long distances. However, the lower sintering temperature and the introduction of more low-hardness phases make it difficult to meet the grinding needs of rough wafers. The mechanisms of porosity elimination, grain boundaries, and volume diffusion of ceramics only play a role when the sintering temperature reaches more than 2000 °C This ultimately makes it difficult to achieve the sintering densification of B4C ceramics with poor plasticity (KIC is about 2 MPa∙m1/2), and its ultra-hard properties are difficult to fully achieve [7,8].
In order to achieve the densification and sintering of B4C-based ceramics, various sintering aids and second phases were introduced to reduce the sintering temperature. It has been found that carbides and borides such as SiC, TiC, HfC, TaB2, TiB2, and ZrB2 benefit the performance improvement of composite ceramics [9,10,11]. Chen et al. [12] found that ZrB2–ZrC–B4C composites have been successfully fabricated by SPRS from B4C and different content of Zr, while the flexure strength and hardness decrease first and then increase with increasing Zr content. It is found that boron carbide containing 66% Zr has the highest fracture toughness (5.83 MPa·m1/2) and flexure strength (386.45 MPa) due to the uniform distribution of ZrB2 and ZrC. Sha et al. [13] explored the effect of carbon content on the mechanical properties of reaction-sintered B4C composite ceramics. When the carbon content is 10%, the composite ceramics have the highest bending strength (444 MPa) and elastic modulus (329 GPa). After adding an appropriate proportion of carbon and silicon in the preparation of B4C/TiB2 composite ceramics, Zhu et al. [14] found that layered graphite appeared at the grain boundary of B4C and it could react with B2O3 to cause volume shrinkage, which significantly improved the mechanical properties of the material. Yan et al. [15] introduced TiSiC2 additive to prepare B4C-TiB2 composite ceramics. Compared with the pure B4C ceramics prepared by the same method, the hardness of the composite ceramics decreased, but the flexural strength and fracture toughness increased significantly. Ren et al. [16,17] used NaCl–KCl as a molten salt medium to prepare the composite powder with Al3BC and TiB2 coating uniformly on the surface of B4C particles. Due to the conductivity of boride coating, the material transportation is dynamically enhanced through SPS sintering.
Spark plasma sintering (SPS) is a new rapid densification technology that uses high-frequency and high currents through the sample to heat the sample and apply mechanical pressure [18,19,20,21]. Yavas et al. [22] found that SPS sintering has higher surface energy than B4C powder with small particle size (HS) at heating rate of 75, 150 and 225 °C·min−1, which improves the sintering driving force and enables the sintering of B4C ceramics to be completed at 1590 °C. Moshtaghioun et al. [23] combined high-energy ball milling, annealing treatment and SPS to provide sinterability for ultrafine B4C powder, and the optimal sintering conditions were 100 °C∙min−1 temperature to 1700 °C for 3 min, which could prepare B4C ceramics with ultra-fine grains with a relative density of more than 98.5%. At the same time, due to the significant decrease of grain size and the increase of transcrystalline fracture mode, the ceramic has an ultra-high hardness of 38 GPa without corresponding ductile loss (~3 MPa∙m1/2). Experiments show that the two-step SPS treatment enables B4C powder to be densified while retaining nanoscale particles, with ultra-high hardness and good toughness [24].
Although many studies have improved the performance of B4C ceramics to varying degrees, there are still problems such as high sintering temperature, poor fracture toughness and lower hardness. In this paper, B4C–(Zr, Ti)B2 composite ceramic with high hardness was prepared using the SPS and using (Zr, Ti)B2–coated B4C powder prepared by molten salt as raw materials. The effects of the ratio of raw materials on the phase composition and microstructure of composite powders and ceramics were discussed, and the hardness of the ceramics was compared.

