Next Article in Journal
Antimicrobial PMMA Bone Cement Containing Long Releasing Multi-Walled Carbon Nanotubes
Previous Article in Journal
Oxide Derived Copper for Electrochemical Reduction of CO2 to C2+ Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reaction Sintering of Machinable TiB2-BN-C Ceramics with In-Situ Formed h-BN Nanostructure

by
Oleksii Popov
1,2,*,
Dmitry V. Shtansky
3,
Vladimir Vishnyakov
4,
Oleksandra Klepko
5,
Sergey Polishchuk
6,
Magzhan K. Kutzhanov
3,
Elizaveta S. Permyakova
3 and
Petro Teselko
1
1
Faculty of Physics, Taras Shevchenko National University of Kiev, 01033 Kyiv, Ukraine
2
Science and Research Centre “Synthesis”, 02139 Kiev, Ukraine
3
Research Laboratory of Inorganic Nanomaterials, National University of Science and Technology “MISiS”, Leninsky Prospect 4, 119049 Moscow, Russia
4
Centre for Engineering Materials, University of Huddersfield, Huddersfield HD1 3DH, UK
5
Inorganic Materials Vapor Phase Department, E. O. Paton Electric Welding Institute, 03150 Kyiv, Ukraine
6
Department of Physics of Dispersed Systems, G. V. Kurdyumov Institute for Metal Physics of the N. A. S. of Ukraine, 03142 Kyiv, Ukraine
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(8), 1379; https://doi.org/10.3390/nano12081379
Submission received: 9 March 2022 / Revised: 24 March 2022 / Accepted: 24 March 2022 / Published: 18 April 2022
(This article belongs to the Section Nanocomposite Materials)

Abstract

:
Soft TiB2-BN-C hetero-modulus ceramics were sintered with the assistance of in-situ reactions during the hot pressing of TiN-B4C precursors. TiB2 formation was observed already after the hot pressing at 1100 °C, remaining the only phase identifiable by XRD even after sintering at 1500 °C. Analysis of reaction kinetics allows us to assume that the most probable reaction controlling stage is boron atoms sublimation and gas phase transfer from B4C to TiN. Reactive sintering route allows almost full densification of TiB2-BN-C composite ceramics at 1900 °C. The processes enable the formation of multilayer h-BN nanosheets inside the TiB2 matrix. The manufactured TiB2-33BN-13C ceramic with K1C = 5.3 MPa·m1/2 and HV = 1.6 GPa is extremely thermal shock-resistant at least up to quenching temperature differential of 800 °C. The sintered UHTC composite can be machined into complex geometry components.

1. Introduction

The covalent nature of the interatomic bonds of metal oxides, nitrides, carbides, and borides not only results in the high melting point of ceramics and oxidation resistance but also shifts the strength loss homologous temperature (the ratio of the current temperature to the material melting point) from 0.2 ÷ 0.3 for metals to 0.5 ÷ 0.7. This makes ceramics a key candidate for high-temperature applications [1,2,3,4,5], despite well-known drawbacks of susceptibility to sharp temperature variations, sintering challenges, and shaping/forming difficulties [6,7].
Hetero-modulus ceramics (HMCs), being a composition of a super hard matrix with soft graphite or h-BN inclusions, demonstrate essential improvements in brittle material thermal shock resistance. It has also been shown that the materials are machinable and the inclusions have no negative impact on their high-temperature properties. Everything mentioned facilitates the industrial use of HMCs [8,9,10,11,12,13,14,15,16]. As is shown in several works, both thermal shock resistance and machinability raise significantly after adding 20 vol.% of the soft phase [9,12,14,17,18]. However, poor sinterability of the graphite and h-BN hinders the material densification and leads to the need for higher sintering parameters [19,20]. E.g., to obtain non-porous TiC with 30 vol.% of C ceramic, Shabalin et al. [15] hot pressed the green body at 2700 °C and 12 MPa. The application of the same parameters for TiC ceramic with 40 vol.% of C resulted in 5% porosity. A similar sintering route has been used by John and Jenkins for C-MeC composites [21]. TiB2 with 50 vol.% of hBN ceramics had been hot-pressed at 1900 °C and 30 MPa for 60 min by Akarsu et al., but the resulting material contained approximately 30% of pores [22]. Reduced porosity in BN-based material was obtained with the addition of B2O3 as a fusible sintering aid when ceramic has been densified up to 92% at 1900 °C and 50 MPa for 60 min [23]. Bearing in mind that TiB2-BN based ceramics are being widely used for aluminum evaporation boats [24] and as a side-dam material for thin-strip casting in steel production [23], the optimization of the high-quality HMC sintering procedure remains an important task.
The reaction sintering technique, which is when all material phases are formed in-situ via the high-temperature chemical interaction between the precursor powders, was shown to intensify material consolidation and lower porosity and allowed for less demanding sintering conditions [25,26,27,28,29,30,31]. In addition, the reaction hot pressing of hetero-modulus ceramics allowed the production of submicron graphite or even graphene platelets [32]. The platelets efficiently blunt a crack tip and by this process increase the ceramic toughness [11,33]. TiB2-C based ceramics were efficiently manufactured by means of the hot pressing of TiC-B4C precursors (e.g., green bodies) by utilizing the following reaction:
2MeC + B4C → 2MeB2 + 3C
The non-porous UHTCs were produced at 1800–1900 °C during 2–16 min of pressing [11,34]. It was shown that the reactively formed submicron graphite platelets improved the ceramic matrix thermal shock resistance. However, the content of the soft phase, which one can obtain in the titanium diboride matrix by utilizing reaction (1), cannot exceed 30 vol.%. Zou et al., showed that the hot pressing of ZrN, B4C and Si powder mixture resulted in ZrB2-BN-SiC composite with an h-BN content of 28 vol.%. The h-BN was homogeneously distributed in the ZrB2-SiC matrix [26]. It is evident that a similar process with no silicon in the green body should prevent the carbon from forming hard SiC crystals and therefore increase the soft phase content. The presented work is the investigation of the possibility to create soft TiB2-BN-C hetero-modulus ceramics using the reaction of the hot pressing of TiN-B4C precursors. The soft phase content increasing should significantly improve the UHT material machinability and thermal shock resistance with no negative impact on high temperature properties. The reactive sintering approach would simplify the ceramic processing.

