Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

ISSN 10634576, Journal of Superhard Materials, 2018, Vol. 40, No. 4, pp. 236–242. © Allerton Press, Inc., 2018.

Original Ukrainian Text © I.V. Solodkyi, I.I. Bogomol, M.Ya. Vterkovs’kyi, P.I. Loboda, 2018, published in Sverkhtverdye Materialy, 2018, Vol. 40, No. 4, pp. 11–19.

PRODUCTION, STRUCTURE, PROPERTIES

LowTemperature Synthesis of Boron Carbide Ceramics


I. V. Solodkyi*, I. I. Bogomol, M. Ya. Vterkovs’kyi, and P. I. Loboda
National Technical University of Ukraine
“Igor Sikorsky Kyiv Polytechnic Institute”,
pr. Peremohy 37, Kiev, 03056 Ukraine
*email: evgen.solodky@gmail.com
Received May 22, 2017

Abstract—This study has been the first to demonstrate the possibility of producing boron carbide ceramics
from coarse (D = 25–150 μm) B4C powder (which is impossible to sinter by conventional methods)
through infiltration with molten silicon and subsequent treatment within the field of the controlled tem
perature gradient. This produces yields a composite ceramics B4C–SiC–Si with a hardness of 26 to
35 GPa and a splitting tensile strength of 110 to 170 MPa. The influence of the velocity of movement of
the temperature gradient on the structure, phase composition, and properties of the prepared composites
has been studied.
DOI: 10.3103/S1063457618040020
Keywords: boron carbide, infiltration, hardness, temperature gradient.

1. INTRODUCTION
Boron carbide (B4C) is a lightweight, hard and strong ceramic material. These characteristics make it
suitable for the use in armor, abrasive and wear resistant materials [1–3]. High mechanical properties of boron
carbide result from its strong covalent bond. However, it is this nature of chemical bonding, which makes dif
ficult the consolidation of boron carbide ceramics. Therefore, in the conventional solidification (pressureless
sintering) the sintering temperature has to be as high as 2200°C and above [4]. Besides the usual process, boron
carbide is sintered through hot pressing (HP) [5], hot isostatic pressing (HIP) [6], spark plasma sintering
(SPS) [7–8]. Generally, hot pressing and pressureless sintering are used for industrialscale production of
boron carbide articles. The use of high temperatures (in the case of the conventional sintering) and a very
expensive equipment (for HP and SPS) are responsible for a considerable increase of the ceramics production
costs. Furthermore, a pure, finegrained (better than 5 μm) starting powder [1–2, 9] has to be used in order to
produce ceramics with a high density and required mechanical properties. This also adds to the ceramics cost.
One of the most promising and costeffective methods of preparing boron carbide ceramics is infiltration
(impregnation) of porous compacts with molten silicon [10–12]. In this process boron carbide can serve as an
alternative source of carbon to form a secondary silicon carbide [13]. The process of infiltration provides boron
carbide composites that contain silicon carbide particles and 5 to 15%1 residual (“free”) silicon [9–13].
A method of zone melting is widely used for producing monocrystals and directionally solidified alloys
[14–17]. Also, it produces materials of increased purity, though the starting materials may be just of commer
cial purity. Therefore, it is fair to assume that a partial use of zone melting, i.e., melting of only one component
(silicon) in the composite material could be effective in reducing “free” silicon amount in the B4C–SiC–Si
system.
Thus, the objective of this work has been to clarify the influence of treatment of the B4C–SiC–Si compos
ite ceramics in the field of controlled temperature gradient on the structure, phase composition, and mechan
ical properties of the composites.

2. EXPERIMENTAL
In this work we used two different boron carbide powders—95% B4C (PJSC “Zaporizhabrasive”,
Ukraine), hereinafter referred to as B4C1, and 96% B4C (Khim Reaktiv Co. Ltd., Ukraine), hereinafter
referred to as B4C2—and 99.0% Si powder (Khim Reaktiv Co. Ltd., Ukraine).
1
Hereinafter the composition of the materials prepared is given in vol %.