2. Materials and Methods

2.1. Materials

The composite powder for preparing composite ceramics was synthesized by molten salt method using B4C (Mudanjiang Diamond Co., Ltd., W1.5, Mudanjiang, China), ZrH2 (ST-NANO, 0.5 μm, Shanghai, China), TiH2 (ST-NANO, 0.5 μm, Shanghai, China) as raw materials, and NaCl (Sinopharm Chemical ReagentCo., Ltd, AR, Shanghai, China), KCl (Macklin, AR, Shanghai, China) as a molten salt medium. First, mixed the raw materials in the mortar for about 25 min (according to the mass ratio of m (mixed salt):m (mixed powders) = 7:3 and m (NaCl):m (KCl) = 3:2. ZrH2 and TiH2 will decompose and precipitate H2 at low temperatures to obtain metal Zr and Ti [25]. The actual reaction occurs between Zr/Ti and B4C. Due to the extremely low H content and the low addition of raw materials in each group of mixed powders, the ratio was calculated according to the molar mass of Zr and Ti. The molar ratio of B4C to Zr/Ti are 8:1:1, 12:1:1 and 16:1:1, named ZT8, ZT12, and ZT16, respectively. Put the mixture into the alumina crucible with a lid, then the crucible containing the mixture was placed in an atmosphere furnace and held to 1100 °C for 2 h under Ar atmosphere. The solid obtained after heat treatment was washed with deionized water after ultrasonic cleaning of 15 min in a water bath at 45 °C (these operations should be repeated three times). Lastly, put the products at 110 °C and dried for 12 h to obtain composite powder.
Accurately weighed composite powder was placed in a special cylindrical graphite crucible (Φ20 mm), and the crucible containing the powder was placed in an SPS sintering furnace for sintering (the inner wall of the crucible, and the composite powder were isolated with clean graphite paper, and the powder was compacted and sealed with a graphite plunger at the upper and lower ports). The sintering system of this experiment was: holding at 1400 °C for 3 min, holding at 1700 °C for 6 min, the heating rate was 100 °C/min, and the sintering pressure was 50 MPa. Pressure was increased gradually during heating and reached maximum pressure before maximum temperature. After the heat preservation, the pressure was gradually removed and the samples demoulded.
After removing the graphite paper from the ceramic samples, the ceramic samples were ground, polished, then characterized and tested for the performance of the samples.

2.2. Characterization and Testing

The phase composition and microstructure of the composite powders and the composite ceramics were characterized using the X-ray diffractometer (XRD, X’Pert Pro, Philips, The Netherlands), field emission scanning electron microscope (SEM, Nova nano 400, FEI, Hillsboro, OR, USA) equipped with an energy dispersive spectrometer (EDS, IE350 Penta FET X-3, Oxford, UK) and high resolution transmission electron microscope (TEM, JEM2100, JEOL, Tokyo, Japan). The element composition and chemical bond bonding state on the surface of the composite powder was analyzed by an X-ray photoelectron spectrometer (XPS, AXIS SUPRA+, Shimadzu-Kratos, Hadano, Kanagawa, Japan). The Vickers hardness of the ceramic samples was measured by an automatic micro/macro hardness tester, the applied load was 0.5 kg and the loading time was 15 s. The indentation diagram is shown in Figure 1.

3. Results and Discussion

3.1. B4C@ZrB2–TiB2 Composite Powder

Figure 2 shows the XRD patterns of the composite powder prepared at 1100 °C. It can be seen from Figure 2 that the main phase of the composite powder is composed of B4C, ZrB2, and a small amount of ZrO2. The oxidation of metal zirconium powder and the reaction of B2O3 on the surface of B4C with Zr powder may cause the diffraction peak of ZrO2. Meanwhile, only a weak diffraction peak of TiB2 was detected in the composite powder, which might be the dissolution of TiB2 in ZrB2 [26], and the significant atomic number of Zr leads to the reduction of X–ray scattering factor of TiB2 [27].
Figure 3 shows the SEM images of the composite powder prepared at 1100 °C. Figure 3a,b show the morphology of the ZT8 composite powder and its elemental mapping analysis. On the one hand, it confirms the generation of TiB2 and ZrB2 on the surface of B4C particles after heat treatment at 1100 °C in molten salt. On the other hand, it illustrates the homogeneous encapsulation of the two phases on the B4C particles. Figure 3c,d show the ZT12 and ZT16 composite powder morphology. The composite powder particles maintain the original shape of the irregular polyhedra of the B4C particles, and the original surface is gradually exposed with a decrease in the proportion of metal powders.
To further verify the formation of ZrB2 and TiB2 on the surface of B4C particles, the chemical composition of ZT8 composite powder was analyzed by XPS. Figure 4a shows the pattern of B1s, which is divided into two peaks at 187.45 eV and 192.55 eV. Since the energies of the B–Zr bond [28,29] of ZrB2 and the B–Ti bond [30] of TiB2 are similar, the superposition of peaks may occur at 187.45 eV. The 192.55 eV corresponds to the B–O bond of B2O3 [31], which is the oxidation on the surface of the composite powder. The pattern of Ti2p is divided into four peaks, which are two peaks generated by the energy level splitting of the Ti–B bond and Ti–O bond in Figure 4b. The Ti2p3/2 sub–peak at 454.38 eV and the Ti2p1/2 sub–peak at 458.95 eV correspond to the Ti–B bond of TiB2 [32], which proves the existence of TiB2 and the coating of the surface of B4C particle. The Ti2p3/2 sub–peak at 459.67 eV and Ti2p1/2 sub–peak at 464.86 eV correspond to the Ti–O bond in TiO2 [33], due to the oxidation on the surface of the composite powder.