2. Experimental Procedure

2.1. Sample Sintering and Preparation

Commercially available powders of TiN (~30 μm) and B4C (~20 μm) (Donetsk Reactive Factory, Donetsk, Ukraine) were used as green body materials. The materials purity, as stated by the manufacturers, was around 99.0 at%. The green body composition was chosen to comply with the stoichiometry of the following reaction:
4TiN + 3B4C → 4TiB2 + 4BN + 3C
Two green body mixtures entitled “reactive” (4TiN + 3B4C) and “nonreactive” (4TiB2 + 4BN + 3C) were prepared according to the reaction’s (2) left and right-hand side, respectively (See Table 1).
The powder mixtures were ground in a high-energy planetary mill (1500 r.p.m.) (Physics Faculty, Taras Shevchenko National University of Kyiv, Kyiv, Ukraine) in a zirconia jar with 10 mm zirconia balls for 30 min and hot-pressed at a pressure of 30 MPa in a graphite die, without a special protective atmosphere. The die assembly was heated with an AC (50 Hz) current in the air with the use of the hot-pressing equipment DCS-0 produced by SRC Synthesis (Kiev, Ukraine). The heating rate (after preheating to 600 °C for 20 min) was approximately 100°/min. The synthesis temperature for the reactive powder ranged from 1000 to 1900 °C. The nonreactive powder has been hot pressed at 1900 °C only. The sample thickness has been calculated from the plunger movement, which has been measured in micrometres. The micrometre accuracy was about 10 μm.

2.2. Material Characterization

The microstructure of the TiB2-BN-C composites was studied on a JSM F7600 (JEOL, Japan) and Tescan Vega3 scanning electron microscope (Tescan, Czech Republic) both equipped with energy-dispersive X-ray (EDX) spectrometers. Raman spectrum was recorded on a Thermo DXR spectrometer (Thermo Fisher Scientific, Waltham, WI, USA) with a 532 nm excitation laser.
The bulk densities of obtained materials were measured using Archimedes’ method. The relative densities were calculated based on the theoretical densities of TiB2 (4.495 g/cm3), BN (2.1 g/cm3), and C (graphite) (2.28 g/cm3) assuming the rule of mixtures. X-ray diffraction (DRON-4M) with Cu-Kα radiation analyzed with Rietveld method according to the procedure presented in [35] was used to identify the phase composition of the specimens. Chemical composition was determined by Energy Dispersive X-ray Spectroscopy (EDS) (Oxford, UK) analysis.
Indentation measurements (hardness and fracture toughness) were performed with a load of 98 N and a 10 s holding at the load. The measurements were taken on the polished sample surfaces. The fracture toughness was estimated by measuring the crack lengths generated by the Vickers indentations. The toughness was calculated according to the formula of Niihara et al. [36].
Thermal shock crack growth was examined according to the indentation-quench method, as described elsewhere [37]. The disk specimens were polished and indented using a Vickers pyramidal indenter as described above. Five indentations were made on each specimen. Four cracks at the places of pyramid ribs were formed after each indentation. This way, 20 cracks per sample were measured. The samples were heated to a preselected set of temperatures: 200, 300, 400, 500, 600, 700, and 800 °C in the air for 10 min and then quenched in a water bath at room temperature (20 °C). The radial crack lengths of each indentation were measured before and after each quenching event by an optical microscope. The crack growth was expressed as a percentage of the initial crack length as:
( c i c 0 ) c 0 × 100 %
with i being the sequential number of the quenching cycle. The average crack growth for each sample was estimated as a function of the quenching temperature.
Room-temperature compressive strength was determined using an AG-Xplus universal tensile testing machine (Shimadzu, Japan). The load increase rate was 8.0 × 10−5 m/s.
To reveal the influence of h-BN on the material resistance to shock-dynamic loads, the TiB2-BN-C and TiB2-C composites were subjected to cyclic impact test at an applied load of 300 N using an impact tester (CemeCon) (CemeCon, Germany). Each sample was exposed to 105 impacts with a frequency of 50 Hz using a WC-Co ball with a diameter of 5 mm. The shape and morphology of impact cavities were observed with an NT1100 optical profiler (WYKO).
The degree of conversion α in chemical reactions was estimated based on amounts of titanium in the green body and in the product phases based on the XRD data. If at time t, an amount of Ti consisted in the product phase (TiB2) was NTi(t) and the total amount of Ti in both product (TiB2) and precursor (TiN) phases was NTi0 then α was estimated as:
α = N T i ( t ) N T i 0
The samples were made from the same stoichiometric green body composition but then hot pressed at different times and temperatures to build the kinetic α = α(t) curves for 1100, 1300, 1400, and 1500 °C. The reaction speed w = dα/dt dependences were drawn. We assumed that the reaction speed was dependent on the sintering temperature due to Arrhenius’ Equation:
w = A e E a k T
where A is a constant depending on the degree of conversion and Ea is the reaction activation energy. Therefore, plotting ln(w) on 1/kT for a given α we found the activation energy as a slope of the corresponding line.
Finally, two samples were machined with standard metal-working tools (drill, files, hand tap, and die) to create M3 nut and screw.