236
LOWTEMPERATURE SYNTHESIS 237

A SEM micrograph (Fig. 1) shows that the B4C1 powder has a wide particle size distribution within the
range between 25 and 150 μm (coarse powder), while the particle size of the B4C2 ranges from 1 to 10 μm.

500 μm

Fig. 1. Microstructure of the starting B4C powder (B4C1).

To prepare a uniform mixture of the B4C1 and B4C2 powders, we added a 2.5% aqueous solution of poly
vinyl alcohol (PVA) to the B4C1 powder, with 10 mL per 100 g of powder, for the purpose of moistening. Only
after a thorough wetting was achieved, the fine powder B4C2 (40%) was added in order to increase the com
pact density upon pressing. The B4C1 and B4C2 powders were intermixed for 30 min in a ball mill using a
plastic container and boron carbide balls. Then, the mixture was dried at 100°C for 12 h to remove excessive
moisture and provide polymerization of PVA. Using this mixture we prepared samples and compacted them
by uniaxial pressing at 25 MPa. The residual porosity of the samples was determined by measuring their geo
metric dimensions.
The porous samples (Table 1) were infiltrated with silicon in a resistanceheating vacuum furnace (vacuum
being 10 Pa), with a heating rate of 10°C/min to a temperature of 1530°C, followed by holding for 2 min. The
infiltration was provided by placing a silicon briquette on the top of porous compacts.
The treatment of the infiltrated samples within the field of controlled temperature gradient was carried out
in a Mod. Kristal 206 (Russian Federation) crystalgrowing system equipped with an induction heater, in an
argon atmosphere with a gage pressure of 1 atm. The speed of movement of the samples through the heating
zone was 5 and 10 mm/min.

Table 1. Initial characteristics of samples infiltrated with molten silicon


Mixture composition, vol %
Sample Initial porosity, % Postinfiltration treatment
B4C1 B4C2
1 100 NA 40 None
2 100 NA 40 CTG, 5 mm/min
3 60 40 34 CTG, 5 mm/min
4 60 40 34 CTG, 10 mm/min

The sample surface to be studied was prepared using diamond disks with a grain size ranging from 3 to
125 μm. The microstructure examination was performed on sample surfaces polished to a mirror finish, by the
scanning electron microscopy (SEM) using a Mod. REM 106I (SELMI, Ukraine) microscope in combina
tion energy dispersive spectroscopy (EDS).
The phase composition of the ceramics produced was studied by means of a Mod. Ultima IV (Rigaku,
Japan) Xray diffraction system in the CuKαradiation. The 2θ measuring range was from 15 to 85° with a step
size of 0.02° and sampling time of 2 s at each point. The diffraction system was calibrated using silicon as a
reference substance; the lattice constants were refined by the Rietveld method using PDXL software.
Density of samples was determined by Archimedes method as per ASTM B 963–08, the Vickers hardness
(HV) was measured by a Mod. MHV1000 microhardness tester (China) with a load of 9.8 N and holding time
of 10 s. The average hardness value was found upon at least ten tests.

JOURNAL OF SUPERHARD MATERIALS Vol. 40 No. 4 2018


238 SOLODKYI et al.

The splitting tensile strength of samples was measured by the “Brazilian” method according to ASTM
D3967–95a. The tests were carried out on cylindrical samples 12 mm in diameter and 10 mm high, the loading
speed was 0.5 mm/min. For the strength testing we took three samples of each batch.
The splitting tensile strength σt, MPa, was calculated by the formula

2P
σt = ,
πWD
where P is the load, W and D are the cylinder height and diameter.