3.2. B4C–(Zr, Ti)B2 Composite Ceramics

Figure 5 shows the XRD patterns of the composite ceramics sintering at 1700 °C. It can be seen from the picture that the primary phases are B4C (PDF #86-1024), (Zr, Ti)B2(PDF #89-3924), and C(PDF #26-1079). The diffraction peaks of ZrB2 (PDF #34-0423) shifts to high-angle azimuth (Figure 5b). As a result of the small size of Ti atoms, the solid solution replaces the Zr atoms into the lattice of ZrB2, causing distortion and compressive stress, resulting in the reduction of cell parameters (0.3100 nm) [27]. The diffraction peaks of ZrO2 disappear after high–temperature sintering. The lowest temperatures required for the complete conversion of Zr and Ti oxides into diborides were 1620 °C and 1690 °C under 0.8 CO ambient pressure. Furthermore, excessive B4C and lower CO partial pressure will further reduce the reaction temperature [34]. Asl et al. [35] found that ZrO2 will react with B4C to form a delicate ZrB2 connecting phase until it completely reacts in the subsequent heating process at 1281 °C.
Figure 6 shows the ternary phase diagram of the B4C–Zr–Ti system at 1700 °C obtained by FactSage calculations. When the B4C content is high, the region above the phase line L is a stable four–phase composition of B4C, ZrB2, TiB2, and C at 1700 °C. Combined with XRD analysis, ZrB2 and TiB2 will generate (Zr, Ti)B2 solid solution after sintering at 1700 °C.
Figure 7 shows the BSE images of the B4C–(Zr, Ti)B2 composite ceramics with different raw material ratios. Due to the larger atomic number of Zr and Ti atoms show a brighter white lining in the backscattering mode compared to B and C atoms. Combining the XRD analysis in Figure 5 and the EDS analysis (in Figure 7b,c), it can be found that the bright white phases in Figure 7 are (Zr, Ti)B2. Figure 7a,b show the BSE image of the ZT8 sample and mapping analysis of the rectangular area. Zr and Ti elements are concentrated and evenly distributed in the bright white area. Furthermore, the same distribution interval of Zr and Ti elements and the shift of the XRD diffraction peak (Figure 5) confirms the generation of the solid solution of ZrB2 and TiB2. Figure 7c,d are the BSE plots of ZT12 and ZT16 samples, respectively. With the decrease of the additional amount of Zr and Ti, the distribution of (Zr, Ti)B2 in the B4C matrix was gradually sparse and dispersed.
Figure 8 shows the TEM image of ZT8 composite ceramic at 1700 °C. Figure 8b shows the elemental mapping analysis of Figure 8a. The surface of the ceramic sample shows the aggregation of B, Zr and Ti. Further illustration of (Zr, Ti)B2 solid solution formation and good bonding with B4C grains. The segregation of the C elements in the grain is mainly due to the influence of carbon supporting film.
Figure 9 shows the Vickers hardness and the indentation morphology of B4C–(Zr, Ti)B2 composite ceramics. It can be seen from Figure 9a that the Vickers hardness of the sample increases at first and then decreases, when n(B4C):n(Zr):n(Ti) = 12:1:1, the Vickers hardness reaches 41.2 GPa. The clear diamond-shaped indentations and extension cracks can be observed in Figure 9b. Combined with the BSE images (Figure 7), when the amount of Zr/Ti is reduced, the generation of (Zr, Ti)B2 is diminished, weakening the pinning effect on the B4C matrix—resulting in the coarsening of B4C grains, the reduction of the grain boundaries, and the reduction of the ability to resist local deformation. Ti2+ solid solution into the ZrB2 lattice will generate compression stress to enhance the internal energy. The active (Zr, Ti)B2 grains formation will block the movement of the crystal world in the sintered process of the SPS, reducing the grain size, which plays the role of good crystal reinforcement, because internal stress will be generated around the grains when the solid solution is formed. Compared with the mechanical properties data of B4C composite ceramics reported in references (Listed in Table 1), such as higher temperature and longer holding time, this work realizes the preparation of high hardness B4C ceramics at low temperatures. The Vickers hardness is better than the data reported in the references due to the solution strengthening effect of ZrB2 and TiB2.
Through the XRD phase analysis and microstructure of the composite powder and composite ceramics, and the ternary phase diagram of the B4C–Zr–Ti system at 1700 °C, the reaction mechanism of composite powder and ceramic was been studied: The mass transfer of Zr and Ti could be promoted because of the existence of a molten salt medium, Zr and Ti would react with B4C to form ZrB2 and TiB2 at high temperature and the product would coat the B4C surface. Thus, the composite powder exhibited a core–shell structure. When the composite powder was sintered by high temperature and pulse activation into composite ceramics, ZrB2 and TiB2 of the powder surface would produce a solid solution (Zr, Ti)B2). These (Zr, Ti)B2 solid solutions would prevent the movement of crystal boundaries during SPS sintering, thereby reducing the grain size of the composite ceramics. The reaction mechanism is shown in Figure 10.