3. Results and Discussion

3.1. Densification and Reaction Kinetics

TiB2 formation in TiN-B4C was observed already after the hot pressing at 1100 °C (Figure 1). The sintering temperature rise intensified the reaction; therefore, no initial phases can be seen after 4 min of soaking at 1500 °C. The low-intensity halo at 2θ between 26 and 27° is the only very weak signal from carbon and boron nitride in the sample hot-pressed at this temperature.
However, the powder processing at 1900 °C revealed both graphite and h-BN in their crystalline forms (See Figure 1). Therefore, the reaction (2) occurred during the hot pressing of the 4TiN + 3B4C green body and resulted in TiB2-BN-C HMC formation.
The phase composition evolution (Figure 2a) shows an essential change between 1300 and 1500 °C with the TiN to TiB2 conversion degree not exceeding 70% after 16 min at 1300 °C, and extending up to 100% after 2 min at 1500 °C. The reaction activation energy raises considerably at α > 40% (Figure 2b), thus revealing the change in the main reaction mechanism and its limiting stage. Based on the similarity between TiN and TiC, both having NaCl-type crystalline structure with the lattice parameters of 0.424 nm and 0.433 nm, correspondingly, it is reasonable to assume the similarity of the reaction (2) to the well-investigated one between titanium and boron carbides [33,38]:
2TiC + B4C → 2TiB2 + 3C
As is shown in [30], the reaction (3) starts at 1100–1200 °C with the boron carbide decomposition [38] and boron diffusion into TiC grains. The process goes along with TiB2 nucleation along (111) plains of the carbide crystals. It is possible to assume that the reaction (2) is progressing in the same way. Considering that the value of B4C enthalpy does not exceed 63 kJ/moll (~0.1 eV per atom) [39], the material decomposition cannot be the reaction (2) limiting stage (whose activation energy, according to Figure 2b, is at and above 4 eV). The same can be stated about B atoms solid-state diffusion via the product (TiB2) phase, whose activation energy is 2.2 eV [40]. This forces us to assume that the reaction is most probably limited with boron atoms transfer through the interphase boundary between B4C and TiN. As is shown in [38], solid-state diffusion cannot provide a considerable transformation degree, so the activation energy growth after α = 40% reflects the reaction limiting stage change. According to [30], the intense sublimation of boron atoms from the B4C surface begins at 1400 °C, thus the gas phase diffusion became a predominant B to TiN transport mechanism. According to [41], boron sublimation enthalpy approximates 6 eV, which in principle correlates with Figure 2b data and confirms the above-mentioned transport model.
Comparison of the “reactive” and “non-reactive” densification kinetics (Figure 3) for two types of green bodies with the same atomic contents of boron, carbon, titanium, and nitrogen, but whose atomic contents are present in different chemical compounds (4TiN + 3B4C and 4TiB2 + 4BN + 3C) (See Table 1) showed the following:
  • Despite the fact that both mixtures were milled in the same way and the samples contained the same green body mass, at the beginning of the process the non-reactive sample was 10% higher (thicker, e.g., had 10% more volume) than the reactive one. It can be explained on the basis that the bulk density of the non-reactive composition (3.42 g/cm3) is 10% lower than that of the reactive one (3.72 g/cm3).
  • The non-reactive charge started shrinking (e.g., densifying) after 600 °C while the reactive one remained almost unchanged up to 1250 °C. At 1250 °C, the green body expands and then begins to shrink at 1350 °C. The behavior of the non-reactive powder is most probably mediated by the powder densification helped by the lubrication provided by h-BN and graphite. The reactive sample growth corresponded to the TiB2 arrival (the degree of conversion raised from zero, Figure 3b) and could be caused by the reaction (2) dilatometry: the product density being lower than that of the precursors.
  • The non-reactive sample densification proceeds at an almost constant rate over a broad span of time (more than 13 min) and sintering temperature (600 to 1900 °C). Even at 1900 °C and after holding at this temperature for 4 min the densification continues. While the reactive sample densification is almost complete within 5 min and in the interval of 1350–1900 °C.
  • The density of sintered reactive sample is approximately 99% which is 8% higher than that of the non-reactive one.
All listed features are connected to the reaction (2) oddities. Therefore, the in-situ reaction leads to faster ceramic consolidation and denser material.