3. RESULTS AND DISCUSSION

3.1. Characterization of Microstructure and Phase Composition


Figure 2a shows a typical microstructure of the B4C ceramics prepared by infiltration with molten sili
con—this is a socalled corerim structure of boron carbide particles measuring 100 to 150 μm. The βSiC
phase shows up as lightgray particles in the shape of irregular polyhedrons. The largest lightcolored areas
correspond to residual silicon. The Xray diffraction study (XRD) (Fig. 3b) has revealed that sample 1 pre
pared by infiltration consists of two boron carbide phases that clearly differ in the lattice constants—B4C and
B12(SiBC)3—as well as βSiC, αSiC, and residual silicon. The B4C phase has the following lattice constants:
a = 5.6149 Å, c = 12.1397 Å, while B12(SiBC)3 has a = 5.6542 Å, c = 12.3426 Å. It was stated by Hayun [11,
18] that the phase with a larger lattice constant is a ternary compound B12(SiBC)3. The findings of the present
work for sample 1 are in good agreement with those given in [11, 18]. Note that the intensity of the diffraction
peak due to B12(SiBC)3 is higher than that of the starting boron carbide phase, suggesting that the content of
B12(SiBC)3 prevails. The formation of the corerim structure is accompanied by the recrystallization process
to eventually formation of a new phase—B12(SiBC)3 with a particle size of about 10 μm (see Fig. 2a). Sample
1 has a larger amount of residual silicon: approximately 21 and 27% by the results of the microstructure exam
ination and XRD data, respectively. A very small amount (ca. 2%) of the αSiC phase was found by the quan
titative XRD analysis.
Figure 2b shows a microstructure of sample 2 upon treatment in the CTG field.
As expected, during the treatment the starting boron carbide grains undergo almost a complete recrystal
lization to form secondary boron carbide and silicon carbide (see Fig. 2b). The grain size of the CTGtreated
boron carbide ceramics varies with a wide range—from 10 to 80 μm.
The XRD data on sample 4 have confirmed the presence of only three phases: B4C, βSiC (about 6%), and
αSiC (approx. 7%) (see Fig. 3a). The residual silicon phase of less than 1.5 and 3% was revealed by SEM and
XRD, respectively. No phases that would correspond to a ternary compound B12(SiBC)3 with larger lattice
constants have been found.
Figure 2c shows a microstructure of sample 3 (see Table 1). The infiltration and subsequent treatment in
the CTG field for the sample comprised of boron carbide powders of different fractions lead to an increase of
the βSiC content and to the formation of a boron carbide framework. This effect can be explained by a fast
recrystallization of fine boron carbide particles, which is accompanied by the nucleation of βSiC crystals
[11, 18] during the infiltration and their growth during the subsequent treatment in the CTG field. Thus, it can
be inferred that the amount of the silicon carbide phase in the composite can be controlled by varying the grain
size distribution of the starting powders. According to the XRD data, the treatment in the CTG field upon
infiltration of the mixture of boron carbide powders of different fractions results in a higher content of the α
SiC phase (ca. 14%). Unfortunately, the microstructure observations fail to separate the αSiC.
The thermal activation is one of the prerequisites for nucleation and growth of crystals through a liquid
phase [19, 20]. However, a driving force for this is the process of “supercooling” that occurs in the CTG field
and is governed by the CTG rate of variation (movement) [21]. A microstructure of the sample treated in the
CTG field with a rate of 10 mm/s is shown in Fig. 2d. It is clearly seen that darkcolored boron carbide grains
are separated by residual silicon and βSiC (lightgray) phase. The starting boron carbide grains are indicated
by dotted circles. The XRD study (see Fig. 3a) has revealed a significant amount (19%) of the αSiC phase in
sample 4. The edged particles measuring approx. 5 μm were identified as αSiC (see Fig. 2d). The edged α
SiC particles are easy to see in the micrograph (Fig. 2e) of the fracture surface of sample 4. Also, the volume
concentration of residual silicon was within 5%. Thus, an increase of the velocity of the sample movement

JOURNAL OF SUPERHARD MATERIALS Vol. 40 No. 4 2018


LOWTEMPERATURE SYNTHESIS 239

through the heating zone reduces the duration of interaction between the molten silicon and the boron carbide
particles and thus shortens the boron carbide recrystallization time.

βSiC B4C

Si Rim βSiC

B4C

Core
200 μm 100 μm
(a) (b)

B13C2
βSiC

βSiC
B13C2

Si
αSiC
100 μm 50 μm
(c) (d)

10 μm
(e)
Fig. 2. Microstructure of the B4Cbased ceramics prepared by infiltration with molten silicon: samples 1 (a), 2 (b), 3 (c), 4 (d);
brittle fracture surface of sample 4 (e).