4. Conclusions

B4C–(Zr, Ti)B2 composite ceramics were obtained by SPS sintering of the composite powder synthesized by the molten salt method, and its microstructure and mechanical properties were studied. The following conclusions were drawn.
(1)
Thermodynamic calculation and phase analysis revealed that Zr and Ti reacted with B4C to form ZrB2 and TiB2 in a B4C-rich environment. Thanks to the heat and mass transfer of molten salt medium, the product is uniformly coated on the surface of B4C particles.
(2)
When the powder is sintered by SPS at 1700 ℃, ZrB2 and TiB2 form solid solution (Zr, Ti)B2 at high temperature, the primary phases of composite ceramics are B4C, (Zr, Ti)B2, and C.
(3)
Due to the synergistic effect of solid solution strengthening and particle toughening, better comprehensive properties can be obtained. When n(B4C):n(Zr):n(Ti) = 12:1:1, the Vickers hardness reach 41.2 GPa.

Author Contributions

J.D., X.L. and C.D.: conceptualization, methodology, investigation; J.D., J.W., H.Y., Z.L., C.Y., X.L., C.D. and H.Z.: data curation, writing—original draft preparation; J.D., X.L.: supervision; J.D. and X.L.: writing—reviewing and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support from the project supported by the Natural Science Foundation of Hubei Province (Grant Nos. 2020BAA028 and 2020CFA038) and the National Natural Science Foundation of China (Grant Nos. 51972242 and U20A20239).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available on request due to restrictions privacy. The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Johnson, W.B.; Claar, T.D.; Schiroky, G.H. Preparation and processing of platelet-reinforced ceramics by the directed reaction of zirconium with boron carbide. Ceram. Eng. Sci. Proc. 1989, 10, 588–598. [Google Scholar]
  2. Johnson, W.B.; Claar, T.D.; Schiroky, G.H. Microstructure and properties of platelet-reinforced ceramics by the directed reaction of zirconium with boron carbide. Ceram. Eng. Sci. Proc. 1989, 10, 599–609. [Google Scholar]
  3. Johnson, W.B.; Nagelberg, A.S.; Breval, E. Kinetics of formation of a platelet-reinforced ceramic composite prepared by the directed reaction of zirconium with boron carbide. J. Am. Ceram. Soc. 2010, 74, 2093–2101. [Google Scholar] [CrossRef]
  4. Guo, W.M.; Wu, L.X.; You, Y.; Lin, H.T.; Zhang, G.J. Three-step reactive hot pressing of B4C-ZrB2 ceramics. J. Eur. Ceram. Soc. 2016, 36, 951–957. [Google Scholar] [CrossRef]
  5. Rehman, S.S.; Ji, W.; Fu, Z.Y.; Wang, W.M.; Wang, H.; Asif, M.; Zhang, J.Y. In situ synthesis and sintering of B4C/ZrB2 composites from B4C and ZrH2 mixtures by spark plasma sintering. J. Eur. Ceram. Soc. 2015, 35, 1139–1145. [Google Scholar] [CrossRef]
  6. Zamora, V.; Guiberteau, F.; Borrero-López, O.; Ortiz, A.L. Ultra-low temperature spark plasma sintering of super wear-resistant hard B4C composites. Scripta Mater. 2022, 211, 114516. [Google Scholar] [CrossRef]