3.2. Atomic Structure Formation

As shown, the formation of titanium diboride starts at 1100 °C, which is almost two thousand degrees lower than the phase melting point. The TiB2 nanocrystals nucleated in such conditions can be observed on the SEM image (Figure 4a) alongside the significantly larger grains which are just the precursor particles. The absence of the rough grains in the sample after hot pressing at 1650 °C (Figure 4b) confirms the completion of TiN to TiB2 conversion at the temperature. Nucleated at lower temperatures, TiB2 nanoparticles evidently recrystallized into the submicron ones. A hot pressing temperature rise to 1950 °C resulted in further composite structure changes: the diboride grains predictably grew to 2–4 μm, and the new type of plate-like particles appeared on the sample fracture surface (Figure 4c). SEM analysis in back-scattered electrons showed that the plate-like grains contained lighter elements which, considering the XRD results (Figure 1), showed them to be either h-BN or graphite.
Sample fracture surface investigation with EDS mapping (Figure 5) showed that the plate-like inclusions are boron nitride. Graphite inclusions also seen on the EDS maps, are not in the platelet shape. As was shown in our previous works, TiB2-C composites reactively sintered from TiC-B4C precursors according to reaction (3), contained two types of carbon particles: the plate-like particles, segregating from TiC after its conversion into TiB2, and sponge-like particles, formed with the carbon atoms that remain after boron carbide decomposition [30,33]. Interestingly, the TiN-B4C to TiB2-BN-C conversion presented here not only repeated the reaction path but also resulted in a similar structure formed via the same reaction mechanism. In which roughly equiaxial 2–4 μm diboride particles nucleated and grew from the parental TiN grains with the displacement of nitrogen by boron and formation of shapeless graphite inclusions after B atoms left B4C crystals. BN platelets appear at higher temperatures when the displaced TiN nitrogen reacts with boron.
It should be noted that BN in TiN-B4C system formation differs from graphite in TiC-B4C segregation. It might be expected that N atoms leaving TiN crystals would form a gaseous N2 phase and escape from the material. However, neither XRD nor EDS indicated any free boron, which would have inevitably remained in case of the nitrogen disappearance. As was already stated, boron carbide decomposition was the first stage of the interaction process. Therefore, N atoms leaving TiN particles during their conversion into TiB2 nucleus react with boron either on the crystal surface or even before reaching it. This probably explains why the BN structures look to grow (Figure 5) as if from inside of the diboride grains perpendicular to their surfaces.
The described growing processes facilitate the formation of BN nanosheets (Figure 6), possibly inside the TiB2-based matrix. The nanosheets do not “cover” the diboride grains, allowing the formation of a continuous TiB2 frame, which was also confirmed by the fact that the sintered samples were electrically conductive. The possibility of in-situ formation of the h-BN nanostructures is essential considering the difficulties of BN nanosheets manufacturing [42].
Raman spectra of the TiB2-BN-C ceramics obtained from the fracture surface are presented in Figure 7. Since the TiB2 phase is not Raman active, only h-BN and C components can be observed. The characteristic Raman G peak at 1360 cm−1 is a fingerprint of h-BN, which appears due to in-plane vibrations between B and N atoms [43]. Relatively high peak intensity suggests a multilayer h-BN configuration. Other observed features in the Raman spectra at 454, 1309, 1580, and 2680 can be assigned to sp- and sp2-bonded carbon. The Raman spectrum of graphite has a main band at 1580 cm−1 (D peak) and a small feature at 2687 cm−1 (denoted as 2D or G’ peak). Multilayer graphene has additional bands at 1333–1349 and 1604–1618 cm−1 [44,45]. These modes are Raman active in defective graphitic materials. In addition, unlike a singlet peak in graphene, the 2D band in multilayer graphene appears as a doublet [46]. The intensity of the second peak in the doublet at approximately 2913 nm−1, which is a characteristic of the disorder [47], is very small. Low-frequency peaks observed at 454 cm−1 can be attributed to bending modes of sp-coordinated carbon nanostructures [48].

3.3. Mechanical Properties

Stress-strain curves for three different samples of the TiB2-BN-C composite is shown in Figure 8. All investigated samples demonstrated elastic behavior up to approximately 180 MPa. Steps on the curves correspond to the formation of microcracks. Unlike the well-known behavior of other ceramic materials, the crack formation did not result in instant sample destruction. As can be seen from Figure 8, the sintered hetero-modulus composite demonstrates quasiplasticity, when the material deformation is being provided with microcracks formation rather than dislocation movement. Such behavior was explained in [15] with an essential crack deflection on the plate-like soft inclusion. It should be noted, however, that the value of TiB2-BN-C composite compressive strength is 2.3 and 1.2 times higher than that of TiC-41C and TiC-33C materials, respectively [15]. Mechanical properties of different soft ceramics used for high-temperature applications are compared in Table 2.
As was already shown [12], the indentation crack growth in the TiB2-31C sample becomes noticeable when the quenching temperature difference exceeded 400 °C and extended to 36% at ΔT = 700 °C (See Figure 9), which bettered the results [37] substantially. The reactively pressed TiB2-33BN-13C material presented even further improvement with the cracks being entirely stable up to quenching temperature differential of 800 °C. The latter coincides with the well-known thermal shock resistance criteria [51,52]:
R = σ f ( 1 ν ) k α E
and
R = K 1 C ( 1 ν ) k α E
where σ f , K 1 C , ν, α, k, and E are the fracture strength, toughness, Poisson’s ratio, thermal expansion coefficient, thermal conductivity and Young’s modulus respectively. The investigated reactively pressed ceramic has about 46% of the low Young’s modulus (graphite and h-BN) phases. This allows us to explain the ceramic enhanced, as compared to TiB2-31C, thermal shock resistance.
Comparison of the dynamic impact test results also showed the TiB2-BN-C material is less susceptible to mechanical failure compared to the TiB2-C counterpart and does not have cracks and damaging zones along the impact cavity edges (Figure 10).
Finally, as can be seen from Figure 11, the sintered composite is quite machinable. This property presents the possibility of complex shape formation.