Noteworthy is that the CTG treatment of samples 2, 3, and 4 is responsible for disappearing of diffraction
peaks of the ternary phase B12(SiBC)3 and forming the starting phase B4C. The mechanism of formation of
the corerim structure is well described by Hayun [11, 12, 18]; based on that publications, it has been inferred
that the “stoichiometric saturation” is the most likely mechanism representing the formation of the corerim
structure in the boron carbide–silicon system. However, the return to the starting composition of boron car
bide is a somewhat strange but an interesting phenomenon that could be attributed to the presence of a tem
perature gradient (with all the related processes) which causes dissolution of fine B12(SiBC)3 particles and
their recrystallization through the molten silicon to form B4C particles.

JOURNAL OF SUPERHARD MATERIALS Vol. 40 No. 4 2018


240 SOLODKYI et al.

B4C

αSiC B4C
B4C β
SiC( αSiC, Si)
B4C
βSiC( αSiC)
B4C B4C
Intensity, rel. units
Si Si B4C Si
αSiC, Si)
βSiC(
B4C αSiC
(a)

B12(SiBC) 3 αSiC B12(SiBC) 3


B12(SiBC) 3
(b) B12(SiBC) 3

20 30 40 50 60 2θ, deg
Fig. 3. Results of XRD study of the B4Cbased ceramics upon infiltration with molten silicon: samples 4 (a), 1 (b).

3.2. Mechanical Properties


Table 2 summarizes mechanical and physical properties of the materials prepared by infiltration and sub
sequent treatment in the CTG field. Hardness of the ceramic sample 2 (approx. 35 GPa) exceeds that of sam
ples 3 and 4 (about 32 and 27 GPa, respectively). Hardness of ceramic samples definitely depends on the con
tent of the boron carbide hard phase.

Table 2. Physical and mechanical properties of ceramics produced by infiltration and CTG treatment
Density, Phase composition (according to XRD data), vol % Splitting tensile
Sample Hardness, GPa*
g/cm3 B12(SiBC)3 B4C βSiC α-SiC Si strength, MPa
1 2.51 45 21 5 2 27 14.1(±1.6) 58 ± 7
2 2.54 NA 85.5 6 7 1.5 35.4(±1.4) 114 ± 14
3 2.66 NA 74 11 14 1 31.8(±1.9) 138 ± 15
4 2.74 NA 62 24 19 5 26.7(±1.2) 169 ± 17
* The measurements were performed at a load of 196 N.

The data in Table 2 clearly show a decrease of hardness of the ceramics with reducing boron carbide con
tent, as was expected. Hardness of sample 2 correlates with that of the boron carbide ceramics produced by
HP [1, 2, 9] or SPS [7, 8, 22, 23]. Hardness of asinfiltrated sample 1 is about 14 GPa at a load of 196 N. A
considerable amount of residual silicon lowers dramatically hardness of the ceramic materials. Unlike hard
ness, which depends mostly on the content of boron carbide hard phase, the strength of these materials exhib
its a different tendency. Specifically, the splitting tensile strength of the ceramic samples treated in the CTG
field (sample 2) is twice as high as that of sample 1. In this case, an improvement in the strength is attributed
to a reduction of the residual silicon content from 27 to 1.5%. The use of a mixture of different fractions of
boron carbide powders (see Table 1) leads to a change in the phase composition and structure of ceramics and
has a beneficial effect on the strength which is as high as 138 ± 15 MPa. An increase of the speed of sample
movement in the CTG field 5 to 10 mm/min makes the structure finer and gives rise to platelike βSiC par
ticles (see Fig. 2d). It is well known that the strength of a polycrystalline material can be increased by decreas
ing the grain size—a socalled Hall–Petch effect [24]. Also, the formation of the αSiC phase in the course
of infiltration and the growth of edged grains in the CTG field provide strengthening of the ceramic materials.
As a result, the splitting tensile strength of sample 4 has been raised to 169 ± 7 MPa.
Figure 4 shows typical fracture surfaces upon strength testing. For all the samples tested, we observed a
combination of intercrystalline and transcrystalline fracture mechanisms.
The fracture surface of sample 1 that has the lowest splitting tensile strength (Table 2) suggests that the
residual silicon phase undergoes fracture through the intercrystalline mechanism (see Fig. 4a). Note that
boron carbide grains have large fracture surfaces due to the transcrystalline fracture mechanism. This is
explained by the formation of regions in the structure, which are made up of several boron carbide grains that
are not separated from each other by residual silicon. The treatment in the CTG field (see Fig. 4b) provides
the formation of a strong framework comprised of fine boron carbide particles (see Fig. 2b) that bring into