  7. Zhang, W.K.; Gao, L.Z.; Lei, Y.; Yang, B.J.; Li, J.; Xiao, L.; Yin, Y.S. TiAl/B4C composite fabricated by high energy ball milling and hot press sintering processes and its mechanical properties. Mater. Sci. Eng. 2010, 527, 7436–7441. [Google Scholar] [CrossRef]
  8. Sun, J.L.; Liu, C.X.; Wang, R.G. Low pressure hot pressing of B4C matrix ceramic composites improved by Al2O3 and TiC additives. Mater. Sci. Eng. A 2009, 519, 27–31. [Google Scholar] [CrossRef]
  9. Guo, W.C.; He, Q.L.; Wang, A.Y.; Tian, T.; Liu, C.; Hu, L.X.; Wang, W.M.; Wang, H.; Fu, Z.Y. Effect of TiB2 particles on microstructure and mechanical properties of B4C–TiB2 ceramics prepared by hot pressing. Ceram. Int. 2022, 48, 25637–25641. [Google Scholar] [CrossRef]
  10. Feng, B.; Martin, H.-P.; Michaelis, A. Preparation and Characterization of B4C-HfB2 Composites as Material for High-Temperature Thermocouples. Crystals 2022, 12, 621. [Google Scholar] [CrossRef]
  11. Liu, G.; Chen, S.; Zhao, Y.; Fu, Y.; Wang, Y. The Effects of Transition Metal Oxides (Me = Ti, Zr, Nb, and Ta) on the Mechanical Properties and Interfaces of B4C Ceramics Fabricated via Pressureless Sintering. Coatings 2020, 10, 1253. [Google Scholar] [CrossRef]
  12. Chen, H.; Zeng, F.H.; Li, W.J.; Liu, J.A.; Gu, Y.; Zhang, F.Q. Densification behavior and mechanical properties of spark plasma reaction sintered ZrB2-ZrC-B4C ceramics from B4C–Zr system. Ceram. Int. 2019, 45, 12122–12129. [Google Scholar] [CrossRef]
  13. Sha, W.H.; Liu, Y.Y.; Zhou, Y.B.; Huang, Y.H.; Huang, Z.R. Effect of Carbon Content on Mechanical Properties of Boron Carbide Ceramics Composites Prepared by Reaction Sintering. Materials 2022, 15, 6028. [Google Scholar] [CrossRef] [PubMed]
  14. Zhu, Y.; Cheng, H.W.; Wang, Y.W.; An, R. Effects of carbon and silicon on microstructure and mechanical properties of pressureless sintered B4C/TiB2 composites. J. Alloys Compd. 2019, 772, 537–545. [Google Scholar] [CrossRef]
  15. Yan, X.H.; Zhou, X.G.; Wang, H.L. Effect of Additive Ti3SiC2 Content on the Mechanical Properties of B4C–TiB2 Composites Ceramics Sintered by Spark Plasma Sintering. Materials 2020, 13, 4616. [Google Scholar] [CrossRef]
  16. Ren, D.L.; Deng, Q.H.; Wang, J.; Li, Y.B.; Li, M.; Ran, S.L.; Du, S.Y.; 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]
  17. Wang, J.; Ren, D.L.; Chen, L.L.; Man, G.A.; Zhang, H.Y.; Zhang, H.P.; Luo, L.H.; Li, W.P.; Pan, Y.B.; Gao, P.F.; et al. Initial investigation of B4C–TiB2 composites as neutron absorption material for nuclear reactors. J. Nucl. Mater. 2020, 539, 152275. [Google Scholar] [CrossRef]
  18. Zamora, V.; Nygren, M.; Guiberteau, F.; Ortiz, A.L. Effect of graphite addition on the spark-plasma sinterability of ZrB2 and ZrB2-SiC ultra-high-temperature ceramics. Ceram. Int. 2014, 40, 11457–11464. [Google Scholar] [CrossRef]