4. Conclusions

New TiB2-BN-C hetero-modulus ceramics with 46% of soft (e.g., low Young’s modulus) phases, h-BN and carbon, were sintered by the in-situ reaction during hot pressing of TiN-B4C precursors. TiB2 formation was observed already at a temperature around 1100 °C. Reaction activation energy at the beginning of the reaction is at around 4 eV but increases as TiN into TiB2 chemical conversion proceeds and reaches values in the region of 6 eV. Nonreactive hot pressing of TiB2, BN and carbon precursors proceeds in a wide sintering temperature region from 1100 to 1900 °C but achieves only 92% of the composite theoretical density. Reactive hot pressing of TiN-B4C precursors leads to almost full densification (ρ = 99%) of TiB2-BN-C hetero-modulus ceramics at 1900 °C. During reactive sintering at temperatures above 1500 °C, multilayer h-BN nanosheets appear. The sintered ceramic exhibits pseudoplastic deformation when microcracks are formed but crack propagation is prevented by the crack tip blunting in the soft inclusions. A high content of low Young’s modulus phases in TiB2-33BN-13C ceramic leads to the extremely high thermal shock-resistance when initiated cracks do not grow at least up to quenching temperature differential of 800 °C. The sintered ceramic is easily machinable and can be machined into engineering components subjected to extreme temperature variations.

Author Contributions

Conceptualization, O.P.; methodology, O.P., D.V.S. and V.V.; investigation, O.P., O.K., S.P., M.K.K., E.S.P. and P.T.; writing—original draft preparation, O.P. 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

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data presented in this study are available as suplementory material (Raw Data_TiN-B4C.zip).

Conflicts of Interest

The authors declare no conflict of interest.