JOURNAL OF SUPERHARD MATERIALS Vol. 40 No. 4 2018


LOWTEMPERATURE SYNTHESIS 241

action the crack deflection mechanism (indicated by a black arrow). The fracture mechanism in sample 4 (see
Fig. 4d) is close to be an intercrystalline one due to the refined structure of the ceramics upon the treatment
in the CTG field.

200 μm 100 μm
(a) (b)

200 μm 50 μm
(c) (d)
Fig. 4. Microstructure of the fracture surface of the B4Cbased ceramics prepared by infiltration with molten silicon: samples
1 (a), 2 (b), 3 (c), 4 (d).

4. CONCLUSIONS
The infiltration of boron carbide with molten silicon and subsequent treatment in the CTG field have made
it possible to prepare fully dense ceramic materials. The application of this methodology has resulted in the
formation of a strong framework made up of fine B4C particles arising owing to recrystallization through the
molten silicon. Also, this reduces the residual silicon content below 5%.
The influence of the velocity of the CTG field movement has been clarified, and it has been demonstrated
that an increase of the velosity from 5 to 10 mm/min results in an alteration of the phase composition and grain
size. The use of CTG with a velocity of 10 mm/min makes the ternary phase B12(SiBC)3 disappear and leads
to the formation of the starting B4C phase with finer grains.
The prepared B4Cbased ceramics has a hardness of 26 to 35 GPa and a splitting tensile strength of 110 to
170 MPa.
A conceptually new method of using CTG procedure offers further opportunities for the lowtemperature
synthesis of hightemperature ceramic materials. The present findings are not limited to materials based on
boron carbide but are expected to be applicable to other covalent compounds.

REFERENCES
1. Suri, A.K., Subramanian, C., Sonber, J.K., and Murthy, T.S.R., Synthesis and consolidation of boron carbide: a
review, Int. Mater. Rev., 2010, vol. 55, pp. 4–40.
2. Thevenot, F., Boron carbide—a comprehensive review, J. Eur. Ceram. Soc., 1990, vol. 6, pp. 205–225.