  19. Wang, S.; Gao, S.B.; Xing, P.F.; Nie, D.; Yan, S.; Zhuang, Y.X. Pressureless liquid-phase sintering of B4C with MoSi2 as a sintering aid. Ceram. Int. 2019, 45, 13502–13508. [Google Scholar] [CrossRef]
  20. Wang, S.; Deng, Y.Y.; Yang, M.S.; Wang, L.Y.; Li, H.Q.; Xing, P.F. Microstructure and mechanical property of B4C-SiC-CrB2 composites fabricated via reactive hot pressing. Ceram. Int. 2020, 46, 29261–29270. [Google Scholar] [CrossRef]
  21. Song, Q.; Zhang, Z.H.; Hu, Z.Y.; Yin, S.P.; Wang, H.; Ma, Z.W. Microstructure and mechanical properties of super-hard B4C ceramic fabricated by spark plasma sintering with (Ti3SiC2+Si) as sintering aid. Ceram. Int. 2019, 45, 8790–8797. [Google Scholar] [CrossRef]
  22. Yavas, B.; Sahin, F.; Yucel, O.; Goller, G. Effect of particle size, heating rate and CNT addition on densification, microstructure and mechanical properties of B4C ceramics. Ceram. Int. 2015, 41, 8936–8944. [Google Scholar] [CrossRef]
  23. Moshtaghioun, B.M.; Cumbrera, F.L.; Ortiz, A.L.; Casitllo-Rodriguez, M.; Gomez-Garcia, D. Additive-free superhard B4C with ultrafine-grained dense microstructures. J. Eur. Ceram. Soc. 2014, 34, 841–848. [Google Scholar] [CrossRef] [Green Version]
  24. Moshtaghioun, B.M.; Ortiz, A.L.; Gomez-Garcia, D.; Dominguez-Rodriguez, A. Densification of B4C nanopowder with nanograin retention by spark-plasma sintering. J. Eur. Ceram. Soc. 2015, 35, 1991–1998. [Google Scholar] [CrossRef]
  25. Zeppelin, F.; Hirscher, M.; Stanzick, H.; Banhart, J. Desorption of hydrogen from blowing agents used for foaming metals. Compos. Sci. Technol. 2003, 63, 2293–2300. [Google Scholar]
  26. Namini, A.S.; Delbari, S.A.; Asl, M.S.; Le, Q.V.; Shokouhimehr, M. Characterization of reactive spark plasma sintered (Zr, Ti)B2–ZrC–SiC composites. J. Taiwan Inst. Chem. Eng. 2021, 119, 187–195. [Google Scholar] [CrossRef]
  27. Akarsu, M.K.; Akin, I. Mechanical properties and oxidation behavior of spark plasma sintered (Zr, Ti)B2 ceramics with graphene nanoplatelets. Ceram. Int. 2020, 46, 26109–26120. [Google Scholar] [CrossRef]
  28. Jin, H.; Meng, S.H.; Zhang, X.H.; Zeng, Q.X.; Niu, J.H. Effects of oxygen partial pressure on the oxidation of ZrB2–SiC–graphite composites at 1800 °C. Ceram. Int. 2016, 42, 6480–6486. [Google Scholar] [CrossRef]
  29. Ga, X.H.; Qiu, X.L.; Li, X.T.; Theiss, W.; Chen, B.H.; Guo, H.X.; Zhao, T.H.; Liu, G. Structure, thermal stability and optical simulation of ZrB2 based spectrally selective solar absorber coatings. Sol. Energy Mater. Sol. Cells 2019, 193, 178–183. [Google Scholar] [CrossRef]
  30. Ding, J.C.; Lee, D.; Mei, H.J.; Zhang, T.F.; Kang, M.C.; Wang, Q.M.; Kim, K.H. Influence of Si addition on structure and properties of TiB2–Si nanocomposite coatings deposited by high–power impulse magnetron sputtering. Ceram. Int. 2019, 45, 6363–6372. [Google Scholar] [CrossRef]