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] [Green Version]
  2. Padture, N.P. Advanced structural ceramics in aerospace propulsion. Nat. Mater. 2016, 15, 804–809. [Google Scholar] [CrossRef] [PubMed]
  3. Sani, E.; Mercatelli, L.; Sans, J.-L.; Silvestroni, L.; Sciti, D. Porous and dense hafnium and zirconium ultra-high temperature ceramics for solar receivers. Opt. Mater. 2013, 36, 163–168. [Google Scholar] [CrossRef]
  4. Sani, E.; Meucci, M.; Mercatelli, L.; Balbo, A.; Musa, C.; Licheri, R.; Orrù, R.; Cao, G. Titanium diboride ceramics for solar thermal absorbers. Sol. Energy Mater. Sol. Cells 2017, 169, 313–319. [Google Scholar] [CrossRef] [Green Version]
  5. Shabalin, I.L. Ultra-High. Temperature Materials, I.; Springer: Amsterdam, The Netherlands, 2014; p. 794. [Google Scholar]
  6. Bannister, M.K.; Swain, M.V. Thermal Shock of a Titanium Di-boride Based Composite. Ceram. Int. 1990, 16, 77–83. [Google Scholar] [CrossRef]
  7. Cheng, T.; Li, W.; Fang, D. Thermal Shock Resistance of Ultra-High-Temperature Ceramics. AIAA J. 2013, 51, 840–848. [Google Scholar] [CrossRef]
  8. Shabalin, I.L.; Tomkinson, D.M.; Shabalin, L.I. High-temperature hot-pressing of titanium carbide–graphite hetero-modulus ceramics. J. Eur. Ceram. Soc. 2007, 27, 2171–2181. [Google Scholar] [CrossRef]
  9. Gao, L.; Jin, X.; Li, J.; Li, Y.; Sun, J. BN/Si3N4 nanocomposite with high strength and good machinability. Mater. Sci. Eng. A 2006, 415, 145–148. [Google Scholar] [CrossRef]
  10. Gömze, L.A.; Gömze, L.N. Hetero-modulus alumina matrix nanoceramics and CMCs with extreme dynamic strength. IOP Conf. Ser. Mater. Sci. Eng. 2011, 18, 082001. [Google Scholar] [CrossRef]
  11. Popov, O.; Vishnyakov, V. Fracture toughness in some hetero-modulus composite carbides: Carbon inclusions and voids. Adv. Appl. Ceram. 2017, 116, 61–70. [Google Scholar] [CrossRef] [Green Version]
  12. Popov, O.; Avramenko, T.; Vishnyakov, V. Thermal conductivity and thermal shock resistance of TiB2-based UHTCs enhanced by graphite platelets. Mater. Today Commun. 2020, 26, 101756. [Google Scholar] [CrossRef]
  13. Yuan, B.; Wang, G. Preparation and properties of Si3N4/BN ceramic composites. Procedia Eng. 2012, 27, 1292–1298. [Google Scholar] [CrossRef] [Green Version]
  14. Jiang, T.; Jin, Z.; Yang, J.; Qiao, G. Mechanical property and R-curve behavior of the B4C/BN ceramics composites. Mater. Sci. Eng. A 2008, 494, 203–216. [Google Scholar] [CrossRef]
  15. Shabalin, I.L.; Wang, Y.; Krynkin, A.V.; Umnova, O.V.; Vishnyakov, V.M.; Shabalin, L.I.; Churkin, V.K. Physicomechanical properties of ultrahigh temperature heteromodulus ceramics based on group 4 transition metal carbides. Adv. Appl. Ceram. 2010, 109, 405–416. [Google Scholar] [CrossRef]
  16. Kovalčíková, A.; Tatarko, P.; Sedlák, R.; Medved, 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]
  17. Kusunose, T.; Sekino, T.; Choa, Y.H.; Niihara, K. Machinability of Silicon Nitride/Boron Nitride Nanocomposites. J. Am. Ceram. Soc. 2002, 85, 2689–2695. [Google Scholar] [CrossRef]
  18. Li, X.; Gao, Y.; Pan, W.; Wang, X.; Song, L.; Zhong, Z.; Wu, S. Fabrication and characterization of B4C-based ceramic composites with different mass fractions of hexagonal boron nitride. Ceram. Int. 2015, 41, 27–36. [Google Scholar] [CrossRef]
  19. BSong, B.; Yang, W.; Liu, X.; Chen, H.; Akhlaghi, M. Microstructural characterization of TiB2–SiC–BN ceramics prepared by hot pressing. Ceram. Int. 2021, 47, 29174–29182. [Google Scholar]
  20. Kusunose, T. Fabrication of Boron Nitride Dispersed Nanocomposites by Chemical Processing and Their Mechanical Properties. J. Ceram. Soc. Jpn. 2006, 114, 167–173. [Google Scholar] [CrossRef] [Green Version]
  21. John, D.; Jenkins, G.M. Hot-working and strengthening in metal carbide-graphite composites. J. Mater. Sci. 1986, 21, 2941–2958. [Google Scholar] [CrossRef]
  22. Akarsu, A.; Gökçe, H.; Boyraz, T.; Ertuğ, B.; Addemir, A.; Öveçoğlu, M. The Characterization of The Mechanical And Thermal Properties of Hot-Pressed Hexagonal Boron Nitride-Titanium Diboride Composites. In Proceedings of the Scientific Works of 3rd International Symposium on SiAlONs and Non-Oxides, Nevşehir, Turkey, 1–4 June 2010. [Google Scholar]
  23. Eichler, J.; Lesniak, C. Boron nitride (BN) and BN composites for high-temperature applications. J. Eur. Ceram. Soc. 2008, 28, 1105–1109. [Google Scholar] [CrossRef]
  24. Ping, H.; Weimin, W.; Yanling, D. Thermal Shock Resistance and Erosion Resistance of TiB2 Multiphase Ceramic Composites. J. Wuhan Univ. Technol. 2006, 21, 117–120. [Google Scholar] [CrossRef]
  25. MSalari, M.A.; Muğlu, G.M.; Rezaei, M.; Kumar, M.S.; Pulikkalparambil, H.; Siengchin, S. In-situ synthesis of TiN and TiB2 compounds during reactive spark plasma sintering of BN–Ti composites. Synth. Sinter. 2021, 1, 48–53. [Google Scholar]
  26. Zou, J.; Liu, J.; Zhao, J.; Zhang, G.J.; Huang, S.; Qian, B.; Vleugels, J.; Biest, O.V.; Shen, J.Z. A top-down ap-proach to densify ZrB2–SiC–BN composites with deeper homogeneity and improved reliability. Chem. Eng. J. 2014, 249, 93–101. [Google Scholar] [CrossRef]