JOURNAL OF SUPERHARD MATERIALS Vol. 40 No. 4 2018


242 SOLODKYI et al.

3. Walley, S.M., Historical review on high strain rate and shock properties of ceramics relevant to their application in
armor, Adv. Appl. Ceram., 2010, vol. 9, pp. 446–466.
4. Lee, H., Speyer, R.F., and Hackenberger, W.S., Sintering of boron carbide heattreated with hydrogen, J. Am. Ceram.
Soc., 2002, vol. 85, no. 8, pp. 2131–2133.
5. Chen, M.W., McCauley, J.W., LaSalvia, J.C., and Hemker, K.J., Microstructural characterization of commercial
hotpressed boron carbide ceramics, J. Am. Ceram. Soc., 2002, vol. 88, no. 7, pp. 1935–1942.
6. Cho, N., Bao, Z., and Speyer, R.F., Density and hardness optimized pressureless sintered and posthot isostatic
pressed B4C, J. Mater. Res., 2005, vol. 20, no. 8, pp. 2110–2116.
7. Vasylkiv, O., Demirskyi, D., Badica, P., Nishimura, T., Tok, A.I.Y., Sakka, Y., and Borodianska, H., Room and high
temperature flexural failure of spark plasma sintered boron carbide, Ceram. Int., 2016, vol. 42, pp. 7001–7013.
8. Badica, P., Borodianska, H., Xie, S., Zhao, T., Demirskyi, D., Li, P.F., Tok, A.I.Y., Sakka, Y., and Vasylkiv, O.,
Toughness control of boron carbide obtained by spark plasma sintering in nitrogen atmosphere, Ceram. Int., 2014,
vol. 40, pp. 3053–3061.
9. Kisly, P.S., Kuzenkova, M.A., Bodnaruk, N.I., and Grabchuk, B.L., Karbid bora (Boron Carbide), Kiev: Nauk.
Dumka, 1988.
10. Taylor, K.M. and Palicka, R.J., US Patent 3 765 300, 1973.
11. Hayun, S., Frage, N., and Dariel, M.P., The morphology of ceramic phases in BxC–SiC–Si infiltrated composites,
J. Solid State Chem., 2006, vol. 179, pp. 2875–2879.
12. Hayun, S., Frage, N., Dilman, H., Tourbabin, V., and Dariel, M.P., Synthesis of dense B4C–SiC–TiB2 composites,
in Medvedovski, E. (Ed.), Ceramic Armor and Armor Systems II, Baltimore, MD, USA: American Ceramic Society,
2006, pp. 37–44.
13. Han, I.S., Lee, K.S., Seo, D.W., and Woo, S.K., A comparative study on SiC–B4C–Si cermet prepared by pressure
less sintering and spark plasma sintering methods, J. Mater. Sci. Lett., 2002, vol. 21, pp. 703–706.
14. Pfann, W.G. Zone Melting, New York: John Wiley & Sons, 1966.
15. Bogomol, I., Nishimura, T., Vasylkiv, O., Sakka, Y., and Loboda, P., Microstructure and hightemperature strength
of B4C–TiB2 composite prepared by a crucibleless zone melting method, J. Alloy. Comp., 2009, vol. 485, pp. 677–
681.
16. Bogomol, I., Badica, P., Shen, Y.Q., Nishimura, T., Loboda, P., and Vasylkiv, O., Room and high temperature tough
ening in directionally solidified B4C–TiB2 eutectic composites by Si doping, J. Alloy. Comp., 2013, vol. 570, pp. 94–
99.
17. Bogomol, I., Nishimura, T., Nesterenko, Yu., Vasylkiv, O., Sakka, Y., and Loboda, P., The bending strength temper
ature dependence of the directionally solidified eutectic LaB6–ZrB2 composite, J. Alloy. Comp., 2011, vol. 509,
pp. 6123–6129.
18. Hayun, S., Weizmann, A., Dariel, M.P., and Frage, N., Microstructural evolution during the infiltration of boron car
bide with molten silicon, J. Eur. Ceram. Soc., 2010, vol. 30, pp. 1007–1014.
19. Morosin, B., Aselage, T.L., and Feigelson, R.S., Crystal structure refinements of rhombohedral symmetry materials
containing boronrich icosahedra, Mater. Res. Symp. Proc., 1987, vol. 97, pp. 145–149.
20. Ashbrook, R.L., Directionally solidified ceramic eutectics, J. Am. Ceram. Soc., 1977, vol. 60, pp. 428–435.
21. Chalmers, B., Principles of Solidification, New York: Wiley & Sons, 1964.
22. Hayun, S., Paris, V., Dariel, M.P., Frage, N., and Zaretzky, E., The morphology of ceramic phases in BxC–SiC–Si
infiltrated composites, J. Eur. Ceram. Soc., 2009, vol. 29, pp. 3395–3400.
23. Xu, C., Cai, Y., Flodström, K., Li, Z., Esmaeilzadeh, S., and Zhang, G.J., Spark plasma sintering of B4C ceramics:
The effects of milling medium and TiB2 addition, Int. J. Refract. Met. Hard Mater., 2012, vol. 30, pp. 139–144.
24. Wang, N., Wang, Z., Aust, K.T., and Erb, U., Effect of grain size on mechanical properties of nanocrystalline mate
rials, Acta Metall. Mater., 1995, vol. 43, pp. 519–528.

Translated by Yu. Kravchenko

JOURNAL OF SUPERHARD MATERIALS Vol. 40 No. 4 2018

You might also like