  31. Yuan, J.Y.; Zhang, Z.Z.; Yang, M.M.; Guo, F.; Men, X.H.; Liu, W.M. TiB2 reinforced hybrid–fabric composites with enhanced thermal and mechanical properties for high–temperature tribological applications. Tribol. Int. 2017, 115, 8–17. [Google Scholar] [CrossRef]
  32. Xia, M.; Feng, Y.; Tian, P.; Song, Z.K.; Zhao, L.; Cai, C.Y. Quantitative Analysis of TiB2 Particles and Properties of Cu-TiB2 Composite Prepared by in Situ Reaction. Rare Metal Mat. Eng. 2017, 46, 3260–3266. [Google Scholar]
  33. Bakhit, B.; Palisaitis, J.; Thörnberg, J.; Rosen, J.; .Persson, P.Å.; Hultman, L.; Petrov, L.; .Greene, J.E.; Greczynski, G. Improving the high-temperature oxidation resistance of TiB2 thin films by alloying with Al. Acta Mater. 2020, 196, 677–689. [Google Scholar] [CrossRef]
  34. Shestakov, V.A.; Gudyma, T.S.; Krutskii, Y.L.; Uvarov, N.F. Determination of the optimal temperature range for synthesis of B4C–TiB2 and B4C–ZrB2 powder composite materials. Mater. Today 2020, 31, 506–508. [Google Scholar] [CrossRef]
  35. Asl, M.S.; Kakroudi, M.G.; Nayebi, B. A fractographical approach to the sintering process in porous ZrB2–B4C binary composites. Ceram. Int. 2015, 41, 379–387. [Google Scholar]
  36. 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. [Google Scholar] [CrossRef]
  37. Lin, X.; Ai, S.H.; Feng, Y.R.; Gao, D.Z.; Guo, X.; Liu, Y.; Xie, B.Y.; Gong, H.Y.; Zhang, Y.J. Fabrication and properties of in–situ pressureless–sintered ZrB2/B4C composites. Ceram. Int. 2017, 43, 15593–15596. [Google Scholar] [CrossRef]
  38. He, R.J.; Jing, L.; Qu, Z.L.; Zhou, Z.L.; Ai, S.G.; Kai, W. Effects of ZrB2 contents on the mechanical properties and thermal shock resistance of B4C–ZrB2 ceramics. Mater. Des. 2015, 71, 56–61. [Google Scholar] [CrossRef]
  39. He, R.J.; Zhou, Z.L.; Qu, Z.L.; Cheng, X.M. High temperature flexural strength and oxidation be-havior of hot–pressed B4C–ZrB2 ceramics with various ZrB2 contents at 1000–1600 °C in air. Int. Refract. Met. H 2016, 57, 125–133. [Google Scholar] [CrossRef] [Green Version]
  40. Huang, S.G.; Vanmeensel, K.; Vleugels, J. Powder synthesis and densification of ultrafine B4C–ZrB2 composite by pulsed electrical current sintering. J. Eur. Ceram. Soc. 2014, 34, 1923–1933. [Google Scholar] [CrossRef]
Figure 1. Indentation diagram.
Figure 1. Indentation diagram.
Materials 16 00882 g001
Figure 2. XRD patterns of the composite powder prepared at 1100 °C.
Figure 2. XRD patterns of the composite powder prepared at 1100 °C.
Materials 16 00882 g002
Figure 3. SEM images of the powder prepared at 1100 °C: (a) ZT8; (b) elemental mapping analysis in (a); (c) ZT12; (d) ZT16.
Figure 3. SEM images of the powder prepared at 1100 °C: (a) ZT8; (b) elemental mapping analysis in (a); (c) ZT12; (d) ZT16.
Materials 16 00882 g003
Figure 4. XPS patterns of the ZT8 powder prepared at 1100 °C: (a) B1s; (b) Ti2p.
Figure 4. XPS patterns of the ZT8 powder prepared at 1100 °C: (a) B1s; (b) Ti2p.