  27. Derakhshandeh, M.R.; Fazili, A.; Golenji, R.B.; Alipour, F.; Eshraghi, M.J.; Nikzad, L. Fabrication of (TixZr1−x)B2-(ZrxTi1−x)N composites by reactive spark plasma sintering of ZrB2-TiN. J. Alloys Compd. 2021, 887, 161403. [Google Scholar] [CrossRef]
  28. Namini, A.S.; Delbari, S.A.; Oh, Y.; Asl, M.S.; Van Le, Q.; Cha, J.H.; Lee, S.-H.; Jang, H.W.; Han, H.N.; Shokouhimehr, M. Role of TiCN addition on the characteristics of reactive spark plasma sintered ZrB2-based novel composites. J. Alloys Compd. 2021, 875, 159901. [Google Scholar] [CrossRef]
  29. Namini, A.S.; Delbari, S.A.; Asl, M.S.; Van Le, Q.; 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]
  30. Popov, O.; Chornobuk, S.; Vishnyakov, V. Structure formation of TiB2-TiC-B4C-C hetero-modulus ceramics via reaction hot pressing. Int. J. Refract. Met. Hard Mater. 2017, 64, 106–112. [Google Scholar] [CrossRef]
  31. Popov, O.; Vishnyakov, V.; Fleming, L.; Podgurskiy, M.; Blunt, L. Reaction sintering of biocompatible Al2O3-hBN ceramics. ACS Omega 2022, 7, 2205–2209. [Google Scholar] [CrossRef] [PubMed]
  32. Popov, O.; Vishnyakov, V.; Poperenko, L.; Yurgelevych, I.; Avramenko, T.; Ovcharenko, A. Reactively sintered TiB2-based heteromodulus UHT ceramics with in-situ formed graphene for machinable concentrated solar light absorbers. Ceram. Int. 2022, in press. [Google Scholar] [CrossRef]
  33. Popov, O.; Vishnyakov, V.; Chornobuk, S.; Totsky, I.; Plyushchay, I. Mechanisms of TiB2 and graphite nucleation during TiC-B4C high temperature interaction. Ceram. Int. 2019, in press. [Google Scholar] [CrossRef]
  34. Chornobuk, S.V.; Popov, A.Y.; Makara, V.A. Structure and mechanical properties of reaction-sintered ceramic composite materials based on titanium and hafnium diborides. J. Superhard Mater. 2009, 31, 86–88. [Google Scholar] [CrossRef]
  35. Evans, J.S.O.; Evans, I.R. Structure Analysis from Powder Diffraction Data: Rietveld Refinement in Excel. J. Chem. Educ. 2021, 98, 495–505. [Google Scholar] [CrossRef]
  36. Niihara, K.; Morena, R.; Hasselman, D.P.H. Evaluation of K1c of brittle solids by the indentation method with low crack-to-indent ratios. J. Mater. Sci. Lett. 1982, 1, 13–16. [Google Scholar] [CrossRef]
  37. Liang, J.; Wang, Y.; Fang, G.; Han, J. Research on thermal shock resistance of ZrB2–SiC–AlN ceramics using an indentation-quench method. J. Alloys Compd. 2010, 493, 695–698. [Google Scholar] [CrossRef]
  38. Popov, O.; Klepko, A.; Lutsak, E. The influence of high pressure on TiC-B4C reaction kinetics. Int. J. Refract. Met. Hard Mater. 2018, 75, 234–237. [Google Scholar] [CrossRef]
  39. NIST. Chemistry WebBook. Available online: http://webbook.nist.gov/chemistry/ (accessed on 1 September 2021).
  40. Schmidt, H.; Borchardt, G. Self-diffusion of boron in TiB2. J. Appl. Phys. 2003, 93, 907–911. [Google Scholar] [CrossRef]
  41. Mar, R.; Bedford, R. Sublimation of boron. High Temp. Sci. 1976, 8, 365–376. [Google Scholar]
  42. Golberg, D.; Bando, Y.; Huang, Y.; Terao, T.; Mitome, M.; Tang, C.; Zhi, C. Boron Nitride Nanotubes and Nanosheets. ACS Nano 2010, 4, 2979–2993. [Google Scholar] [CrossRef] [PubMed]
  43. Bergman, L.; Nemanich, R.J. Raman spectroscopy characterization of hard, wide-bandgap semiconductors: Diamond, GaN, GaAlN, AlN, BN. Annu. Rev. Mater. Sci. 1996, 26, 551–579. [Google Scholar] [CrossRef]
  44. Bokobza, L.; Bruneel, J.-L.; Couzi, M. Raman spectra of carbon-based materials (from graphite to carbon black) and of some silicone composites. C-J. Carbon Res. 2015, 1, 77–94. [Google Scholar] [CrossRef] [Green Version]
  45. Bokobza, L.; Bruneel, J.-L.; Couzi, M. Raman spectroscopy as a tool for the analysis of carbon-based materials (highly oriented pyrolitic graphite, multilayer graphene and multiwall carbon nanotubes) and of some of their elastomeric composites. Vib. Spectrosc. 2014, 74, 57–63. [Google Scholar] [CrossRef]
  46. Ferrari, A.C.; Basko, D.M. Raman spectroscopy as a versatile tool for studying the properties of graphene. Nat. Nanotechnol. 2013, 8, 235–246. [Google Scholar] [CrossRef] [Green Version]
  47. Saito, R.; Hofmann, M.; Dresselhaus, G.; Jorio, A.; Dresselhaus, M.S. Raman scattering of graphene and carbon nanotubes. Adv. Phys. 2011, 60, 413–550. [Google Scholar] [CrossRef]
  48. Casari, C.S.; Bassi, A.L.; Baserga, A.; Ravagnan, L.; Piseri, P.; Lenardi, C.; Tommasini, M.; Milani, A.; Fazzi, D.; Bottani, C.E.; et al. Low-frequency modes in the Raman spectrum of sp-sp2 nanostructured carbon. Phys. Rev. B 2008, 77, 195444. [Google Scholar] [CrossRef] [Green Version]
  49. Tian, S.; Xing, W.; Wang, W.; FU, Z.; Wang, H.; Zhang, J. Influence of particle size of TiB2 on densification and properties of BN-TiB2 composite ceramics. J. Chin. Ceram. Soc. 2001, 39, 1757–1762. [Google Scholar]
  50. Kitiwan, M.; Ito, A.; Goto, T. Spark plasma sintering of TiN–TiB2–hBN composites and their properties. Ceram. Int. 2015, 41, 4498–4503. [Google Scholar] [CrossRef]
  51. Li, W.; Li, D.; Cheng, T.; Fang, D. Temperature-damage-dependent thermal shock resistance model for ultra-high temperature ceramics. Eng. Fract. Mech. 2012, 82, 9–16. [Google Scholar] [CrossRef]
  52. Lu, T.; Fleck, N. The thermal shock resistance of solids. Acta Mater. 1998, 46, 4755–4768. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Normalized XRD spectra of 4TiN + 3B4C green body after 4-min hot pressing at different temperatures. Peak identification is completed using by Rietveld method.
Figure 1. Normalized XRD spectra of 4TiN + 3B4C green body after 4-min hot pressing at different temperatures. Peak identification is completed using by Rietveld method.
Nanomaterials 12 01379 g001
Figure 2. The dependences of the reaction (2) kinetic parameters: the degree of conversion α on time (a), and the activation energy (Ea) on α (b).
Figure 2. The dependences of the reaction (2) kinetic parameters: the degree of conversion α on time (a), and the activation energy (Ea) on α (b).
Nanomaterials 12 01379 g002
Figure 3. The sample heights (H) (thickness) evolution during hot pressing of the “reactive” (reaction (2) precursor stoichiometry) and “non-reactive” (reaction (2) products) powder mixtures (Table 1) alongside the process temperature (a), and the degree of conversion (b) for the reactive sample.
Figure 3. The sample heights (H) (thickness) evolution during hot pressing of the “reactive” (reaction (2) precursor stoichiometry) and “non-reactive” (reaction (2) products) powder mixtures (Table 1) alongside the process temperature (a), and the degree of conversion (b) for the reactive sample.
Nanomaterials 12 01379 g003
Figure 4. SEM images of the fracture surfaces of TiB2-BN-C samples. Hot pressing from TiN-B4C powder precursors at: (a) 1200 °C (secondary electrons), (b) 1650 °C, (secondary electrons), (c) 1900 °C (secondary electrons), (d) 1900 °C (back scattered electrons).
Figure 4. SEM images of the fracture surfaces of TiB2-BN-C samples. Hot pressing from TiN-B4C powder precursors at: (a) 1200 °C (secondary electrons), (b) 1650 °C, (secondary electrons), (c) 1900 °C (secondary electrons), (d) 1900 °C (back scattered electrons).
Nanomaterials 12 01379 g004
Figure 5. SEM (secondary electrons) and EDS elemental maps of the fractured TiB2-BN-C surface when the sample is hot pressed from TiN-B4C green body at 1900 °C.
Figure 5. SEM (secondary electrons) and EDS elemental maps of the fractured TiB2-BN-C surface when the sample is hot pressed from TiN-B4C green body at 1900 °C.
Nanomaterials 12 01379 g005
Figure 6. SEM image (secondary electrons) of the BN nanosheets formed near the TiB2 grain.
Figure 6. SEM image (secondary electrons) of the BN nanosheets formed near the TiB2 grain.
Nanomaterials 12 01379 g006
Figure 7. Raman shift spectra of TiB2-BN-C ceramics.
Figure 7. Raman shift spectra of TiB2-BN-C ceramics.
Nanomaterials 12 01379 g007
Figure 8. Stress-strain diagram for three reactively pressed TiB2-BN-C samples.
Figure 8. Stress-strain diagram for three reactively pressed TiB2-BN-C samples.
Nanomaterials 12 01379 g008
Figure 9. The dependence of the indentation crack growth after quenching from different temperatures of TiB2-C (asterisk, [12]) and TiB2-BN-C (this work, circles) samples.
Figure 9. The dependence of the indentation crack growth after quenching from different temperatures of TiB2-C (asterisk, [12]) and TiB2-BN-C (this work, circles) samples.
Nanomaterials 12 01379 g009
Figure 10. The impact crater for TiB2-BN-C (a) and TiB2-C (b) samples.
Figure 10. The impact crater for TiB2-BN-C (a) and TiB2-C (b) samples.
Nanomaterials 12 01379 g010
Figure 11. M3 nut and screw produced from the TiB2-BN-C ceramic sample.
Figure 11. M3 nut and screw produced from the TiB2-BN-C ceramic sample.
Nanomaterials 12 01379 g011
Table 1. The green body compositions.
Table 1. The green body compositions.
TitleTiN, wt.%B4C, wt.%TiB2, wt.%BN, wt.%C, wt.%
Reactive60.439.6---
Nonreactive--71.120.38.6
Table 2. Some mechanical properties of different soft ceramics for high-temperature applications.
Table 2. Some mechanical properties of different soft ceramics for high-temperature applications.
Composition, vol.%Ref.Sintering Routeρ/ρth, %HV, GPaK1C, MPa·m1/2σflex, MPaσcomp, MPa
TiB2-33BN-13CThis workRHP, 30 MPa, 1900 °C, 4 min991.65.3 420
TiB2-31C[12]RHP, 30MPa, 1800 °C, 8 min1004.35
TiC-41C[15]HP, 8 MPa, 2700 °C, 30 min95.70.85.550180
TiC-33C[15]HP, 8 MPa, 2700 °C, 30 min99.71.55.5120350
TiB2-60BN[49]HP, 30 MPa, 1850 °C, 90 min83.9 85
TiB2-50BN[22]HP, 30 MPa, 1900 °C, 60 min711.2
TiB2-21TiN-30BN[50]SPS, 100 MPa, 1700 °C, 5 min968<3.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Popov, O.; Shtansky, D.V.; Vishnyakov, V.; Klepko, O.; Polishchuk, S.; Kutzhanov, M.K.; Permyakova, E.S.; Teselko, P. Reaction Sintering of Machinable TiB2-BN-C Ceramics with In-Situ Formed h-BN Nanostructure. Nanomaterials 2022, 12, 1379. https://doi.org/10.3390/nano12081379

AMA Style

Popov O, Shtansky DV, Vishnyakov V, Klepko O, Polishchuk S, Kutzhanov MK, Permyakova ES, Teselko P. Reaction Sintering of Machinable TiB2-BN-C Ceramics with In-Situ Formed h-BN Nanostructure. Nanomaterials. 2022; 12(8):1379. https://doi.org/10.3390/nano12081379

Chicago/Turabian Style

Popov, Oleksii, Dmitry V. Shtansky, Vladimir Vishnyakov, Oleksandra Klepko, Sergey Polishchuk, Magzhan K. Kutzhanov, Elizaveta S. Permyakova, and Petro Teselko. 2022. "Reaction Sintering of Machinable TiB2-BN-C Ceramics with In-Situ Formed h-BN Nanostructure" Nanomaterials 12, no. 8: 1379. https://doi.org/10.3390/nano12081379

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