Materials 16 00882 g004
Figure 5. XRD patterns of the composite ceramics: (a) different samples; (b) enlarge image.
Figure 5. XRD patterns of the composite ceramics: (a) different samples; (b) enlarge image.
Materials 16 00882 g005
Figure 6. The ternary phase diagram of the B4C–Zr–Ti system at 1700 °C.
Figure 6. The ternary phase diagram of the B4C–Zr–Ti system at 1700 °C.
Materials 16 00882 g006
Figure 7. BSE images of the B4C–(Zr, Ti)B2 composite ceramics: (a) ZT8; (b) elemental mapping analysis of rectangular area in (a); (c) ZT12; (d) ZT16.
Figure 7. BSE images of the B4C–(Zr, Ti)B2 composite ceramics: (a) ZT8; (b) elemental mapping analysis of rectangular area in (a); (c) ZT12; (d) ZT16.
Materials 16 00882 g007
Figure 8. TEM images of composite ceramic: (a) ZT8; (b) elemental mapping analysis of (a).
Figure 8. TEM images of composite ceramic: (a) ZT8; (b) elemental mapping analysis of (a).
Materials 16 00882 g008
Figure 9. Vickers hardness and schematic image of the indentation of composite ceramics: (a) Vickers hardness, (b) schematic image.
Figure 9. Vickers hardness and schematic image of the indentation of composite ceramics: (a) Vickers hardness, (b) schematic image.
Materials 16 00882 g009
Figure 10. The mechanism diagram of the composite powder and ceramics.
Figure 10. The mechanism diagram of the composite powder and ceramics.
Materials 16 00882 g010
Table 1. Mechanical properties data of B4C ceramics reported in references.
Table 1. Mechanical properties data of B4C ceramics reported in references.
Sintering ProcessVickers Hardness (GPa)References
Pressureless–sintered, 2100 °C, 60 min30.2[36]
Pressureless–sintered, 2100 °C, 60 min34.2[37]
Pressureless–sintered, 2150 °C, 60 min33.2[12]
HP, 1950 °C, 30 MPa, 60 min34.6[38]
HP, 1950 °C, 30 MPa, 60 min[39]
HP, 1950 °C, 30 MPa, 30 min35.22[9]
HP, 2000 °C, 30 MPa, 60 min33[4]
SPS, 1700 °C, 32 MPa, 10 min31.28[5]
SPS, 2000 °C, 30 MPa, 3 min32.33[40]
SPS, 1700 °C, 50 MPa, 6 min41.2This work
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

Ding, J.; Wang, J.; Yang, H.; Liu, Z.; Yu, C.; Li, X.; Deng, C.; Zhu, H. Improvement of Mechanical Properties of Composites with Surface Modified B4C for Precision Machining. Materials 2023, 16, 882. https://doi.org/10.3390/ma16020882

AMA Style

Ding J, Wang J, Yang H, Liu Z, Yu C, Li X, Deng C, Zhu H. Improvement of Mechanical Properties of Composites with Surface Modified B4C for Precision Machining. Materials. 2023; 16(2):882. https://doi.org/10.3390/ma16020882

Chicago/Turabian Style

Ding, Jun, Jintao Wang, Hao Yang, Zhenglong Liu, Chao Yu, Xiangcheng Li, Chengji Deng, and Hongxi Zhu. 2023. "Improvement of Mechanical Properties of Composites with Surface Modified B4C for Precision Machining" Materials 16, no. 2: 882. https://doi.org/10.3390/ma16020882

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